example_library.bib 1.6 MB

12345678910111213141516171819202122232425262728293031323334353637383940414243444546474849505152535455565758596061626364656667686970717273747576777879808182838485868788899091929394959697989910010110210310410510610710810911011111211311411511611711811912012112212312412512612712812913013113213313413513613713813914014114214314414514614714814915015115215315415515615715815916016116216316416516616716816917017117217317417517617717817918018118218318418518618718818919019119219319419519619719819920020120220320420520620720820921021121221321421521621721821922022122222322422522622722822923023123223323423523623723823924024124224324424524624724824925025125225325425525625725825926026126226326426526626726826927027127227327427527627727827928028128228328428528628728828929029129229329429529629729829930030130230330430530630730830931031131231331431531631731831932032132232332432532632732832933033133233333433533633733833934034134234334434534634734834935035135235335435535635735835936036136236336436536636736836937037137237337437537637737837938038138238338438538638738838939039139239339439539639739839940040140240340440540640740840941041141241341441541641741841942042142242342442542642742842943043143243343443543643743843944044144244344444544644744844945045145245345445545645745845946046146246346446546646746846947047147247347447547647747847948048148248348448548648748848949049149249349449549649749849950050150250350450550650750850951051151251351451551651751851952052152252352452552652752852953053153253353453553653753853954054154254354454554654754854955055155255355455555655755855956056156256356456556656756856957057157257357457557657757857958058158258358458558658758858959059159259359459559659759859960060160260360460560660760860961061161261361461561661761861962062162262362462562662762862963063163263363463563663763863964064164264364464564664764864965065165265365465565665765865966066166266366466566666766866967067167267367467567667767867968068168268368468568668768868969069169269369469569669769869970070170270370470570670770870971071171271371471571671771871972072172272372472572672772872973073173273373473573673773873974074174274374474574674774874975075175275375475575675775875976076176276376476576676776876977077177277377477577677777877978078178278378478578678778878979079179279379479579679779879980080180280380480580680780880981081181281381481581681781881982082182282382482582682782882983083183283383483583683783883984084184284384484584684784884985085185285385485585685785885986086186286386486586686786886987087187287387487587687787887988088188288388488588688788888989089189289389489589689789889990090190290390490590690790890991091191291391491591691791891992092192292392492592692792892993093193293393493593693793893994094194294394494594694794894995095195295395495595695795895996096196296396496596696796896997097197297397497597697797897998098198298398498598698798898999099199299399499599699799899910001001100210031004100510061007100810091010101110121013101410151016101710181019102010211022102310241025102610271028102910301031103210331034103510361037103810391040104110421043104410451046104710481049105010511052105310541055105610571058105910601061106210631064106510661067106810691070107110721073107410751076107710781079108010811082108310841085108610871088108910901091109210931094109510961097109810991100110111021103110411051106110711081109111011111112111311141115111611171118111911201121112211231124112511261127112811291130113111321133113411351136113711381139114011411142114311441145114611471148114911501151115211531154115511561157115811591160116111621163116411651166116711681169117011711172117311741175117611771178117911801181118211831184118511861187118811891190119111921193119411951196119711981199120012011202120312041205120612071208120912101211121212131214121512161217121812191220122112221223122412251226122712281229123012311232123312341235123612371238123912401241124212431244124512461247124812491250125112521253125412551256125712581259126012611262126312641265126612671268126912701271127212731274127512761277127812791280128112821283128412851286128712881289129012911292129312941295129612971298129913001301130213031304130513061307130813091310131113121313131413151316131713181319132013211322132313241325132613271328132913301331133213331334133513361337133813391340134113421343134413451346134713481349135013511352135313541355135613571358135913601361136213631364136513661367136813691370137113721373137413751376137713781379138013811382138313841385138613871388138913901391139213931394139513961397139813991400140114021403140414051406140714081409141014111412141314141415141614171418141914201421142214231424142514261427142814291430143114321433143414351436143714381439144014411442144314441445144614471448144914501451145214531454145514561457145814591460146114621463146414651466146714681469147014711472147314741475147614771478147914801481148214831484148514861487148814891490149114921493149414951496149714981499150015011502150315041505150615071508150915101511151215131514151515161517151815191520152115221523152415251526152715281529153015311532153315341535153615371538153915401541154215431544154515461547154815491550155115521553155415551556155715581559156015611562156315641565156615671568156915701571157215731574157515761577157815791580158115821583158415851586158715881589159015911592159315941595159615971598159916001601160216031604160516061607160816091610161116121613161416151616161716181619162016211622162316241625162616271628162916301631163216331634163516361637163816391640164116421643164416451646164716481649165016511652165316541655165616571658165916601661166216631664166516661667166816691670167116721673167416751676167716781679168016811682168316841685168616871688168916901691169216931694169516961697169816991700170117021703170417051706170717081709171017111712171317141715171617171718171917201721172217231724172517261727172817291730173117321733173417351736173717381739174017411742174317441745174617471748174917501751175217531754175517561757175817591760176117621763176417651766176717681769177017711772177317741775177617771778177917801781178217831784178517861787178817891790179117921793179417951796179717981799180018011802180318041805180618071808180918101811181218131814181518161817181818191820182118221823182418251826182718281829183018311832183318341835183618371838183918401841184218431844184518461847184818491850185118521853185418551856185718581859186018611862186318641865186618671868186918701871187218731874187518761877187818791880188118821883188418851886188718881889189018911892189318941895189618971898189919001901190219031904190519061907190819091910191119121913191419151916191719181919192019211922192319241925192619271928192919301931193219331934193519361937193819391940194119421943194419451946194719481949195019511952195319541955195619571958195919601961196219631964196519661967196819691970197119721973197419751976197719781979198019811982198319841985198619871988198919901991199219931994199519961997199819992000200120022003200420052006200720082009201020112012201320142015201620172018201920202021202220232024202520262027202820292030203120322033203420352036203720382039204020412042204320442045204620472048204920502051205220532054205520562057205820592060206120622063206420652066206720682069207020712072207320742075207620772078207920802081208220832084208520862087208820892090209120922093209420952096209720982099210021012102210321042105210621072108210921102111211221132114211521162117211821192120212121222123212421252126212721282129213021312132213321342135213621372138213921402141214221432144214521462147214821492150215121522153215421552156215721582159216021612162216321642165216621672168216921702171217221732174217521762177217821792180218121822183218421852186218721882189219021912192219321942195219621972198219922002201220222032204220522062207220822092210221122122213221422152216221722182219222022212222222322242225222622272228222922302231223222332234223522362237223822392240224122422243224422452246224722482249225022512252225322542255225622572258225922602261226222632264226522662267226822692270227122722273227422752276227722782279228022812282228322842285228622872288228922902291229222932294229522962297229822992300230123022303230423052306230723082309231023112312231323142315231623172318231923202321232223232324232523262327232823292330233123322333233423352336233723382339234023412342234323442345234623472348234923502351235223532354235523562357235823592360236123622363236423652366236723682369237023712372237323742375237623772378237923802381238223832384238523862387238823892390239123922393239423952396239723982399240024012402240324042405240624072408240924102411241224132414241524162417241824192420242124222423242424252426242724282429243024312432243324342435243624372438243924402441244224432444244524462447244824492450245124522453245424552456245724582459246024612462246324642465246624672468246924702471247224732474247524762477247824792480248124822483248424852486248724882489249024912492249324942495249624972498249925002501250225032504250525062507250825092510251125122513251425152516251725182519252025212522252325242525252625272528252925302531253225332534253525362537253825392540254125422543254425452546254725482549255025512552255325542555255625572558255925602561256225632564256525662567256825692570257125722573257425752576257725782579258025812582258325842585258625872588258925902591259225932594259525962597259825992600260126022603260426052606260726082609261026112612261326142615261626172618261926202621262226232624262526262627262826292630263126322633263426352636263726382639264026412642264326442645264626472648264926502651265226532654265526562657265826592660266126622663266426652666266726682669267026712672267326742675267626772678267926802681268226832684268526862687268826892690269126922693269426952696269726982699270027012702270327042705270627072708270927102711271227132714271527162717271827192720272127222723272427252726272727282729273027312732273327342735273627372738273927402741274227432744274527462747274827492750275127522753275427552756275727582759276027612762276327642765276627672768276927702771277227732774277527762777277827792780278127822783278427852786278727882789279027912792279327942795279627972798279928002801280228032804280528062807280828092810281128122813281428152816281728182819282028212822282328242825282628272828282928302831283228332834283528362837283828392840284128422843284428452846284728482849285028512852285328542855285628572858285928602861286228632864286528662867286828692870287128722873287428752876287728782879288028812882288328842885288628872888288928902891289228932894289528962897289828992900290129022903290429052906290729082909291029112912291329142915291629172918291929202921292229232924292529262927292829292930293129322933293429352936293729382939294029412942294329442945294629472948294929502951295229532954295529562957295829592960296129622963296429652966296729682969297029712972297329742975297629772978297929802981298229832984298529862987298829892990299129922993299429952996299729982999300030013002300330043005300630073008300930103011301230133014301530163017301830193020302130223023302430253026302730283029303030313032303330343035303630373038303930403041304230433044304530463047304830493050305130523053305430553056305730583059306030613062306330643065306630673068306930703071307230733074307530763077307830793080308130823083308430853086308730883089309030913092309330943095309630973098309931003101310231033104310531063107310831093110311131123113311431153116311731183119312031213122312331243125312631273128312931303131313231333134313531363137313831393140314131423143314431453146314731483149315031513152315331543155315631573158315931603161316231633164316531663167316831693170317131723173317431753176317731783179318031813182318331843185318631873188318931903191319231933194319531963197319831993200320132023203320432053206320732083209321032113212321332143215321632173218321932203221322232233224322532263227322832293230323132323233323432353236323732383239324032413242324332443245324632473248324932503251325232533254325532563257325832593260326132623263326432653266326732683269327032713272327332743275327632773278327932803281328232833284328532863287328832893290329132923293329432953296329732983299330033013302330333043305330633073308330933103311331233133314331533163317331833193320332133223323332433253326332733283329333033313332333333343335333633373338333933403341334233433344334533463347334833493350335133523353335433553356335733583359336033613362336333643365336633673368336933703371337233733374337533763377337833793380338133823383338433853386338733883389339033913392339333943395339633973398339934003401340234033404340534063407340834093410341134123413341434153416341734183419342034213422342334243425342634273428342934303431343234333434343534363437343834393440344134423443344434453446344734483449345034513452345334543455345634573458345934603461346234633464346534663467346834693470347134723473347434753476347734783479348034813482348334843485348634873488348934903491349234933494349534963497349834993500350135023503350435053506350735083509351035113512351335143515351635173518351935203521352235233524352535263527352835293530353135323533353435353536353735383539354035413542354335443545354635473548354935503551355235533554355535563557355835593560356135623563356435653566356735683569357035713572357335743575357635773578357935803581358235833584358535863587358835893590359135923593359435953596359735983599360036013602360336043605360636073608360936103611361236133614361536163617361836193620362136223623362436253626362736283629363036313632363336343635363636373638363936403641364236433644364536463647364836493650365136523653365436553656365736583659366036613662366336643665366636673668366936703671367236733674367536763677367836793680368136823683368436853686368736883689369036913692369336943695369636973698369937003701370237033704370537063707370837093710371137123713371437153716371737183719372037213722372337243725372637273728372937303731373237333734373537363737373837393740374137423743374437453746374737483749375037513752375337543755375637573758375937603761376237633764376537663767376837693770377137723773377437753776377737783779378037813782378337843785378637873788378937903791379237933794379537963797379837993800380138023803380438053806380738083809381038113812381338143815381638173818381938203821382238233824382538263827382838293830383138323833383438353836383738383839384038413842384338443845384638473848384938503851385238533854385538563857385838593860386138623863386438653866386738683869387038713872387338743875387638773878387938803881388238833884388538863887388838893890389138923893389438953896389738983899390039013902390339043905390639073908390939103911391239133914391539163917391839193920392139223923392439253926392739283929393039313932393339343935393639373938393939403941394239433944394539463947394839493950395139523953395439553956395739583959396039613962396339643965396639673968396939703971397239733974397539763977397839793980398139823983398439853986398739883989399039913992399339943995399639973998399940004001400240034004400540064007400840094010401140124013401440154016401740184019402040214022402340244025402640274028402940304031403240334034403540364037403840394040404140424043404440454046404740484049405040514052405340544055405640574058405940604061406240634064406540664067406840694070407140724073407440754076407740784079408040814082408340844085408640874088408940904091409240934094409540964097409840994100410141024103410441054106410741084109411041114112411341144115411641174118411941204121412241234124412541264127412841294130413141324133413441354136413741384139414041414142414341444145414641474148414941504151415241534154415541564157415841594160416141624163416441654166416741684169417041714172417341744175417641774178417941804181418241834184418541864187418841894190419141924193419441954196419741984199420042014202420342044205420642074208420942104211421242134214421542164217421842194220422142224223422442254226422742284229423042314232423342344235423642374238423942404241424242434244424542464247424842494250425142524253425442554256425742584259426042614262426342644265426642674268426942704271427242734274427542764277427842794280428142824283428442854286428742884289429042914292429342944295429642974298429943004301430243034304430543064307430843094310431143124313431443154316431743184319432043214322432343244325432643274328432943304331433243334334433543364337433843394340434143424343434443454346434743484349435043514352435343544355435643574358435943604361436243634364436543664367436843694370437143724373437443754376437743784379438043814382438343844385438643874388438943904391439243934394439543964397439843994400440144024403440444054406440744084409441044114412441344144415441644174418441944204421442244234424442544264427442844294430443144324433443444354436443744384439444044414442444344444445444644474448444944504451445244534454445544564457445844594460446144624463446444654466446744684469447044714472447344744475447644774478447944804481448244834484448544864487448844894490449144924493449444954496449744984499450045014502450345044505450645074508450945104511451245134514451545164517451845194520452145224523452445254526452745284529453045314532453345344535453645374538453945404541454245434544454545464547454845494550455145524553455445554556455745584559456045614562456345644565456645674568456945704571457245734574457545764577457845794580458145824583458445854586458745884589459045914592459345944595459645974598459946004601460246034604460546064607460846094610461146124613461446154616461746184619462046214622462346244625462646274628462946304631463246334634463546364637463846394640464146424643464446454646464746484649465046514652465346544655465646574658465946604661466246634664466546664667466846694670467146724673467446754676467746784679468046814682468346844685468646874688468946904691469246934694469546964697469846994700470147024703470447054706470747084709471047114712471347144715471647174718471947204721472247234724472547264727472847294730473147324733473447354736473747384739474047414742474347444745474647474748474947504751475247534754475547564757475847594760476147624763476447654766476747684769477047714772477347744775477647774778477947804781478247834784478547864787478847894790479147924793479447954796479747984799480048014802480348044805480648074808480948104811481248134814481548164817481848194820482148224823482448254826482748284829483048314832483348344835483648374838483948404841484248434844484548464847484848494850485148524853485448554856485748584859486048614862486348644865486648674868486948704871487248734874487548764877487848794880488148824883488448854886488748884889489048914892489348944895489648974898489949004901490249034904490549064907490849094910491149124913491449154916491749184919492049214922492349244925492649274928492949304931493249334934493549364937493849394940494149424943494449454946494749484949495049514952495349544955495649574958495949604961496249634964496549664967496849694970497149724973497449754976497749784979498049814982498349844985498649874988498949904991499249934994499549964997499849995000500150025003500450055006500750085009501050115012501350145015501650175018501950205021502250235024502550265027502850295030503150325033503450355036503750385039504050415042504350445045504650475048504950505051505250535054505550565057505850595060506150625063506450655066506750685069507050715072507350745075507650775078507950805081508250835084508550865087508850895090509150925093509450955096509750985099510051015102510351045105510651075108510951105111511251135114511551165117511851195120512151225123512451255126512751285129513051315132513351345135513651375138513951405141514251435144514551465147514851495150515151525153515451555156515751585159516051615162516351645165516651675168516951705171517251735174517551765177517851795180518151825183518451855186518751885189519051915192519351945195519651975198519952005201520252035204520552065207520852095210521152125213521452155216521752185219522052215222522352245225522652275228522952305231523252335234523552365237523852395240524152425243524452455246524752485249525052515252525352545255525652575258525952605261526252635264526552665267526852695270527152725273527452755276527752785279528052815282528352845285528652875288528952905291529252935294529552965297529852995300530153025303530453055306530753085309531053115312531353145315531653175318531953205321532253235324532553265327532853295330533153325333533453355336533753385339534053415342534353445345534653475348534953505351535253535354535553565357535853595360536153625363536453655366536753685369537053715372537353745375537653775378537953805381538253835384538553865387538853895390539153925393539453955396539753985399540054015402540354045405540654075408540954105411541254135414541554165417541854195420542154225423542454255426542754285429543054315432543354345435543654375438543954405441544254435444544554465447544854495450545154525453545454555456545754585459546054615462546354645465546654675468546954705471547254735474547554765477547854795480548154825483548454855486548754885489549054915492549354945495549654975498549955005501550255035504550555065507550855095510551155125513551455155516551755185519552055215522552355245525552655275528552955305531553255335534553555365537553855395540554155425543554455455546554755485549555055515552555355545555555655575558555955605561556255635564556555665567556855695570557155725573557455755576557755785579558055815582558355845585558655875588558955905591559255935594559555965597559855995600560156025603560456055606560756085609561056115612561356145615561656175618561956205621562256235624562556265627562856295630563156325633563456355636563756385639564056415642564356445645564656475648564956505651565256535654565556565657565856595660566156625663566456655666566756685669567056715672567356745675567656775678567956805681568256835684568556865687568856895690569156925693569456955696569756985699570057015702570357045705570657075708570957105711571257135714571557165717571857195720572157225723572457255726572757285729573057315732573357345735573657375738573957405741574257435744574557465747574857495750575157525753575457555756575757585759576057615762576357645765576657675768576957705771577257735774577557765777577857795780578157825783578457855786578757885789579057915792579357945795579657975798579958005801580258035804580558065807580858095810581158125813581458155816581758185819582058215822582358245825582658275828582958305831583258335834583558365837583858395840584158425843584458455846584758485849585058515852585358545855585658575858585958605861586258635864586558665867586858695870587158725873587458755876587758785879588058815882588358845885588658875888588958905891589258935894589558965897589858995900590159025903590459055906590759085909591059115912591359145915591659175918591959205921592259235924592559265927592859295930593159325933593459355936593759385939594059415942594359445945594659475948594959505951595259535954595559565957595859595960596159625963596459655966596759685969597059715972597359745975597659775978597959805981598259835984598559865987598859895990599159925993599459955996599759985999600060016002600360046005600660076008600960106011601260136014601560166017601860196020602160226023602460256026602760286029603060316032603360346035603660376038603960406041604260436044604560466047604860496050605160526053605460556056605760586059606060616062606360646065606660676068606960706071607260736074607560766077607860796080608160826083608460856086608760886089609060916092609360946095609660976098609961006101610261036104610561066107610861096110611161126113611461156116611761186119612061216122612361246125612661276128612961306131613261336134613561366137613861396140614161426143614461456146614761486149615061516152615361546155615661576158615961606161616261636164616561666167616861696170617161726173617461756176617761786179618061816182618361846185618661876188618961906191619261936194619561966197619861996200620162026203620462056206620762086209621062116212621362146215621662176218621962206221622262236224622562266227622862296230623162326233623462356236623762386239624062416242624362446245624662476248624962506251625262536254625562566257625862596260626162626263626462656266626762686269627062716272627362746275627662776278627962806281628262836284628562866287628862896290629162926293629462956296629762986299630063016302630363046305630663076308630963106311631263136314631563166317631863196320632163226323632463256326632763286329633063316332633363346335633663376338633963406341634263436344634563466347634863496350635163526353635463556356635763586359636063616362636363646365636663676368636963706371637263736374637563766377637863796380638163826383638463856386638763886389639063916392639363946395639663976398639964006401640264036404640564066407640864096410641164126413641464156416641764186419642064216422642364246425642664276428642964306431643264336434643564366437643864396440644164426443644464456446644764486449645064516452645364546455645664576458645964606461646264636464646564666467646864696470647164726473647464756476647764786479648064816482648364846485648664876488648964906491649264936494649564966497649864996500650165026503650465056506650765086509651065116512651365146515651665176518651965206521652265236524652565266527652865296530653165326533653465356536653765386539654065416542654365446545654665476548654965506551655265536554655565566557655865596560656165626563656465656566656765686569657065716572657365746575657665776578657965806581658265836584658565866587658865896590659165926593659465956596659765986599660066016602660366046605660666076608660966106611661266136614661566166617661866196620662166226623662466256626662766286629663066316632663366346635663666376638663966406641664266436644664566466647664866496650665166526653665466556656665766586659666066616662666366646665666666676668666966706671667266736674667566766677667866796680668166826683668466856686668766886689669066916692669366946695669666976698669967006701670267036704670567066707670867096710671167126713671467156716671767186719672067216722672367246725672667276728672967306731673267336734673567366737673867396740674167426743674467456746674767486749675067516752675367546755675667576758675967606761676267636764676567666767676867696770677167726773677467756776677767786779678067816782678367846785678667876788678967906791679267936794679567966797679867996800680168026803680468056806680768086809681068116812681368146815681668176818681968206821682268236824682568266827682868296830683168326833683468356836683768386839684068416842684368446845684668476848684968506851685268536854685568566857685868596860686168626863686468656866686768686869687068716872687368746875687668776878687968806881688268836884688568866887688868896890689168926893689468956896689768986899690069016902690369046905690669076908690969106911691269136914691569166917691869196920692169226923692469256926692769286929693069316932693369346935693669376938693969406941694269436944694569466947694869496950695169526953695469556956695769586959696069616962696369646965696669676968696969706971697269736974697569766977697869796980698169826983698469856986698769886989699069916992699369946995699669976998699970007001700270037004700570067007700870097010701170127013701470157016701770187019702070217022702370247025702670277028702970307031703270337034703570367037703870397040704170427043704470457046704770487049705070517052705370547055705670577058705970607061706270637064706570667067706870697070707170727073707470757076707770787079708070817082708370847085708670877088708970907091709270937094709570967097709870997100710171027103710471057106710771087109711071117112711371147115711671177118711971207121712271237124712571267127712871297130713171327133713471357136713771387139714071417142714371447145714671477148714971507151715271537154715571567157715871597160716171627163716471657166716771687169717071717172717371747175717671777178717971807181718271837184718571867187718871897190719171927193719471957196719771987199720072017202720372047205720672077208720972107211721272137214721572167217721872197220722172227223722472257226722772287229723072317232723372347235723672377238723972407241724272437244724572467247724872497250725172527253725472557256725772587259726072617262726372647265726672677268726972707271727272737274727572767277727872797280728172827283728472857286728772887289729072917292729372947295729672977298729973007301730273037304730573067307730873097310731173127313731473157316731773187319732073217322732373247325732673277328732973307331733273337334733573367337733873397340734173427343734473457346734773487349735073517352735373547355735673577358735973607361736273637364736573667367736873697370737173727373737473757376737773787379738073817382738373847385738673877388738973907391739273937394739573967397739873997400740174027403740474057406740774087409741074117412741374147415741674177418741974207421742274237424742574267427742874297430743174327433743474357436743774387439744074417442744374447445744674477448744974507451745274537454745574567457745874597460746174627463746474657466746774687469747074717472747374747475747674777478747974807481748274837484748574867487748874897490749174927493749474957496749774987499750075017502750375047505750675077508750975107511751275137514751575167517751875197520752175227523752475257526752775287529753075317532753375347535753675377538753975407541754275437544754575467547754875497550755175527553755475557556755775587559756075617562756375647565756675677568756975707571757275737574757575767577757875797580758175827583758475857586758775887589759075917592759375947595759675977598759976007601760276037604760576067607760876097610761176127613761476157616761776187619762076217622762376247625762676277628762976307631763276337634763576367637763876397640764176427643764476457646764776487649765076517652765376547655765676577658765976607661766276637664766576667667766876697670767176727673767476757676767776787679768076817682768376847685768676877688768976907691769276937694769576967697769876997700770177027703770477057706770777087709771077117712771377147715771677177718771977207721772277237724772577267727772877297730773177327733773477357736773777387739774077417742774377447745774677477748774977507751775277537754775577567757775877597760776177627763776477657766776777687769777077717772777377747775777677777778777977807781778277837784778577867787778877897790779177927793779477957796779777987799780078017802780378047805780678077808780978107811781278137814781578167817781878197820782178227823782478257826782778287829783078317832783378347835783678377838783978407841784278437844784578467847784878497850785178527853785478557856785778587859786078617862786378647865786678677868786978707871787278737874787578767877787878797880788178827883788478857886788778887889789078917892789378947895789678977898789979007901790279037904790579067907790879097910791179127913791479157916791779187919792079217922792379247925792679277928792979307931793279337934793579367937793879397940794179427943794479457946794779487949795079517952795379547955795679577958795979607961796279637964796579667967796879697970797179727973797479757976797779787979798079817982798379847985798679877988798979907991799279937994799579967997799879998000800180028003800480058006800780088009801080118012801380148015801680178018801980208021802280238024802580268027802880298030803180328033803480358036803780388039804080418042804380448045804680478048804980508051805280538054805580568057805880598060806180628063806480658066806780688069807080718072807380748075807680778078807980808081808280838084808580868087808880898090809180928093809480958096809780988099810081018102810381048105810681078108810981108111811281138114811581168117811881198120812181228123812481258126812781288129813081318132813381348135813681378138813981408141814281438144814581468147814881498150815181528153815481558156815781588159816081618162816381648165816681678168816981708171817281738174817581768177817881798180818181828183818481858186818781888189819081918192819381948195819681978198819982008201820282038204820582068207820882098210821182128213821482158216821782188219822082218222822382248225822682278228822982308231823282338234823582368237823882398240824182428243824482458246824782488249825082518252825382548255825682578258825982608261826282638264826582668267826882698270827182728273827482758276827782788279828082818282828382848285828682878288828982908291829282938294829582968297829882998300830183028303830483058306830783088309831083118312831383148315831683178318831983208321832283238324832583268327832883298330833183328333833483358336833783388339834083418342834383448345834683478348834983508351835283538354835583568357835883598360836183628363836483658366836783688369837083718372837383748375837683778378837983808381838283838384838583868387838883898390839183928393839483958396839783988399840084018402840384048405840684078408840984108411841284138414841584168417841884198420842184228423842484258426842784288429843084318432843384348435843684378438843984408441844284438444844584468447844884498450845184528453845484558456845784588459846084618462846384648465846684678468846984708471847284738474847584768477847884798480848184828483848484858486848784888489849084918492849384948495849684978498849985008501850285038504850585068507850885098510851185128513851485158516851785188519852085218522852385248525852685278528852985308531853285338534853585368537853885398540854185428543854485458546854785488549855085518552855385548555855685578558855985608561856285638564856585668567856885698570857185728573857485758576857785788579858085818582858385848585858685878588858985908591859285938594859585968597859885998600860186028603860486058606860786088609861086118612861386148615861686178618861986208621862286238624862586268627862886298630863186328633863486358636863786388639864086418642864386448645864686478648864986508651865286538654865586568657865886598660866186628663866486658666866786688669867086718672867386748675867686778678867986808681868286838684868586868687868886898690869186928693869486958696869786988699870087018702870387048705870687078708870987108711871287138714871587168717871887198720872187228723872487258726872787288729873087318732873387348735873687378738873987408741874287438744874587468747874887498750875187528753875487558756875787588759876087618762876387648765876687678768876987708771877287738774877587768777877887798780878187828783878487858786878787888789879087918792879387948795879687978798879988008801880288038804880588068807880888098810881188128813881488158816881788188819882088218822882388248825882688278828882988308831883288338834883588368837883888398840884188428843884488458846884788488849885088518852885388548855885688578858885988608861886288638864886588668867886888698870887188728873887488758876887788788879888088818882888388848885888688878888888988908891889288938894889588968897889888998900890189028903890489058906890789088909891089118912891389148915891689178918891989208921892289238924892589268927892889298930893189328933893489358936893789388939894089418942894389448945894689478948894989508951895289538954895589568957895889598960896189628963896489658966896789688969897089718972897389748975897689778978897989808981898289838984898589868987898889898990899189928993899489958996899789988999900090019002900390049005900690079008900990109011901290139014901590169017901890199020902190229023902490259026902790289029903090319032903390349035903690379038903990409041904290439044904590469047904890499050905190529053905490559056905790589059906090619062906390649065906690679068906990709071907290739074907590769077907890799080908190829083908490859086908790889089909090919092909390949095909690979098909991009101910291039104910591069107910891099110911191129113911491159116911791189119912091219122912391249125912691279128912991309131913291339134913591369137913891399140914191429143914491459146914791489149915091519152915391549155915691579158915991609161916291639164916591669167916891699170917191729173917491759176917791789179918091819182918391849185918691879188918991909191919291939194919591969197919891999200920192029203920492059206920792089209921092119212921392149215921692179218921992209221922292239224922592269227922892299230923192329233923492359236923792389239924092419242924392449245924692479248924992509251925292539254925592569257925892599260926192629263926492659266926792689269927092719272927392749275927692779278927992809281928292839284928592869287928892899290929192929293929492959296929792989299930093019302930393049305930693079308930993109311931293139314931593169317931893199320932193229323932493259326932793289329933093319332933393349335933693379338933993409341934293439344934593469347934893499350935193529353935493559356935793589359936093619362936393649365936693679368936993709371937293739374937593769377937893799380938193829383938493859386938793889389939093919392939393949395939693979398939994009401940294039404940594069407940894099410941194129413941494159416941794189419942094219422942394249425942694279428942994309431943294339434943594369437943894399440944194429443944494459446944794489449945094519452945394549455945694579458945994609461946294639464946594669467946894699470947194729473947494759476947794789479948094819482948394849485948694879488948994909491949294939494949594969497949894999500950195029503950495059506950795089509951095119512951395149515951695179518951995209521952295239524952595269527952895299530953195329533953495359536953795389539954095419542954395449545954695479548954995509551955295539554955595569557955895599560956195629563956495659566956795689569957095719572957395749575957695779578957995809581958295839584958595869587958895899590959195929593959495959596959795989599960096019602960396049605960696079608960996109611961296139614961596169617961896199620962196229623962496259626962796289629963096319632963396349635963696379638963996409641964296439644964596469647964896499650965196529653965496559656965796589659966096619662966396649665966696679668966996709671967296739674967596769677967896799680968196829683968496859686968796889689969096919692969396949695969696979698969997009701970297039704970597069707970897099710971197129713971497159716971797189719972097219722972397249725972697279728972997309731973297339734973597369737973897399740974197429743974497459746974797489749975097519752975397549755975697579758975997609761976297639764976597669767976897699770977197729773977497759776977797789779978097819782978397849785978697879788978997909791979297939794979597969797979897999800980198029803980498059806980798089809981098119812981398149815981698179818981998209821982298239824982598269827982898299830983198329833983498359836983798389839984098419842984398449845984698479848984998509851985298539854985598569857985898599860986198629863986498659866986798689869987098719872987398749875987698779878987998809881988298839884988598869887988898899890989198929893989498959896989798989899990099019902990399049905990699079908990999109911991299139914991599169917991899199920992199229923992499259926992799289929993099319932993399349935993699379938993999409941994299439944994599469947994899499950995199529953995499559956995799589959996099619962996399649965996699679968996999709971997299739974997599769977997899799980998199829983998499859986998799889989999099919992999399949995999699979998999910000100011000210003100041000510006100071000810009100101001110012100131001410015100161001710018100191002010021100221002310024100251002610027100281002910030100311003210033100341003510036100371003810039100401004110042100431004410045100461004710048100491005010051100521005310054100551005610057100581005910060100611006210063100641006510066100671006810069100701007110072100731007410075100761007710078100791008010081100821008310084100851008610087100881008910090100911009210093100941009510096100971009810099101001010110102101031010410105101061010710108101091011010111101121011310114101151011610117101181011910120101211012210123101241012510126101271012810129101301013110132101331013410135101361013710138101391014010141101421014310144101451014610147101481014910150101511015210153101541015510156101571015810159101601016110162101631016410165101661016710168101691017010171101721017310174101751017610177101781017910180101811018210183101841018510186101871018810189101901019110192101931019410195101961019710198101991020010201102021020310204102051020610207102081020910210102111021210213102141021510216102171021810219102201022110222102231022410225102261022710228102291023010231102321023310234102351023610237102381023910240102411024210243102441024510246102471024810249102501025110252102531025410255102561025710258102591026010261102621026310264102651026610267102681026910270102711027210273102741027510276102771027810279102801028110282102831028410285102861028710288102891029010291102921029310294102951029610297102981029910300103011030210303103041030510306103071030810309103101031110312103131031410315103161031710318103191032010321103221032310324103251032610327103281032910330103311033210333103341033510336103371033810339103401034110342103431034410345103461034710348103491035010351103521035310354103551035610357103581035910360103611036210363103641036510366103671036810369103701037110372103731037410375103761037710378103791038010381103821038310384103851038610387103881038910390103911039210393103941039510396103971039810399104001040110402104031040410405104061040710408104091041010411104121041310414104151041610417104181041910420104211042210423104241042510426104271042810429104301043110432104331043410435104361043710438104391044010441104421044310444104451044610447104481044910450104511045210453104541045510456104571045810459104601046110462104631046410465104661046710468104691047010471104721047310474104751047610477104781047910480104811048210483104841048510486104871048810489104901049110492104931049410495104961049710498104991050010501105021050310504105051050610507105081050910510105111051210513105141051510516105171051810519105201052110522105231052410525105261052710528105291053010531105321053310534105351053610537105381053910540105411054210543105441054510546105471054810549105501055110552105531055410555105561055710558105591056010561105621056310564105651056610567105681056910570105711057210573105741057510576105771057810579105801058110582105831058410585105861058710588105891059010591105921059310594105951059610597105981059910600106011060210603106041060510606106071060810609106101061110612106131061410615106161061710618106191062010621106221062310624106251062610627106281062910630106311063210633106341063510636106371063810639106401064110642106431064410645106461064710648106491065010651106521065310654106551065610657106581065910660106611066210663106641066510666106671066810669106701067110672106731067410675106761067710678106791068010681106821068310684106851068610687106881068910690106911069210693106941069510696106971069810699107001070110702107031070410705107061070710708107091071010711107121071310714107151071610717107181071910720107211072210723107241072510726107271072810729107301073110732107331073410735107361073710738107391074010741107421074310744107451074610747107481074910750107511075210753107541075510756107571075810759107601076110762107631076410765107661076710768107691077010771107721077310774107751077610777107781077910780107811078210783107841078510786107871078810789107901079110792107931079410795107961079710798107991080010801108021080310804108051080610807108081080910810108111081210813108141081510816108171081810819108201082110822108231082410825108261082710828108291083010831108321083310834108351083610837108381083910840108411084210843108441084510846108471084810849108501085110852108531085410855108561085710858108591086010861108621086310864108651086610867108681086910870108711087210873108741087510876108771087810879108801088110882108831088410885108861088710888108891089010891108921089310894108951089610897108981089910900109011090210903109041090510906109071090810909109101091110912109131091410915109161091710918109191092010921109221092310924109251092610927109281092910930109311093210933109341093510936109371093810939109401094110942109431094410945109461094710948109491095010951109521095310954109551095610957109581095910960109611096210963109641096510966109671096810969109701097110972109731097410975109761097710978109791098010981109821098310984109851098610987109881098910990109911099210993109941099510996109971099810999110001100111002110031100411005110061100711008110091101011011110121101311014110151101611017110181101911020110211102211023110241102511026110271102811029110301103111032110331103411035110361103711038110391104011041110421104311044110451104611047110481104911050110511105211053110541105511056110571105811059110601106111062110631106411065110661106711068110691107011071110721107311074110751107611077110781107911080110811108211083110841108511086110871108811089110901109111092110931109411095110961109711098110991110011101111021110311104111051110611107111081110911110111111111211113111141111511116111171111811119111201112111122111231112411125111261112711128111291113011131111321113311134111351113611137111381113911140111411114211143111441114511146111471114811149111501115111152111531115411155111561115711158111591116011161111621116311164111651116611167111681116911170111711117211173111741117511176111771117811179111801118111182111831118411185111861118711188111891119011191111921119311194111951119611197111981119911200112011120211203112041120511206112071120811209112101121111212112131121411215112161121711218112191122011221112221122311224112251122611227112281122911230112311123211233112341123511236112371123811239112401124111242112431124411245112461124711248112491125011251112521125311254112551125611257112581125911260112611126211263112641126511266112671126811269112701127111272112731127411275112761127711278112791128011281112821128311284112851128611287112881128911290112911129211293112941129511296112971129811299113001130111302113031130411305113061130711308113091131011311113121131311314113151131611317113181131911320113211132211323113241132511326113271132811329113301133111332113331133411335113361133711338113391134011341113421134311344113451134611347113481134911350113511135211353113541135511356113571135811359113601136111362113631136411365113661136711368113691137011371113721137311374113751137611377113781137911380113811138211383113841138511386113871138811389113901139111392113931139411395113961139711398113991140011401114021140311404114051140611407114081140911410114111141211413114141141511416114171141811419114201142111422114231142411425114261142711428114291143011431114321143311434114351143611437114381143911440114411144211443114441144511446114471144811449114501145111452114531145411455114561145711458114591146011461114621146311464114651146611467114681146911470114711147211473114741147511476114771147811479114801148111482114831148411485114861148711488114891149011491114921149311494114951149611497114981149911500115011150211503115041150511506115071150811509115101151111512115131151411515115161151711518115191152011521115221152311524115251152611527115281152911530115311153211533115341153511536115371153811539115401154111542115431154411545115461154711548115491155011551115521155311554115551155611557115581155911560115611156211563115641156511566115671156811569115701157111572115731157411575115761157711578115791158011581115821158311584115851158611587115881158911590115911159211593115941159511596115971159811599116001160111602116031160411605116061160711608116091161011611116121161311614116151161611617116181161911620116211162211623116241162511626116271162811629116301163111632116331163411635116361163711638116391164011641116421164311644116451164611647116481164911650116511165211653116541165511656116571165811659116601166111662116631166411665116661166711668116691167011671116721167311674116751167611677116781167911680116811168211683116841168511686116871168811689116901169111692116931169411695116961169711698116991170011701117021170311704117051170611707117081170911710117111171211713117141171511716117171171811719117201172111722117231172411725117261172711728117291173011731117321173311734117351173611737117381173911740117411174211743117441174511746117471174811749117501175111752117531175411755117561175711758117591176011761117621176311764117651176611767117681176911770117711177211773117741177511776117771177811779117801178111782117831178411785117861178711788117891179011791117921179311794117951179611797117981179911800118011180211803118041180511806118071180811809118101181111812118131181411815118161181711818118191182011821118221182311824118251182611827118281182911830118311183211833118341183511836118371183811839118401184111842118431184411845118461184711848118491185011851118521185311854118551185611857118581185911860118611186211863118641186511866118671186811869118701187111872118731187411875118761187711878118791188011881118821188311884118851188611887118881188911890118911189211893118941189511896118971189811899119001190111902119031190411905119061190711908119091191011911119121191311914119151191611917119181191911920119211192211923119241192511926119271192811929119301193111932119331193411935119361193711938119391194011941119421194311944119451194611947119481194911950119511195211953119541195511956119571195811959119601196111962119631196411965119661196711968119691197011971119721197311974119751197611977119781197911980119811198211983119841198511986119871198811989119901199111992119931199411995119961199711998119991200012001120021200312004120051200612007120081200912010120111201212013120141201512016120171201812019120201202112022120231202412025120261202712028120291203012031120321203312034120351203612037120381203912040120411204212043120441204512046120471204812049120501205112052120531205412055120561205712058120591206012061120621206312064120651206612067120681206912070120711207212073120741207512076120771207812079120801208112082120831208412085120861208712088120891209012091120921209312094120951209612097120981209912100121011210212103121041210512106121071210812109121101211112112121131211412115121161211712118121191212012121121221212312124121251212612127121281212912130121311213212133121341213512136121371213812139121401214112142121431214412145121461214712148121491215012151121521215312154121551215612157121581215912160121611216212163121641216512166121671216812169121701217112172121731217412175121761217712178121791218012181121821218312184121851218612187121881218912190121911219212193121941219512196121971219812199122001220112202122031220412205122061220712208122091221012211122121221312214122151221612217122181221912220122211222212223122241222512226122271222812229122301223112232122331223412235122361223712238122391224012241122421224312244122451224612247122481224912250122511225212253122541225512256122571225812259122601226112262122631226412265122661226712268122691227012271122721227312274122751227612277122781227912280122811228212283122841228512286122871228812289122901229112292122931229412295122961229712298122991230012301123021230312304123051230612307123081230912310123111231212313123141231512316123171231812319123201232112322123231232412325123261232712328123291233012331123321233312334123351233612337123381233912340123411234212343123441234512346123471234812349123501235112352123531235412355123561235712358123591236012361123621236312364123651236612367123681236912370123711237212373123741237512376123771237812379123801238112382123831238412385123861238712388123891239012391123921239312394123951239612397123981239912400124011240212403124041240512406124071240812409124101241112412124131241412415124161241712418124191242012421124221242312424124251242612427124281242912430124311243212433124341243512436124371243812439124401244112442124431244412445124461244712448124491245012451124521245312454124551245612457124581245912460124611246212463124641246512466124671246812469124701247112472124731247412475124761247712478124791248012481124821248312484124851248612487124881248912490124911249212493124941249512496124971249812499125001250112502125031250412505125061250712508125091251012511125121251312514125151251612517125181251912520125211252212523125241252512526125271252812529125301253112532125331253412535125361253712538125391254012541125421254312544125451254612547125481254912550125511255212553125541255512556125571255812559125601256112562125631256412565125661256712568125691257012571125721257312574125751257612577125781257912580125811258212583125841258512586125871258812589125901259112592125931259412595125961259712598125991260012601126021260312604126051260612607126081260912610126111261212613126141261512616126171261812619126201262112622126231262412625126261262712628126291263012631126321263312634126351263612637126381263912640126411264212643126441264512646126471264812649126501265112652126531265412655126561265712658126591266012661126621266312664126651266612667126681266912670126711267212673126741267512676126771267812679126801268112682126831268412685126861268712688126891269012691126921269312694126951269612697126981269912700127011270212703127041270512706127071270812709127101271112712127131271412715127161271712718127191272012721127221272312724127251272612727127281272912730127311273212733127341273512736127371273812739127401274112742127431274412745127461274712748127491275012751127521275312754127551275612757127581275912760127611276212763127641276512766127671276812769127701277112772127731277412775127761277712778127791278012781127821278312784127851278612787127881278912790127911279212793127941279512796127971279812799128001280112802128031280412805128061280712808128091281012811128121281312814128151281612817128181281912820128211282212823128241282512826128271282812829128301283112832128331283412835128361283712838128391284012841128421284312844128451284612847128481284912850128511285212853128541285512856128571285812859128601286112862128631286412865128661286712868128691287012871128721287312874128751287612877128781287912880128811288212883128841288512886128871288812889128901289112892128931289412895128961289712898128991290012901129021290312904129051290612907129081290912910129111291212913129141291512916129171291812919129201292112922129231292412925129261292712928129291293012931129321293312934129351293612937129381293912940129411294212943129441294512946129471294812949129501295112952129531295412955129561295712958129591296012961129621296312964129651296612967129681296912970129711297212973129741297512976129771297812979129801298112982129831298412985129861298712988129891299012991129921299312994129951299612997129981299913000130011300213003130041300513006130071300813009130101301113012130131301413015130161301713018130191302013021130221302313024130251302613027130281302913030130311303213033130341303513036130371303813039130401304113042130431304413045130461304713048130491305013051130521305313054130551305613057130581305913060130611306213063130641306513066130671306813069130701307113072130731307413075130761307713078130791308013081130821308313084130851308613087130881308913090130911309213093130941309513096130971309813099131001310113102131031310413105131061310713108131091311013111131121311313114131151311613117131181311913120131211312213123131241312513126131271312813129131301313113132131331313413135131361313713138131391314013141131421314313144131451314613147131481314913150131511315213153131541315513156131571315813159131601316113162131631316413165131661316713168131691317013171131721317313174131751317613177131781317913180131811318213183131841318513186131871318813189131901319113192131931319413195131961319713198131991320013201132021320313204132051320613207132081320913210132111321213213132141321513216132171321813219132201322113222132231322413225132261322713228132291323013231132321323313234132351323613237132381323913240132411324213243132441324513246132471324813249132501325113252132531325413255132561325713258132591326013261132621326313264132651326613267132681326913270132711327213273132741327513276132771327813279132801328113282132831328413285132861328713288132891329013291132921329313294132951329613297132981329913300133011330213303133041330513306133071330813309133101331113312133131331413315133161331713318133191332013321133221332313324133251332613327133281332913330133311333213333133341333513336133371333813339133401334113342133431334413345133461334713348133491335013351133521335313354133551335613357133581335913360133611336213363133641336513366133671336813369133701337113372133731337413375133761337713378133791338013381133821338313384133851338613387133881338913390133911339213393133941339513396133971339813399134001340113402134031340413405134061340713408134091341013411134121341313414134151341613417134181341913420134211342213423134241342513426134271342813429134301343113432134331343413435134361343713438134391344013441134421344313444134451344613447134481344913450134511345213453134541345513456134571345813459134601346113462134631346413465134661346713468134691347013471134721347313474134751347613477134781347913480134811348213483134841348513486134871348813489134901349113492134931349413495134961349713498134991350013501135021350313504135051350613507135081350913510135111351213513135141351513516135171351813519135201352113522135231352413525135261352713528135291353013531135321353313534135351353613537135381353913540135411354213543135441354513546135471354813549135501355113552135531355413555135561355713558135591356013561135621356313564135651356613567135681356913570135711357213573135741357513576135771357813579135801358113582135831358413585135861358713588135891359013591135921359313594135951359613597135981359913600136011360213603136041360513606136071360813609136101361113612136131361413615136161361713618136191362013621136221362313624136251362613627136281362913630136311363213633136341363513636136371363813639136401364113642136431364413645136461364713648136491365013651136521365313654136551365613657136581365913660136611366213663136641366513666136671366813669136701367113672136731367413675136761367713678136791368013681136821368313684136851368613687136881368913690136911369213693136941369513696136971369813699137001370113702137031370413705137061370713708137091371013711137121371313714137151371613717137181371913720137211372213723137241372513726137271372813729137301373113732137331373413735137361373713738137391374013741137421374313744137451374613747137481374913750137511375213753137541375513756137571375813759137601376113762137631376413765137661376713768137691377013771137721377313774137751377613777137781377913780137811378213783137841378513786137871378813789137901379113792137931379413795137961379713798137991380013801138021380313804138051380613807138081380913810138111381213813138141381513816138171381813819138201382113822138231382413825138261382713828138291383013831138321383313834138351383613837138381383913840138411384213843138441384513846138471384813849138501385113852138531385413855138561385713858138591386013861138621386313864138651386613867138681386913870138711387213873138741387513876138771387813879138801388113882138831388413885138861388713888138891389013891138921389313894138951389613897138981389913900139011390213903139041390513906139071390813909139101391113912139131391413915139161391713918139191392013921139221392313924139251392613927139281392913930139311393213933139341393513936139371393813939139401394113942139431394413945139461394713948139491395013951139521395313954139551395613957139581395913960139611396213963139641396513966139671396813969139701397113972139731397413975139761397713978139791398013981139821398313984139851398613987139881398913990139911399213993139941399513996139971399813999140001400114002140031400414005140061400714008140091401014011140121401314014140151401614017140181401914020140211402214023140241402514026140271402814029140301403114032140331403414035140361403714038140391404014041140421404314044140451404614047140481404914050140511405214053140541405514056140571405814059140601406114062140631406414065140661406714068140691407014071140721407314074140751407614077140781407914080140811408214083140841408514086140871408814089140901409114092140931409414095140961409714098140991410014101141021410314104141051410614107141081410914110141111411214113141141411514116141171411814119141201412114122141231412414125141261412714128141291413014131141321413314134141351413614137141381413914140141411414214143141441414514146141471414814149141501415114152141531415414155141561415714158141591416014161141621416314164141651416614167141681416914170141711417214173141741417514176141771417814179141801418114182141831418414185141861418714188141891419014191141921419314194141951419614197141981419914200142011420214203142041420514206142071420814209142101421114212142131421414215142161421714218142191422014221142221422314224142251422614227142281422914230142311423214233142341423514236142371423814239142401424114242142431424414245142461424714248142491425014251142521425314254142551425614257142581425914260142611426214263142641426514266142671426814269142701427114272142731427414275142761427714278142791428014281142821428314284142851428614287142881428914290142911429214293142941429514296142971429814299143001430114302143031430414305143061430714308143091431014311143121431314314143151431614317143181431914320143211432214323143241432514326143271432814329143301433114332143331433414335143361433714338143391434014341143421434314344143451434614347143481434914350143511435214353143541435514356143571435814359143601436114362143631436414365143661436714368143691437014371143721437314374143751437614377143781437914380143811438214383143841438514386143871438814389143901439114392143931439414395143961439714398143991440014401144021440314404144051440614407144081440914410144111441214413144141441514416144171441814419144201442114422144231442414425144261442714428144291443014431144321443314434144351443614437144381443914440144411444214443144441444514446144471444814449144501445114452144531445414455144561445714458144591446014461144621446314464144651446614467144681446914470144711447214473144741447514476144771447814479144801448114482144831448414485144861448714488144891449014491144921449314494144951449614497144981449914500145011450214503145041450514506145071450814509145101451114512145131451414515145161451714518145191452014521145221452314524145251452614527145281452914530145311453214533145341453514536145371453814539145401454114542145431454414545145461454714548145491455014551145521455314554145551455614557145581455914560145611456214563145641456514566145671456814569145701457114572145731457414575145761457714578145791458014581145821458314584145851458614587145881458914590145911459214593145941459514596145971459814599146001460114602146031460414605146061460714608146091461014611146121461314614146151461614617146181461914620146211462214623146241462514626146271462814629146301463114632146331463414635146361463714638146391464014641146421464314644146451464614647146481464914650146511465214653146541465514656146571465814659146601466114662146631466414665146661466714668146691467014671146721467314674146751467614677146781467914680146811468214683146841468514686146871468814689146901469114692146931469414695146961469714698146991470014701147021470314704147051470614707147081470914710147111471214713147141471514716147171471814719147201472114722147231472414725147261472714728147291473014731147321473314734147351473614737147381473914740147411474214743147441474514746147471474814749147501475114752147531475414755147561475714758147591476014761147621476314764147651476614767147681476914770147711477214773147741477514776147771477814779147801478114782147831478414785147861478714788147891479014791147921479314794147951479614797147981479914800148011480214803148041480514806148071480814809148101481114812148131481414815148161481714818148191482014821148221482314824148251482614827148281482914830148311483214833148341483514836148371483814839148401484114842148431484414845148461484714848148491485014851148521485314854148551485614857148581485914860148611486214863148641486514866148671486814869148701487114872148731487414875148761487714878148791488014881148821488314884148851488614887148881488914890148911489214893148941489514896148971489814899149001490114902149031490414905149061490714908149091491014911149121491314914149151491614917149181491914920149211492214923149241492514926149271492814929149301493114932149331493414935149361493714938149391494014941149421494314944149451494614947149481494914950149511495214953149541495514956149571495814959149601496114962149631496414965149661496714968149691497014971149721497314974149751497614977149781497914980149811498214983149841498514986149871498814989149901499114992149931499414995149961499714998149991500015001150021500315004150051500615007150081500915010150111501215013150141501515016150171501815019150201502115022150231502415025150261502715028150291503015031150321503315034150351503615037150381503915040150411504215043150441504515046150471504815049150501505115052150531505415055150561505715058150591506015061150621506315064150651506615067150681506915070150711507215073150741507515076150771507815079150801508115082150831508415085150861508715088150891509015091150921509315094150951509615097150981509915100151011510215103151041510515106151071510815109151101511115112151131511415115151161511715118151191512015121151221512315124151251512615127151281512915130151311513215133151341513515136151371513815139151401514115142151431514415145151461514715148151491515015151151521515315154151551515615157151581515915160151611516215163151641516515166151671516815169151701517115172151731517415175151761517715178151791518015181151821518315184151851518615187151881518915190151911519215193151941519515196151971519815199152001520115202152031520415205152061520715208152091521015211152121521315214152151521615217152181521915220152211522215223152241522515226152271522815229152301523115232152331523415235152361523715238152391524015241152421524315244152451524615247152481524915250152511525215253152541525515256152571525815259152601526115262152631526415265152661526715268152691527015271152721527315274152751527615277152781527915280152811528215283152841528515286152871528815289152901529115292152931529415295152961529715298152991530015301153021530315304153051530615307153081530915310153111531215313153141531515316153171531815319153201532115322153231532415325153261532715328153291533015331153321533315334153351533615337153381533915340153411534215343153441534515346153471534815349153501535115352153531535415355153561535715358153591536015361153621536315364153651536615367153681536915370153711537215373153741537515376153771537815379153801538115382153831538415385153861538715388153891539015391153921539315394153951539615397153981539915400154011540215403154041540515406154071540815409154101541115412154131541415415154161541715418154191542015421154221542315424154251542615427154281542915430154311543215433154341543515436154371543815439154401544115442154431544415445154461544715448154491545015451154521545315454154551545615457154581545915460154611546215463154641546515466154671546815469154701547115472154731547415475154761547715478154791548015481154821548315484154851548615487154881548915490154911549215493154941549515496154971549815499155001550115502155031550415505155061550715508155091551015511155121551315514155151551615517155181551915520155211552215523155241552515526155271552815529155301553115532155331553415535155361553715538155391554015541155421554315544155451554615547155481554915550155511555215553155541555515556155571555815559155601556115562155631556415565155661556715568155691557015571155721557315574155751557615577155781557915580155811558215583155841558515586155871558815589155901559115592155931559415595155961559715598155991560015601156021560315604156051560615607156081560915610156111561215613156141561515616156171561815619156201562115622156231562415625156261562715628156291563015631156321563315634156351563615637156381563915640156411564215643156441564515646156471564815649156501565115652156531565415655156561565715658156591566015661156621566315664156651566615667156681566915670156711567215673156741567515676156771567815679156801568115682156831568415685156861568715688156891569015691156921569315694156951569615697156981569915700157011570215703157041570515706157071570815709157101571115712157131571415715157161571715718157191572015721157221572315724157251572615727157281572915730157311573215733157341573515736157371573815739157401574115742157431574415745157461574715748157491575015751157521575315754157551575615757157581575915760157611576215763157641576515766157671576815769157701577115772157731577415775157761577715778157791578015781157821578315784157851578615787157881578915790157911579215793157941579515796157971579815799158001580115802158031580415805158061580715808158091581015811158121581315814158151581615817158181581915820158211582215823158241582515826158271582815829158301583115832158331583415835158361583715838158391584015841158421584315844158451584615847158481584915850158511585215853158541585515856158571585815859158601586115862158631586415865158661586715868158691587015871158721587315874158751587615877158781587915880158811588215883158841588515886158871588815889158901589115892158931589415895158961589715898158991590015901159021590315904159051590615907159081590915910159111591215913159141591515916159171591815919159201592115922159231592415925159261592715928159291593015931159321593315934159351593615937159381593915940159411594215943159441594515946159471594815949159501595115952159531595415955159561595715958159591596015961159621596315964159651596615967159681596915970159711597215973159741597515976159771597815979159801598115982159831598415985159861598715988159891599015991159921599315994159951599615997159981599916000160011600216003160041600516006160071600816009160101601116012160131601416015160161601716018160191602016021160221602316024160251602616027160281602916030160311603216033160341603516036160371603816039160401604116042160431604416045160461604716048160491605016051160521605316054160551605616057160581605916060160611606216063160641606516066160671606816069160701607116072160731607416075160761607716078160791608016081160821608316084160851608616087160881608916090160911609216093160941609516096160971609816099161001610116102161031610416105161061610716108161091611016111161121611316114161151611616117161181611916120161211612216123161241612516126161271612816129161301613116132161331613416135161361613716138161391614016141161421614316144161451614616147161481614916150161511615216153161541615516156161571615816159161601616116162161631616416165161661616716168161691617016171161721617316174161751617616177161781617916180161811618216183161841618516186161871618816189161901619116192161931619416195161961619716198161991620016201162021620316204162051620616207162081620916210162111621216213162141621516216162171621816219162201622116222162231622416225162261622716228162291623016231162321623316234162351623616237162381623916240162411624216243162441624516246162471624816249162501625116252162531625416255162561625716258162591626016261162621626316264162651626616267162681626916270162711627216273162741627516276162771627816279162801628116282162831628416285162861628716288162891629016291162921629316294162951629616297162981629916300163011630216303163041630516306163071630816309163101631116312163131631416315163161631716318163191632016321163221632316324163251632616327163281632916330163311633216333163341633516336163371633816339163401634116342163431634416345163461634716348163491635016351163521635316354163551635616357163581635916360163611636216363163641636516366163671636816369163701637116372163731637416375163761637716378163791638016381163821638316384163851638616387163881638916390163911639216393163941639516396163971639816399164001640116402164031640416405164061640716408164091641016411164121641316414164151641616417164181641916420164211642216423164241642516426164271642816429164301643116432164331643416435164361643716438164391644016441164421644316444164451644616447164481644916450164511645216453164541645516456164571645816459164601646116462164631646416465164661646716468164691647016471164721647316474164751647616477164781647916480164811648216483164841648516486164871648816489164901649116492164931649416495164961649716498164991650016501165021650316504165051650616507165081650916510165111651216513165141651516516165171651816519165201652116522165231652416525165261652716528165291653016531165321653316534165351653616537165381653916540165411654216543165441654516546165471654816549165501655116552165531655416555165561655716558165591656016561165621656316564165651656616567165681656916570165711657216573165741657516576165771657816579165801658116582165831658416585165861658716588165891659016591165921659316594165951659616597165981659916600166011660216603166041660516606166071660816609166101661116612166131661416615166161661716618166191662016621166221662316624166251662616627166281662916630166311663216633166341663516636166371663816639166401664116642166431664416645166461664716648166491665016651166521665316654166551665616657166581665916660166611666216663166641666516666166671666816669166701667116672166731667416675166761667716678166791668016681166821668316684166851668616687166881668916690166911669216693166941669516696166971669816699167001670116702167031670416705167061670716708167091671016711167121671316714167151671616717167181671916720167211672216723167241672516726167271672816729167301673116732167331673416735167361673716738167391674016741167421674316744167451674616747167481674916750167511675216753167541675516756167571675816759167601676116762167631676416765167661676716768167691677016771167721677316774167751677616777167781677916780167811678216783167841678516786167871678816789167901679116792167931679416795167961679716798167991680016801168021680316804168051680616807168081680916810168111681216813168141681516816168171681816819168201682116822168231682416825168261682716828168291683016831168321683316834168351683616837168381683916840168411684216843168441684516846168471684816849168501685116852168531685416855168561685716858168591686016861168621686316864168651686616867168681686916870168711687216873168741687516876168771687816879168801688116882168831688416885168861688716888168891689016891168921689316894168951689616897168981689916900169011690216903169041690516906169071690816909169101691116912169131691416915169161691716918169191692016921169221692316924169251692616927169281692916930169311693216933169341693516936169371693816939169401694116942169431694416945169461694716948169491695016951169521695316954169551695616957169581695916960169611696216963169641696516966169671696816969169701697116972169731697416975169761697716978169791698016981169821698316984169851698616987169881698916990169911699216993169941699516996169971699816999170001700117002170031700417005170061700717008170091701017011170121701317014170151701617017170181701917020170211702217023170241702517026170271702817029170301703117032170331703417035170361703717038170391704017041170421704317044170451704617047170481704917050170511705217053170541705517056170571705817059170601706117062170631706417065170661706717068170691707017071170721707317074170751707617077170781707917080170811708217083170841708517086170871708817089170901709117092170931709417095170961709717098170991710017101171021710317104171051710617107171081710917110171111711217113171141711517116171171711817119171201712117122171231712417125171261712717128171291713017131171321713317134171351713617137171381713917140171411714217143171441714517146171471714817149171501715117152171531715417155171561715717158171591716017161171621716317164171651716617167171681716917170171711717217173171741717517176171771717817179171801718117182171831718417185171861718717188171891719017191171921719317194171951719617197171981719917200172011720217203172041720517206172071720817209172101721117212172131721417215172161721717218172191722017221172221722317224172251722617227172281722917230172311723217233172341723517236172371723817239172401724117242172431724417245172461724717248172491725017251172521725317254172551725617257172581725917260172611726217263172641726517266172671726817269172701727117272172731727417275172761727717278172791728017281172821728317284172851728617287172881728917290172911729217293172941729517296172971729817299173001730117302173031730417305173061730717308173091731017311173121731317314173151731617317173181731917320173211732217323173241732517326173271732817329173301733117332173331733417335173361733717338173391734017341173421734317344173451734617347173481734917350173511735217353173541735517356173571735817359173601736117362173631736417365173661736717368173691737017371173721737317374173751737617377173781737917380173811738217383173841738517386173871738817389173901739117392173931739417395173961739717398173991740017401174021740317404174051740617407174081740917410174111741217413174141741517416174171741817419174201742117422174231742417425174261742717428174291743017431174321743317434174351743617437174381743917440174411744217443174441744517446174471744817449174501745117452174531745417455174561745717458174591746017461174621746317464174651746617467174681746917470174711747217473174741747517476174771747817479174801748117482174831748417485174861748717488174891749017491174921749317494174951749617497174981749917500175011750217503175041750517506175071750817509175101751117512175131751417515175161751717518175191752017521175221752317524175251752617527175281752917530175311753217533175341753517536175371753817539175401754117542175431754417545175461754717548175491755017551175521755317554175551755617557175581755917560175611756217563175641756517566175671756817569175701757117572175731757417575175761757717578175791758017581175821758317584175851758617587175881758917590175911759217593175941759517596175971759817599176001760117602176031760417605176061760717608176091761017611176121761317614176151761617617176181761917620176211762217623176241762517626176271762817629176301763117632176331763417635176361763717638176391764017641176421764317644176451764617647176481764917650176511765217653176541765517656176571765817659176601766117662176631766417665176661766717668176691767017671176721767317674176751767617677176781767917680176811768217683176841768517686176871768817689176901769117692176931769417695176961769717698176991770017701177021770317704177051770617707177081770917710177111771217713177141771517716177171771817719177201772117722177231772417725177261772717728177291773017731177321773317734177351773617737177381773917740177411774217743177441774517746177471774817749177501775117752177531775417755177561775717758177591776017761177621776317764177651776617767177681776917770177711777217773177741777517776177771777817779177801778117782177831778417785177861778717788177891779017791177921779317794177951779617797177981779917800178011780217803178041780517806178071780817809178101781117812178131781417815178161781717818178191782017821178221782317824178251782617827178281782917830178311783217833178341783517836178371783817839178401784117842178431784417845178461784717848178491785017851178521785317854178551785617857178581785917860178611786217863178641786517866178671786817869178701787117872178731787417875178761787717878178791788017881178821788317884178851788617887178881788917890178911789217893178941789517896178971789817899179001790117902179031790417905179061790717908179091791017911179121791317914179151791617917179181791917920179211792217923179241792517926179271792817929179301793117932179331793417935179361793717938179391794017941179421794317944179451794617947179481794917950179511795217953179541795517956179571795817959179601796117962179631796417965179661796717968179691797017971179721797317974179751797617977179781797917980179811798217983179841798517986179871798817989179901799117992179931799417995179961799717998179991800018001180021800318004180051800618007180081800918010180111801218013180141801518016180171801818019180201802118022180231802418025180261802718028180291803018031180321803318034180351803618037180381803918040180411804218043180441804518046180471804818049180501805118052180531805418055180561805718058180591806018061180621806318064180651806618067180681806918070180711807218073180741807518076180771807818079180801808118082180831808418085180861808718088180891809018091180921809318094180951809618097180981809918100181011810218103181041810518106181071810818109181101811118112181131811418115181161811718118181191812018121181221812318124181251812618127181281812918130181311813218133181341813518136181371813818139181401814118142181431814418145181461814718148181491815018151181521815318154181551815618157181581815918160181611816218163181641816518166181671816818169181701817118172181731817418175181761817718178181791818018181181821818318184181851818618187181881818918190181911819218193181941819518196181971819818199182001820118202182031820418205182061820718208182091821018211182121821318214182151821618217182181821918220182211822218223182241822518226182271822818229182301823118232182331823418235182361823718238182391824018241182421824318244182451824618247182481824918250182511825218253182541825518256182571825818259182601826118262182631826418265182661826718268182691827018271182721827318274182751827618277182781827918280182811828218283182841828518286182871828818289182901829118292182931829418295182961829718298182991830018301183021830318304183051830618307183081830918310183111831218313183141831518316183171831818319183201832118322183231832418325183261832718328183291833018331183321833318334183351833618337183381833918340183411834218343183441834518346183471834818349183501835118352183531835418355183561835718358183591836018361183621836318364183651836618367183681836918370183711837218373183741837518376183771837818379183801838118382183831838418385183861838718388183891839018391183921839318394183951839618397183981839918400184011840218403184041840518406184071840818409184101841118412184131841418415184161841718418184191842018421184221842318424184251842618427184281842918430184311843218433184341843518436184371843818439184401844118442184431844418445184461844718448184491845018451184521845318454184551845618457184581845918460184611846218463184641846518466184671846818469184701847118472184731847418475184761847718478184791848018481184821848318484184851848618487184881848918490184911849218493184941849518496184971849818499185001850118502185031850418505185061850718508185091851018511185121851318514185151851618517185181851918520185211852218523185241852518526185271852818529185301853118532185331853418535185361853718538185391854018541185421854318544185451854618547185481854918550185511855218553185541855518556185571855818559185601856118562185631856418565185661856718568185691857018571185721857318574185751857618577185781857918580185811858218583185841858518586185871858818589185901859118592185931859418595185961859718598185991860018601186021860318604186051860618607186081860918610186111861218613186141861518616186171861818619186201862118622186231862418625186261862718628186291863018631186321863318634186351863618637186381863918640186411864218643186441864518646186471864818649186501865118652186531865418655186561865718658186591866018661186621866318664186651866618667186681866918670186711867218673186741867518676186771867818679186801868118682186831868418685186861868718688186891869018691186921869318694186951869618697186981869918700187011870218703187041870518706187071870818709187101871118712187131871418715187161871718718187191872018721187221872318724187251872618727187281872918730187311873218733187341873518736187371873818739187401874118742187431874418745187461874718748187491875018751187521875318754187551875618757187581875918760187611876218763187641876518766187671876818769187701877118772187731877418775187761877718778187791878018781187821878318784187851878618787187881878918790187911879218793187941879518796187971879818799188001880118802188031880418805188061880718808188091881018811188121881318814188151881618817188181881918820188211882218823188241882518826188271882818829188301883118832188331883418835188361883718838188391884018841188421884318844188451884618847188481884918850188511885218853188541885518856188571885818859188601886118862188631886418865188661886718868188691887018871188721887318874188751887618877188781887918880188811888218883188841888518886188871888818889188901889118892188931889418895188961889718898188991890018901189021890318904189051890618907189081890918910189111891218913189141891518916189171891818919189201892118922189231892418925189261892718928189291893018931189321893318934189351893618937189381893918940189411894218943189441894518946189471894818949189501895118952189531895418955189561895718958189591896018961189621896318964189651896618967189681896918970189711897218973189741897518976189771897818979189801898118982189831898418985189861898718988189891899018991189921899318994189951899618997189981899919000190011900219003190041900519006190071900819009190101901119012190131901419015190161901719018190191902019021190221902319024190251902619027190281902919030190311903219033190341903519036190371903819039190401904119042190431904419045190461904719048190491905019051190521905319054190551905619057190581905919060190611906219063190641906519066190671906819069190701907119072190731907419075190761907719078190791908019081190821908319084190851908619087190881908919090190911909219093190941909519096190971909819099191001910119102191031910419105191061910719108191091911019111191121911319114191151911619117191181911919120191211912219123191241912519126191271912819129191301913119132191331913419135191361913719138191391914019141191421914319144191451914619147191481914919150191511915219153191541915519156191571915819159191601916119162191631916419165191661916719168191691917019171191721917319174191751917619177191781917919180191811918219183191841918519186191871918819189191901919119192191931919419195191961919719198191991920019201192021920319204192051920619207192081920919210192111921219213192141921519216192171921819219192201922119222192231922419225192261922719228192291923019231192321923319234192351923619237192381923919240192411924219243192441924519246192471924819249192501925119252192531925419255192561925719258192591926019261192621926319264192651926619267192681926919270192711927219273192741927519276192771927819279192801928119282192831928419285192861928719288192891929019291192921929319294192951929619297192981929919300193011930219303193041930519306193071930819309193101931119312193131931419315193161931719318193191932019321193221932319324193251932619327193281932919330193311933219333193341933519336193371933819339193401934119342193431934419345193461934719348193491935019351193521935319354193551935619357193581935919360193611936219363193641936519366193671936819369193701937119372193731937419375193761937719378193791938019381193821938319384193851938619387193881938919390193911939219393193941939519396193971939819399194001940119402194031940419405194061940719408194091941019411194121941319414194151941619417194181941919420194211942219423194241942519426194271942819429194301943119432194331943419435194361943719438194391944019441194421944319444194451944619447194481944919450194511945219453194541945519456194571945819459194601946119462194631946419465194661946719468194691947019471194721947319474194751947619477194781947919480194811948219483194841948519486194871948819489194901949119492194931949419495194961949719498194991950019501195021950319504195051950619507195081950919510195111951219513195141951519516195171951819519195201952119522195231952419525195261952719528195291953019531195321953319534195351953619537195381953919540195411954219543195441954519546195471954819549195501955119552195531955419555195561955719558195591956019561195621956319564195651956619567195681956919570195711957219573195741957519576195771957819579195801958119582195831958419585195861958719588195891959019591195921959319594195951959619597195981959919600196011960219603196041960519606196071960819609196101961119612196131961419615196161961719618196191962019621196221962319624196251962619627196281962919630196311963219633196341963519636196371963819639196401964119642196431964419645196461964719648196491965019651196521965319654196551965619657196581965919660196611966219663196641966519666196671966819669196701967119672196731967419675196761967719678196791968019681196821968319684196851968619687196881968919690196911969219693196941969519696196971969819699197001970119702197031970419705197061970719708197091971019711197121971319714197151971619717197181971919720197211972219723197241972519726197271972819729197301973119732197331973419735197361973719738197391974019741197421974319744197451974619747197481974919750197511975219753197541975519756197571975819759197601976119762197631976419765197661976719768197691977019771197721977319774197751977619777197781977919780197811978219783197841978519786197871978819789197901979119792197931979419795197961979719798197991980019801198021980319804198051980619807198081980919810198111981219813198141981519816198171981819819198201982119822198231982419825198261982719828198291983019831198321983319834198351983619837198381983919840198411984219843198441984519846198471984819849198501985119852198531985419855198561985719858198591986019861198621986319864198651986619867198681986919870198711987219873198741987519876198771987819879198801988119882198831988419885198861988719888198891989019891198921989319894198951989619897198981989919900199011990219903199041990519906199071990819909199101991119912199131991419915199161991719918199191992019921199221992319924199251992619927199281992919930199311993219933199341993519936199371993819939199401994119942199431994419945199461994719948199491995019951199521995319954199551995619957199581995919960199611996219963199641996519966199671996819969199701997119972199731997419975199761997719978199791998019981199821998319984199851998619987199881998919990199911999219993199941999519996199971999819999200002000120002200032000420005200062000720008200092001020011200122001320014200152001620017200182001920020200212002220023200242002520026200272002820029200302003120032200332003420035200362003720038200392004020041200422004320044200452004620047200482004920050200512005220053200542005520056200572005820059200602006120062200632006420065200662006720068200692007020071200722007320074200752007620077200782007920080200812008220083200842008520086200872008820089200902009120092200932009420095200962009720098200992010020101201022010320104201052010620107201082010920110201112011220113201142011520116201172011820119201202012120122201232012420125201262012720128201292013020131201322013320134201352013620137201382013920140201412014220143201442014520146201472014820149201502015120152201532015420155201562015720158201592016020161201622016320164201652016620167201682016920170201712017220173201742017520176201772017820179201802018120182201832018420185201862018720188201892019020191201922019320194201952019620197201982019920200202012020220203202042020520206202072020820209202102021120212202132021420215202162021720218202192022020221202222022320224202252022620227202282022920230202312023220233202342023520236202372023820239202402024120242202432024420245202462024720248202492025020251202522025320254202552025620257202582025920260202612026220263202642026520266202672026820269202702027120272202732027420275202762027720278202792028020281202822028320284202852028620287202882028920290202912029220293202942029520296202972029820299203002030120302203032030420305203062030720308203092031020311203122031320314203152031620317203182031920320203212032220323203242032520326203272032820329203302033120332203332033420335203362033720338203392034020341203422034320344203452034620347203482034920350203512035220353203542035520356203572035820359203602036120362203632036420365203662036720368203692037020371203722037320374203752037620377203782037920380203812038220383203842038520386203872038820389203902039120392203932039420395203962039720398203992040020401204022040320404204052040620407204082040920410204112041220413204142041520416204172041820419204202042120422204232042420425204262042720428204292043020431204322043320434204352043620437204382043920440204412044220443204442044520446204472044820449204502045120452204532045420455204562045720458204592046020461204622046320464204652046620467204682046920470204712047220473204742047520476204772047820479204802048120482204832048420485204862048720488204892049020491204922049320494204952049620497204982049920500205012050220503205042050520506205072050820509205102051120512205132051420515205162051720518205192052020521205222052320524205252052620527205282052920530205312053220533205342053520536205372053820539205402054120542205432054420545205462054720548205492055020551205522055320554205552055620557205582055920560205612056220563205642056520566205672056820569205702057120572205732057420575205762057720578205792058020581205822058320584205852058620587205882058920590205912059220593205942059520596205972059820599206002060120602206032060420605206062060720608206092061020611206122061320614206152061620617206182061920620206212062220623206242062520626206272062820629206302063120632206332063420635206362063720638206392064020641206422064320644206452064620647206482064920650206512065220653206542065520656206572065820659206602066120662206632066420665206662066720668206692067020671206722067320674206752067620677206782067920680206812068220683206842068520686206872068820689206902069120692206932069420695206962069720698206992070020701207022070320704207052070620707207082070920710207112071220713207142071520716207172071820719207202072120722207232072420725207262072720728207292073020731207322073320734207352073620737207382073920740207412074220743207442074520746207472074820749207502075120752207532075420755207562075720758207592076020761207622076320764207652076620767207682076920770207712077220773207742077520776207772077820779207802078120782207832078420785207862078720788207892079020791207922079320794207952079620797207982079920800208012080220803208042080520806208072080820809208102081120812208132081420815208162081720818208192082020821208222082320824208252082620827208282082920830208312083220833208342083520836208372083820839208402084120842208432084420845208462084720848208492085020851208522085320854208552085620857208582085920860208612086220863208642086520866208672086820869208702087120872208732087420875208762087720878208792088020881208822088320884208852088620887208882088920890208912089220893208942089520896208972089820899209002090120902209032090420905209062090720908209092091020911209122091320914209152091620917209182091920920209212092220923209242092520926209272092820929209302093120932209332093420935209362093720938209392094020941209422094320944209452094620947209482094920950209512095220953209542095520956209572095820959209602096120962209632096420965209662096720968209692097020971209722097320974209752097620977209782097920980209812098220983209842098520986209872098820989209902099120992209932099420995209962099720998209992100021001210022100321004210052100621007210082100921010210112101221013210142101521016210172101821019210202102121022210232102421025210262102721028210292103021031210322103321034210352103621037210382103921040210412104221043210442104521046210472104821049210502105121052210532105421055210562105721058210592106021061210622106321064210652106621067210682106921070210712107221073210742107521076210772107821079210802108121082210832108421085210862108721088210892109021091210922109321094210952109621097210982109921100211012110221103211042110521106211072110821109211102111121112211132111421115211162111721118211192112021121211222112321124211252112621127211282112921130211312113221133211342113521136211372113821139211402114121142211432114421145211462114721148211492115021151211522115321154211552115621157211582115921160211612116221163211642116521166211672116821169211702117121172211732117421175211762117721178211792118021181211822118321184211852118621187211882118921190211912119221193211942119521196211972119821199212002120121202212032120421205212062120721208212092121021211212122121321214212152121621217212182121921220212212122221223212242122521226212272122821229212302123121232212332123421235212362123721238212392124021241212422124321244212452124621247212482124921250212512125221253212542125521256212572125821259212602126121262212632126421265212662126721268212692127021271212722127321274212752127621277212782127921280212812128221283212842128521286212872128821289212902129121292212932129421295212962129721298212992130021301213022130321304213052130621307213082130921310213112131221313213142131521316213172131821319213202132121322213232132421325213262132721328213292133021331213322133321334213352133621337213382133921340213412134221343213442134521346213472134821349213502135121352213532135421355213562135721358213592136021361213622136321364213652136621367213682136921370213712137221373213742137521376213772137821379213802138121382213832138421385213862138721388213892139021391213922139321394213952139621397213982139921400214012140221403214042140521406214072140821409214102141121412214132141421415214162141721418214192142021421214222142321424214252142621427214282142921430214312143221433214342143521436214372143821439214402144121442214432144421445214462144721448214492145021451214522145321454214552145621457214582145921460214612146221463214642146521466214672146821469214702147121472214732147421475214762147721478214792148021481214822148321484214852148621487214882148921490214912149221493214942149521496214972149821499215002150121502215032150421505215062150721508215092151021511215122151321514215152151621517215182151921520215212152221523215242152521526215272152821529215302153121532215332153421535215362153721538215392154021541215422154321544215452154621547215482154921550215512155221553215542155521556215572155821559215602156121562215632156421565215662156721568215692157021571215722157321574215752157621577215782157921580215812158221583215842158521586215872158821589215902159121592215932159421595215962159721598215992160021601216022160321604216052160621607216082160921610216112161221613216142161521616216172161821619216202162121622216232162421625216262162721628216292163021631216322163321634216352163621637216382163921640216412164221643216442164521646216472164821649216502165121652216532165421655216562165721658216592166021661216622166321664216652166621667216682166921670216712167221673216742167521676216772167821679216802168121682216832168421685216862168721688216892169021691216922169321694216952169621697216982169921700217012170221703217042170521706217072170821709217102171121712217132171421715217162171721718217192172021721217222172321724217252172621727217282172921730217312173221733217342173521736217372173821739217402174121742217432174421745217462174721748217492175021751217522175321754217552175621757217582175921760217612176221763217642176521766217672176821769217702177121772217732177421775217762177721778217792178021781217822178321784217852178621787217882178921790217912179221793217942179521796217972179821799218002180121802218032180421805218062180721808218092181021811218122181321814218152181621817218182181921820218212182221823218242182521826218272182821829218302183121832218332183421835218362183721838218392184021841218422184321844218452184621847218482184921850218512185221853218542185521856218572185821859218602186121862218632186421865218662186721868218692187021871218722187321874218752187621877218782187921880218812188221883218842188521886218872188821889218902189121892218932189421895218962189721898218992190021901219022190321904219052190621907219082190921910219112191221913219142191521916219172191821919219202192121922219232192421925219262192721928219292193021931219322193321934219352193621937219382193921940219412194221943219442194521946219472194821949219502195121952219532195421955219562195721958219592196021961219622196321964219652196621967219682196921970219712197221973219742197521976219772197821979219802198121982219832198421985219862198721988219892199021991219922199321994219952199621997219982199922000220012200222003220042200522006220072200822009220102201122012220132201422015220162201722018220192202022021220222202322024220252202622027220282202922030220312203222033220342203522036220372203822039220402204122042220432204422045220462204722048220492205022051220522205322054220552205622057220582205922060220612206222063220642206522066220672206822069220702207122072220732207422075220762207722078220792208022081220822208322084220852208622087220882208922090220912209222093220942209522096220972209822099221002210122102221032210422105221062210722108221092211022111221122211322114221152211622117221182211922120221212212222123221242212522126221272212822129221302213122132221332213422135221362213722138221392214022141221422214322144221452214622147221482214922150221512215222153221542215522156221572215822159221602216122162221632216422165221662216722168221692217022171221722217322174221752217622177221782217922180221812218222183221842218522186221872218822189221902219122192221932219422195221962219722198221992220022201222022220322204222052220622207222082220922210222112221222213222142221522216222172221822219222202222122222222232222422225222262222722228222292223022231222322223322234222352223622237222382223922240222412224222243222442224522246222472224822249222502225122252222532225422255222562225722258222592226022261222622226322264222652226622267222682226922270222712227222273222742227522276222772227822279222802228122282222832228422285222862228722288222892229022291222922229322294222952229622297222982229922300223012230222303223042230522306223072230822309223102231122312223132231422315223162231722318223192232022321223222232322324223252232622327223282232922330223312233222333223342233522336223372233822339223402234122342223432234422345223462234722348223492235022351223522235322354223552235622357223582235922360223612236222363223642236522366223672236822369223702237122372223732237422375223762237722378223792238022381223822238322384223852238622387223882238922390223912239222393223942239522396223972239822399224002240122402224032240422405224062240722408224092241022411224122241322414224152241622417224182241922420224212242222423224242242522426224272242822429224302243122432224332243422435224362243722438224392244022441224422244322444224452244622447224482244922450224512245222453224542245522456224572245822459224602246122462224632246422465224662246722468224692247022471224722247322474224752247622477224782247922480224812248222483224842248522486224872248822489224902249122492224932249422495224962249722498224992250022501225022250322504225052250622507225082250922510225112251222513225142251522516225172251822519225202252122522225232252422525225262252722528225292253022531225322253322534225352253622537225382253922540225412254222543225442254522546225472254822549225502255122552225532255422555225562255722558225592256022561225622256322564225652256622567225682256922570225712257222573225742257522576225772257822579225802258122582225832258422585225862258722588225892259022591225922259322594225952259622597225982259922600226012260222603226042260522606226072260822609226102261122612226132261422615226162261722618226192262022621226222262322624226252262622627226282262922630226312263222633226342263522636226372263822639226402264122642226432264422645226462264722648226492265022651226522265322654226552265622657226582265922660226612266222663226642266522666226672266822669226702267122672226732267422675226762267722678226792268022681226822268322684226852268622687226882268922690226912269222693226942269522696226972269822699227002270122702227032270422705227062270722708227092271022711227122271322714227152271622717227182271922720227212272222723227242272522726227272272822729227302273122732227332273422735227362273722738227392274022741227422274322744227452274622747227482274922750227512275222753227542275522756227572275822759227602276122762227632276422765227662276722768227692277022771227722277322774227752277622777227782277922780227812278222783227842278522786227872278822789227902279122792227932279422795227962279722798227992280022801228022280322804228052280622807228082280922810228112281222813228142281522816228172281822819228202282122822228232282422825228262282722828228292283022831228322283322834228352283622837228382283922840228412284222843228442284522846228472284822849228502285122852228532285422855228562285722858228592286022861228622286322864228652286622867228682286922870228712287222873228742287522876228772287822879228802288122882228832288422885228862288722888228892289022891228922289322894228952289622897228982289922900229012290222903229042290522906229072290822909229102291122912229132291422915229162291722918229192292022921229222292322924229252292622927229282292922930229312293222933229342293522936229372293822939229402294122942229432294422945229462294722948229492295022951229522295322954229552295622957229582295922960229612296222963229642296522966229672296822969229702297122972229732297422975229762297722978229792298022981229822298322984229852298622987229882298922990229912299222993229942299522996229972299822999230002300123002230032300423005230062300723008230092301023011230122301323014230152301623017230182301923020230212302223023230242302523026230272302823029230302303123032230332303423035230362303723038230392304023041230422304323044230452304623047230482304923050230512305223053230542305523056230572305823059230602306123062230632306423065230662306723068230692307023071230722307323074230752307623077230782307923080230812308223083230842308523086230872308823089230902309123092230932309423095230962309723098230992310023101231022310323104231052310623107231082310923110231112311223113231142311523116231172311823119231202312123122231232312423125231262312723128231292313023131231322313323134231352313623137231382313923140231412314223143231442314523146231472314823149231502315123152231532315423155231562315723158231592316023161231622316323164231652316623167231682316923170231712317223173231742317523176231772317823179231802318123182231832318423185231862318723188231892319023191231922319323194231952319623197231982319923200232012320223203232042320523206232072320823209232102321123212232132321423215232162321723218232192322023221232222322323224232252322623227232282322923230232312323223233232342323523236232372323823239232402324123242232432324423245232462324723248232492325023251232522325323254232552325623257232582325923260232612326223263232642326523266232672326823269232702327123272232732327423275232762327723278232792328023281232822328323284232852328623287232882328923290232912329223293232942329523296232972329823299233002330123302233032330423305233062330723308233092331023311233122331323314233152331623317233182331923320233212332223323233242332523326233272332823329233302333123332233332333423335233362333723338233392334023341233422334323344233452334623347233482334923350233512335223353233542335523356233572335823359233602336123362233632336423365233662336723368233692337023371233722337323374233752337623377233782337923380233812338223383233842338523386233872338823389233902339123392233932339423395233962339723398233992340023401234022340323404234052340623407234082340923410234112341223413234142341523416234172341823419234202342123422234232342423425234262342723428234292343023431234322343323434234352343623437234382343923440234412344223443234442344523446234472344823449234502345123452234532345423455234562345723458234592346023461234622346323464234652346623467234682346923470234712347223473234742347523476234772347823479234802348123482234832348423485234862348723488234892349023491234922349323494234952349623497234982349923500235012350223503235042350523506235072350823509235102351123512235132351423515235162351723518235192352023521235222352323524235252352623527235282352923530235312353223533235342353523536235372353823539235402354123542235432354423545235462354723548235492355023551235522355323554235552355623557235582355923560235612356223563235642356523566235672356823569235702357123572235732357423575235762357723578235792358023581235822358323584235852358623587235882358923590235912359223593235942359523596235972359823599236002360123602236032360423605236062360723608236092361023611236122361323614236152361623617236182361923620236212362223623236242362523626236272362823629236302363123632236332363423635236362363723638236392364023641236422364323644236452364623647236482364923650236512365223653236542365523656236572365823659236602366123662236632366423665236662366723668236692367023671236722367323674236752367623677236782367923680236812368223683236842368523686236872368823689236902369123692236932369423695236962369723698236992370023701237022370323704237052370623707237082370923710237112371223713237142371523716237172371823719237202372123722237232372423725237262372723728237292373023731237322373323734237352373623737237382373923740237412374223743237442374523746237472374823749237502375123752237532375423755237562375723758237592376023761237622376323764237652376623767237682376923770237712377223773237742377523776237772377823779237802378123782237832378423785237862378723788237892379023791237922379323794237952379623797237982379923800238012380223803238042380523806238072380823809238102381123812238132381423815238162381723818238192382023821238222382323824238252382623827238282382923830238312383223833238342383523836238372383823839238402384123842238432384423845238462384723848238492385023851238522385323854238552385623857238582385923860238612386223863238642386523866238672386823869238702387123872238732387423875238762387723878238792388023881238822388323884238852388623887238882388923890238912389223893238942389523896238972389823899239002390123902239032390423905239062390723908239092391023911239122391323914239152391623917239182391923920239212392223923239242392523926239272392823929239302393123932239332393423935239362393723938239392394023941239422394323944239452394623947239482394923950239512395223953239542395523956239572395823959239602396123962239632396423965239662396723968239692397023971239722397323974239752397623977239782397923980239812398223983239842398523986239872398823989239902399123992239932399423995239962399723998239992400024001240022400324004240052400624007240082400924010240112401224013240142401524016240172401824019240202402124022240232402424025240262402724028240292403024031240322403324034240352403624037240382403924040240412404224043240442404524046240472404824049240502405124052240532405424055240562405724058240592406024061240622406324064240652406624067240682406924070240712407224073240742407524076240772407824079240802408124082240832408424085240862408724088240892409024091240922409324094240952409624097240982409924100241012410224103241042410524106241072410824109241102411124112241132411424115241162411724118241192412024121241222412324124241252412624127241282412924130241312413224133241342413524136241372413824139241402414124142241432414424145241462414724148241492415024151241522415324154241552415624157241582415924160241612416224163241642416524166241672416824169241702417124172241732417424175241762417724178241792418024181241822418324184241852418624187241882418924190241912419224193241942419524196241972419824199242002420124202242032420424205242062420724208242092421024211242122421324214242152421624217242182421924220242212422224223242242422524226242272422824229242302423124232242332423424235242362423724238242392424024241242422424324244242452424624247242482424924250242512425224253242542425524256242572425824259242602426124262242632426424265242662426724268242692427024271242722427324274242752427624277242782427924280242812428224283242842428524286242872428824289242902429124292242932429424295242962429724298242992430024301243022430324304243052430624307243082430924310243112431224313243142431524316243172431824319243202432124322243232432424325243262432724328243292433024331243322433324334243352433624337243382433924340243412434224343243442434524346243472434824349243502435124352243532435424355243562435724358243592436024361243622436324364243652436624367243682436924370243712437224373243742437524376243772437824379243802438124382243832438424385243862438724388243892439024391243922439324394243952439624397243982439924400244012440224403244042440524406244072440824409244102441124412244132441424415244162441724418244192442024421244222442324424244252442624427244282442924430244312443224433244342443524436244372443824439244402444124442244432444424445244462444724448244492445024451244522445324454244552445624457244582445924460244612446224463244642446524466244672446824469244702447124472244732447424475244762447724478244792448024481244822448324484244852448624487244882448924490244912449224493244942449524496244972449824499245002450124502245032450424505245062450724508245092451024511245122451324514245152451624517245182451924520245212452224523245242452524526245272452824529245302453124532245332453424535245362453724538245392454024541245422454324544245452454624547245482454924550245512455224553245542455524556245572455824559245602456124562245632456424565245662456724568245692457024571245722457324574245752457624577245782457924580245812458224583245842458524586245872458824589245902459124592245932459424595245962459724598245992460024601246022460324604246052460624607246082460924610246112461224613246142461524616246172461824619246202462124622246232462424625246262462724628246292463024631246322463324634246352463624637246382463924640246412464224643246442464524646246472464824649246502465124652246532465424655246562465724658246592466024661246622466324664246652466624667246682466924670246712467224673246742467524676246772467824679246802468124682246832468424685246862468724688246892469024691246922469324694246952469624697246982469924700247012470224703247042470524706247072470824709247102471124712247132471424715247162471724718247192472024721247222472324724247252472624727247282472924730247312473224733247342473524736247372473824739247402474124742247432474424745247462474724748247492475024751247522475324754247552475624757247582475924760247612476224763247642476524766247672476824769247702477124772247732477424775247762477724778247792478024781247822478324784247852478624787247882478924790247912479224793247942479524796247972479824799248002480124802248032480424805248062480724808248092481024811248122481324814248152481624817248182481924820248212482224823248242482524826248272482824829248302483124832248332483424835248362483724838248392484024841248422484324844248452484624847248482484924850248512485224853248542485524856248572485824859248602486124862248632486424865248662486724868248692487024871248722487324874248752487624877248782487924880248812488224883248842488524886248872488824889248902489124892248932489424895248962489724898248992490024901249022490324904249052490624907249082490924910249112491224913249142491524916249172491824919249202492124922249232492424925249262492724928249292493024931249322493324934249352493624937249382493924940249412494224943249442494524946249472494824949249502495124952249532495424955249562495724958249592496024961249622496324964249652496624967249682496924970249712497224973249742497524976249772497824979249802498124982249832498424985249862498724988249892499024991249922499324994249952499624997249982499925000250012500225003250042500525006250072500825009250102501125012250132501425015250162501725018250192502025021250222502325024250252502625027250282502925030250312503225033250342503525036250372503825039250402504125042250432504425045250462504725048250492505025051250522505325054250552505625057250582505925060250612506225063250642506525066250672506825069250702507125072250732507425075250762507725078250792508025081250822508325084250852508625087250882508925090250912509225093250942509525096250972509825099251002510125102251032510425105251062510725108251092511025111251122511325114251152511625117251182511925120251212512225123251242512525126251272512825129251302513125132251332513425135251362513725138251392514025141251422514325144251452514625147251482514925150251512515225153251542515525156251572515825159251602516125162251632516425165251662516725168251692517025171251722517325174251752517625177251782517925180251812518225183251842518525186251872518825189251902519125192251932519425195251962519725198251992520025201252022520325204252052520625207252082520925210252112521225213252142521525216252172521825219252202522125222252232522425225252262522725228252292523025231252322523325234252352523625237252382523925240252412524225243252442524525246252472524825249252502525125252252532525425255252562525725258252592526025261252622526325264252652526625267252682526925270252712527225273252742527525276252772527825279252802528125282252832528425285252862528725288252892529025291252922529325294252952529625297252982529925300253012530225303253042530525306253072530825309253102531125312253132531425315253162531725318253192532025321253222532325324253252532625327253282532925330253312533225333253342533525336253372533825339253402534125342253432534425345253462534725348253492535025351253522535325354253552535625357253582535925360253612536225363253642536525366253672536825369253702537125372253732537425375253762537725378253792538025381253822538325384253852538625387253882538925390253912539225393253942539525396253972539825399254002540125402254032540425405254062540725408254092541025411254122541325414254152541625417254182541925420254212542225423254242542525426254272542825429254302543125432254332543425435254362543725438254392544025441254422544325444254452544625447254482544925450254512545225453254542545525456254572545825459254602546125462254632546425465254662546725468254692547025471254722547325474254752547625477254782547925480254812548225483254842548525486254872548825489254902549125492254932549425495254962549725498254992550025501255022550325504255052550625507255082550925510255112551225513255142551525516255172551825519255202552125522255232552425525255262552725528255292553025531255322553325534255352553625537255382553925540255412554225543255442554525546255472554825549255502555125552255532555425555255562555725558255592556025561255622556325564255652556625567255682556925570255712557225573255742557525576255772557825579255802558125582255832558425585255862558725588255892559025591255922559325594255952559625597255982559925600256012560225603256042560525606256072560825609256102561125612256132561425615256162561725618256192562025621256222562325624256252562625627256282562925630256312563225633256342563525636256372563825639256402564125642256432564425645256462564725648256492565025651256522565325654256552565625657256582565925660256612566225663256642566525666256672566825669256702567125672256732567425675256762567725678256792568025681256822568325684256852568625687256882568925690256912569225693256942569525696256972569825699257002570125702257032570425705257062570725708257092571025711257122571325714257152571625717257182571925720257212572225723257242572525726257272572825729257302573125732257332573425735257362573725738257392574025741257422574325744257452574625747257482574925750257512575225753257542575525756257572575825759257602576125762257632576425765257662576725768257692577025771257722577325774257752577625777257782577925780257812578225783257842578525786257872578825789257902579125792257932579425795257962579725798257992580025801258022580325804258052580625807258082580925810258112581225813258142581525816258172581825819258202582125822258232582425825258262582725828258292583025831258322583325834258352583625837258382583925840258412584225843258442584525846258472584825849258502585125852258532585425855258562585725858258592586025861258622586325864258652586625867258682586925870258712587225873258742587525876258772587825879258802588125882258832588425885258862588725888258892589025891258922589325894258952589625897258982589925900259012590225903259042590525906259072590825909259102591125912259132591425915259162591725918259192592025921259222592325924259252592625927259282592925930259312593225933259342593525936259372593825939259402594125942259432594425945259462594725948259492595025951259522595325954259552595625957259582595925960259612596225963259642596525966259672596825969259702597125972259732597425975259762597725978259792598025981259822598325984259852598625987259882598925990259912599225993259942599525996259972599825999260002600126002260032600426005260062600726008260092601026011260122601326014260152601626017260182601926020260212602226023260242602526026260272602826029260302603126032260332603426035260362603726038260392604026041260422604326044260452604626047260482604926050260512605226053260542605526056260572605826059260602606126062260632606426065260662606726068260692607026071260722607326074260752607626077260782607926080260812608226083260842608526086260872608826089260902609126092260932609426095260962609726098260992610026101261022610326104261052610626107261082610926110261112611226113261142611526116261172611826119261202612126122261232612426125261262612726128261292613026131261322613326134261352613626137261382613926140261412614226143261442614526146261472614826149261502615126152261532615426155261562615726158261592616026161261622616326164261652616626167261682616926170261712617226173261742617526176261772617826179261802618126182261832618426185261862618726188261892619026191261922619326194261952619626197261982619926200262012620226203262042620526206262072620826209262102621126212262132621426215262162621726218262192622026221262222622326224262252622626227262282622926230262312623226233262342623526236262372623826239262402624126242262432624426245262462624726248262492625026251262522625326254262552625626257262582625926260262612626226263262642626526266262672626826269262702627126272262732627426275262762627726278262792628026281262822628326284262852628626287262882628926290262912629226293262942629526296262972629826299263002630126302263032630426305263062630726308263092631026311263122631326314263152631626317263182631926320263212632226323263242632526326263272632826329263302633126332263332633426335263362633726338263392634026341263422634326344263452634626347263482634926350263512635226353263542635526356263572635826359263602636126362263632636426365263662636726368263692637026371263722637326374263752637626377263782637926380263812638226383263842638526386263872638826389263902639126392263932639426395263962639726398263992640026401264022640326404264052640626407264082640926410264112641226413264142641526416264172641826419264202642126422264232642426425264262642726428264292643026431264322643326434264352643626437264382643926440264412644226443264442644526446264472644826449264502645126452264532645426455264562645726458264592646026461264622646326464264652646626467264682646926470264712647226473264742647526476264772647826479264802648126482264832648426485264862648726488264892649026491264922649326494264952649626497264982649926500265012650226503265042650526506265072650826509265102651126512265132651426515265162651726518265192652026521265222652326524265252652626527265282652926530265312653226533265342653526536265372653826539265402654126542265432654426545265462654726548265492655026551265522655326554265552655626557265582655926560265612656226563265642656526566265672656826569265702657126572265732657426575265762657726578265792658026581265822658326584265852658626587265882658926590265912659226593265942659526596265972659826599266002660126602266032660426605266062660726608266092661026611266122661326614266152661626617266182661926620266212662226623266242662526626266272662826629266302663126632266332663426635266362663726638266392664026641266422664326644266452664626647266482664926650266512665226653266542665526656266572665826659266602666126662266632666426665266662666726668266692667026671266722667326674266752667626677266782667926680266812668226683266842668526686266872668826689266902669126692266932669426695266962669726698266992670026701267022670326704267052670626707267082670926710267112671226713267142671526716267172671826719267202672126722267232672426725267262672726728267292673026731267322673326734267352673626737267382673926740267412674226743267442674526746267472674826749267502675126752267532675426755267562675726758267592676026761267622676326764267652676626767267682676926770267712677226773267742677526776267772677826779267802678126782267832678426785267862678726788267892679026791267922679326794267952679626797267982679926800268012680226803268042680526806268072680826809268102681126812268132681426815268162681726818268192682026821268222682326824268252682626827268282682926830268312683226833268342683526836268372683826839268402684126842268432684426845268462684726848268492685026851268522685326854268552685626857268582685926860268612686226863268642686526866268672686826869268702687126872268732687426875268762687726878268792688026881268822688326884268852688626887268882688926890268912689226893268942689526896268972689826899269002690126902269032690426905269062690726908269092691026911269122691326914269152691626917269182691926920269212692226923269242692526926269272692826929269302693126932269332693426935269362693726938269392694026941269422694326944269452694626947269482694926950269512695226953269542695526956269572695826959269602696126962269632696426965269662696726968269692697026971269722697326974269752697626977269782697926980269812698226983269842698526986269872698826989269902699126992269932699426995269962699726998269992700027001270022700327004270052700627007270082700927010270112701227013270142701527016270172701827019270202702127022270232702427025270262702727028270292703027031270322703327034270352703627037270382703927040270412704227043270442704527046270472704827049270502705127052270532705427055270562705727058270592706027061270622706327064270652706627067270682706927070270712707227073270742707527076270772707827079270802708127082270832708427085270862708727088270892709027091270922709327094270952709627097270982709927100271012710227103271042710527106271072710827109271102711127112271132711427115271162711727118271192712027121271222712327124271252712627127271282712927130271312713227133271342713527136271372713827139271402714127142271432714427145271462714727148271492715027151271522715327154271552715627157271582715927160271612716227163271642716527166271672716827169271702717127172271732717427175271762717727178271792718027181271822718327184271852718627187271882718927190271912719227193271942719527196271972719827199272002720127202272032720427205272062720727208272092721027211272122721327214272152721627217272182721927220272212722227223272242722527226272272722827229272302723127232272332723427235272362723727238272392724027241272422724327244272452724627247272482724927250272512725227253272542725527256272572725827259272602726127262272632726427265272662726727268272692727027271272722727327274272752727627277272782727927280272812728227283272842728527286272872728827289272902729127292272932729427295272962729727298272992730027301273022730327304273052730627307273082730927310273112731227313273142731527316273172731827319273202732127322273232732427325273262732727328273292733027331273322733327334273352733627337273382733927340273412734227343273442734527346273472734827349273502735127352273532735427355273562735727358273592736027361273622736327364273652736627367273682736927370273712737227373273742737527376273772737827379273802738127382273832738427385273862738727388273892739027391273922739327394273952739627397273982739927400274012740227403274042740527406274072740827409274102741127412274132741427415274162741727418274192742027421274222742327424274252742627427274282742927430274312743227433274342743527436274372743827439274402744127442274432744427445274462744727448274492745027451274522745327454274552745627457274582745927460274612746227463274642746527466274672746827469274702747127472274732747427475274762747727478274792748027481274822748327484274852748627487274882748927490274912749227493274942749527496274972749827499275002750127502275032750427505275062750727508275092751027511275122751327514275152751627517275182751927520275212752227523275242752527526275272752827529275302753127532275332753427535275362753727538275392754027541275422754327544275452754627547275482754927550275512755227553275542755527556275572755827559275602756127562275632756427565275662756727568275692757027571275722757327574275752757627577275782757927580275812758227583275842758527586275872758827589275902759127592275932759427595275962759727598275992760027601276022760327604276052760627607276082760927610276112761227613276142761527616276172761827619276202762127622276232762427625276262762727628276292763027631276322763327634276352763627637276382763927640276412764227643276442764527646276472764827649276502765127652276532765427655276562765727658276592766027661276622766327664276652766627667276682766927670276712767227673276742767527676276772767827679276802768127682276832768427685276862768727688276892769027691276922769327694276952769627697276982769927700277012770227703277042770527706277072770827709277102771127712277132771427715277162771727718277192772027721277222772327724277252772627727277282772927730277312773227733277342773527736277372773827739277402774127742277432774427745277462774727748277492775027751277522775327754277552775627757277582775927760277612776227763277642776527766277672776827769277702777127772277732777427775277762777727778277792778027781277822778327784277852778627787277882778927790277912779227793277942779527796277972779827799278002780127802278032780427805278062780727808278092781027811278122781327814278152781627817278182781927820278212782227823278242782527826278272782827829278302783127832278332783427835278362783727838278392784027841278422784327844278452784627847278482784927850278512785227853278542785527856278572785827859278602786127862278632786427865278662786727868278692787027871278722787327874278752787627877278782787927880278812788227883278842788527886278872788827889278902789127892278932789427895278962789727898278992790027901279022790327904279052790627907279082790927910279112791227913279142791527916279172791827919279202792127922279232792427925279262792727928279292793027931279322793327934279352793627937279382793927940279412794227943279442794527946279472794827949279502795127952279532795427955279562795727958279592796027961279622796327964279652796627967279682796927970279712797227973279742797527976279772797827979279802798127982279832798427985279862798727988279892799027991279922799327994279952799627997279982799928000280012800228003280042800528006280072800828009280102801128012280132801428015280162801728018280192802028021280222802328024280252802628027280282802928030280312803228033280342803528036280372803828039280402804128042280432804428045280462804728048280492805028051280522805328054280552805628057280582805928060280612806228063280642806528066280672806828069280702807128072280732807428075280762807728078280792808028081280822808328084280852808628087280882808928090280912809228093280942809528096280972809828099281002810128102281032810428105281062810728108281092811028111281122811328114281152811628117281182811928120281212812228123281242812528126281272812828129281302813128132281332813428135281362813728138281392814028141281422814328144281452814628147281482814928150281512815228153281542815528156281572815828159281602816128162281632816428165281662816728168281692817028171281722817328174281752817628177281782817928180281812818228183281842818528186281872818828189281902819128192281932819428195281962819728198281992820028201282022820328204282052820628207282082820928210282112821228213282142821528216282172821828219282202822128222282232822428225282262822728228282292823028231282322823328234282352823628237282382823928240282412824228243282442824528246282472824828249282502825128252282532825428255282562825728258282592826028261282622826328264282652826628267282682826928270282712827228273282742827528276282772827828279282802828128282282832828428285282862828728288282892829028291282922829328294282952829628297282982829928300283012830228303283042830528306283072830828309283102831128312283132831428315283162831728318283192832028321283222832328324283252832628327283282832928330283312833228333283342833528336283372833828339283402834128342283432834428345283462834728348283492835028351283522835328354283552835628357283582835928360283612836228363283642836528366283672836828369283702837128372283732837428375283762837728378283792838028381283822838328384283852838628387283882838928390283912839228393283942839528396283972839828399284002840128402284032840428405284062840728408284092841028411284122841328414284152841628417284182841928420284212842228423284242842528426284272842828429284302843128432284332843428435284362843728438284392844028441284422844328444284452844628447284482844928450284512845228453284542845528456284572845828459284602846128462284632846428465284662846728468284692847028471284722847328474284752847628477284782847928480284812848228483284842848528486284872848828489284902849128492284932849428495284962849728498284992850028501285022850328504285052850628507285082850928510285112851228513285142851528516285172851828519285202852128522285232852428525285262852728528285292853028531285322853328534285352853628537285382853928540285412854228543285442854528546285472854828549285502855128552285532855428555285562855728558285592856028561285622856328564285652856628567285682856928570285712857228573285742857528576285772857828579285802858128582285832858428585285862858728588285892859028591285922859328594285952859628597285982859928600286012860228603286042860528606286072860828609286102861128612286132861428615286162861728618286192862028621286222862328624286252862628627286282862928630286312863228633286342863528636286372863828639286402864128642286432864428645286462864728648286492865028651286522865328654286552865628657286582865928660286612866228663286642866528666286672866828669286702867128672286732867428675286762867728678286792868028681286822868328684286852868628687286882868928690286912869228693286942869528696286972869828699287002870128702287032870428705287062870728708287092871028711287122871328714287152871628717287182871928720287212872228723287242872528726287272872828729287302873128732287332873428735287362873728738287392874028741287422874328744287452874628747287482874928750287512875228753287542875528756287572875828759287602876128762287632876428765287662876728768287692877028771287722877328774287752877628777287782877928780287812878228783287842878528786287872878828789287902879128792287932879428795287962879728798287992880028801288022880328804288052880628807288082880928810288112881228813288142881528816288172881828819288202882128822288232882428825288262882728828288292883028831288322883328834288352883628837288382883928840288412884228843288442884528846288472884828849288502885128852288532885428855288562885728858288592886028861288622886328864288652886628867288682886928870288712887228873288742887528876288772887828879288802888128882288832888428885288862888728888288892889028891288922889328894288952889628897288982889928900289012890228903289042890528906289072890828909289102891128912289132891428915289162891728918289192892028921289222892328924289252892628927289282892928930289312893228933289342893528936289372893828939289402894128942289432894428945289462894728948289492895028951289522895328954289552895628957289582895928960289612896228963289642896528966289672896828969289702897128972289732897428975289762897728978289792898028981289822898328984289852898628987289882898928990289912899228993289942899528996289972899828999290002900129002290032900429005290062900729008290092901029011290122901329014290152901629017290182901929020290212902229023290242902529026290272902829029290302903129032290332903429035290362903729038290392904029041290422904329044290452904629047290482904929050290512905229053290542905529056290572905829059290602906129062290632906429065290662906729068290692907029071290722907329074290752907629077290782907929080290812908229083290842908529086290872908829089290902909129092290932909429095290962909729098290992910029101291022910329104291052910629107291082910929110291112911229113291142911529116291172911829119291202912129122291232912429125291262912729128291292913029131291322913329134291352913629137291382913929140291412914229143291442914529146291472914829149291502915129152291532915429155291562915729158291592916029161291622916329164291652916629167291682916929170291712917229173291742917529176291772917829179291802918129182291832918429185291862918729188291892919029191291922919329194291952919629197291982919929200292012920229203292042920529206292072920829209292102921129212292132921429215292162921729218292192922029221292222922329224292252922629227292282922929230292312923229233292342923529236292372923829239292402924129242292432924429245292462924729248292492925029251292522925329254292552925629257292582925929260292612926229263292642926529266292672926829269292702927129272292732927429275292762927729278292792928029281292822928329284292852928629287292882928929290292912929229293292942929529296292972929829299293002930129302293032930429305293062930729308293092931029311293122931329314293152931629317293182931929320293212932229323293242932529326293272932829329293302933129332293332933429335293362933729338293392934029341293422934329344293452934629347293482934929350293512935229353293542935529356293572935829359293602936129362293632936429365293662936729368293692937029371293722937329374293752937629377293782937929380293812938229383293842938529386293872938829389293902939129392293932939429395293962939729398293992940029401294022940329404294052940629407294082940929410294112941229413294142941529416294172941829419294202942129422294232942429425294262942729428294292943029431294322943329434294352943629437294382943929440294412944229443294442944529446294472944829449294502945129452294532945429455294562945729458294592946029461294622946329464294652946629467294682946929470294712947229473294742947529476294772947829479294802948129482294832948429485294862948729488294892949029491294922949329494294952949629497294982949929500295012950229503295042950529506295072950829509295102951129512295132951429515295162951729518295192952029521295222952329524295252952629527295282952929530295312953229533295342953529536295372953829539295402954129542295432954429545295462954729548295492955029551295522955329554295552955629557295582955929560295612956229563295642956529566295672956829569295702957129572295732957429575295762957729578295792958029581295822958329584295852958629587295882958929590295912959229593295942959529596295972959829599296002960129602296032960429605296062960729608296092961029611296122961329614296152961629617296182961929620296212962229623296242962529626296272962829629296302963129632296332963429635296362963729638296392964029641296422964329644
  1. % Generated by Paperpile. Check out https://paperpile.com for more information.
  2. % BibTeX export options can be customized via Settings -> BibTeX.
  3. @UNPUBLISHED{Okazawa2021-qd,
  4. title = "The geometry of the representation of decision variable and
  5. stimulus difficulty in the parietal cortex",
  6. author = "Okazawa, Gouki and Hatch, Christina E and Mancoo, Allan and
  7. Machens, Christian K and Kiani, Roozbeh",
  8. abstract = "Lateral intraparietal (LIP) neurons represent formation of
  9. perceptual decisions involving eye movements. In circuit models
  10. for these decisions, neural ensembles that encode actions compete
  11. to form decisions. Consequently, decision variables (DVs) are
  12. represented as partially potentiated action plans, where
  13. ensembles increase their average responses for stronger evidence
  14. supporting their preferred actions. As another consequence, DV
  15. representation and readout are implemented similarly for
  16. decisions with identical competing actions, irrespective of input
  17. and task context differences. Here, we challenge those core
  18. principles using a novel face-discrimination task, where LIP
  19. firing rates decrease with supporting evidence, contrary to
  20. conventional motion-discrimination tasks. These opposite response
  21. patterns arise from similar mechanisms in which decisions form
  22. along curved population-response manifolds misaligned with action
  23. representations. These manifolds rotate in state space based on
  24. task context, necessitating distinct readouts. We show similar
  25. manifolds in lateral and medial prefrontal cortices, suggesting a
  26. ubiquitous representational geometry across decision-making
  27. circuits. \#\#\# Competing Interest Statement The authors have
  28. declared no competing interest.",
  29. journal = "Cold Spring Harbor Laboratory",
  30. pages = "2021.01.04.425244",
  31. month = jan,
  32. year = 2021,
  33. keywords = "To Read",
  34. language = "en"
  35. }
  36. @ARTICLE{Murakami2014-lh,
  37. title = "Neural antecedents of self-initiated actions in secondary motor
  38. cortex",
  39. author = "Murakami, Masayoshi and Vicente, M In{\^e}s and Costa, Gil M and
  40. Mainen, Zachary F",
  41. abstract = "The neural origins of spontaneous or self-initiated actions are
  42. not well understood and their interpretation is controversial.
  43. To address these issues, we used a task in which rats decide
  44. when to abort waiting for a delayed tone. We recorded neurons in
  45. the secondary motor cortex (M2) and interpreted our findings in
  46. light of an integration-to-bound decision model. A first
  47. population of M2 neurons ramped to a constant threshold at rates
  48. proportional to waiting time, strongly resembling integrator
  49. output. A second population, which we propose provide input to
  50. the integrator, fired in sequences and showed trial-to-trial
  51. rate fluctuations correlated with waiting times. An integration
  52. model fit to these data also quantitatively predicted the
  53. observed inter-neuronal correlations. Together, these results
  54. reinforce the generality of the integration-to-bound model of
  55. decision-making. These models identify the initial intention to
  56. act as the moment of threshold crossing while explaining how
  57. antecedent subthreshold neural activity can influence an action
  58. without implying a decision.",
  59. journal = "Nat. Neurosci.",
  60. publisher = "nature.com",
  61. volume = 17,
  62. number = 11,
  63. pages = "1574--1582",
  64. month = nov,
  65. year = 2014,
  66. language = "en"
  67. }
  68. @ARTICLE{Gremel2013-im,
  69. title = "Premotor cortex is critical for goal-directed actions",
  70. author = "Gremel, Christina M and Costa, Rui M",
  71. abstract = "Shifting between motor plans is often necessary for adaptive
  72. behavior. When faced with changing consequences of one's
  73. actions, it is often imperative to switch from automatic actions
  74. to deliberative and controlled actions. The pre-supplementary
  75. motor area (pre-SMA) in primates, akin to the premotor cortex
  76. (M2) in mice, has been implicated in motor learning and
  77. planning, and action switching. We hypothesized that M2 would be
  78. differentially involved in goal-directed actions, which are
  79. controlled by their consequences vs. habits, which are more
  80. dependent on their past reinforcement history and less on their
  81. consequences. To investigate this, we performed M2 lesions in
  82. mice and then concurrently trained them to press the same lever
  83. for the same food reward using two different schedules of
  84. reinforcement that differentially bias towards the use of
  85. goal-directed versus habitual action strategies. We then probed
  86. whether actions were dependent on their expected consequence
  87. through outcome revaluation testing. We uncovered that M2
  88. lesions did not affect the acquisition of lever-pressing.
  89. However, in mice with M2 lesions, lever-pressing was insensitive
  90. to changes in expected outcome value following goal-directed
  91. training. However, habitual actions were intact. We confirmed a
  92. role for M2 in goal-directed but not habitual actions in
  93. separate groups of mice trained on the individual schedules
  94. biasing towards goal-directed versus habitual actions. These
  95. data indicate that M2 is critical for actions to be updated
  96. based on their consequences, and suggest that habitual action
  97. strategies may not require processing by M2 and the updating of
  98. motor plans.",
  99. journal = "Front. Comput. Neurosci.",
  100. publisher = "frontiersin.org",
  101. volume = 7,
  102. pages = "110",
  103. month = aug,
  104. year = 2013,
  105. keywords = "action selection; goal-directed actions; habitual actions;
  106. premotor cortex; value-based decision making;To Read",
  107. language = "en"
  108. }
  109. @ARTICLE{Erlich2011-rn,
  110. title = "A cortical substrate for memory-guided orienting in the rat",
  111. author = "Erlich, Jeffrey C and Bialek, Max and Brody, Carlos D",
  112. abstract = "Anatomical, stimulation, and lesion data have suggested a
  113. homology between the rat frontal orienting fields (FOF)
  114. (centered at +2 AP, $\pm$1.3 ML mm from Bregma) and primate
  115. frontal cortices such as the frontal or supplementary eye
  116. fields. We investigated the functional role of the FOF using
  117. rats trained to perform a memory-guided orienting task, in which
  118. there was a delay period between the end of a sensory stimulus
  119. instructing orienting direction and the time of the allowed
  120. motor response. Unilateral inactivation of the FOF resulted in
  121. impaired contralateral responses. Extracellular recordings of
  122. single units revealed that 37\% of FOF neurons had delay period
  123. firing rates that predicted the direction of the rats' later
  124. orienting motion. Our data provide the first
  125. electrophysiological and pharmacological evidence supporting the
  126. existence in the rat, as in the primate, of a frontal cortical
  127. area involved in the preparation and/or planning of orienting
  128. responses.",
  129. journal = "Neuron",
  130. publisher = "Elsevier",
  131. volume = 72,
  132. number = 2,
  133. pages = "330--343",
  134. month = oct,
  135. year = 2011,
  136. keywords = "To Read",
  137. language = "en"
  138. }
  139. @ARTICLE{Jeong2016-dq,
  140. title = "Comparative three-dimensional connectome map of motor cortical
  141. projections in the mouse brain",
  142. author = "Jeong, Minju and Kim, Yongsoo and Kim, Jeongjin and Ferrante,
  143. Daniel D and Mitra, Partha P and Osten, Pavel and Kim, Daesoo",
  144. abstract = "The motor cortex orchestrates simple to complex motor behaviors
  145. through its output projections to target areas. The primary (MOp)
  146. and secondary (MOs) motor cortices are known to produce specific
  147. output projections that are targeted to both similar and
  148. different target areas. These projections are further divided
  149. into layer 5 and 6 neuronal outputs, thereby producing four
  150. cortical outputs that may target other areas in a combinatorial
  151. manner. However, the precise network structure that integrates
  152. these four projections remains poorly understood. Here, we
  153. constructed a whole-brain, three-dimensional (3D) map showing the
  154. tract pathways and targeting locations of these four motor
  155. cortical outputs in mice. Remarkably, these motor cortical
  156. projections showed unique and separate tract pathways despite
  157. targeting similar areas. Within target areas, various
  158. combinations of these four projections were defined based on
  159. specific 3D spatial patterns, reflecting anterior-posterior,
  160. dorsal-ventral, and core-capsular relationships. This 3D
  161. topographic map ultimately provides evidence for the relevance of
  162. comparative connectomics: motor cortical projections known to be
  163. convergent are actually segregated in many target areas with
  164. unique targeting patterns, a finding that has anatomical value
  165. for revealing functional subdomains that have not been classified
  166. by conventional methods.",
  167. journal = "Sci. Rep.",
  168. volume = 6,
  169. pages = "20072",
  170. month = feb,
  171. year = 2016,
  172. keywords = "To Read",
  173. language = "en"
  174. }
  175. % The entry below contains non-ASCII chars that could not be converted
  176. % to a LaTeX equivalent.
  177. @ARTICLE{Gradinaru2007-qv,
  178. title = "Targeting and readout strategies for fast optical neural control
  179. in vitro and in vivo",
  180. author = "Gradinaru, Viviana and Thompson, Kimberly R and Zhang, Feng and
  181. Mogri, Murtaza and Kay, Kenneth and Schneider, M Bret and
  182. Deisseroth, Karl",
  183. abstract = "Major obstacles faced by neuroscientists in attempting to
  184. unravel the complexity of brain function include both the
  185. heterogeneity of brain tissue (with a multitude of cell types
  186. present in vivo) and the high speed of brain information
  187. processing (with behaviorally relevant millisecondscale
  188. electrical activity patterns). To address different aspects of
  189. these technical constraints, genetically targetable neural
  190. modulation tools have been developed by a number of groups
  191. (Zemelman et al., 2002; Banghart et al., 2004; Karpova et al.,
  192. 2005; Lima …",
  193. journal = "J. Neurosci.",
  194. publisher = "Soc Neuroscience",
  195. volume = 27,
  196. number = 52,
  197. pages = "14231--14238",
  198. month = dec,
  199. year = 2007,
  200. keywords = "Locomotion;To Read",
  201. language = "en"
  202. }
  203. @ARTICLE{Magno2019-qz,
  204. title = "Optogenetic Stimulation of the {M2} Cortex Reverts Motor
  205. Dysfunction in a Mouse Model of Parkinson's Disease",
  206. author = "Magno, Luiz Alexandre Viana and Tenza-Ferrer, Helia and
  207. Collodetti, M{\'e}lcar and Aguiar, Matheus Felipe Guimar{\~a}es
  208. and Rodrigues, Ana Paula Carneiro and da Silva, Rodrigo Souza and
  209. Silva, Joice do Prado and Nicolau, Nycolle Ferreira and Rosa,
  210. Daniela Valad{\~a}o Freitas and Birbrair, Alexander and Miranda,
  211. D{\'e}bora Marques and Romano-Silva, Marco Aur{\'e}lio",
  212. abstract = "Neuromodulation of deep brain structures (deep brain stimulation)
  213. is the current surgical procedure for treatment of Parkinson's
  214. disease (PD). Less studied is the stimulation of cortical motor
  215. areas to treat PD symptoms, although also known to alleviate
  216. motor disturbances in PD. We were able to show that optogenetic
  217. activation of secondary (M2) motor cortex improves motor
  218. functions in dopamine-depleted male mice. The stimulated M2
  219. cortex harbors glutamatergic pyramidal neurons that project to
  220. subcortical structures, critically involved in motor control, and
  221. makes synaptic contacts with dopaminergic neurons. Strikingly,
  222. optogenetic activation of M2 neurons or axons into the
  223. dorsomedial striatum increases striatal levels of dopamine and
  224. evokes locomotor activity. We found that dopamine
  225. neurotransmission sensitizes the locomotor behavior elicited by
  226. activation of M2 neurons. Furthermore, combination of intranigral
  227. infusion of glutamatergic antagonists and circuit specific
  228. optogenetic stimulation revealed that behavioral response
  229. depended on the activity of M2 neurons projecting to SNc.
  230. Interestingly, repeated M2 stimulation combined with l-DOPA
  231. treatment produced an unanticipated improvement in working memory
  232. performance, which was absent in control mice under l-DOPA
  233. treatment only. Therefore, the M2-basal ganglia circuit is
  234. critical for the assembly of the motor and cognitive function,
  235. and this study demonstrates a therapeutic mechanism for cortical
  236. stimulation in PD that involves recruitment of long-range
  237. glutamatergic projection neurons.SIGNIFICANCE STATEMENT Some
  238. patients with Parkinson's disease are offered treatment through
  239. surgery, which consists of delivering electrical current to
  240. regions deep within the brain. This study shows that stimulation
  241. of an area located on the brain surface, known as the secondary
  242. motor cortex, can also reverse movement disorders in mice.
  243. Authors have used a brain stimulation technique called
  244. optogenetics, which allowed targeting a specific type of surface
  245. neuron that communicates with the deep part of the brain involved
  246. in movement control. The study also shows that a combination of
  247. this stimulation with drug treatment might be useful to treat
  248. memory impairment, a kind of cognitive problem in Parkinson's
  249. disease.",
  250. journal = "J. Neurosci.",
  251. volume = 39,
  252. number = 17,
  253. pages = "3234--3248",
  254. month = apr,
  255. year = 2019,
  256. keywords = "Parkinson's disorder; brain stimulation; cognition; movement;
  257. optogenetics; prefrontal cortex;Locomotion;To Read",
  258. language = "en"
  259. }
  260. @ARTICLE{Schiemann2015-th,
  261. title = "Cellular mechanisms underlying behavioral state-dependent
  262. bidirectional modulation of motor cortex output",
  263. author = "Schiemann, Julia and Puggioni, Paolo and Dacre, Joshua and
  264. Pelko, Miha and Domanski, Aleksander and van Rossum, Mark C W
  265. and Duguid, Ian",
  266. abstract = "Neuronal activity in primary motor cortex (M1) correlates with
  267. behavioral state, but the cellular mechanisms underpinning
  268. behavioral state-dependent modulation of M1 output remain
  269. largely unresolved. Here, we performed in vivo patch-clamp
  270. recordings from layer 5B (L5B) pyramidal neurons in awake mice
  271. during quiet wakefulness and self-paced, voluntary movement. We
  272. show that L5B output neurons display bidirectional (i.e.,
  273. enhanced or suppressed) firing rate changes during movement,
  274. mediated via two opposing subthreshold mechanisms: (1) a global
  275. decrease in membrane potential variability that reduced L5B
  276. firing rates (L5Bsuppressed neurons), and (2) a coincident
  277. noradrenaline-mediated increase in excitatory drive to a
  278. subpopulation of L5B neurons (L5Benhanced neurons) that elevated
  279. firing rates. Blocking noradrenergic receptors in forelimb M1
  280. abolished the bidirectional modulation of M1 output during
  281. movement and selectively impaired contralateral forelimb motor
  282. coordination. Together, our results provide a mechanism for how
  283. noradrenergic neuromodulation and network-driven input changes
  284. bidirectionally modulate M1 output during motor behavior.",
  285. journal = "Cell Rep.",
  286. publisher = "Elsevier",
  287. volume = 11,
  288. number = 8,
  289. pages = "1319--1330",
  290. month = may,
  291. year = 2015,
  292. keywords = "To Read",
  293. language = "en"
  294. }
  295. @ARTICLE{Ebbesen2017-cm,
  296. title = "Motor cortex - to act or not to act?",
  297. author = "Ebbesen, Christian Laut and Brecht, Michael",
  298. abstract = "The motor cortex is a large frontal structure in the cerebral
  299. cortex of eutherian mammals. A vast array of evidence implicates
  300. the motor cortex in the volitional control of motor output, but
  301. how does the motor cortex exert this 'control'? Historically,
  302. ideas regarding motor cortex function have been shaped by the
  303. discovery of cortical 'motor maps' - that is, ordered
  304. representations of stimulation-evoked movements in anaesthetized
  305. animals. Volitional control, however, entails the initiation of
  306. movements and the ability to suppress undesired movements. In
  307. this article, we highlight classic and recent findings that
  308. emphasize that motor cortex neurons have a role in both
  309. processes.",
  310. journal = "Nat. Rev. Neurosci.",
  311. volume = 18,
  312. number = 11,
  313. pages = "694--705",
  314. month = oct,
  315. year = 2017,
  316. language = "en"
  317. }
  318. @ARTICLE{Calton2009-hj,
  319. title = "Where am {I} and how will {I} get there from here? A role for
  320. posterior parietal cortex in the integration of spatial
  321. information and route planning",
  322. author = "Calton, Jeffrey L and Taube, Jeffrey S",
  323. abstract = "The ability of an organism to accurately navigate from one place
  324. to another requires integration of multiple spatial constructs,
  325. including the determination of one's position and direction in
  326. space relative to allocentric landmarks, movement velocity, and
  327. the perceived location of the goal of the movement. In this
  328. review, we propose that while limbic areas are important for the
  329. sense of spatial orientation, the posterior parietal cortex is
  330. responsible for relating this sense with the location of a
  331. navigational goal and in formulating a plan to attain it. Hence,
  332. the posterior parietal cortex is important for the computation of
  333. the correct trajectory or route to be followed while navigating.
  334. Prefrontal and motor areas are subsequently responsible for
  335. executing the planned movement. Using this theory, we are able to
  336. bridge the gap between the rodent and primate literatures by
  337. suggesting that the allocentric role of the rodent PPC is largely
  338. analogous to the egocentric role typically emphasized in
  339. primates, that is, the integration of spatial orientation with
  340. potential goals in the planning of goal-directed movements.",
  341. journal = "Neurobiol. Learn. Mem.",
  342. volume = 91,
  343. number = 2,
  344. pages = "186--196",
  345. month = feb,
  346. year = 2009,
  347. keywords = "navigation;To Read",
  348. language = "en"
  349. }
  350. @ARTICLE{Cho2001-nw,
  351. title = "Head direction, place, and movement correlates for cells in the
  352. rat retrosplenial cortex",
  353. author = "Cho, J and Sharp, P E",
  354. abstract = "The retrosplenial cortex is strongly connected with brain
  355. regions involved in spatial signaling. To test whether it also
  356. codes space, single cells were recorded while rats navigated in
  357. an open field. As in earlier work (L. L. Chen, L. H. Lin, C. A.
  358. Barnes, \& B. L. McNaughton, 1994; L. L. Chen, L. H. Lin, E. J.
  359. Green, C. A. Barnes, \& B. L. McNaughton, 1994), the authors
  360. found head direction cells with properties similar to those in
  361. other areas. These cells were slightly anticipatory. Another
  362. cell type fired to particular combinations of location,
  363. direction, and movement, which suggested that they may fire
  364. whenever the rat approaches a particular location, using a
  365. particular locomotor behavior. The remaining cells could not be
  366. clearly categorized but also showed a significant correlation
  367. with one or more of the spatial-movement variables examined. The
  368. fact that the retrosplenial cortex contains spatial and
  369. movement-related signals and is connected with the motor cortex
  370. suggests that it may play a role in path integration or
  371. navigational motor planning.",
  372. journal = "Behav. Neurosci.",
  373. publisher = "psycnet.apa.org",
  374. volume = 115,
  375. number = 1,
  376. pages = "3--25",
  377. month = feb,
  378. year = 2001,
  379. keywords = "navigation;To Read",
  380. language = "en"
  381. }
  382. @ARTICLE{Li2015-tj,
  383. title = "A motor cortex circuit for motor planning and movement",
  384. author = "Li, Nuo and Chen, Tsai-Wen and Guo, Zengcai V and Gerfen, Charles
  385. R and Svoboda, Karel",
  386. abstract = "Activity in motor cortex predicts specific movements seconds
  387. before they occur, but how this preparatory activity relates to
  388. upcoming movements is obscure. We dissected the conversion of
  389. preparatory activity to movement within a structured motor cortex
  390. circuit. An anterior lateral region of the mouse cortex (a
  391. possible homologue of premotor cortex in primates) contains equal
  392. proportions of intermingled neurons predicting ipsi- or
  393. contralateral movements, yet unilateral inactivation of this
  394. cortical region during movement planning disrupts contralateral
  395. movements. Using cell-type-specific electrophysiology, cellular
  396. imaging and optogenetic perturbation, we show that layer 5
  397. neurons projecting within the cortex have unbiased laterality.
  398. Activity with a contralateral population bias arises specifically
  399. in layer 5 neurons projecting to the brainstem, and only late
  400. during movement planning. These results reveal the transformation
  401. of distributed preparatory activity into movement commands within
  402. hierarchically organized cortical circuits.",
  403. journal = "Nature",
  404. volume = 519,
  405. number = 7541,
  406. pages = "51--56",
  407. month = mar,
  408. year = 2015,
  409. keywords = "To Read",
  410. language = "en"
  411. }
  412. @ARTICLE{McNaughton1994-hv,
  413. title = "Cortical representation of motion during unrestrained spatial
  414. navigation in the rat",
  415. author = "McNaughton, B L and Mizumori, S J and Barnes, C A and Leonard, B
  416. J and Marquis, M and Green, E J",
  417. abstract = "Neural activity related to unrestrained movement through space
  418. was studied in rat sensorimotor and posterior parietal cortices
  419. during performance of an eight-arm, radial maze task. Nearly half
  420. of the cells exhibited movement-related activity that
  421. discriminated among three basic modes of locomotion: left turns,
  422. right turns, and forward motion. Correlates ranged from strong
  423. excitation (relative to the still condition) to strong
  424. inhibition, and were distributed among the movement modes in a
  425. variety of different ways. For example, cells that discriminated
  426. between clockwise and counterclockwise turns did so with either
  427. antagonistic responses or simple excitation or inhibition. Others
  428. showed either excitation or inhibition relative to both turning
  429. and the still condition, and hence were selective for forward
  430. motion. Many cells exhibited somatosensory responsiveness;
  431. however, in agreement with findings of others, motion correlates
  432. could rarely be sensibly explained by the somatosensory response.
  433. Moreover, movement correlates sometimes varied considerably with
  434. spatial context. Some cells exhibited more complex motion
  435. correlates, such as an apparent dependence on the nature of the
  436. preceding movement. Irrespective of the specific sensory or motor
  437. determinants of cell activity, which varied considerably among
  438. cells, the posterior neocortex of the rat appears to generate a
  439. robust and redundant internal representation of body motion
  440. through space. Such a representation could be useful in
  441. constructing ``cognitive maps'' of the environment.",
  442. journal = "Cereb. Cortex",
  443. volume = 4,
  444. number = 1,
  445. pages = "27--39",
  446. month = jan,
  447. year = 1994,
  448. keywords = "navigation",
  449. language = "en"
  450. }
  451. @ARTICLE{Kiehn2006-wi,
  452. title = "Locomotor circuits in the mammalian spinal cord",
  453. author = "Kiehn, Ole",
  454. abstract = "Intrinsic spinal networks, known as central pattern generators
  455. (CPGs), control the timing and pattern of the muscle activity
  456. underlying locomotion in mammals. This review discusses new
  457. advances in understanding the mammalian CPGs with a focus on
  458. experiments that address the overall network structure as well
  459. as the identification of CPG neurons. I address the
  460. identification of excitatory CPG neurons and their role in
  461. rhythm generation, the organization of flexor-extensor networks,
  462. and the diverse role of commissural interneurons in coordinating
  463. left-right movements. Molecular and genetic approaches that have
  464. the potential to elucidate the function of populations of CPG
  465. interneurons are also discussed.",
  466. journal = "Annu. Rev. Neurosci.",
  467. publisher = "annualreviews.org",
  468. volume = 29,
  469. pages = "279--306",
  470. year = 2006,
  471. keywords = "Locomotion",
  472. language = "en"
  473. }
  474. @ARTICLE{Fukuoka2015-ks,
  475. title = "A simple rule for quadrupedal gait generation determined by leg
  476. loading feedback: a modeling study",
  477. author = "Fukuoka, Yasuhiro and Habu, Yasushi and Fukui, Takahiro",
  478. abstract = "We discovered a specific rule for generating typical quadrupedal
  479. gaits (the order of the movement of four legs) through a
  480. simulated quadrupedal locomotion, in which unprogrammed gaits
  481. (diagonal/lateral sequence walks, left/right-lead canters, and
  482. left/right-lead transverse gallops) spontaneously emerged because
  483. of leg loading feedbacks to the CPGs hard-wired to produce a
  484. default trot. Additionally, all gaits transitioned according to
  485. speed, as seen in animals. We have therefore hypothesized that
  486. various gaits derive from a trot because of posture control
  487. through leg loading feedback. The body tilt on the two support
  488. legs of each diagonal pair during trotting was classified into
  489. three types (level, tilted up, or tilted down) according to
  490. speed. The load difference between the two legs led to the phase
  491. difference between their CPGs via the loading feedbacks,
  492. resulting in nine gaits (3(2): three tilts to the power of two
  493. diagonal pairs) including the aforementioned.",
  494. journal = "Sci. Rep.",
  495. volume = 5,
  496. pages = "8169",
  497. month = feb,
  498. year = 2015,
  499. keywords = "Locomotion",
  500. language = "en"
  501. }
  502. @ARTICLE{Inagaki_undated-th,
  503. title = "A midbrain - thalamus - cortex circuit reorganizes cortical
  504. dynamics to initiate planned movement",
  505. author = "Inagaki, Hidehiko K and Chen, Susu and Ridder, Margreet C and
  506. Sah, Pankaj and Li, Nuo and Yang, Zidan and Hasanbegovic, Hana
  507. and Gao, Zhenyu and Gerfen, Charles R and Svoboda, Karel",
  508. keywords = "To Read"
  509. }
  510. @UNPUBLISHED{Roseberry2019-iz,
  511. title = "Locomotor suppression by a monosynaptic amygdala to brainstem
  512. circuit",
  513. author = "Roseberry, Thomas K and Lalive, Arnaud L and Margolin, Benjamin D
  514. and Kreitzer, Anatol C",
  515. abstract = "Abstract The control of locomotion is fundamental to vertebrate
  516. animal survival. Defensive situations require an animal to
  517. rapidly decide whether to run away or suppress locomotor activity
  518. to avoid detection. While much of the neural circuitry involved
  519. in defensive action selection has been elucidated, top-down
  520. modulation of brainstem locomotor circuitry remains unclear. Here
  521. we provide evidence for the existence and functionality of a
  522. monosynaptic connection from the central amygdala (CeA) to the
  523. mesencephalic locomotor region (MLR) that inhibits locomotion in
  524. unconditioned and conditioned defensive behavior in mice. We show
  525. that locomotion stimulated by airpuff coincides with increased
  526. activity of MLR glutamatergic neurons. Using retrograde tracing
  527. and ex vivo electrophysiology, we find that the CeA makes a
  528. monosynaptic connection with the MLR. In the open field, in vivo
  529. stimulation of this projection suppressed spontaneous locomotion,
  530. whereas inhibition of this projection had no effect. However,
  531. inhibiting CeA terminals within the MLR increased both neural
  532. activity and locomotor responses to airpuff. Finally, using a
  533. conditioned avoidance paradigm known to activate CeA neurons, we
  534. find that inhibition of the CeA projection increased successful
  535. escape, whereas activating the projection reduced escape.
  536. Together these results provide evidence for a new circuit
  537. substrate influencing locomotion and defensive behaviors.",
  538. journal = "Cold Spring Harbor Laboratory",
  539. pages = "724252",
  540. month = aug,
  541. year = 2019,
  542. keywords = "Locomotion",
  543. language = "en"
  544. }
  545. @ARTICLE{Carvalho2020-pw,
  546. title = "A Brainstem Locomotor Circuit Drives the Activity of Speed Cells
  547. in the Medial Entorhinal Cortex",
  548. author = "Carvalho, Miguel M and Tanke, Nouk and Kropff, Emilio and Witter,
  549. Menno P and Moser, May-Britt and Moser, Edvard I",
  550. abstract = "Locomotion activates an array of sensory inputs that may help
  551. build the self-position map of the medial entorhinal cortex
  552. (MEC). In this map, speed-coding neurons are thought to
  553. dynamically update representations of the animal's position. A
  554. possible origin for the entorhinal speed signal is the
  555. mesencephalic locomotor region (MLR), which is critically
  556. involved in the activation of locomotor programs. Here, we
  557. describe, in rats, a circuit connecting the pedunculopontine
  558. tegmental nucleus (PPN) of the MLR to the MEC via the horizontal
  559. limb of the diagonal band of Broca (HDB). At each level of this
  560. pathway, locomotion speed is linearly encoded in neuronal firing
  561. rates. Optogenetic activation of PPN cells drives locomotion and
  562. modulates activity of speed-modulated neurons in HDB and MEC. Our
  563. results provide evidence for a pathway by which brainstem speed
  564. signals can reach cortical structures implicated in navigation
  565. and higher-order dynamic representations of space.",
  566. journal = "Cell Rep.",
  567. volume = 32,
  568. number = 10,
  569. pages = "108123",
  570. month = sep,
  571. year = 2020,
  572. keywords = "diagonal band of Broca; medial entorhinal cortex; mesencephalic
  573. locomotor region; pedunculopontine tegmental nucleus; speed
  574. cells;Locomotion",
  575. language = "en"
  576. }
  577. @UNPUBLISHED{Dautan2020-lv,
  578. title = "Modulation of motor behavior by the mesencephalic locomotor
  579. region",
  580. author = "Dautan, Daniel and Kov{\'a}cs, Adrienn and Bayasgalan,
  581. Tsogbadrakh and Diaz-Acevedo, Miguel A and Pal, Balazs and
  582. Mena-Segovia, Juan",
  583. abstract = "The mesencephalic locomotor region (MLR) serves as an interface
  584. between higher-order motor systems and lower motor neurons. The
  585. excitatory module of the MLR is composed of the pedunculopontine
  586. nucleus (PPN) and the cuneiform nucleus (CnF), and their
  587. activation has been proposed to elicit different modalities of
  588. movement, but how the differences in connectivity and
  589. physiological properties explain their contributions to motor
  590. activity is not known. Here we report that CnF glutamatergic
  591. neurons are electrophysiologically homogeneous and have
  592. short-range axonal projections, whereas PPN glutamatergic neurons
  593. are heterogeneous and maintain long-range connections, most
  594. notably with the basal ganglia. Optogenetic activation of CnF
  595. neurons produced fast-onset, involuntary motor activity mediated
  596. by short-lasting muscle activation. In contrast, activation of
  597. PPN neurons produced long-lasting increases in muscle tone that
  598. reduced motor activity and disrupted gait. Our results thus
  599. reveal a differential contribution to motor behavior by the
  600. structures that compose the MLR. \#\#\# Competing Interest
  601. Statement The authors have declared no competing interest.",
  602. journal = "Cold Spring Harbor Laboratory",
  603. pages = "2020.06.25.172296",
  604. month = jun,
  605. year = 2020,
  606. keywords = "Locomotion",
  607. language = "en"
  608. }
  609. @ARTICLE{Ruder2016-iv,
  610. title = "{Long-Distance} Descending Spinal Neurons Ensure Quadrupedal
  611. Locomotor Stability",
  612. author = "Ruder, Ludwig and Takeoka, Aya and Arber, Silvia",
  613. abstract = "Locomotion is an essential animal behavior used for
  614. translocation. The spinal cord acts as key executing center, but
  615. how it coordinates many body parts located across distance
  616. remains poorly understood. Here we employed mouse genetic and
  617. viral approaches to reveal organizational principles of
  618. long-projecting spinal circuits and their role in quadrupedal
  619. locomotion. Using neurotransmitter identity, developmental
  620. origin, and projection patterns as criteria, we uncover that
  621. spinal segments controlling forelimbs and hindlimbs are
  622. bidirectionally connected by symmetrically organized direct
  623. synaptic pathways that encompass multiple genetically tractable
  624. neuronal subpopulations. We demonstrate that selective ablation
  625. of descending spinal neurons linking cervical to lumbar segments
  626. impairs coherent locomotion, by reducing postural stability and
  627. speed during exploratory locomotion, as well as perturbing
  628. interlimb coordination during reinforced high-speed stepping.
  629. Together, our results implicate a highly organized long-distance
  630. projection system of spinal origin in the control of postural
  631. body stabilization and reliability during quadrupedal
  632. locomotion.",
  633. journal = "Neuron",
  634. publisher = "Elsevier",
  635. volume = 92,
  636. number = 5,
  637. pages = "1063--1078",
  638. month = dec,
  639. year = 2016,
  640. keywords = "genetic identity; interlimb coordination; locomotion; motor
  641. control; posture; spinal cord;Locomotion",
  642. language = "en"
  643. }
  644. @ARTICLE{Drew2015-sv,
  645. title = "Taking the next step: cortical contributions to the control of
  646. locomotion",
  647. author = "Drew, Trevor and Marigold, Daniel S",
  648. abstract = "The planning and execution of both discrete voluntary movements
  649. and visually guided locomotion depends on the contribution of
  650. multiple cortical areas. In this review, we discuss recent
  651. experiments that address the contribution of the posterior
  652. parietal cortex (PPC) and the motor cortex to the control of
  653. locomotion. The results from these experiments show that the PPC
  654. contributes to the planning of locomotion by providing an
  655. estimate of the position of an animal with respect to objects in
  656. its path. In contrast, the motor cortex contributes primarily to
  657. the execution of gait modifications by modulating the activity
  658. of groups of synergistic muscles active at different times
  659. during the gait cycle.",
  660. journal = "Curr. Opin. Neurobiol.",
  661. publisher = "Elsevier",
  662. volume = 33,
  663. pages = "25--33",
  664. month = aug,
  665. year = 2015,
  666. keywords = "Locomotion",
  667. language = "en"
  668. }
  669. @ARTICLE{Ueno2011-tt,
  670. title = "Kinematic analyses reveal impaired locomotion following injury
  671. of the motor cortex in mice",
  672. author = "Ueno, Masaki and Yamashita, Toshihide",
  673. abstract = "Brain injury in the motor cortex can result in deleterious
  674. functional deficits of skilled and fine motor functions.
  675. However, in contrast to humans, the destruction of cortex and
  676. its descending fibers has been thought not to cause remarkable
  677. deficits in simple locomotion in quadropedal animals. In the
  678. present study, we aimed to investigate in detail how lesion of
  679. the sensorimotor cortex affected locomotion ability in mice
  680. using the KinemaTracer system, a novel video-based kinematic
  681. analyzer. We found that traumatic injury to the left
  682. sensorimotor cortex induced several apparent deficits in the
  683. movement of contralesional right limbs during treadmill
  684. locomotion. The step length of right limbs decreased, and the
  685. speed in the forward direction was abrogated in the swing phase.
  686. The coordinates and angle of each joint were also changed after
  687. the injury. Some of the abnormal values in these parameters
  688. gradually recovered near the control level. The number of
  689. cFos-expressing neurons following locomotion significantly
  690. decreased in the right side of the spinal cord in injured mice,
  691. suggesting a role for cortex and descending fibers in
  692. locomotion. In contrast, interlimb coordination did not change
  693. remarkably even after the injury, supporting the notion that the
  694. basic locomotor pattern was determined by intraspinal neural
  695. circuits. These results indicate that the motor cortex and its
  696. descending fibers regulate several aspects of fine limb movement
  697. during locomotion. Our findings provide practical parameters to
  698. assess motor deficits and recovery following cortical injury in
  699. mice.",
  700. journal = "Exp. Neurol.",
  701. publisher = "Elsevier",
  702. volume = 230,
  703. number = 2,
  704. pages = "280--290",
  705. month = aug,
  706. year = 2011,
  707. keywords = "Locomotion;To Read",
  708. language = "en"
  709. }
  710. @ARTICLE{Holmes2006-fn,
  711. title = "The Dynamics of Legged Locomotion: Models, Analyses, and
  712. Challenges",
  713. author = "Holmes, Philip and Full, Robert J and Koditschek, Dan and
  714. Guckenheimer, John",
  715. abstract = "Cheetahs and beetles run, dolphins and salmon swim, and bees and
  716. birds fly with grace and economy surpassing our technology.
  717. Evolution has shaped the breathtaking abilities of animals,
  718. leaving us the challenge of reconstructing their targets of
  719. control and mechanisms of dexterity. In this review we explore a
  720. corner of this fascinating world. We describe mathematical
  721. models for legged animal locomotion, focusing on rapidly running
  722. insects and highlighting past achievements and challenges that
  723. remain. Newtonian body--limb dynamics are most naturally
  724. formulated as piecewise-holonomic rigid body mechanical systems,
  725. whose constraints change as legs touch down or lift off. Central
  726. pattern generators and proprioceptive sensing require models of
  727. spiking neurons and simplified phase oscillator descriptions of
  728. ensembles of them. A full neuromechanical model of a running
  729. animal requires integration of these elements, along with
  730. proprioceptive feedback and models of goal-oriented sensing,
  731. planning, and learning. We outline relevant background material
  732. from biomechanics and neurobiology, explain key properties of
  733. the hybrid dynamical systems that underlie legged locomotion
  734. models, and provide numerous examples of such models, from the
  735. simplest, completely soluble ``peg-leg walker'' to complex
  736. neuromuscular subsystems that are yet to be assembled into
  737. models of behaving animals. This final integration in a
  738. tractable and illuminating model is an outstanding challenge.",
  739. journal = "SIAM Rev.",
  740. publisher = "Society for Industrial and Applied Mathematics",
  741. volume = 48,
  742. number = 2,
  743. pages = "207--304",
  744. month = jan,
  745. year = 2006,
  746. keywords = "Locomotion;To Read"
  747. }
  748. @ARTICLE{Schwenkgrub2020-yl,
  749. title = "Deep imaging in the brainstem reveals functional heterogeneity
  750. in V2a neurons controlling locomotion",
  751. author = "Schwenkgrub, Joanna and Harrell, Evan R and Bathellier, Brice
  752. and Bouvier, Julien",
  753. abstract = "V2a neurons are a genetically defined cell class that forms a
  754. major excitatory descending pathway from the brainstem reticular
  755. formation to the spinal cord. Their activation has been linked
  756. to the termination of locomotor activity based on broad
  757. optogenetic manipulations. However, because of the difficulties
  758. involved in accessing brainstem structures for in vivo cell
  759. type-specific recordings, V2a neuron function has never been
  760. directly observed during natural behaviors. Here, we imaged the
  761. activity of V2a neurons using micro-endoscopy in freely moving
  762. mice. We find that as many as half of the V2a neurons are
  763. excited at locomotion arrest and with low reliability. Other V2a
  764. neurons are inhibited at locomotor arrests and/or activated
  765. during other behaviors such as locomotion initiation or
  766. stationary grooming. Our results establish that V2a neurons not
  767. only drive stops as suggested by bulk optogenetics but also are
  768. stratified into subpopulations that likely contribute to diverse
  769. motor patterns.",
  770. journal = "Sci Adv",
  771. publisher = "advances.sciencemag.org",
  772. volume = 6,
  773. number = 49,
  774. month = dec,
  775. year = 2020,
  776. keywords = "Locomotion",
  777. language = "en"
  778. }
  779. @ARTICLE{Lemieux2019-yc,
  780. title = "Glutamatergic neurons of the gigantocellular reticular nucleus
  781. shape locomotor pattern and rhythm in the freely behaving mouse",
  782. author = "Lemieux, Maxime and Bretzner, Frederic",
  783. abstract = "Because of their intermediate position between supraspinal
  784. locomotor centers and spinal circuits, gigantocellular reticular
  785. nucleus (GRN) neurons play a key role in motor command. However,
  786. the functional contribution of glutamatergic GRN neurons in
  787. initiating, maintaining, and stopping locomotion is still
  788. unclear. Combining electromyographic recordings with optogenetic
  789. manipulations in freely behaving mice, we investigate the
  790. functional contribution of glutamatergic brainstem neurons of
  791. the GRN to motor and locomotor activity. Short-pulse
  792. photostimulation of one side of the glutamatergic GRN did not
  793. elicit locomotion but evoked distinct motor responses in flexor
  794. and extensor muscles at rest and during locomotion.
  795. Glutamatergic GRN outputs to the spinal cord appear to be gated
  796. according to the spinal locomotor network state. Increasing the
  797. duration of photostimulation increased motor and postural tone
  798. at rest and reset locomotor rhythm during ongoing locomotion. In
  799. contrast, photoinhibition impaired locomotor pattern and rhythm.
  800. We conclude that unilateral activation of glutamatergic GRN
  801. neurons triggered motor activity and modified ongoing locomotor
  802. pattern and rhythm.",
  803. journal = "PLoS Biol.",
  804. publisher = "journals.plos.org",
  805. volume = 17,
  806. number = 4,
  807. pages = "e2003880",
  808. month = apr,
  809. year = 2019,
  810. keywords = "Locomotion",
  811. language = "en"
  812. }
  813. @ARTICLE{Karadimas2020-ub,
  814. title = "Sensory cortical control of movement",
  815. author = "Karadimas, Spyridon K and Satkunendrarajah, Kajana and
  816. Laliberte, Alex M and Ringuette, Dene and Weisspapir, Iliya and
  817. Li, Lijun and Gosgnach, Simon and Fehlings, Michael G",
  818. abstract = "Walking in our complex environment requires continual higher
  819. order integrated spatiotemporal information. This information is
  820. processed in the somatosensory cortex, and it has long been
  821. presumed that it influences movement via descending tracts
  822. originating from the motor cortex. Here we show that neuronal
  823. activity in the primary somatosensory cortex tightly correlates
  824. with the onset and speed of locomotion in freely moving mice.
  825. Using optogenetics and pharmacogenetics in combination with in
  826. vivo and in vitro electrophysiology, we provide evidence for a
  827. direct corticospinal pathway from the primary somatosensory
  828. cortex that synapses with cervical excitatory neurons and
  829. modulates the lumbar locomotor network independently of the
  830. motor cortex and other supraspinal locomotor centers.
  831. Stimulation of this pathway enhances speed of locomotion, while
  832. inhibition decreases locomotor speed and ultimately terminates
  833. stepping. Our findings reveal a novel pathway for neural control
  834. of movement whereby the somatosensory cortex directly influences
  835. motor behavior, possibly in response to environmental cues.",
  836. journal = "Nat. Neurosci.",
  837. publisher = "nature.com",
  838. volume = 23,
  839. number = 1,
  840. pages = "75--84",
  841. month = jan,
  842. year = 2020,
  843. keywords = "Locomotion",
  844. language = "en"
  845. }
  846. @ARTICLE{Bouvier2015-nm,
  847. title = "Descending Command Neurons in the Brainstem that Halt Locomotion",
  848. author = "Bouvier, Julien and Caggiano, Vittorio and Leiras, Roberto and
  849. Caldeira, Vanessa and Bellardita, Carmelo and Balueva, Kira and
  850. Fuchs, Andrea and Kiehn, Ole",
  851. abstract = "The episodic nature of locomotion is thought to be controlled by
  852. descending inputs from the brainstem. Most studies have largely
  853. attributed this control to initiating excitatory signals, but
  854. little is known about putative commands that may specifically
  855. determine locomotor offset. To link identifiable brainstem
  856. populations to a potential locomotor stop signal, we used
  857. developmental genetics and considered a discrete neuronal
  858. population in the reticular formation: the V2a neurons. We find
  859. that those neurons constitute a major excitatory pathway to
  860. locomotor areas of the ventral spinal cord. Selective activation
  861. of V2a neurons of the rostral medulla stops ongoing locomotor
  862. activity, owing to an inhibition of premotor locomotor networks
  863. in the spinal cord. Moreover, inactivation of such neurons
  864. decreases spontaneous stopping in vivo. Therefore, the V2a ``stop
  865. neurons'' represent a glutamatergic descending pathway that
  866. favors immobility and may thus help control the episodic nature
  867. of locomotion.",
  868. journal = "Cell",
  869. volume = 163,
  870. number = 5,
  871. pages = "1191--1203",
  872. month = nov,
  873. year = 2015,
  874. keywords = "Locomotion",
  875. language = "en"
  876. }
  877. @ARTICLE{Caggiano2018-td,
  878. title = "Midbrain circuits that set locomotor speed and gait selection",
  879. author = "Caggiano, V and Leiras, R and Go{\~n}i-Erro, H and Masini, D and
  880. Bellardita, C and Bouvier, J and Caldeira, V and Fisone, G and
  881. Kiehn, O",
  882. abstract = "Locomotion is a fundamental motor function common to the animal
  883. kingdom. It is implemented episodically and adapted to
  884. behavioural needs, including exploration, which requires slow
  885. locomotion, and escape behaviour, which necessitates faster
  886. speeds. The control of these functions originates in brainstem
  887. structures, although the neuronal substrate(s) that support them
  888. have not yet been elucidated. Here we show in mice that speed and
  889. gait selection are controlled by glutamatergic excitatory neurons
  890. (GlutNs) segregated in two distinct midbrain nuclei: the
  891. cuneiform nucleus (CnF) and the pedunculopontine nucleus (PPN).
  892. GlutNs in both of these regions contribute to the control of
  893. slower, alternating-gait locomotion, whereas only GlutNs in the
  894. CnF are able to elicit high-speed, synchronous-gait locomotion.
  895. Additionally, both the activation dynamics and the input and
  896. output connectivity matrices of GlutNs in the PPN and the CnF
  897. support explorative and escape locomotion, respectively. Our
  898. results identify two regions in the midbrain that act in
  899. conjunction to select context-dependent locomotor behaviours.",
  900. journal = "Nature",
  901. volume = 553,
  902. number = 7689,
  903. pages = "455--460",
  904. month = jan,
  905. year = 2018,
  906. keywords = "Locomotion",
  907. language = "en"
  908. }
  909. @ARTICLE{Usseglio2020-nl,
  910. title = "Control of Orienting Movements and Locomotion by
  911. {Projection-Defined} Subsets of Brainstem V2a Neurons",
  912. author = "Usseglio, Giovanni and Gatier, Edwin and Heuz{\'e}, Aur{\'e}lie
  913. and H{\'e}rent, Coralie and Bouvier, Julien",
  914. abstract = "Spatial orientation requires the execution of lateralized
  915. movements and a change in the animal's heading in response to
  916. multiple sensory modalities. While much research has focused on
  917. the circuits for sensory integration, chiefly to the midbrain
  918. superior colliculus (SC), the downstream cells and circuits that
  919. engage adequate motor actions have remained elusive. Furthermore,
  920. the mechanisms supporting trajectory changes are still
  921. speculative. Here, using transneuronal viral tracings in mice, we
  922. show that brainstem V2a neurons, a genetically defined subtype of
  923. glutamatergic neurons of the reticular formation, receive
  924. putative synaptic inputs from the contralateral SC. This makes
  925. them a candidate relay of lateralized orienting commands. We next
  926. show that unilateral optogenetic activations of brainstem V2a
  927. neurons in vivo evoked ipsilateral orienting-like responses of
  928. the head and the nose tip on stationary mice. When animals are
  929. walking, similar stimulations impose a transient locomotor arrest
  930. followed by a change of trajectory. Third, we reveal that these
  931. distinct motor actions are controlled by dedicated V2a subsets
  932. each projecting to a specific spinal cord segment, with at least
  933. (1) a lumbar-projecting subset whose unilateral activation
  934. specifically controls locomotor speed but neither impacts
  935. trajectory nor evokes orienting movements, and (2) a
  936. cervical-projecting subset dedicated to head orientation, but not
  937. to locomotor speed. Activating the latter subset suffices to
  938. steer the animals' directional heading, placing the head
  939. orientation as the prime driver of locomotor trajectory. V2a
  940. neurons and their modular organization may therefore underlie the
  941. orchestration of multiple motor actions during multi-faceted
  942. orienting behaviors.",
  943. journal = "Curr. Biol.",
  944. volume = 30,
  945. number = 23,
  946. pages = "4665--4681.e6",
  947. month = dec,
  948. year = 2020,
  949. keywords = "V2a neurons; brainstem; circuit tracings; locomotion; motor
  950. control; mouse; optogenetics; orientation; reticulospinal
  951. neurons; spinal cord;Locomotion",
  952. language = "en"
  953. }
  954. @ARTICLE{Taylor1982-pi,
  955. title = "Energetics and mechanics of terrestrial locomotion. I. Metabolic
  956. energy consumption as a function of speed and body size in birds
  957. and mammals",
  958. author = "Taylor, C R and Heglund, N C and Maloiy, G M",
  959. abstract = "This series of four papers investigates the link between the
  960. energetics and the mechanics of terrestrial locomotion. Two
  961. experimental variables are used throughout the study: speed and
  962. body size. Mass-specific metabolic rates of running animals can
  963. be varied by about tenfold using either variable. This first
  964. paper considers metabolic energy consumed during terrestrial
  965. locomotion. New data relating rate of oxygen consumption and
  966. speed are reported for: eight species of wild and domestic
  967. artiodactyls; seven species of carnivores; four species of
  968. primates; and one species of rodent. These are combined with
  969. previously published data to formulate a new allometric equation
  970. relating mass-specific rates of oxygen consumed (VO2/Mb) during
  971. locomotion at a constant speed to speed and body mass (based on
  972. data from 62 avian and mammalian species): VO2/Mb = 0.533
  973. Mb-0.316.vg + 0.300 Mb-0.303 where VO2/Mb has the units ml O2
  974. s-1 kg-1; Mb is in kg; and vg is in m s-1. This equation can be
  975. expressed in terms of mass-specific rates of energy consumption
  976. (Emetab/Mb) using the energetic equivalent of 1 ml O2 = 20.1 J
  977. because the contribution of anaerobic glycolysis was negligible:
  978. Emetab/Mb = 10.7 Mb-0.316.vg + 6.03 Mb-0.303 where Emetab/Mb has
  979. the units watts/kg. This new relationship applies equally well
  980. to bipeds and quadrupeds and differs little from the allometric
  981. equation reported 12 years ago by Taylor, Schmid-Nielsen \& Raab
  982. (1970). Ninety per cent of the values calculated from this
  983. genera equation for the diverse assortment of avian and
  984. mammalian species included in this regression fall within 25\%
  985. of the observed values at the middle of the speed range where
  986. measurements were made. This agreement is impressive when one
  987. considers that mass-specific rates of oxygen consumption
  988. differed by more than 1400\% over this size range of animals.",
  989. journal = "J. Exp. Biol.",
  990. publisher = "jeb.biologists.org",
  991. volume = 97,
  992. pages = "1--21",
  993. month = apr,
  994. year = 1982,
  995. keywords = "Locomotion",
  996. language = "en"
  997. }
  998. @INCOLLECTION{Matsuyama2004-fv,
  999. title = "Locomotor role of the corticoreticular--reticulospinal--spinal
  1000. interneuronal system",
  1001. booktitle = "Progress in Brain Research",
  1002. author = "Matsuyama, Kiyoji and Mori, Futoshi and Nakajima, Katsumi and
  1003. Drew, Trevor and Aoki, Mamoru and Mori, Shigemi",
  1004. abstract = "In vertebrates, the descending reticulospinal pathway is the
  1005. primary means of conveying locomotor command signals from higher
  1006. motor centers to spinal interneuronal circuits, the latter
  1007. including the central pattern generators for locomotion. The
  1008. pathway is morphologically heterogeneous, being composed of
  1009. various types of in-parallel-descending axons, which terminate
  1010. with different arborization patterns in the spinal cord. Such
  1011. morphology suggests that this pathway and its target spinal
  1012. interneurons comprise varying types of functional subunits,
  1013. which have a wide variety of functional roles, as dictated by
  1014. command signals from the higher motor centers. Corticoreticular
  1015. fibers are one of the major output pathways from the motor
  1016. cortex to the brainstem. They project widely and diffusely
  1017. within the pontomedullary reticular formation. Such a diffuse
  1018. projection pattern seems well suited to combining and
  1019. integrating the function of the various types of reticulospinal
  1020. neurons, which are widely scattered throughout the
  1021. pontomedullary reticular formation. The
  1022. corticoreticular--reticulospinal--spinal interneuronal
  1023. connections appear to operate as a cohesive, yet flexible,
  1024. control system for the elaboration of a wide variety of
  1025. movements, including those that combine goal-directed locomotion
  1026. with other motor actions.",
  1027. publisher = "Elsevier",
  1028. volume = 143,
  1029. pages = "239--249",
  1030. month = jan,
  1031. year = 2004,
  1032. keywords = "Locomotion"
  1033. }
  1034. % The entry below contains non-ASCII chars that could not be converted
  1035. % to a LaTeX equivalent.
  1036. @INCOLLECTION{Velagic2008-nb,
  1037. title = "Nonlinear motion control of mobile robot dynamic model",
  1038. booktitle = "Motion planning",
  1039. author = "Velagic, Jasmin and Lacevic, Bakir and Osmic, Nedim",
  1040. abstract = "The problem of motion planning and control of mobile robots has
  1041. attracted the interest of researchers in view of its theoretical
  1042. challenges because of their obvious relevance in applications.
  1043. From a control viewpoint, the peculiar nature of nonholonomic
  1044. kinematics and dynamic complexity of the mobile robot makes that
  1045. feedback stabilization at a given posture cannot be achieved via
  1046. smooth time-invariant control (Oriolo et al., 2002). This
  1047. indicates that the problem is truly nonlinear; linear control is
  1048. ineffective, and innovative design techniques …",
  1049. publisher = "IntechOpen",
  1050. year = 2008,
  1051. keywords = "control"
  1052. }
  1053. @UNPUBLISHED{Kao2020-dl,
  1054. title = "Optimal anticipatory control as a theory of motor preparation: a
  1055. thalamo-cortical circuit model",
  1056. author = "Kao, Ta-Chu and Sadabadi, Mahdieh S and Hennequin, Guillaume",
  1057. abstract = "Summary Across a range of motor and cognitive tasks, cortical
  1058. activity can be accurately described by low-dimensional dynamics
  1059. unfolding from specific initial conditions on every trial. These
  1060. ``preparatory states'' largely determine the subsequent evolution
  1061. of both neural activity and behaviour, and their importance
  1062. raises questions regarding how they are --- or ought to be ---
  1063. set. Here, we formulate motor preparation as optimal prospective
  1064. control of future movements. The solution is a form of internal
  1065. control of cortical circuit dynamics, which can be implemented as
  1066. a thalamo-cortical loop gated by the basal ganglia. Critically,
  1067. optimal control predicts selective quenching of variability in
  1068. components of preparatory population activity that have future
  1069. motor consequences, but not in others. This is consistent with
  1070. recent perturbation experiments performed in mice, and with our
  1071. novel analysis of monkey motor cortex activity during reaching.
  1072. Together, these results suggest optimal anticipatory control of
  1073. movement.",
  1074. journal = "Cold Spring Harbor Laboratory",
  1075. pages = "2020.02.02.931246",
  1076. month = feb,
  1077. year = 2020,
  1078. keywords = "control",
  1079. language = "en"
  1080. }
  1081. @ARTICLE{Marshall2020-rp,
  1082. title = "Continuous {Whole-Body} {3D} Kinematic Recordings across the
  1083. Rodent Behavioral Repertoire",
  1084. author = "Marshall, Jesse D and Aldarondo, Diego E and Dunn, Timothy W and
  1085. Wang, William L and Berman, Gordon J and {\"O}lveczky, Bence P",
  1086. abstract = "In mammalian animal models, high-resolution kinematic tracking is
  1087. restricted to brief sessions in constrained environments,
  1088. limiting our ability to probe naturalistic behaviors and their
  1089. neural underpinnings. To address this, we developed CAPTURE
  1090. (Continuous Appendicular and Postural Tracking Using
  1091. Retroreflector Embedding), a behavioral monitoring system that
  1092. combines motion capture and deep learning to continuously track
  1093. the 3D kinematics of a rat's head, trunk, and limbs for week-long
  1094. timescales in freely behaving animals. CAPTURE realizes 10- to
  1095. 100-fold gains in precision and robustness compared with existing
  1096. convolutional network approaches to behavioral tracking. We
  1097. demonstrate CAPTURE's ability to comprehensively profile the
  1098. kinematics and sequential organization of natural rodent
  1099. behavior, its variation across individuals, and its perturbation
  1100. by drugs and disease, including identifying perseverative
  1101. grooming states in a rat model of fragile X syndrome. CAPTURE
  1102. significantly expands the range of behaviors and contexts that
  1103. can be quantitatively investigated, opening the door to a new
  1104. understanding of natural behavior and its neural basis.",
  1105. journal = "Neuron",
  1106. month = dec,
  1107. year = 2020,
  1108. keywords = "animal tracking; autism; behavior; computational ethology;
  1109. grooming; individuality; motion capture; phenotyping",
  1110. language = "en"
  1111. }
  1112. @ARTICLE{Simmons2009-ax,
  1113. title = "Comparing histological data from different brains: sources of
  1114. error and strategies for minimizing them",
  1115. author = "Simmons, Donna M and Swanson, Larry W",
  1116. abstract = "The recent development of brain atlases with computer graphics
  1117. templates, and of huge databases of neurohistochemical data on
  1118. the internet, has forced a systematic re-examination of errors
  1119. associated with comparing histological features between adjacent
  1120. sections of the same brain, between brains treated in the same
  1121. way, and between brains from groups treated in different ways.
  1122. The long-term goal is to compare as accurately as possible a
  1123. broad array of data from experimental brains within the
  1124. framework of reference atlases. Main sources of error, each of
  1125. which ideally should be measured and minimized, include
  1126. intrinsic biological variation, linear and nonlinear distortion
  1127. of histological sections, plane of section differences between
  1128. each brain, section alignment problems, and sampling errors.
  1129. These variables are discussed, along with approaches to error
  1130. estimation and minimization in terms of a specific example-the
  1131. distribution of neuroendocrine neurons in the rat
  1132. paraventricular nucleus. Based on the strategy developed here,
  1133. the main conclusion is that the best long-term solution is a
  1134. high-resolution 3D computer graphics model of the brain that can
  1135. be sliced in any plane and used as the framework for
  1136. quantitative neuroanatomy, databases, knowledge management
  1137. systems, and structure-function modeling. However, any approach
  1138. to the automatic annotation of neuroanatomical data-relating its
  1139. spatial distribution to a reference atlas-should deal
  1140. systematically with these sources of error, which reduce
  1141. localization reliability.",
  1142. journal = "Brain Res. Rev.",
  1143. publisher = "Elsevier",
  1144. volume = 60,
  1145. number = 2,
  1146. pages = "349--367",
  1147. month = may,
  1148. year = 2009,
  1149. language = "en"
  1150. }
  1151. @ARTICLE{Hahn2020-rj,
  1152. title = "An open access mouse brain flatmap and upgraded rat and human
  1153. brain flatmaps based on current reference atlases",
  1154. author = "Hahn, Joel D and Swanson, Larry W and Bowman, Ian and Foster,
  1155. Nicholas N and Zingg, Brian and Bienkowski, Michael S and
  1156. Hintiryan, Houri and Dong, Hong-Wei",
  1157. abstract = "Here we present a flatmap of the mouse central nervous system
  1158. (CNS) (brain) and substantially enhanced flatmaps of the rat and
  1159. human brain. Also included are enhanced representations of
  1160. nervous system white matter tracts, ganglia, and nerves, and an
  1161. enhanced series of 10 flatmaps showing different stages of rat
  1162. brain development. The adult mouse and rat brain flatmaps
  1163. provide layered diagrammatic representation of CNS divisions,
  1164. according to their arrangement in corresponding reference
  1165. atlases: Brain Maps 4.0 (BM4, rat) (Swanson, The Journal of
  1166. Comparative Neurology, 2018, 526, 935-943), and the first
  1167. version of the Allen Reference Atlas (mouse) (Dong, The Allen
  1168. reference atlas, (book + CD-ROM): A digital color brain atlas of
  1169. the C57BL/6J male mouse, 2007). To facilitate comparative
  1170. analysis, both flatmaps are scaled equally, and the divisional
  1171. hierarchy of gray matter follows a topographic arrangement used
  1172. in BM4. Also included with the mouse and rat brain flatmaps are
  1173. cerebral cortex atlas level contours based on the reference
  1174. atlases, and direct graphical and tabular comparison of regional
  1175. parcellation. To encourage use of the brain flatmaps, they were
  1176. designed and organized, with supporting reference tables, for
  1177. ease-of-use and to be amenable to computational applications. We
  1178. demonstrate how they can be adapted to represent novel
  1179. parcellations resulting from experimental data, and we provide a
  1180. proof-of-concept for how they could form the basis of a
  1181. web-based graphical data viewer and analysis platform. The
  1182. mouse, rat, and human brain flatmap vector graphics files (Adobe
  1183. Reader/Acrobat viewable and Adobe Illustrator editable) and
  1184. supporting tables are provided open access; they constitute a
  1185. broadly applicable neuroscience toolbox resource for researchers
  1186. seeking to map and perform comparative analysis of brain data.",
  1187. journal = "J. Comp. Neurol.",
  1188. publisher = "Wiley",
  1189. number = "cne.24966",
  1190. month = jun,
  1191. year = 2020,
  1192. keywords = "brain atlases; brain flatmap; brain mapping; computer graphics;
  1193. human; mouse; rat",
  1194. copyright = "http://onlinelibrary.wiley.com/termsAndConditions\#vor",
  1195. language = "en"
  1196. }
  1197. @ARTICLE{Maheswaranathan2020-fy,
  1198. title = "How recurrent networks implement contextual processing in
  1199. sentiment analysis",
  1200. author = "Maheswaranathan, Niru and Sussillo, David",
  1201. abstract = "Neural networks have a remarkable capacity for contextual
  1202. processing--using recent or nearby inputs to modify
  1203. processing of current input. For example, in natural
  1204. language, contextual processing is necessary to correctly
  1205. interpret negation (e.g. phrases such as ``not bad'').
  1206. However, our ability to understand how networks process
  1207. context is limited. Here, we propose general methods for
  1208. reverse engineering recurrent neural networks (RNNs) to
  1209. identify and elucidate contextual processing. We apply these
  1210. methods to understand RNNs trained on sentiment
  1211. classification. This analysis reveals inputs that induce
  1212. contextual effects, quantifies the strength and timescale of
  1213. these effects, and identifies sets of these inputs with
  1214. similar properties. Additionally, we analyze contextual
  1215. effects related to differential processing of the beginning
  1216. and end of documents. Using the insights learned from the
  1217. RNNs we improve baseline Bag-of-Words models with simple
  1218. extensions that incorporate contextual modification,
  1219. recovering greater than 90\% of the RNN's performance
  1220. increase over the baseline. This work yields a new
  1221. understanding of how RNNs process contextual information,
  1222. and provides tools that should provide similar insight more
  1223. broadly.",
  1224. month = apr,
  1225. year = 2020,
  1226. keywords = "RNN;RNN To read",
  1227. archivePrefix = "arXiv",
  1228. primaryClass = "cs.CL",
  1229. eprint = "2004.08013"
  1230. }
  1231. @ARTICLE{Madhav2020-qs,
  1232. title = "The Synergy Between Neuroscience and Control Theory: The Nervous
  1233. System as Inspiration for Hard Control Challenges",
  1234. author = "Madhav, Manu S and Cowan, Noah J",
  1235. abstract = "Here, we review the role of control theory in modeling neural
  1236. control systems through a top-down analysis approach.
  1237. Specifically, we examine the role of the brain and central
  1238. nervous system as the controller in the organism, connected to
  1239. but isolated from the rest of the animal through insulated
  1240. interfaces. Though biological and engineering control systems
  1241. operate on similar principles, they differ in several critical
  1242. features, which makes drawing inspiration from biology for
  1243. engineering controllers challenging but worthwhile. We also
  1244. outline a procedure that the control theorist can use to draw
  1245. inspiration from the biological controller: starting from the
  1246. intact, behaving animal; designing experiments to deconstruct
  1247. and model hierarchies of feedback; modifying feedback
  1248. topologies; perturbing inputs and plant dynamics; using the
  1249. resultant outputs to perform system identification; and tuning
  1250. and validating the resultant control-theoretic model using
  1251. specially engineered robophysical models.",
  1252. journal = "Annu. Rev. Control Robot. Auton. Syst.",
  1253. publisher = "Annual Reviews",
  1254. volume = 3,
  1255. number = 1,
  1256. pages = "243--267",
  1257. month = may,
  1258. year = 2020
  1259. }
  1260. @ARTICLE{Fieseler2020-ne,
  1261. title = "Unsupervised learning of control signals and their encodings
  1262. in $\textit{C. elegans}$ whole-brain recordings",
  1263. author = "Fieseler, Charles and Zimmer, Manuel and Nathan Kutz, J",
  1264. abstract = "Recent whole brain imaging experiments on $\textit\{C.
  1265. elegans\}$ has revealed that the neural population dynamics
  1266. encode motor commands and stereotyped transitions between
  1267. behaviors on low dimensional manifolds. Efforts to
  1268. characterize the dynamics on this manifold have used
  1269. piecewise linear models to describe the entire state space,
  1270. but it is unknown how a single, global dynamical model can
  1271. generate the observed dynamics. Here, we propose a control
  1272. framework to achieve such a global model of the dynamics,
  1273. whereby underlying linear dynamics is actuated by sparse
  1274. control signals. This method learns the control signals in
  1275. an unsupervised way from data, then uses $\textit\{ Dynamic
  1276. Mode Decomposition with control\}$ (DMDc) to create the
  1277. first global, linear dynamical system that can reconstruct
  1278. whole-brain imaging data. These control signals are shown to
  1279. be implicated in transitions between behaviors. In addition,
  1280. we analyze the time-delay encoding of these control signals,
  1281. showing that these transitions can be predicted from neurons
  1282. previously implicated in behavioral transitions, but also
  1283. additional neurons previously unidentified. Moreover, our
  1284. decomposition method allows one to understand the observed
  1285. nonlinear global dynamics instead as linear dynamics with
  1286. control. The proposed mathematical framework is generic and
  1287. can be generalized to other neurosensory systems,
  1288. potentially revealing transitions and their encodings in a
  1289. completely unsupervised way.",
  1290. month = jan,
  1291. year = 2020,
  1292. archivePrefix = "arXiv",
  1293. primaryClass = "q-bio.QM",
  1294. eprint = "2001.08346"
  1295. }
  1296. @ARTICLE{Jude2020-mq,
  1297. title = "Hippocampal representations emerge when training recurrent
  1298. neural networks on a memory dependent maze navigation task",
  1299. author = "Jude, Justin and Hennig, Matthias H",
  1300. abstract = "Can neural networks learn goal-directed behaviour using
  1301. similar strategies to the brain, by combining the
  1302. relationships between the current state of the organism and
  1303. the consequences of future actions? Recent work has shown
  1304. that recurrent neural networks trained on goal based tasks
  1305. can develop representations resembling those found in the
  1306. brain, entorhinal cortex grid cells, for instance. Here we
  1307. explore the evolution of the dynamics of their internal
  1308. representations and compare this with experimental data. We
  1309. observe that once a recurrent network is trained to learn
  1310. the structure of its environment solely based on sensory
  1311. prediction, an attractor based landscape forms in the
  1312. network's representation, which parallels hippocampal place
  1313. cells in structure and function. Next, we extend the
  1314. predictive objective to include Q-learning for a reward
  1315. task, where rewarding actions are dependent on delayed cue
  1316. modulation. Mirroring experimental findings in hippocampus
  1317. recordings in rodents performing the same task, this
  1318. training paradigm causes nonlocal neural activity to sweep
  1319. forward in space at decision points, anticipating the future
  1320. path to a rewarded location. Moreover, prevalent choice and
  1321. cue-selective neurons form in this network, again
  1322. recapitulating experimental findings. Together, these
  1323. results indicate that combining predictive, unsupervised
  1324. learning of the structure of an environment with
  1325. reinforcement learning can help understand the formation of
  1326. hippocampus-like representations containing both spatial and
  1327. task-relevant information.",
  1328. month = dec,
  1329. year = 2020,
  1330. keywords = "RNN;RNN To read",
  1331. archivePrefix = "arXiv",
  1332. primaryClass = "q-bio.NC",
  1333. eprint = "2012.01328"
  1334. }
  1335. @UNPUBLISHED{Schaeffer2020-qv,
  1336. title = "Reverse-engineering Recurrent Neural Network solutions to a
  1337. hierarchical inference task for mice",
  1338. author = "Schaeffer, Rylan and Khona, Mikail and Meshulam, Leenoy and
  1339. {International Brain Laboratory} and Fiete, Ila Rani",
  1340. abstract = "We study how recurrent neural networks (RNNs) solve a
  1341. hierarchical inference task involving two latent variables and
  1342. disparate timescales separated by 1-2 orders of magnitude. The
  1343. task is of interest to the International Brain Laboratory, a
  1344. global collaboration of experimental and theoretical
  1345. neuroscientists studying how the mammalian brain generates
  1346. behavior. We make four discoveries. First, RNNs learn behavior
  1347. that is quantitatively similar to ideal Bayesian baselines.
  1348. Second, RNNs perform inference by learning a two-dimensional
  1349. subspace defining beliefs about the latent variables. Third, the
  1350. geometry of RNN dynamics reflects an induced coupling between the
  1351. two separate inference processes necessary to solve the task.
  1352. Fourth, we perform model compression through a novel form of
  1353. knowledge distillation on hidden representations --
  1354. Representations and Dynamics Distillation (RADD)-- to reduce the
  1355. RNN dynamics to a low-dimensional, highly interpretable model.
  1356. This technique promises a useful tool for interpretability of
  1357. high dimensional nonlinear dynamical systems. Altogether, this
  1358. work yields predictions to guide exploration and analysis of
  1359. mouse neural data and circuity. \#\#\# Competing Interest
  1360. Statement The authors have declared no competing interest.",
  1361. journal = "Cold Spring Harbor Laboratory",
  1362. pages = "2020.06.09.142745",
  1363. month = jun,
  1364. year = 2020,
  1365. keywords = "RNN",
  1366. language = "en"
  1367. }
  1368. % The entry below contains non-ASCII chars that could not be converted
  1369. % to a LaTeX equivalent.
  1370. @UNPUBLISHED{Van_der_Zouwen2020-zn,
  1371. title = "Freely behaving mice can brake and turn during optogenetic
  1372. stimulation of the Mesencephalic Locomotor Region",
  1373. author = "van der Zouwen, Cornelis Immanuel and Boutin, Jo{\"e}l and
  1374. Foug{\`e}re, Maxime and Flaive, Aur{\'e}lie and Vivancos,
  1375. M{\'e}lanie and Santuz, Alessandro and Akay, Turgay and Sarret,
  1376. Philippe and Ryczko, Dimitri",
  1377. abstract = "Background Stimulation of the Mesencephalic Locomotor Region (
  1378. MLR ) is increasingly considered as a target to improve locomotor
  1379. function in Parkinson's disease, spinal cord injury and stroke. A
  1380. key function of the MLR is to control the speed of forward
  1381. symmetrical locomotor movements. However, the ability of freely
  1382. moving mammals to integrate environmental cues to brake and turn
  1383. during MLR stimulation is poorly documented. Objective/hypothesis
  1384. We investigated whether freely behaving mice could brake or turn
  1385. based on environmental cues during MLR stimulation. Methods We
  1386. stimulated the cuneiform nucleus in mice expressing
  1387. channelrhodopsin in Vglut2-positive neurons in a Cre-dependent
  1388. manner (Vglut2-ChR2-EYFP) using optogenetics. We detected
  1389. locomotor movements using deep learning. We used patch-clamp
  1390. recordings to validate the functional expression of
  1391. channelrhodopsin and neuroanatomy to visualize the stimulation
  1392. sites. Results Optogenetic stimulation of the MLR evoked
  1393. locomotion and increasing laser power increased locomotor speed.
  1394. Gait diagram and limb kinematics were similar during spontaneous
  1395. and optogenetic-evoked locomotion. Mice could brake and make
  1396. sharp turns (∼90⁰) when approaching a corner during MLR
  1397. stimulation in an open-field arena. The speed during the turn was
  1398. scaled with the speed before the turn, and with the turn angle.
  1399. In a reporter mouse, many Vglut2-ZsGreen neurons were
  1400. immunopositive for glutamate in the MLR. Patch-clamp recordings
  1401. in Vglut2-ChR2-EYFP mice show that blue light evoked short
  1402. latency spiking in MLR neurons. Conclusion MLR glutamatergic
  1403. neurons are a relevant target to improve locomotor activity
  1404. without impeding the ability to brake and turn when approaching
  1405. an obstacle, thus ensuring smooth and adaptable navigation.
  1406. Highlights \#\#\# Competing Interest Statement The authors have
  1407. declared no competing interest.",
  1408. journal = "Cold Spring Harbor Laboratory",
  1409. pages = "2020.11.30.404525",
  1410. month = dec,
  1411. year = 2020,
  1412. keywords = "Locomotion",
  1413. language = "en"
  1414. }
  1415. @UNPUBLISHED{Harris2020-im,
  1416. title = "Nonsense correlations in neuroscience",
  1417. author = "Harris, Kenneth D",
  1418. abstract = "Most neurophysiological signals exhibit slow continuous trends
  1419. over time. Because standard correlation analyses assume that all
  1420. samples are independent, they can yield apparently significant
  1421. ``nonsense correlations'' even for signals that are completely
  1422. unrelated. Here we compare the performance of several methods for
  1423. assessing correlations between timeseries, using simulated slowly
  1424. drifting signals with and without genuine correlations. The best
  1425. performance was obtained from a ``pseudosession method'', which
  1426. relies on one of the signals being randomly generated by the
  1427. experimenter, or a ``session perturbation'' method which requires
  1428. multiple recordings under the same conditions. If neither of
  1429. these is applicable, we find that a ``linear shift method can
  1430. work well, but only when one of the signals is stationary.
  1431. Methods based on cross-validation, circular shifting, phase
  1432. randomization, or detrending gave up to 100\% false positive
  1433. rates in our simulations. We conclude that analysis of neural
  1434. timeseries is best performed when stationarity and randomization
  1435. is built into the experimental design. \#\#\# Competing Interest
  1436. Statement The authors have declared no competing interest.",
  1437. journal = "Cold Spring Harbor Laboratory",
  1438. pages = "2020.11.29.402719",
  1439. month = nov,
  1440. year = 2020,
  1441. language = "en"
  1442. }
  1443. @UNPUBLISHED{Michaels2020-ut,
  1444. title = "A modular neural network model of grasp movement generation",
  1445. author = "Michaels, Jonathan A and Schaffelhofer, Stefan and Agudelo-Toro,
  1446. Andres and Scherberger, Hansj{\"o}rg",
  1447. abstract = "Summary One of the primary ways we interact with the world is
  1448. using our hands. In macaques, the circuit spanning the anterior
  1449. intraparietal area, the hand area of the ventral premotor cortex,
  1450. and the primary motor cortex is necessary for transforming visual
  1451. information into grasping movements. We hypothesized that a
  1452. recurrent neural network mimicking the multi-area structure of
  1453. the anatomical circuit and using visual features to generate the
  1454. required muscle dynamics to grasp objects would explain the
  1455. neural and computational basis of the grasping circuit. Modular
  1456. networks with object feature input and sparse inter-module
  1457. connectivity outperformed other models at explaining neural data
  1458. and the inter-area relationships present in the biological
  1459. circuit, despite the absence of neural data during network
  1460. training. Network dynamics were governed by simple rules, and
  1461. targeted lesioning of modules produced deficits similar to those
  1462. observed in lesion studies, providing a potential explanation for
  1463. how grasping movements are generated.",
  1464. journal = "Cold Spring Harbor Laboratory",
  1465. pages = "742189",
  1466. month = feb,
  1467. year = 2020,
  1468. keywords = "RNN;RNN To read;To Read",
  1469. language = "en"
  1470. }
  1471. @ARTICLE{Harris2020-oh,
  1472. title = "Array programming with {NumPy}",
  1473. author = "Harris, Charles R and Millman, K Jarrod and van der Walt,
  1474. St{\'e}fan J and Gommers, Ralf and Virtanen, Pauli and
  1475. Cournapeau, David and Wieser, Eric and Taylor, Julian and Berg,
  1476. Sebastian and Smith, Nathaniel J and Kern, Robert and Picus,
  1477. Matti and Hoyer, Stephan and van Kerkwijk, Marten H and Brett,
  1478. Matthew and Haldane, Allan and Del R{\'\i}o, Jaime Fern{\'a}ndez
  1479. and Wiebe, Mark and Peterson, Pearu and G{\'e}rard-Marchant,
  1480. Pierre and Sheppard, Kevin and Reddy, Tyler and Weckesser,
  1481. Warren and Abbasi, Hameer and Gohlke, Christoph and Oliphant,
  1482. Travis E",
  1483. abstract = "Array programming provides a powerful, compact and expressive
  1484. syntax for accessing, manipulating and operating on data in
  1485. vectors, matrices and higher-dimensional arrays. NumPy is the
  1486. primary array programming library for the Python language. It
  1487. has an essential role in research analysis pipelines in fields
  1488. as diverse as physics, chemistry, astronomy, geoscience,
  1489. biology, psychology, materials science, engineering, finance and
  1490. economics. For example, in astronomy, NumPy was an important
  1491. part of the software stack used in the discovery of
  1492. gravitational waves1 and in the first imaging of a black hole2.
  1493. Here we review how a few fundamental array concepts lead to a
  1494. simple and powerful programming paradigm for organizing,
  1495. exploring and analysing scientific data. NumPy is the foundation
  1496. upon which the scientific Python ecosystem is constructed. It is
  1497. so pervasive that several projects, targeting audiences with
  1498. specialized needs, have developed their own NumPy-like
  1499. interfaces and array objects. Owing to its central position in
  1500. the ecosystem, NumPy increasingly acts as an interoperability
  1501. layer between such array computation libraries and, together
  1502. with its application programming interface (API), provides a
  1503. flexible framework to support the next decade of scientific and
  1504. industrial analysis.",
  1505. journal = "Nature",
  1506. publisher = "nature.com",
  1507. volume = 585,
  1508. number = 7825,
  1509. pages = "357--362",
  1510. month = sep,
  1511. year = 2020,
  1512. language = "en"
  1513. }
  1514. @UNPUBLISHED{BRAIN_Initiative_Cell_Census_Network_BICCN2020-cp,
  1515. title = "A multimodal cell census and atlas of the mammalian primary motor
  1516. cortex",
  1517. author = "{BRAIN Initiative Cell Census Network (BICCN)} and Adkins, Ricky
  1518. S and Aldridge, Andrew I and Allen, Shona and Ament, Seth A and
  1519. An, Xu and Armand, Ethan and Ascoli, Giorgio A and Bakken, Trygve
  1520. E and Bandrowski, Anita and Banerjee, Samik and Barkas, Nikolaos
  1521. and Bartlett, Anna and Bateup, Helen S and Margarita Behrens, M
  1522. and Berens, Philipp and Berg, Jim and Bernabucci, Matteo and
  1523. Bernaerts, Yves and Bertagnolli, Darren and Biancalani, Tommaso
  1524. and Boggeman, Lara and Sina Booeshaghi, A and Bowman, Ian and
  1525. Bravo, H{\'e}ctor Corrada and Cadwell, Cathryn Ren{\'e} and
  1526. Callaway, Edward M and Carlin, Benjamin and O'Connor, Carolyn and
  1527. Carter, Robert and Casper, Tamara and Castanon, Rosa G and
  1528. Castro, Jesus Ramon and Chance, Rebecca K and Chatterjee, Apaala
  1529. and Chen, Huaming and Chun, Jerold and Colantuoni, Carlo and
  1530. Crabtree, Jonathan and Creasy, Heather and Crichton, Kirsten and
  1531. Crow, Megan and D'Orazi, Florence D and Daigle, Tanya L and
  1532. Dalley, Rachel and Dee, Nick and Degatano, Kylee and Dichter,
  1533. Benjamin and Diep, Dinh and Ding, Liya and Ding, Song-Lin and
  1534. Dominguez, Bertha and Dong, Hong-Wei and Dong, Weixiu and
  1535. Dougherty, Elizabeth L and Dudoit, Sandrine and Ecker, Joseph R
  1536. and Eichhorn, Stephen W and Fang, Rongxin and Felix, Victor and
  1537. Feng, Guoping and Feng, Zhao and Fischer, Stephan and
  1538. Fitzpatrick, Conor and Fong, Olivia and Foster, Nicholas N and
  1539. Galbavy, William and Gee, James C and Ghosh, Satrajit S and
  1540. Giglio, Michelle and Gillespie, Thomas H and Gillis, Jesse and
  1541. Goldman, Melissa and Goldy, Jeff and Gong, Hui and Gou, Lin and
  1542. Grauer, Michael and Halchenko, Yaroslav O and Harris, Julie A and
  1543. Hartmanis, Leonard and Hatfield, Joshua T and Hawrylycz, Mike and
  1544. Helba, Brian and Herb, Brian R and Hertzano, Ronna and Hintiryan,
  1545. Houri and Hirokawa, Karla E and Hockemeyer, Dirk and Hodge,
  1546. Rebecca D and Hood, Greg and Horwitz, Gregory D and Hou, Xiaomeng
  1547. and Hu, Lijuan and Hu, Qiwen and Josh Huang, Z and Huo, Bingxing
  1548. and Ito-Cole, Tony and Jacobs, Matthew and Jia, Xueyan and Jiang,
  1549. Shengdian and Jiang, Tao and Jiang, Xiaolong and Jin, Xin and
  1550. Jorstad, Nikolas L and Kalmbach, Brian E and Kancherla, Jayaram
  1551. and Dirk Keene, C and Kelly, Kathleen and Khajouei, Farzaneh and
  1552. Kharchenko, Peter V and Kim, Gukhan and Ko, Andrew L and Kobak,
  1553. Dmitry and Konwar, Kishori and Kramer, Daniel J and Krienen,
  1554. Fenna M and Kroll, Matthew and Kuang, Xiuli and Kuo, Hsien-Chi
  1555. and Lake, Blue B and Larsen, Rachael and Lathia, Kanan and
  1556. Laturnus, Sophie and Lee, Angus Y and Lee, Cheng-Ta and Lee,
  1557. Kuo-Fen and Lein, Ed S and Lesnar, Phil and Li, Anan and Li,
  1558. Xiangning and Li, Xu and Li, Yang Eric and Li, Yaoyao and Li,
  1559. Yuanyuan and Lim, Byungkook and Linnarsson, Sten and Liu,
  1560. Christine S and Liu, Hanqing and Liu, Lijuan and Lucero, Jacinta
  1561. D and Luo, Chongyuan and Luo, Qingming and Macosko, Evan Z and
  1562. Mahurkar, Anup and Martone, Maryann E and Matho, Katherine S and
  1563. McCarroll, Steven A and McCracken, Carrie and McMillen, Delissa
  1564. and Miranda, Elanine and Mitra, Partha P and Miyazaki, Paula
  1565. Assakura and Mizrachi, Judith and Mok, Stephanie and Mukamel,
  1566. Eran A and Mulherkar, Shalaka and Nadaf, Naeem M and Naeemi,
  1567. Maitham and Narasimhan, Arun and Nery, Joseph R and Ng, Lydia and
  1568. Ngai, John and Nguyen, Thuc Nghi and Nickel, Lance and Nicovich,
  1569. Philip R and Niu, Sheng-Yong and Ntranos, Vasilis and Nunn,
  1570. Michael and Olley, Dustin and Orvis, Joshua and Osteen, Julia K
  1571. and Osten, Pavel and Owen, Scott F and Pachter, Lior and
  1572. Palaniswamy, Ramesh and Palmer, Carter R and Pang, Yan and Peng,
  1573. Hanchuan and Pham, Thanh and Pinto-Duarte, Antonio and
  1574. Plongthongkum, Nongluk and Poirion, Olivier and Preissl,
  1575. Sebastian and Purdom, Elizabeth and Qu, Lei and Rashid, Mohammad
  1576. and Reed, Nora M and Regev, Aviv and Ren, Bing and Ren, Miao and
  1577. Rimorin, Christine and Risso, Davide and Rivkin, Angeline C and
  1578. Mu{\~n}oz-Casta{\~n}eda, Rodrigo and Romanow, William J and
  1579. Ropelewski, Alexander J and de B{\'e}zieux, Hector Roux and Ruan,
  1580. Zongcai and Sandberg, Rickard and Savoia, Steven and Scala,
  1581. Federico and Schor, Michael and Shen, Elise and Siletti, Kimberly
  1582. and Smith, Jared B and Smith, Kimberly and Somasundaram, Saroja
  1583. and Song, Yuanyuan and Sorensen, Staci A and Stafford, David A
  1584. and Street, Kelly and Sulc, Josef and Sunkin, Susan and Svensson,
  1585. Valentine and Tan, Pengcheng and Tan, Zheng Huan and Tasic,
  1586. Bosiljka and Thompson, Carol and Tian, Wei and Tickle, Timothy L
  1587. and Tieu, Michael and Ting, Jonathan T and Tolias, Andreas Savas
  1588. and Torkelson, Amy and Tung, Herman and Vaishnav, Eeshit Dhaval
  1589. and Van den Berge, Koen and van Velthoven, Cindy T J and
  1590. Vanderburg, Charles R and Veldman, Matthew B and Vu, Minh and
  1591. Wakeman, Wayne and Wang, Peng and Wang, Quanxin and Wang, Xinxin
  1592. and Wang, Yimin and Wang, Yun and Welch, Joshua D and White, Owen
  1593. and Williams, Elora and Xie, Fangming and Xie, Peng and Xiong,
  1594. Feng and William Yang, X and Yanny, Anna Marie and Yao, Zizhen
  1595. and Yin, Lulu and Yu, Yang and Yuan, Jing and Zeng, Hongkui and
  1596. Zhang, Kun and Zhang, Meng and Zhang, Zhuzhu and Zhao, Sujun and
  1597. Zhao, Xuan and Zhou, Jingtian and Zhuang, Xiaowei and Zingg,
  1598. Brian",
  1599. abstract = "We report the generation of a multimodal cell census and atlas of
  1600. the mammalian primary motor cortex (MOp or M1) as the initial
  1601. product of the BRAIN Initiative Cell Census Network (BICCN). This
  1602. was achieved by coordinated large-scale analyses of single-cell
  1603. transcriptomes, chromatin accessibility, DNA methylomes,
  1604. spatially resolved single-cell transcriptomes, morphological and
  1605. electrophysiological properties, and cellular resolution
  1606. input-output mapping, integrated through cross-modal
  1607. computational analysis. Together, our results advance the
  1608. collective knowledge and understanding of brain cell type
  1609. organization: First, our study reveals a unified molecular
  1610. genetic landscape of cortical cell types that congruently
  1611. integrates their transcriptome, open chromatin and DNA
  1612. methylation maps. Second, cross-species analysis achieves a
  1613. unified taxonomy of transcriptomic types and their hierarchical
  1614. organization that are conserved from mouse to marmoset and human.
  1615. Third, cross-modal analysis provides compelling evidence for the
  1616. epigenomic, transcriptomic, and gene regulatory basis of neuronal
  1617. phenotypes such as their physiological and anatomical properties,
  1618. demonstrating the biological validity and genomic underpinning of
  1619. neuron types and subtypes. Fourth, in situ single-cell
  1620. transcriptomics provides a spatially-resolved cell type atlas of
  1621. the motor cortex. Fifth, integrated transcriptomic, epigenomic
  1622. and anatomical analyses reveal the correspondence between neural
  1623. circuits and transcriptomic cell types. We further present an
  1624. extensive genetic toolset for targeting and fate mapping
  1625. glutamatergic projection neuron types toward linking their
  1626. developmental trajectory to their circuit function. Together, our
  1627. results establish a unified and mechanistic framework of neuronal
  1628. cell type organization that integrates multi-layered molecular
  1629. genetic and spatial information with multi-faceted phenotypic
  1630. properties. \#\#\# Competing Interest Statement The competing
  1631. interests are detailed in the Competing Interests section in the
  1632. manuscript file.",
  1633. journal = "Cold Spring Harbor Laboratory",
  1634. pages = "2020.10.19.343129",
  1635. month = oct,
  1636. year = 2020,
  1637. language = "en"
  1638. }
  1639. @ARTICLE{Humphries2020-nf,
  1640. title = "Strong and weak principles of neural dimension reduction",
  1641. author = "Humphries, Mark D",
  1642. abstract = "If spikes are the medium, what is the message? Answering
  1643. that question is driving the development of large-scale,
  1644. single neuron resolution recordings from behaving animals,
  1645. on the scale of thousands of neurons. But these data are
  1646. inherently high-dimensional, with as many dimensions as
  1647. neurons - so how do we make sense of them? For many the
  1648. answer is to reduce the number of dimensions. Here I argue
  1649. we can distinguish weak and strong principles of neural
  1650. dimension reduction. The weak principle is that dimension
  1651. reduction is a convenient tool for making sense of complex
  1652. neural data. The strong principle is that dimension
  1653. reduction shows us how neural circuits actually operate and
  1654. compute. Elucidating these principles is crucial, for which
  1655. we subscribe to provides radically different interpretations
  1656. of the same neural activity data. I show how we could make
  1657. either the weak or strong principles appear to be true based
  1658. on innocuous looking decisions about how we use dimension
  1659. reduction on our data. To counteract these confounds, I
  1660. outline the experimental evidence for the strong principle
  1661. that do not come from dimension reduction; but also show
  1662. there are a number of neural phenomena that the strong
  1663. principle fails to address. To reconcile these conflicting
  1664. data, I suggest that the brain has both principles at play.",
  1665. month = nov,
  1666. year = 2020,
  1667. archivePrefix = "arXiv",
  1668. primaryClass = "q-bio.NC",
  1669. eprint = "2011.08088"
  1670. }
  1671. @ARTICLE{Sussillo2015-xp,
  1672. title = "A neural network that finds a naturalistic solution for the
  1673. production of muscle activity",
  1674. author = "Sussillo, David and Churchland, Mark M and Kaufman, Matthew T and
  1675. Shenoy, Krishna V",
  1676. abstract = "It remains an open question how neural responses in motor cortex
  1677. relate to movement. We explored the hypothesis that motor cortex
  1678. reflects dynamics appropriate for generating temporally patterned
  1679. outgoing commands. To formalize this hypothesis, we trained
  1680. recurrent neural networks to reproduce the muscle activity of
  1681. reaching monkeys. Models had to infer dynamics that could
  1682. transform simple inputs into temporally and spatially complex
  1683. patterns of muscle activity. Analysis of trained models revealed
  1684. that the natural dynamical solution was a low-dimensional
  1685. oscillator that generated the necessary multiphasic commands.
  1686. This solution closely resembled, at both the single-neuron and
  1687. population levels, what was observed in neural recordings from
  1688. the same monkeys. Notably, data and simulations agreed only when
  1689. models were optimized to find simple solutions. An appealing
  1690. interpretation is that the empirically observed dynamics of motor
  1691. cortex may reflect a simple solution to the problem of generating
  1692. temporally patterned descending commands.",
  1693. journal = "Nat. Neurosci.",
  1694. volume = 18,
  1695. number = 7,
  1696. pages = "1025--1033",
  1697. month = jul,
  1698. year = 2015,
  1699. keywords = "RNN;RNN To read",
  1700. language = "en"
  1701. }
  1702. @ARTICLE{Young2020-du,
  1703. title = "{Whole-Brain} Image Analysis and Anatomical Atlas {3D} Generation
  1704. Using {MagellanMapper}",
  1705. author = "Young, David M and Duhn, Clif and Gilson, Michael and Nojima, Mai
  1706. and Yuruk, Deniz and Kumar, Aparna and Yu, Weimiao and Sanders,
  1707. Stephan J",
  1708. abstract = "MagellanMapper is a software suite designed for visual inspection
  1709. and end-to-end automated processing of large-volume, 3D brain
  1710. imaging datasets in a memory-efficient manner. The rapidly
  1711. growing number of large-volume, high-resolution datasets
  1712. necessitates visualization of raw data at both macro- and
  1713. microscopic levels to assess the quality of data, as well as
  1714. automated processing to quantify data in an unbiased manner for
  1715. comparison across a large number of samples. To facilitate these
  1716. analyses, MagellanMapper provides both a graphical user interface
  1717. for manual inspection and a command-line interface for automated
  1718. image processing. At the macroscopic level, the graphical
  1719. interface allows researchers to view full volumetric images
  1720. simultaneously in each dimension and to annotate anatomical label
  1721. placements. At the microscopic level, researchers can inspect
  1722. regions of interest at high resolution to build ground truth data
  1723. of cellular locations such as nuclei positions. Using the
  1724. command-line interface, researchers can automate cell detection
  1725. across volumetric images, refine anatomical atlas labels to fit
  1726. underlying histology, register these atlases to sample images,
  1727. and perform statistical analyses by anatomical region.
  1728. MagellanMapper leverages established open-source computer vision
  1729. libraries and is itself open source and freely available for
  1730. download and extension. \copyright{} 2020 Wiley Periodicals LLC.
  1731. Basic Protocol 1: MagellanMapper installation Alternate Protocol:
  1732. Alternative methods for MagellanMapper installation Basic
  1733. Protocol 2: Import image files into MagellanMapper Basic Protocol
  1734. 3: Region of interest visualization and annotation Basic Protocol
  1735. 4: Explore an atlas along all three dimensions and register to a
  1736. sample brain Basic Protocol 5: Automated 3D anatomical atlas
  1737. construction Basic Protocol 6: Whole-tissue cell detection and
  1738. quantification by anatomical label Support Protocol: Import a
  1739. tiled microscopy image in proprietary format into MagellanMapper.",
  1740. journal = "Curr. Protoc. Neurosci.",
  1741. volume = 94,
  1742. number = 1,
  1743. pages = "e104",
  1744. month = dec,
  1745. year = 2020,
  1746. keywords = "3D atlas; graphical interface; image processing; microscopy
  1747. images; tissue clearing",
  1748. language = "en"
  1749. }
  1750. @UNPUBLISHED{Jin2019-xr,
  1751. title = "{SMART}: An open source extension of {WholeBrain} for {iDISCO+}
  1752. {LSFM} intact mouse brain registration and segmentation",
  1753. author = "Jin, Michelle and Nguyen, Joseph D and Weber, Sophia J and
  1754. Mejias-Aponte, Carlos A and Madangopal, Rajtarun and Golden, Sam
  1755. A",
  1756. abstract = "Abstract Mapping immediate early gene (IEG) expression across
  1757. intact brains is becoming a popular approach for identifying the
  1758. brain-wide activity patterns underlying behavior. Registering
  1759. whole brains to an anatomical atlas presents a technical
  1760. challenge that has predominantly been tackled using automated
  1761. voxel-based registration methods; however, these methods may fail
  1762. when brains are damaged or only partially imaged, can be
  1763. challenging to correct, and require substantial computational
  1764. power. Here we present an open source package in R called SMART
  1765. (semi-manual alignment to reference templates) as an extension to
  1766. the WholeBrain framework for automated segmentation and
  1767. semi-automated registration of experimental images to vectorized
  1768. atlas plates from the Allen Brain Institute Mouse Common
  1769. Coordinate Framework (CCF).The SMART package was created with the
  1770. novice programmer in mind and introduces a streamlined pipeline
  1771. for aligning, registering, and segmenting large LSFM volumetric
  1772. datasets with the CCF across the anterior-posterior axis, using a
  1773. simple `choice game' and interactive user-friendly menus. SMART
  1774. further provides the flexibility to register partial brains or
  1775. discrete user-chosen experimental images across the CCF, making
  1776. it compatible with analysis of traditionally sectioned coronal
  1777. brain slices. In addition to SMART, we introduce a modified
  1778. tissue clearing protocol based on the iDISCO+ procedure that is
  1779. optimized for uniform Fos antibody labeling and tissue clearing
  1780. across whole intact mouse brains. Here we demonstrate the utility
  1781. of the SMART-WholeBrain pipeline, in conjunction with the
  1782. modified iDISCO+ Fos procedure, by providing example datasets
  1783. alongside a full user tutorial. Finally, we present a subset of
  1784. these data online in an interactive web applet. The complete
  1785. SMART package is available for download on GitHub.",
  1786. journal = "Cold Spring Harbor Laboratory",
  1787. pages = "727529",
  1788. month = aug,
  1789. year = 2019,
  1790. language = "en"
  1791. }
  1792. @ARTICLE{Song2020-sg,
  1793. title = "Precise Mapping of Single Neurons by Calibrated {3D}
  1794. Reconstruction of Brain Slices Reveals Topographic Projection in
  1795. Mouse Visual Cortex",
  1796. author = "Song, Jun Ho and Choi, Woochul and Song, You-Hyang and Kim,
  1797. Jae-Hyun and Jeong, Daun and Lee, Seung-Hee and Paik, Se-Bum",
  1798. abstract = "Recent breakthroughs in neuroanatomical tracing methods have
  1799. helped unravel complicated neural connectivity in whole-brain
  1800. tissue at single-cell resolution. However, in most cases,
  1801. analysis of brain images remains dependent on highly subjective
  1802. and sample-specific manual processing, preventing precise
  1803. comparison across sample animals. In the present study, we
  1804. introduce AMaSiNe, software for automated mapping of single
  1805. neurons in the standard mouse brain atlas with annotated regions.
  1806. AMaSiNe automatically calibrates misaligned and deformed slice
  1807. samples to locate labeled neuronal positions from multiple brain
  1808. samples into the standardized 3D Allen Mouse Brain Reference
  1809. Atlas. We exploit the high fidelity and reliability of AMaSiNe to
  1810. investigate the topographic structures of feedforward projections
  1811. from the lateral geniculate nucleus to the primary visual area by
  1812. reconstructing rabies-virus-injected brain slices in 3D space.
  1813. Our results demonstrate that distinct organization of neural
  1814. projections can be precisely mapped using AMaSiNe.",
  1815. journal = "Cell Rep.",
  1816. volume = 31,
  1817. number = 8,
  1818. pages = "107682",
  1819. month = may,
  1820. year = 2020,
  1821. keywords = "Allen Mouse Brain Reference Atlas; automated brain mapping; brain
  1822. image registration; brain slice calibration; mouse brain slice;
  1823. retrograde tracing; single-neuron mapping; standard 3D brain
  1824. atlas; topographic projection; visual cortex",
  1825. language = "en"
  1826. }
  1827. @UNPUBLISHED{Mano2020-dx,
  1828. title = "{CUBIC-Cloud}: An Integrative Computational Framework Towards
  1829. Community-driven {Whole-Mouse-Brain} Mapping",
  1830. author = "Mano, Tomoyuki and Murata, Ken and Kon, Kazuhiro and Shimizu,
  1831. Chika and Ono, Hiroaki and Shi, Shoi and Yamada, Rikuhiro G and
  1832. Miyamichi, Kazunari and Susaki, Etsuo A and Touhara, Kazushige
  1833. and Ueda, Hiroki R",
  1834. abstract = "Recent advancements in tissue clearing technologies have offered
  1835. unparalleled opportunities for researchers to explore the whole
  1836. mouse brain at cellular resolution. With the expansion of this
  1837. experimental technique, however, a scalable and easy-to-use
  1838. computational tool is in demand to effectively analyze and
  1839. integrate whole-brain mapping datasets. To that end, here we
  1840. present CUBIC-Cloud, a cloud-based framework to quantify,
  1841. visualize and integrate whole mouse brain data. CUBIC-Cloud is a
  1842. fully automated system where users can upload their whole-brain
  1843. data, run analysis and publish the results. We demonstrate the
  1844. generality of CUBIC-Cloud by a variety of applications. First, we
  1845. investigated brain-wide distribution of PV, Sst, ChAT, Th and
  1846. Iba1 expressing cells. Second, A$\beta$ plaque deposition in AD
  1847. model mouse brains were quantified. Third, we reconstructed
  1848. neuronal activity profile under LPS-induced inflammation by c-Fos
  1849. immunostaining. Last, we show brain-wide connectivity mapping by
  1850. pseudo-typed Rabies virus. Together, CUBIC-Cloud provides an
  1851. integrative platform to advance scalable and collaborative
  1852. whole-brain mapping. \#\#\# Competing Interest Statement T.M. and
  1853. H.R.U. filed a patent application regarding the CUBIC-Cloud
  1854. software. CUBIC-Cloud web service is provided and maintained by
  1855. CUBICStars Inc.",
  1856. journal = "Cold Spring Harbor Laboratory",
  1857. pages = "2020.08.28.271031",
  1858. month = aug,
  1859. year = 2020,
  1860. language = "en"
  1861. }
  1862. @ARTICLE{Goubran2019-je,
  1863. title = "Multimodal image registration and connectivity analysis for
  1864. integration of connectomic data from microscopy to {MRI}",
  1865. author = "Goubran, Maged and Leuze, Christoph and Hsueh, Brian and Aswendt,
  1866. Markus and Ye, Li and Tian, Qiyuan and Cheng, Michelle Y and
  1867. Crow, Ailey and Steinberg, Gary K and McNab, Jennifer A and
  1868. Deisseroth, Karl and Zeineh, Michael",
  1869. abstract = "3D histology, slice-based connectivity atlases, and diffusion MRI
  1870. are common techniques to map brain wiring. While there are many
  1871. modality-specific tools to process these data, there is a lack of
  1872. integration across modalities. We develop an automated resource
  1873. that combines histologically cleared volumes with connectivity
  1874. atlases and MRI, enabling the analysis of histological features
  1875. across multiple fiber tracts and networks, and their correlation
  1876. with in-vivo biomarkers. We apply our pipeline in a murine stroke
  1877. model, demonstrating not only strong correspondence between MRI
  1878. abnormalities and CLARITY-tissue staining, but also uncovering
  1879. acute cellular effects in areas connected to the ischemic core.
  1880. We provide improved maps of connectivity by quantifying
  1881. projection terminals from CLARITY viral injections, and integrate
  1882. diffusion MRI with CLARITY viral tracing to compare connectivity
  1883. maps across scales. Finally, we demonstrate tract-level
  1884. histological changes of stroke through this multimodal
  1885. integration. This resource can propel investigations of network
  1886. alterations underlying neurological disorders.",
  1887. journal = "Nat. Commun.",
  1888. volume = 10,
  1889. number = 1,
  1890. pages = "5504",
  1891. month = dec,
  1892. year = 2019,
  1893. language = "en"
  1894. }
  1895. @ARTICLE{Renier2016-cx,
  1896. title = "Mapping of Brain Activity by Automated Volume Analysis of
  1897. Immediate Early Genes",
  1898. author = "Renier, Nicolas and Adams, Eliza L and Kirst, Christoph and Wu,
  1899. Zhuhao and Azevedo, Ricardo and Kohl, Johannes and Autry, Anita E
  1900. and Kadiri, Lolahon and Umadevi Venkataraju, Kannan and Zhou, Yu
  1901. and Wang, Victoria X and Tang, Cheuk Y and Olsen, Olav and Dulac,
  1902. Catherine and Osten, Pavel and Tessier-Lavigne, Marc",
  1903. abstract = "Understanding how neural information is processed in
  1904. physiological and pathological states would benefit from precise
  1905. detection, localization, and quantification of the activity of
  1906. all neurons across the entire brain, which has not, to date, been
  1907. achieved in the mammalian brain. We introduce a pipeline for
  1908. high-speed acquisition of brain activity at cellular resolution
  1909. through profiling immediate early gene expression using
  1910. immunostaining and light-sheet fluorescence imaging, followed by
  1911. automated mapping and analysis of activity by an open-source
  1912. software program we term ClearMap. We validate the pipeline first
  1913. by analysis of brain regions activated in response to
  1914. haloperidol. Next, we report new cortical regions downstream of
  1915. whisker-evoked sensory processing during active exploration.
  1916. Last, we combine activity mapping with axon tracing to uncover
  1917. new brain regions differentially activated during parenting
  1918. behavior. This pipeline is widely applicable to different
  1919. experimental paradigms, including animal species for which
  1920. transgenic activity reporters are not readily available.",
  1921. journal = "Cell",
  1922. volume = 165,
  1923. number = 7,
  1924. pages = "1789--1802",
  1925. month = jun,
  1926. year = 2016,
  1927. language = "en"
  1928. }
  1929. @ARTICLE{Carlsson2020-lg,
  1930. title = "Topological methods for data modelling",
  1931. author = "Carlsson, Gunnar",
  1932. abstract = "The analysis of large and complex data sets is one of the most
  1933. important problems facing the scientific community, and physics
  1934. in particular. One response to this challenge has been the
  1935. development of topological data analysis (TDA), which models
  1936. data by graphs or networks rather than by linear algebraic
  1937. (matrix) methods or cluster analysis. TDA represents the shape
  1938. of the data (suitably defined) in a combinatorial fashion.
  1939. Methods for measuring shape have been developed within
  1940. mathematics, providing a toolkit referred to as homology. In
  1941. working with data, one can use this kind of modelling to obtain
  1942. an understanding of the overall structure of the data set. There
  1943. is a suite of methods for constructing vector representations of
  1944. various kinds of unstructured data. In this Review, we sketch
  1945. the basics of TDA and provide examples where this kind of
  1946. analysis has been carried out. The rapidly developing field of
  1947. topological data analysis represents data via graphs rather than
  1948. as solutions to equations or as decompositions into clusters.
  1949. This Review discusses the methods and provides examples from
  1950. physics and other sciences.",
  1951. journal = "Nature Reviews Physics",
  1952. publisher = "Nature Publishing Group",
  1953. pages = "1--12",
  1954. month = nov,
  1955. year = 2020,
  1956. keywords = "RNN",
  1957. language = "en"
  1958. }
  1959. @UNPUBLISHED{Kalidindi2020-wd,
  1960. title = "Rotational dynamics in motor cortex are consistent with a
  1961. feedback controller",
  1962. author = "Kalidindi, Hari Teja and Cross, Kevin P and Lillicrap, Timothy P
  1963. and Omrani, Mohsen and Falotico, Egidio and Sabes, Philip N and
  1964. Scott, Stephen H",
  1965. abstract = "Recent studies hypothesize that motor cortical (MC) dynamics are
  1966. generated largely through its recurrent connections based on
  1967. observations that MC activity exhibits rotational structure.
  1968. However, behavioural and neurophysiological studies suggest that
  1969. MC behaves like a feedback controller where continuous sensory
  1970. feedback and interactions with other brain areas contribute
  1971. substantially to MC processing. We investigated these apparently
  1972. conflicting theories by building recurrent neural networks that
  1973. controlled a model arm and received sensory feedback about the
  1974. limb. Networks were trained to counteract perturbations to the
  1975. limb and to reach towards spatial targets. Network activities and
  1976. sensory feedback signals to the network exhibited rotational
  1977. structure even when the recurrent connections were removed.
  1978. Furthermore, neural recordings in monkeys performing similar
  1979. tasks also exhibited rotational structure not only in MC but also
  1980. in somatosensory cortex. Our results argue that rotational
  1981. structure may reflect dynamics throughout voluntary motor
  1982. circuits involved in online control of motor actions. \#\#\#
  1983. Competing Interest Statement SHS is co-founder and CSO of Kinarm
  1984. which commercializes the robotic technology used in the present
  1985. study.",
  1986. journal = "Cold Spring Harbor Laboratory",
  1987. pages = "2020.11.17.387043",
  1988. month = nov,
  1989. year = 2020,
  1990. keywords = "RNN",
  1991. language = "en"
  1992. }
  1993. @UNPUBLISHED{Reis2020-bp,
  1994. title = "Dorsal Periaqueductal gray ensembles represent approach and
  1995. avoidance states",
  1996. author = "Reis, Fernando Mcv and Lee, Johannes Y and Maesta-Pereira, Sandra
  1997. and Schuette, Peter J and Chakerian, Meghmik and Liu, Jinhan and
  1998. La-Vu, Mimi Q and Tobias, Brooke C and Canteras, Newton and Kao,
  1999. Jonathan C and Adhikari, Avishek",
  2000. abstract = "Animals must balance needs to approach threats for
  2001. risk-assessment and to avoid danger. The dorsal periaqueductal
  2002. gray (dPAG) controls defensive behaviors, but it is unknown how
  2003. it represents states associated with threat approach and
  2004. avoidance. We identified a dPAG threat-avoidance ensemble in mice
  2005. that showed higher activity far from threats such as the open
  2006. arms of the elevated plus maze and a live predator. These cells
  2007. were also more active during threat-avoidance behaviors such as
  2008. escape and freezing, even though these behaviors have
  2009. antagonistic motor output. Conversely, the threat-approach
  2010. ensemble was more active during risk-assessment behaviors and
  2011. near threats. Furthermore, unsupervised methods showed
  2012. approach/avoidance states were encoded with shared activity
  2013. patterns across threats. Lastly, the relative number of cells in
  2014. each ensemble predicted threat-avoidance across mice. Thus, dPAG
  2015. ensembles dynamically encode threat approach and avoidance
  2016. states, providing a flexible mechanism to balance risk-assessment
  2017. and danger avoidance.",
  2018. journal = "Cold Spring Harbor Laboratory",
  2019. pages = "2020.11.19.389486",
  2020. month = nov,
  2021. year = 2020,
  2022. language = "en"
  2023. }
  2024. @ARTICLE{Machado2015-ig,
  2025. title = "A quantitative framework for whole-body coordination reveals
  2026. specific deficits in freely walking ataxic mice",
  2027. author = "Machado, Ana S and Darmohray, Dana M and Fayad, Jo{\~a}o and
  2028. Marques, Hugo G and Carey, Megan R",
  2029. abstract = "The coordination of movement across the body is a fundamental,
  2030. yet poorly understood aspect of motor control. Mutant mice with
  2031. cerebellar circuit defects exhibit characteristic impairments in
  2032. locomotor coordination; however, the fundamental features of this
  2033. gait ataxia have not been effectively isolated. Here we describe
  2034. a novel system (LocoMouse) for analyzing limb, head, and tail
  2035. kinematics of freely walking mice. Analysis of visibly ataxic
  2036. Purkinje cell degeneration (pcd) mice reveals that while
  2037. differences in the forward motion of individual paws are fully
  2038. accounted for by changes in walking speed and body size, more
  2039. complex 3D trajectories and, especially, inter-limb and
  2040. whole-body coordination are specifically impaired. Moreover, the
  2041. coordination deficits in pcd are consistent with a failure to
  2042. predict and compensate for the consequences of movement across
  2043. the body. These results isolate specific impairments in
  2044. whole-body coordination in mice and provide a quantitative
  2045. framework for understanding cerebellar contributions to
  2046. coordinated locomotion.",
  2047. journal = "Elife",
  2048. volume = 4,
  2049. month = oct,
  2050. year = 2015,
  2051. keywords = "Purkinje cell; ataxia; cerebellum; locomotion; mouse;
  2052. neuroscience;Locomotion",
  2053. language = "en"
  2054. }
  2055. @ARTICLE{Russo2018-me,
  2056. title = "Motor Cortex Embeds Muscle-like Commands in an Untangled
  2057. Population Response",
  2058. author = "Russo, Abigail A and Bittner, Sean R and Perkins, Sean M and
  2059. Seely, Jeffrey S and London, Brian M and Lara, Antonio H and
  2060. Miri, Andrew and Marshall, Najja J and Kohn, Adam and Jessell,
  2061. Thomas M and Abbott, Laurence F and Cunningham, John P and
  2062. Churchland, Mark M",
  2063. abstract = "Primate motor cortex projects to spinal interneurons and
  2064. motoneurons, suggesting that motor cortex activity may be
  2065. dominated by muscle-like commands. Observations during reaching
  2066. lend support to this view, but evidence remains ambiguous and
  2067. much debated. To provide a different perspective, we employed a
  2068. novel behavioral paradigm that facilitates comparison between
  2069. time-evolving neural and muscle activity. We found that single
  2070. motor cortex neurons displayed many muscle-like properties, but
  2071. the structure of population activity was not muscle-like. Unlike
  2072. muscle activity, neural activity was structured to avoid
  2073. ``tangling'': moments where similar activity patterns led to
  2074. dissimilar future patterns. Avoidance of tangling was present
  2075. across tasks and species. Network models revealed a potential
  2076. reason for this consistent feature: low tangling confers noise
  2077. robustness. Finally, we were able to predict motor cortex
  2078. activity from muscle activity by leveraging the hypothesis that
  2079. muscle-like commands are embedded in additional structure that
  2080. yields low tangling.",
  2081. journal = "Neuron",
  2082. volume = 97,
  2083. number = 4,
  2084. pages = "953--966.e8",
  2085. month = feb,
  2086. year = 2018,
  2087. keywords = "motor control; motor cortex; movement generation; neural
  2088. dynamics; neural network; pattern generation; rhythmic
  2089. movement;RNN;RNN To read",
  2090. language = "en"
  2091. }
  2092. @UNPUBLISHED{Russo2019-sw,
  2093. title = "Neural trajectories in the supplementary motor area and primary
  2094. motor cortex exhibit distinct geometries, compatible with
  2095. different classes of computation",
  2096. author = "Russo, Abigail A and Khajeh, Ramin and Bittner, Sean R and
  2097. Perkins, Sean M and Cunningham, John P and Abbott, Laurence F and
  2098. Churchland, Mark M",
  2099. abstract = "Abstract The supplementary motor area (SMA) is believed to
  2100. contribute to higher-order aspects of motor control. To examine
  2101. this contribution, we employed a novel cycling task and leveraged
  2102. an emerging strategy: testing whether population trajectories
  2103. possess properties necessary for a hypothesized class of
  2104. computations. We found that, at the single-neuron level, SMA
  2105. exhibited multiple response features absent in M1. We
  2106. hypothesized that these diverse features might contribute, at the
  2107. population level, to avoidance of `population trajectory
  2108. divergence' -- ensuring that two trajectories never followed the
  2109. same path before separating. Trajectory divergence was indeed
  2110. avoided in SMA but not in M1. Network simulations confirmed that
  2111. low trajectory divergence is necessary when guidance of future
  2112. action depends upon internally tracking contextual factors.
  2113. Furthermore, the empirical trajectory geometry -- helical in SMA
  2114. versus elliptical in M1 -- was naturally reproduced by networks
  2115. that did, versus did not, internally track context.",
  2116. journal = "Cold Spring Harbor Laboratory",
  2117. pages = "650002",
  2118. month = may,
  2119. year = 2019,
  2120. keywords = "RNN;RNN To read",
  2121. language = "en"
  2122. }
  2123. @ARTICLE{Elsayed2017-ru,
  2124. title = "Structure in neural population recordings: an expected byproduct
  2125. of simpler phenomena?",
  2126. author = "Elsayed, Gamaleldin F and Cunningham, John P",
  2127. abstract = "Neuroscientists increasingly analyze the joint activity of
  2128. multineuron recordings to identify population-level structures
  2129. believed to be significant and scientifically novel. Claims of
  2130. significant population structure support hypotheses in many
  2131. brain areas. However, these claims require first investigating
  2132. the possibility that the population structure in question is an
  2133. expected byproduct of simpler features known to exist in data.
  2134. Classically, this critical examination can be either intuited or
  2135. addressed with conventional controls. However, these approaches
  2136. fail when considering population data, raising concerns about
  2137. the scientific merit of population-level studies. Here we
  2138. develop a framework to test the novelty of population-level
  2139. findings against simpler features such as correlations across
  2140. times, neurons and conditions. We apply this framework to test
  2141. two recent population findings in prefrontal and motor cortices,
  2142. providing essential context to those studies. More broadly, the
  2143. methodologies we introduce provide a general neural population
  2144. control for many population-level hypotheses.",
  2145. journal = "Nat. Neurosci.",
  2146. publisher = "nature.com",
  2147. volume = 20,
  2148. number = 9,
  2149. pages = "1310--1318",
  2150. month = sep,
  2151. year = 2017,
  2152. keywords = "RNN To read;RNN",
  2153. language = "en"
  2154. }
  2155. @ARTICLE{Rivkind2017-pf,
  2156. title = "Local Dynamics in Trained Recurrent Neural Networks",
  2157. author = "Rivkind, Alexander and Barak, Omri",
  2158. abstract = "Learning a task induces connectivity changes in neural circuits,
  2159. thereby changing their dynamics. To elucidate task-related neural
  2160. dynamics, we study trained recurrent neural networks. We develop
  2161. a mean field theory for reservoir computing networks trained to
  2162. have multiple fixed point attractors. Our main result is that the
  2163. dynamics of the network's output in the vicinity of attractors is
  2164. governed by a low-order linear ordinary differential equation.
  2165. The stability of the resulting equation can be assessed,
  2166. predicting training success or failure. As a consequence,
  2167. networks of rectified linear units and of sigmoidal
  2168. nonlinearities are shown to have diametrically different
  2169. properties when it comes to learning attractors. Furthermore, a
  2170. characteristic time constant, which remains finite at the edge of
  2171. chaos, offers an explanation of the network's output robustness
  2172. in the presence of variability of the internal neural dynamics.
  2173. Finally, the proposed theory predicts state-dependent frequency
  2174. selectivity in the network response.",
  2175. journal = "Phys. Rev. Lett.",
  2176. volume = 118,
  2177. number = 25,
  2178. pages = "258101",
  2179. month = jun,
  2180. year = 2017,
  2181. keywords = "RNN",
  2182. language = "en"
  2183. }
  2184. @ARTICLE{Sussillo2016-zn,
  2185. title = "{LFADS} - Latent Factor Analysis via Dynamical Systems",
  2186. author = "Sussillo, David and Jozefowicz, Rafal and Abbott, L F and
  2187. Pandarinath, Chethan",
  2188. abstract = "Neuroscience is experiencing a data revolution in which many
  2189. hundreds or thousands of neurons are recorded
  2190. simultaneously. Currently, there is little consensus on how
  2191. such data should be analyzed. Here we introduce LFADS
  2192. (Latent Factor Analysis via Dynamical Systems), a method to
  2193. infer latent dynamics from simultaneously recorded,
  2194. single-trial, high-dimensional neural spiking data. LFADS is
  2195. a sequential model based on a variational auto-encoder. By
  2196. making a dynamical systems hypothesis regarding the
  2197. generation of the observed data, LFADS reduces observed
  2198. spiking to a set of low-dimensional temporal factors,
  2199. per-trial initial conditions, and inferred inputs. We
  2200. compare LFADS to existing methods on synthetic data and show
  2201. that it significantly out-performs them in inferring neural
  2202. firing rates and latent dynamics.",
  2203. month = aug,
  2204. year = 2016,
  2205. keywords = "RNN To read;RNN",
  2206. archivePrefix = "arXiv",
  2207. primaryClass = "cs.LG",
  2208. eprint = "1608.06315"
  2209. }
  2210. @ARTICLE{Saxe2020-mi,
  2211. title = "If deep learning is the answer, what is the question?",
  2212. author = "Saxe, Andrew and Nelli, Stephanie and Summerfield, Christopher",
  2213. abstract = "Neuroscience research is undergoing a minor revolution. Recent
  2214. advances in machine learning and artificial intelligence research
  2215. have opened up new ways of thinking about neural computation.
  2216. Many researchers are excited by the possibility that deep neural
  2217. networks may offer theories of perception, cognition and action
  2218. for biological brains. This approach has the potential to
  2219. radically reshape our approach to understanding neural systems,
  2220. because the computations performed by deep networks are learned
  2221. from experience, and not endowed by the researcher. If so, how
  2222. can neuroscientists use deep networks to model and understand
  2223. biological brains? What is the outlook for neuroscientists who
  2224. seek to characterize computations or neural codes, or who wish to
  2225. understand perception, attention, memory and executive functions?
  2226. In this Perspective, our goal is to offer a road map for systems
  2227. neuroscience research in the age of deep learning. We discuss the
  2228. conceptual and methodological challenges of comparing behaviour,
  2229. learning dynamics and neural representations in artificial and
  2230. biological systems, and we highlight new research questions that
  2231. have emerged for neuroscience as a direct consequence of recent
  2232. advances in machine learning.",
  2233. journal = "Nat. Rev. Neurosci.",
  2234. month = nov,
  2235. year = 2020,
  2236. keywords = "RNN"
  2237. }
  2238. @ARTICLE{Chon2019-ka,
  2239. title = "Enhanced and unified anatomical labeling for a common mouse
  2240. brain atlas",
  2241. author = "Chon, Uree and Vanselow, Daniel J and Cheng, Keith C and Kim,
  2242. Yongsoo",
  2243. abstract = "Anatomical atlases in standard coordinates are necessary for the
  2244. interpretation and integration of research findings in a common
  2245. spatial context. However, the two most-used mouse brain atlases,
  2246. the Franklin-Paxinos (FP) and the common coordinate framework
  2247. (CCF) from the Allen Institute for Brain Science, have
  2248. accumulated inconsistencies in anatomical delineations and
  2249. nomenclature, creating confusion among neuroscientists. To
  2250. overcome these issues, we adopt here the FP labels into the CCF
  2251. to merge the labels in the single atlas framework. We use cell
  2252. type-specific transgenic mice and an MRI atlas to adjust and
  2253. further segment our labels. Moreover, detailed segmentations are
  2254. added to the dorsal striatum using cortico-striatal connectivity
  2255. data. Lastly, we digitize our anatomical labels based on the
  2256. Allen ontology, create a web-interface for visualization, and
  2257. provide tools for comprehensive comparisons between the CCF and
  2258. FP labels. Our open-source labels signify a key step towards a
  2259. unified mouse brain atlas.",
  2260. journal = "Nat. Commun.",
  2261. publisher = "nature.com",
  2262. volume = 10,
  2263. number = 1,
  2264. pages = "5067",
  2265. month = nov,
  2266. year = 2019,
  2267. language = "en"
  2268. }
  2269. @ARTICLE{Hanwell2015-wl,
  2270. title = "The Visualization Toolkit ({VTK)}: Rewriting the rendering code
  2271. for modern graphics cards",
  2272. author = "Hanwell, Marcus D and Martin, Kenneth M and Chaudhary, Aashish
  2273. and Avila, Lisa S",
  2274. abstract = "The Visualization Toolkit (VTK) is an open source, permissively
  2275. licensed, cross-platform toolkit for scientific data processing,
  2276. visualization, and data analysis. It is over two decades old,
  2277. originally developed for a very different graphics card
  2278. architecture. Modern graphics cards feature fully programmable,
  2279. highly parallelized architectures with large core counts. VTK's
  2280. rendering code was rewritten to take advantage of modern graphics
  2281. cards, maintaining most of the toolkit's programming interfaces.
  2282. This offers the opportunity to compare the performance of old and
  2283. new rendering code on the same systems/cards. Significant
  2284. improvements in rendering speeds and memory footprints mean that
  2285. scientific data can be visualized in greater detail than ever
  2286. before. The widespread use of VTK means that these improvements
  2287. will reap significant benefits.",
  2288. journal = "SoftwareX",
  2289. volume = "1-2",
  2290. pages = "9--12",
  2291. month = sep,
  2292. year = 2015,
  2293. keywords = "Visualization; Toolkit; Data analysis; Scientific data"
  2294. }
  2295. @ARTICLE{Golub2018-il,
  2296. title = "{FixedPointFinder}: A Tensorflow toolbox for identifying and
  2297. characterizing fixed points in recurrent neural networks",
  2298. author = "Golub, Matthew D and Sussillo, David",
  2299. journal = "Journal of Open Source Software",
  2300. volume = 3,
  2301. number = 31,
  2302. pages = "1003",
  2303. year = 2018
  2304. }
  2305. @ARTICLE{Barak2017-uh,
  2306. title = "Recurrent neural networks as versatile tools of neuroscience
  2307. research",
  2308. author = "Barak, Omri",
  2309. abstract = "Recurrent neural networks (RNNs) are a class of computational
  2310. models that are often used as a tool to explain neurobiological
  2311. phenomena, considering anatomical, electrophysiological and
  2312. computational constraints. RNNs can either be designed to
  2313. implement a certain dynamical principle, or they can be trained
  2314. by input-output examples. Recently, there has been large progress
  2315. in utilizing trained RNNs both for computational tasks, and as
  2316. explanations of neural phenomena. I will review how combining
  2317. trained RNNs with reverse engineering can provide an alternative
  2318. framework for modeling in neuroscience, potentially serving as a
  2319. powerful hypothesis generation tool. Despite the recent progress
  2320. and potential benefits, there are many fundamental gaps towards a
  2321. theory of these networks. I will discuss these challenges and
  2322. possible methods to attack them.",
  2323. journal = "Curr. Opin. Neurobiol.",
  2324. volume = 46,
  2325. pages = "1--6",
  2326. month = oct,
  2327. year = 2017,
  2328. keywords = "RNN",
  2329. language = "en"
  2330. }
  2331. @ARTICLE{Mastrogiuseppe2018-ss,
  2332. title = "Linking Connectivity, Dynamics, and Computations in {Low-Rank}
  2333. Recurrent Neural Networks",
  2334. author = "Mastrogiuseppe, Francesca and Ostojic, Srdjan",
  2335. abstract = "Large-scale neural recordings have established that the
  2336. transformation of sensory stimuli into motor outputs relies on
  2337. low-dimensional dynamics at the population level, while
  2338. individual neurons exhibit complex selectivity. Understanding how
  2339. low-dimensional computations on mixed, distributed
  2340. representations emerge from the structure of the recurrent
  2341. connectivity and inputs to cortical networks is a major
  2342. challenge. Here, we study a class of recurrent network models in
  2343. which the connectivity is a sum of a random part and a minimal,
  2344. low-dimensional structure. We show that, in such networks, the
  2345. dynamics are low dimensional and can be directly inferred from
  2346. connectivity using a geometrical approach. We exploit this
  2347. understanding to determine minimal connectivity required to
  2348. implement specific computations and find that the dynamical range
  2349. and computational capacity quickly increase with the
  2350. dimensionality of the connectivity structure. This framework
  2351. produces testable experimental predictions for the relationship
  2352. between connectivity, low-dimensional dynamics, and computational
  2353. features of recorded neurons.",
  2354. journal = "Neuron",
  2355. volume = 99,
  2356. number = 3,
  2357. pages = "609--623.e29",
  2358. month = aug,
  2359. year = 2018,
  2360. keywords = "low dimensional dynamics; mixed selectivity; neural computations;
  2361. recurrent neural networks;RNN",
  2362. language = "en"
  2363. }
  2364. @ARTICLE{Mastrogiuseppe2019-xu,
  2365. title = "A Geometrical Analysis of Global Stability in Trained Feedback
  2366. Networks",
  2367. author = "Mastrogiuseppe, Francesca and Ostojic, Srdjan",
  2368. abstract = "Recurrent neural networks have been extensively studied in the
  2369. context of neuroscience and machine learning due to their ability
  2370. to implement complex computations. While substantial progress in
  2371. designing effective learning algorithms has been achieved, a full
  2372. understanding of trained recurrent networks is still lacking.
  2373. Specifically, the mechanisms that allow computations to emerge
  2374. from the underlying recurrent dynamics are largely unknown. Here
  2375. we focus on a simple yet underexplored computational setup: a
  2376. feedback architecture trained to associate a stationary output to
  2377. a stationary input. As a starting point, we derive an approximate
  2378. analytical description of global dynamics in trained networks,
  2379. which assumes uncorrelated connectivity weights in the feedback
  2380. and in the random bulk. The resulting mean-field theory suggests
  2381. that the task admits several classes of solutions, which imply
  2382. different stability properties. Different classes are
  2383. characterized in terms of the geometrical arrangement of the
  2384. readout with respect to the input vectors, defined in the
  2385. high-dimensional space spanned by the network population. We find
  2386. that such an approximate theoretical approach can be used to
  2387. understand how standard training techniques implement the
  2388. input-output task in finite-size feedback networks. In
  2389. particular, our simplified description captures the local and the
  2390. global stability properties of the target solution, and thus
  2391. predicts training performance.",
  2392. journal = "Neural Comput.",
  2393. volume = 31,
  2394. number = 6,
  2395. pages = "1139--1182",
  2396. month = jun,
  2397. year = 2019,
  2398. keywords = "RNN;RNN To read",
  2399. language = "en"
  2400. }
  2401. % The entry below contains non-ASCII chars that could not be converted
  2402. % to a LaTeX equivalent.
  2403. @ARTICLE{Schuessler2020-ug,
  2404. title = "The interplay between randomness and structure during learning
  2405. in {RNNs}",
  2406. author = "Schuessler, F and Mastrogiuseppe, F and Dubreuil, A and {others}",
  2407. abstract = "Training recurrent neural networks ( RNNs ) on low-dimensional
  2408. tasks has been widely used to model functional biological
  2409. networks. However, the solutions found by learning and the
  2410. effect of initial connectivity are not well understood. Here, we
  2411. examine RNNs trained using …",
  2412. journal = "Adv. Neural Inf. Process. Syst.",
  2413. publisher = "papers.nips.cc",
  2414. year = 2020,
  2415. keywords = "RNN"
  2416. }
  2417. % The entry below contains non-ASCII chars that could not be converted
  2418. % to a LaTeX equivalent.
  2419. @ARTICLE{Dubreuil2020-fk,
  2420. title = "Complementary roles of dimensionality and population structure
  2421. in neural computations",
  2422. author = "Dubreuil, A and Valente, A and Beiran, M and Mastrogiuseppe, F
  2423. and {others}",
  2424. abstract = "Neural computations are currently investigated using two
  2425. competing approaches: sorting neurons into functional classes,
  2426. or examining the low-dimensional dynamics of collective
  2427. activity. Whether and how these two aspects interact to shape
  2428. computations is currently …",
  2429. journal = "bioRxiv",
  2430. publisher = "biorxiv.org",
  2431. year = 2020,
  2432. keywords = "RNN;RNN To read"
  2433. }
  2434. % The entry below contains non-ASCII chars that could not be converted
  2435. % to a LaTeX equivalent.
  2436. @ARTICLE{Schuessler2020-jm,
  2437. title = "Dynamics of random recurrent networks with correlated low-rank
  2438. structure",
  2439. author = "Schuessler, F and Dubreuil, A and Mastrogiuseppe, F and {others}",
  2440. abstract = "A given neural network in the brain is involved in many
  2441. different tasks. This implies that, when considering a specific
  2442. task, the network's connectivity contains a component which is
  2443. related to the task and another component which can be
  2444. considered random. Understanding …",
  2445. journal = "Physical Review",
  2446. publisher = "APS",
  2447. year = 2020,
  2448. keywords = "RNN;RNN To read"
  2449. }
  2450. @ARTICLE{Beiran2020-tf,
  2451. title = "Shaping dynamics with multiple populations in low-rank
  2452. recurrent networks",
  2453. author = "Beiran, Manuel and Dubreuil, Alexis and Valente, Adrian and
  2454. Mastrogiuseppe, Francesca and Ostojic, Srdjan",
  2455. abstract = "An emerging paradigm proposes that neural computations can
  2456. be understood at the level of dynamical systems that govern
  2457. low-dimensional trajectories of collective neural activity.
  2458. How the connectivity structure of a network determines the
  2459. emergent dynamical system however remains to be clarified.
  2460. Here we consider a novel class of models, Gaussian-mixture
  2461. low-rank recurrent networks, in which the rank of the
  2462. connectivity matrix and the number of statistically-defined
  2463. populations are independent hyper-parameters. We show that
  2464. the resulting collective dynamics form a dynamical system,
  2465. where the rank sets the dimensionality and the population
  2466. structure shapes the dynamics. In particular, the collective
  2467. dynamics can be described in terms of a simplified effective
  2468. circuit of interacting latent variables. While having a
  2469. single, global population strongly restricts the possible
  2470. dynamics, we demonstrate that if the number of populations
  2471. is large enough, a rank $R$ network can approximate any
  2472. $R$-dimensional dynamical system.",
  2473. month = jul,
  2474. year = 2020,
  2475. keywords = "RNN",
  2476. archivePrefix = "arXiv",
  2477. primaryClass = "q-bio.NC",
  2478. eprint = "2007.02062"
  2479. }
  2480. @ARTICLE{Maheswaranathan2019-ue,
  2481. title = "Reverse engineering recurrent networks for sentiment
  2482. classification reveals line attractor dynamics",
  2483. author = "Maheswaranathan, Niru and Williams, Alex H and Golub, Matthew D
  2484. and Ganguli, Surya and Sussillo, David",
  2485. abstract = "Recurrent neural networks (RNNs) are a widely used tool for
  2486. modeling sequential data, yet they are often treated as
  2487. inscrutable black boxes. Given a trained recurrent network, we
  2488. would like to reverse engineer it-to obtain a quantitative,
  2489. interpretable description of how it solves a particular task.
  2490. Even for simple tasks, a detailed understanding of how recurrent
  2491. networks work, or a prescription for how to develop such an
  2492. understanding, remains elusive. In this work, we use tools from
  2493. dynamical systems analysis to reverse engineer recurrent
  2494. networks trained to perform sentiment classification, a
  2495. foundational natural language processing task. Given a trained
  2496. network, we find fixed points of the recurrent dynamics and
  2497. linearize the nonlinear system around these fixed points.
  2498. Despite their theoretical capacity to implement complex,
  2499. high-dimensional computations, we find that trained networks
  2500. converge to highly interpretable, low-dimensional
  2501. representations. In particular, the topological structure of the
  2502. fixed points and corresponding linearized dynamics reveal an
  2503. approximate line attractor within the RNN, which we can use to
  2504. quantitatively understand how the RNN solves the sentiment
  2505. analysis task. Finally, we find this mechanism present across
  2506. RNN architectures (including LSTMs, GRUs, and vanilla RNNs)
  2507. trained on multiple datasets, suggesting that our findings are
  2508. not unique to a particular architecture or dataset. Overall,
  2509. these results demonstrate that surprisingly universal and human
  2510. interpretable computations can arise across a range of recurrent
  2511. networks.",
  2512. journal = "Adv. Neural Inf. Process. Syst.",
  2513. publisher = "papers.nips.cc",
  2514. volume = 32,
  2515. pages = "15696--15705",
  2516. month = dec,
  2517. year = 2019,
  2518. keywords = "RNN To read;RNN",
  2519. language = "en"
  2520. }
  2521. @ARTICLE{Chang2019-bd,
  2522. title = "{AntisymmetricRNN}: A Dynamical System View on Recurrent
  2523. Neural Networks",
  2524. author = "Chang, Bo and Chen, Minmin and Haber, Eldad and Chi, Ed H",
  2525. abstract = "Recurrent neural networks have gained widespread use in
  2526. modeling sequential data. Learning long-term dependencies
  2527. using these models remains difficult though, due to
  2528. exploding or vanishing gradients. In this paper, we draw
  2529. connections between recurrent networks and ordinary
  2530. differential equations. A special form of recurrent networks
  2531. called the AntisymmetricRNN is proposed under this
  2532. theoretical framework, which is able to capture long-term
  2533. dependencies thanks to the stability property of its
  2534. underlying differential equation. Existing approaches to
  2535. improving RNN trainability often incur significant
  2536. computation overhead. In comparison, AntisymmetricRNN
  2537. achieves the same goal by design. We showcase the advantage
  2538. of this new architecture through extensive simulations and
  2539. experiments. AntisymmetricRNN exhibits much more predictable
  2540. dynamics. It outperforms regular LSTM models on tasks
  2541. requiring long-term memory and matches the performance on
  2542. tasks where short-term dependencies dominate despite being
  2543. much simpler.",
  2544. month = feb,
  2545. year = 2019,
  2546. keywords = "RNN",
  2547. archivePrefix = "arXiv",
  2548. primaryClass = "stat.ML",
  2549. eprint = "1902.09689"
  2550. }
  2551. @UNPUBLISHED{Dahmen2020-so,
  2552. title = "Strong coupling and local control of dimensionality across brain
  2553. areas",
  2554. author = "Dahmen, David and Recanatesi, Stefano and Ocker, Gabriel Koch and
  2555. Jia, Xiaoxuan and Helias, Moritz and Shea-Brown, Eric",
  2556. abstract = "The dimensionality of a network's collective activity is the
  2557. number of modes into which it is organized. This quantity is of
  2558. great interest in neural coding: small dimensionality suggests a
  2559. compressed neural code and possibly high robustness and
  2560. generalizability, while high dimensionality suggests expansion of
  2561. input features to enable flexible downstream computation. Here,
  2562. for recurrent neural circuits operating in the ubiquitous
  2563. balanced regime, we show how dimensionality arises
  2564. mechanistically via perhaps the most basic property of neural
  2565. circuits: a single number characterizing the net strength of
  2566. their connectivity. Our results combine novel theoretical
  2567. approaches with new analyses of high-density neuropixels
  2568. recordings and high-throughput synaptic physiology datasets. The
  2569. analysis of electrophysiological recordings identifies bounds on
  2570. the dimensionality of neural responses across brain regions,
  2571. showing that it is on the order of hundreds -- striking a balance
  2572. between high and low-dimensional codes. Furthermore, focusing on
  2573. the visual stream, we show that dimensionality expands from
  2574. primary to deeper visual areas and similarly within an area from
  2575. layer 2/3 to layer 5. We interpret these results via a novel
  2576. theoretical result which links dimensionality to a single measure
  2577. of net connectivity strength. This requires calculations that
  2578. extend beyond traditional mean-field approaches to neural
  2579. networks. Our result suggests that areas across the brain operate
  2580. in a strongly coupled regime where dimensionality is under
  2581. sensitive control by net connectivity strength; moreover, we show
  2582. how this net connectivity strength is regulated by local
  2583. connectivity features, or synaptic motifs. This enables us to
  2584. interpret changes in dimensionality in terms of changes in
  2585. coupling among pairs and triplets of neurons. Analysis of
  2586. large-scale synaptic physiology datasets from both mouse and
  2587. human cortex then reveal the presence of synaptic coupling motifs
  2588. capable of substantially regulating this dimensionality. \#\#\#
  2589. Competing Interest Statement The authors have declared no
  2590. competing interest.",
  2591. journal = "Cold Spring Harbor Laboratory",
  2592. pages = "2020.11.02.365072",
  2593. month = nov,
  2594. year = 2020,
  2595. keywords = "RNN",
  2596. language = "en"
  2597. }
  2598. @ARTICLE{Arganda-Carreras2018-pm,
  2599. title = "A Statistically Representative Atlas for Mapping Neuronal
  2600. Circuits in the Drosophila Adult Brain",
  2601. author = "Arganda-Carreras, Ignacio and Manoliu, Tudor and Mazuras, Nicolas
  2602. and Schulze, Florian and Iglesias, Juan E and B{\"u}hler, Katja
  2603. and Jenett, Arnim and Rouyer, Fran{\c c}ois and Andrey, Philippe",
  2604. abstract = "Imaging the expression patterns of reporter constructs is a
  2605. powerful tool to dissect the neuronal circuits of perception and
  2606. behavior in the adult brain of Drosophila, one of the major
  2607. models for studying brain functions. To date, several Drosophila
  2608. brain templates and digital atlases have been built to
  2609. automatically analyze and compare collections of expression
  2610. pattern images. However, there has been no systematic comparison
  2611. of performances between alternative atlasing strategies and
  2612. registration algorithms. Here, we objectively evaluated the
  2613. performance of different strategies for building adult Drosophila
  2614. brain templates and atlases. In addition, we used
  2615. state-of-the-art registration algorithms to generate a new
  2616. group-wise inter-sex atlas. Our results highlight the benefit of
  2617. statistical atlases over individual ones and show that the newly
  2618. proposed inter-sex atlas outperformed existing solutions for
  2619. automated registration and annotation of expression patterns.
  2620. Over 3,000 images from the Janelia Farm FlyLight collection were
  2621. registered using the proposed strategy. These registered
  2622. expression patterns can be searched and compared with a new
  2623. version of the BrainBaseWeb system and BrainGazer software. We
  2624. illustrate the validity of our methodology and brain atlas with
  2625. registration-based predictions of expression patterns in a subset
  2626. of clock neurons. The described registration framework should
  2627. benefit to brain studies in Drosophila and other insect species.",
  2628. journal = "Front. Neuroinform.",
  2629. volume = 12,
  2630. pages = "13",
  2631. month = mar,
  2632. year = 2018,
  2633. keywords = "Drosophila adult brain; anatomical atlas; atlas-based image
  2634. segmentation; average brain template; brain mapping; confocal
  2635. microscopy; diffeomorphic image registration",
  2636. language = "en"
  2637. }
  2638. @ARTICLE{Genkin2020-jr,
  2639. title = "Moving beyond generalization to accurate interpretation of
  2640. flexible models",
  2641. author = "Genkin, Mikhail and Engel, Tatiana A",
  2642. abstract = "Machine learning optimizes flexible models to predict data. In
  2643. scientific applications, there is a rising interest in
  2644. interpreting these flexible models to derive hypotheses from
  2645. data. However, it is unknown whether good data prediction
  2646. guarantees the accurate interpretation of flexible models. Here,
  2647. we test this connection using a flexible, yet intrinsically
  2648. interpretable framework for modelling neural dynamics. We find
  2649. that many models discovered during optimization predict data
  2650. equally well, yet they fail to match the correct hypothesis. We
  2651. develop an alternative approach that identifies models with
  2652. correct interpretation by comparing model features across data
  2653. samples to separate true features from noise. We illustrate our
  2654. findings using recordings of spiking activity from the visual
  2655. cortex of monkeys performing a fixation task. Our results reveal
  2656. that good predictions cannot substitute for accurate
  2657. interpretation of flexible models and offer a principled approach
  2658. to identify models with correct interpretation.",
  2659. journal = "Nature Machine Intelligence",
  2660. month = oct,
  2661. year = 2020,
  2662. keywords = "RNN To read;RNN"
  2663. }
  2664. @UNPUBLISHED{Claudi2020-tb,
  2665. title = "Brainrender. A python based software for visualisation of
  2666. neuroanatomical and morphological data",
  2667. author = "Claudi, Federico and Tyson, Adam L and Branco, Tiago",
  2668. abstract = "Abstract Here we present brainrender, an open source python
  2669. package for rendering three-dimensional neuroanatomical data
  2670. aligned to the Allen Mouse Atlas. Brainrender can be used to
  2671. explore, visualise and compare data from publicly available
  2672. datasets (e.g. from the Mouse Light project from Janelia) as well
  2673. as data generated within individual laboratories. Brainrender
  2674. facilitates the exploration of neuroanatomical data with
  2675. three-dimensional renderings, aiding the design and
  2676. interpretation of experiments and the dissemination of anatomical
  2677. findings. Additionally, brainrender can also be used to generate
  2678. high-quality, publication-ready, figures for scientific
  2679. publications.",
  2680. journal = "Cold Spring Harbor Laboratory",
  2681. pages = "2020.02.23.961748",
  2682. month = feb,
  2683. year = 2020,
  2684. language = "en"
  2685. }
  2686. @MISC{Musy2019-vb,
  2687. title = "marcomusy/vtkplotter: vtkplotter",
  2688. author = "Musy, Marco and Dalmasso, Giovanni and Sullivan, Bane",
  2689. month = feb,
  2690. year = 2019
  2691. }
  2692. @UNPUBLISHED{Pachitariu2017-be,
  2693. title = "Suite2p: beyond 10,000 neurons with standard two-photon
  2694. microscopy",
  2695. author = "Pachitariu, M and Stringer, C and Dipoppa, M and Schr{\"o}der, S
  2696. and Rossi, L F and Dalgleish, H and Carandini, M and Harris, K D",
  2697. abstract = "Two-photon microscopy of calcium-dependent sensors has enabled
  2698. unprecedented recordings from vast populations of neurons. While
  2699. the sensors and microscopes have matured over several generations
  2700. of development, computational methods to process the resulting
  2701. movies remain inefficient and can give results that are hard to
  2702. interpret. Here we introduce Suite2p: a fast, accurate and
  2703. complete pipeline that registers raw movies, detects active
  2704. cells, extracts their calcium traces and infers their spike
  2705. times. Suite2p runs on standard workstations, operates faster
  2706. than real time, and recovers ~2 times more cells than the
  2707. previous state-of-the-art method. Its low computational load
  2708. allows routine detection of ~10,000 cells simultaneously with
  2709. standard two-photon resonant-scanning microscopes. Recordings at
  2710. this scale promise to reveal the fine structure of activity in
  2711. large populations of neurons or large populations of subcellular
  2712. structures such as synaptic boutons.",
  2713. journal = "bioRxiv",
  2714. pages = "30",
  2715. month = jul,
  2716. year = 2017,
  2717. language = "en"
  2718. }
  2719. @ARTICLE{Mathis2018-cn,
  2720. title = "{DeepLabCut}: markerless pose estimation of user-defined body
  2721. parts with deep learning",
  2722. author = "Mathis, Alexander and Mamidanna, Pranav and Cury, Kevin M and
  2723. Abe, Taiga and Murthy, Venkatesh N and Mathis, Mackenzie
  2724. Weygandt and Bethge, Matthias",
  2725. abstract = "Quantifying behavior is crucial for many applications in
  2726. neuroscience. Videography provides easy methods for the
  2727. observation and recording of animal behavior in diverse
  2728. settings, yet extracting particular aspects of a behavior for
  2729. further analysis can be highly time consuming. In motor control
  2730. studies, humans or other animals are often marked with
  2731. reflective markers to assist with computer-based tracking, but
  2732. markers are intrusive, and the number and location of the
  2733. markers must be determined a priori. Here we present an
  2734. efficient method for markerless pose estimation based on
  2735. transfer learning with deep neural networks that achieves
  2736. excellent results with minimal training data. We demonstrate the
  2737. versatility of this framework by tracking various body parts in
  2738. multiple species across a broad collection of behaviors.
  2739. Remarkably, even when only a small number of frames are labeled
  2740. (~200), the algorithm achieves excellent tracking performance on
  2741. test frames that is comparable to human accuracy.",
  2742. journal = "Nat. Neurosci.",
  2743. publisher = "nature.com",
  2744. volume = 21,
  2745. number = 9,
  2746. pages = "1281--1289",
  2747. month = sep,
  2748. year = 2018,
  2749. language = "en"
  2750. }
  2751. % The entry below contains non-ASCII chars that could not be converted
  2752. % to a LaTeX equivalent.
  2753. @ARTICLE{Ding2016-vn,
  2754. title = "Comprehensive cellular-resolution atlas of the adult human brain",
  2755. author = "Ding, Song-Lin and Royall, Joshua J and Sunkin, Susan M and Ng,
  2756. Lydia and Facer, Benjamin A C and Lesnar, Phil and
  2757. Guillozet-Bongaarts, Angie and McMurray, Bergen and Szafer, Aaron
  2758. and Dolbeare, Tim A and Stevens, Allison and Tirrell, Lee and
  2759. Benner, Thomas and Caldejon, Shiella and Dalley, Rachel A and
  2760. Dee, Nick and Lau, Christopher and Nyhus, Julie and Reding,
  2761. Melissa and Riley, Zackery L and Sandman, David and Shen, Elaine
  2762. and van der Kouwe, Andre and Varjabedian, Ani and Wright,
  2763. Michelle and Z{\"o}llei, Lilla and Dang, Chinh and Knowles, James
  2764. A and Koch, Christof and Phillips, John W and Sestan, Nenad and
  2765. Wohnoutka, Paul and Zielke, H Ronald and Hohmann, John G and
  2766. Jones, Allan R and Bernard, Amy and Hawrylycz, Michael J and Hof,
  2767. Patrick R and Fischl, Bruce and Lein, Ed S",
  2768. abstract = "Detailed anatomical understanding of the human brain is essential
  2769. for unraveling its functional architecture, yet current reference
  2770. atlases have major limitations such as lack of whole-brain
  2771. coverage, relatively low image resolution, and sparse structural
  2772. annotation. We present the first digital human brain atlas to
  2773. incorporate neuroimaging, high-resolution histology, and
  2774. chemoarchitecture across a complete adult female brain,
  2775. consisting of magnetic resonance imaging (MRI),
  2776. diffusion-weighted imaging (DWI), and 1,356 large-format cellular
  2777. resolution (1 µm/pixel) Nissl and immunohistochemistry anatomical
  2778. plates. The atlas is comprehensively annotated for 862
  2779. structures, including 117 white matter tracts and several novel
  2780. cyto- and chemoarchitecturally defined structures, and these
  2781. annotations were transferred onto the matching MRI dataset.
  2782. Neocortical delineations were done for sulci, gyri, and modified
  2783. Brodmann areas to link macroscopic anatomical and microscopic
  2784. cytoarchitectural parcellations. Correlated neuroimaging and
  2785. histological structural delineation allowed fine feature
  2786. identification in MRI data and subsequent structural
  2787. identification in MRI data from other brains. This interactive
  2788. online digital atlas is integrated with existing Allen Institute
  2789. for Brain Science gene expression atlases and is publicly
  2790. accessible as a resource for the neuroscience community. J. Comp.
  2791. Neurol. 524:3127-3481, 2016. \copyright{} 2016 The Authors The
  2792. Journal of Comparative Neurology Published by Wiley Periodicals,
  2793. Inc.",
  2794. journal = "J. Comp. Neurol.",
  2795. volume = 524,
  2796. number = 16,
  2797. pages = "3127--3481",
  2798. month = nov,
  2799. year = 2016,
  2800. keywords = "AB\_2314904; DWI; MRI; RRIDs: AB\_10000343; SCR\_014329;
  2801. amygdala; brain atlas; brainstem; cerebellum; cerebral cortex;
  2802. cytoarchitecture; hippocampal formation; hypothalamus;
  2803. neurofilament protein; parvalbumin; thalamus",
  2804. language = "en"
  2805. }
  2806. @UNPUBLISHED{Tyson2020-mq,
  2807. title = "A deep learning algorithm for {3D} cell detection in whole mouse
  2808. brain image datasets",
  2809. author = "Tyson, Adam L and Rousseau, Charly V and Niedworok, Christian J
  2810. and Keshavarzi, Sepiedeh and Tsitoura, Chryssanthi and Margrie,
  2811. Troy W",
  2812. abstract = "Understanding the function of the nervous system necessitates
  2813. mapping the spatial distributions of its constituent cells
  2814. defined by function, anatomy or gene expression. Recently,
  2815. developments in tissue preparation and microscopy allow cellular
  2816. populations to be imaged throughout the entire rodent brain.
  2817. However, mapping these neurons manually is prone to bias and is
  2818. often impractically time consuming. Here we present an
  2819. open-source algorithm for fully automated 3D detection of
  2820. neuronal somata in mouse whole-brain microscopy images using
  2821. standard desktop computer hardware. We demonstrate the
  2822. applicability and power of our approach by mapping the brain-wide
  2823. locations of large populations of cells labeled with cytoplasmic
  2824. fluorescent proteins expressed via retrograde trans-synaptic
  2825. viral infection. \#\#\# Competing Interest Statement The authors
  2826. have declared no competing interest.",
  2827. journal = "Cold Spring Harbor Laboratory",
  2828. pages = "2020.10.21.348771",
  2829. month = oct,
  2830. year = 2020,
  2831. language = "en"
  2832. }
  2833. @ARTICLE{Claudi2020-go,
  2834. title = "{BrainGlobe} Atlas {API}: a common interface for neuroanatomical
  2835. atlases",
  2836. author = "Claudi, Federico and Petrucco, Luigi and Tyson, Adam and Branco,
  2837. Tiago and Margrie, Troy and Portugues, Ruben",
  2838. abstract = "Software archive",
  2839. journal = "JOSS",
  2840. volume = 5,
  2841. number = 54,
  2842. pages = "2668",
  2843. month = oct,
  2844. year = 2020
  2845. }
  2846. @ARTICLE{Kunst2019-dy,
  2847. title = "A {Cellular-Resolution} Atlas of the Larval Zebrafish Brain",
  2848. author = "Kunst, Michael and Laurell, Eva and Mokayes, Nouwar and Kramer,
  2849. Anna and Kubo, Fumi and Fernandes, Ant{\'o}nio M and F{\"o}rster,
  2850. Dominique and Dal Maschio, Marco and Baier, Herwig",
  2851. abstract = "Understanding brain-wide neuronal dynamics requires a detailed
  2852. map of the underlying circuit architecture. We built an
  2853. interactive cellular-resolution atlas of the zebrafish brain at 6
  2854. days post-fertilization (dpf) based on the reconstructions of
  2855. over 2,000 individually GFP-labeled neurons. We clustered our
  2856. dataset in ``morphotypes,'' establishing a unique database of
  2857. quantitatively described neuronal morphologies together with
  2858. their spatial coordinates in vivo. Over 100 transgene expression
  2859. patterns were imaged separately and co-registered with the
  2860. single-neuron atlas. By annotating 72 non-overlapping brain
  2861. regions, we generated from our dataset an inter-areal wiring
  2862. diagram of the larval brain, which serves as ground truth for
  2863. synapse-scale, electron microscopic reconstructions.
  2864. Interrogating our atlas by ``virtual tract tracing'' has already
  2865. revealed previously unknown wiring principles in the tectum and
  2866. the cerebellum. In conclusion, we present here an evolving
  2867. computational resource and visualization tool, which will be
  2868. essential to map function to structure in a vertebrate brain.
  2869. VIDEO ABSTRACT.",
  2870. journal = "Neuron",
  2871. volume = 103,
  2872. number = 1,
  2873. pages = "21--38.e5",
  2874. month = jul,
  2875. year = 2019,
  2876. keywords = "brain networks; cerebellum; connectomics; digital atlas;
  2877. neuroanatomy; single-cell tracing; tectum; tissue clearing",
  2878. language = "en"
  2879. }
  2880. @ARTICLE{Bates2020-fa,
  2881. title = "The natverse, a versatile toolbox for combining and analysing
  2882. neuroanatomical data",
  2883. author = "Bates, Alexander Shakeel and Manton, James D and Jagannathan,
  2884. Sridhar R and Costa, Marta and Schlegel, Philipp and Rohlfing,
  2885. Torsten and Jefferis, Gregory Sxe",
  2886. abstract = "To analyse neuron data at scale, neuroscientists expend
  2887. substantial effort reading documentation, installing dependencies
  2888. and moving between analysis and visualisation environments. To
  2889. facilitate this, we have developed a suite of interoperable
  2890. open-source R packages called the natverse. The natverse allows
  2891. users to read local and remote data, perform popular analyses
  2892. including visualisation and clustering and graph-theoretic
  2893. analysis of neuronal branching. Unlike most tools, the natverse
  2894. enables comparison across many neurons of morphology and
  2895. connectivity after imaging or co-registration within a common
  2896. template space. The natverse also enables transformations between
  2897. different template spaces and imaging modalities. We demonstrate
  2898. tools that integrate the vast majority of Drosophila
  2899. neuroanatomical light microscopy and electron microscopy
  2900. connectomic datasets. The natverse is an easy-to-use environment
  2901. for neuroscientists to solve complex, large-scale analysis
  2902. challenges as well as an open platform to create new code and
  2903. packages to share with the community.",
  2904. journal = "Elife",
  2905. volume = 9,
  2906. month = apr,
  2907. year = 2020,
  2908. keywords = "D. melanogaster; analysis software; computational biology;
  2909. connectomics; mouse; neural circuits; neuroanatomy; neuronal
  2910. morphology; neuroscience; open-source; systems biology; zebrafish",
  2911. language = "en"
  2912. }
  2913. @MISC{Tyson2020-tt,
  2914. title = "brainreg: automated {3D} brain registration with support for
  2915. multiple species and atlases",
  2916. author = "Tyson, Adam L and Rousseau, Charly V and Margrie, Troy W",
  2917. month = aug,
  2918. year = 2020
  2919. }
  2920. @ARTICLE{Wang2020-ee,
  2921. title = "The Allen Mouse Brain Common Coordinate Framework: A {3D}
  2922. Reference Atlas",
  2923. author = "Wang, Quanxin and Ding, Song-Lin and Li, Yang and Royall, Josh
  2924. and Feng, David and Lesnar, Phil and Graddis, Nile and Naeemi,
  2925. Maitham and Facer, Benjamin and Ho, Anh and Dolbeare, Tim and
  2926. Blanchard, Brandon and Dee, Nick and Wakeman, Wayne and Hirokawa,
  2927. Karla E and Szafer, Aaron and Sunkin, Susan M and Oh, Seung Wook
  2928. and Bernard, Amy and Phillips, John W and Hawrylycz, Michael and
  2929. Koch, Christof and Zeng, Hongkui and Harris, Julie A and Ng,
  2930. Lydia",
  2931. abstract = "Summary Recent large-scale collaborations are generating major
  2932. surveys of cell types and connections in the mouse brain,
  2933. collecting large amounts of data across modalities, spatial
  2934. scales, and brain areas. Successful integration of these data
  2935. requires a standard 3D reference atlas. Here, we present the
  2936. Allen Mouse Brain Common Coordinate Framework (CCFv3) as such a
  2937. resource. We constructed an average template brain at 10 $\mu$m
  2938. voxel resolution by interpolating high resolution in-plane serial
  2939. two-photon tomography images with 100 $\mu$m z-sampling from
  2940. 1,675 young adult C57BL/6J mice. Then, using multimodal reference
  2941. data, we parcellated the entire brain directly in 3D, labeling
  2942. every voxel with a brain structure spanning 43 isocortical areas
  2943. and their layers, 329 subcortical gray matter structures, 81
  2944. fiber tracts, and 8 ventricular structures. CCFv3 can be used to
  2945. analyze, visualize, and integrate multimodal and multiscale
  2946. datasets in 3D and is openly accessible
  2947. (https://atlas.brain-map.org/).",
  2948. journal = "Cell",
  2949. volume = 181,
  2950. number = 4,
  2951. pages = "936--953.e20",
  2952. month = may,
  2953. year = 2020,
  2954. keywords = "average mouse brain; reference atlas; 3D brain atlas; brain
  2955. parcellation; brain anatomy; mouse cortex; common coordinate
  2956. framework; CCFv3; fiber tracts; transgenic mice"
  2957. }
  2958. @ARTICLE{Kleinman_undated-cx,
  2959. title = "Recurrent neural network models of multi-area computation
  2960. underlying decision-making",
  2961. author = "Kleinman, Michael and Chandrasekaran, Chandramouli and Kao,
  2962. Jonathan C"
  2963. }
  2964. @ARTICLE{Maheswaranathan2019-ux,
  2965. title = "Universality and individuality in neural dynamics across large
  2966. populations of recurrent networks",
  2967. author = "Maheswaranathan, Niru and Williams, Alex H and Golub, Matthew D
  2968. and Ganguli, Surya and Sussillo, David",
  2969. abstract = "Task-based modeling with recurrent neural networks (RNNs) has
  2970. emerged as a popular way to infer the computational function of
  2971. different brain regions. These models are quantitatively assessed
  2972. by comparing the low-dimensional neural representations of the
  2973. model with the brain, for example using canonical correlation
  2974. analysis (CCA). However, the nature of the detailed
  2975. neurobiological inferences one can draw from such efforts remains
  2976. elusive. For example, to what extent does training neural
  2977. networks to solve common tasks uniquely determine the network
  2978. dynamics, independent of modeling architectural choices? Or
  2979. alternatively, are the learned dynamics highly sensitive to
  2980. different model choices? Knowing the answer to these questions
  2981. has strong implications for whether and how we should use
  2982. task-based RNN modeling to understand brain dynamics. To address
  2983. these foundational questions, we study populations of thousands
  2984. of networks, with commonly used RNN architectures, trained to
  2985. solve neuroscientifically motivated tasks and characterize their
  2986. nonlinear dynamics. We find the geometry of the RNN
  2987. representations can be highly sensitive to different network
  2988. architectures, yielding a cautionary tale for measures of
  2989. similarity that rely on representational geometry, such as CCA.
  2990. Moreover, we find that while the geometry of neural dynamics can
  2991. vary greatly across architectures, the underlying computational
  2992. scaffold-the topological structure of fixed points, transitions
  2993. between them, limit cycles, and linearized dynamics-often appears
  2994. universal across all architectures.",
  2995. journal = "Adv. Neural Inf. Process. Syst.",
  2996. volume = 2019,
  2997. pages = "15629--15641",
  2998. month = dec,
  2999. year = 2019,
  3000. keywords = "RNN",
  3001. language = "en"
  3002. }
  3003. @ARTICLE{Vyas2020-pw,
  3004. title = "Computation Through Neural Population Dynamics",
  3005. author = "Vyas, Saurabh and Golub, Matthew D and Sussillo, David and
  3006. Shenoy, Krishna V",
  3007. abstract = "Significant experimental, computational, and theoretical work has
  3008. identified rich structure within the coordinated activity of
  3009. interconnected neural populations. An emerging challenge now is
  3010. to uncover the nature of the associated computations, how they
  3011. are implemented, and what role they play in driving behavior. We
  3012. term this computation through neural population dynamics. If
  3013. successful, this framework will reveal general motifs of neural
  3014. population activity and quantitatively describe how neural
  3015. population dynamics implement computations necessary for driving
  3016. goal-directed behavior. Here, we start with a mathematical primer
  3017. on dynamical systems theory and analytical tools necessary to
  3018. apply this perspective to experimental data. Next, we highlight
  3019. some recent discoveries resulting from successful application of
  3020. dynamical systems. We focus on studies spanning motor control,
  3021. timing, decision-making, and working memory. Finally, we briefly
  3022. discuss promising recent lines of investigation and future
  3023. directions for the computation through neural population dynamics
  3024. framework.",
  3025. journal = "Annu. Rev. Neurosci.",
  3026. volume = 43,
  3027. pages = "249--275",
  3028. month = jul,
  3029. year = 2020,
  3030. keywords = "dynamical systems; neural computation; neural population
  3031. dynamics; state spaces;RNN",
  3032. language = "en"
  3033. }
  3034. @ARTICLE{Bellardita2015-ut,
  3035. title = "Phenotypic characterization of speed-associated gait changes in
  3036. mice reveals modular organization of locomotor networks",
  3037. author = "Bellardita, Carmelo and Kiehn, Ole",
  3038. abstract = "Studies of locomotion in mice suggest that circuits controlling
  3039. the alternating between left and right limbs may have a modular
  3040. organization with distinct locomotor circuits being recruited at
  3041. different speeds. It is not clear, however, whether such a
  3042. modular organization reflects specific behavioral outcomes
  3043. expressed at different speeds of locomotion. Here, we use
  3044. detailed kinematic analyses to search for signatures of a
  3045. modular organization of locomotor circuits in intact and
  3046. genetically modified mice moving at different speeds of
  3047. locomotion. We show that wild-type mice display three distinct
  3048. gaits: two alternating, walk and trot, and one synchronous,
  3049. bound. Each gait is expressed in distinct ranges of speed with
  3050. phenotypic inter-limb and intra-limb coordination. A fourth
  3051. gait, gallop, closely resembled bound in most of the locomotor
  3052. parameters but expressed diverse inter-limb coordination.
  3053. Genetic ablation of commissural V0V neurons completely removed
  3054. the expression of one alternating gait, trot, but left intact
  3055. walk, gallop, and bound. Ablation of commissural V0V and V0D
  3056. neurons led to a loss of walk, trot, and gallop, leaving bound
  3057. as the default gait. Our study provides a benchmark for studies
  3058. of the neuronal control of locomotion in the full range of
  3059. speeds. It provides evidence that gait expression depends upon
  3060. selection of different modules of neuronal ensembles.",
  3061. journal = "Curr. Biol.",
  3062. publisher = "Elsevier",
  3063. volume = 25,
  3064. number = 11,
  3065. pages = "1426--1436",
  3066. month = jun,
  3067. year = 2015,
  3068. keywords = "Locomotion",
  3069. language = "en"
  3070. }
  3071. @ARTICLE{Carmelo_Bellardita_undated-rj,
  3072. title = "Phenotypic Characterization of {Speed-Associated} Gait Changes in
  3073. Mice Reveals Modular Organization of Locomotor Networks",
  3074. author = "Carmelo Bellardita, Ole Kiehn",
  3075. keywords = "Locomotion"
  3076. }
  3077. @ARTICLE{Vyas2020-cr,
  3078. title = "Computation Through Neural Population Dynamics",
  3079. author = "Vyas, Saurabh and Golub, Matthew D and Sussillo, David and
  3080. Shenoy, Krishna V",
  3081. abstract = "Significant experimental, computational, and theoretical work has
  3082. identified rich structure within the coordinated activity of
  3083. interconnected neural populations. An emerging challenge now is
  3084. to uncover the nature of the associated computations, how they
  3085. are implemented, and what role they play in driving behavior. We
  3086. term this computation through neural population dynamics. If
  3087. successful, this framework will reveal general motifs of neural
  3088. population activity and quantitatively describe how neural
  3089. population dynamics implement computations necessary for driving
  3090. goal-directed behavior. Here, we start with a mathematical primer
  3091. on dynamical systems theory and analytical tools necessary to
  3092. apply this perspective to experimental data. Next, we highlight
  3093. some recent discoveries resulting from successful application of
  3094. dynamical systems. We focus on studies spanning motor control,
  3095. timing, decision-making, and working memory. Finally, we briefly
  3096. discuss promising recent lines of investigation and future
  3097. directions for the computation through neural population dynamics
  3098. framework.",
  3099. journal = "Annu. Rev. Neurosci.",
  3100. volume = 43,
  3101. pages = "249--275",
  3102. month = jul,
  3103. year = 2020,
  3104. keywords = "dynamical systems; neural computation; neural population
  3105. dynamics; state spaces",
  3106. language = "en"
  3107. }
  3108. @UNPUBLISHED{De_Cothi2020-kd,
  3109. title = "Predictive Maps in Rats and Humans for Spatial Navigation",
  3110. author = "de Cothi, William and Nyberg, Nils and Griesbauer, Eva-Maria and
  3111. Ghaname, Carole and Zisch, Fiona and Fletcher, Lydia and Newton,
  3112. Charlotte and Renaudineau, Sophie and Bendor, Daniel and Grieves,
  3113. Roddy and Duvelle, Eleonore and Barry, Caswell and Spiers, Hugo J",
  3114. abstract = "Much of our understanding of navigation has come from the study
  3115. of rats, humans and simulated artificial agents. To date little
  3116. attempt has been made to integrate these approaches into a common
  3117. framework to understand mechanisms that may be shared across
  3118. mammals and the extent to which different instantiations of
  3119. agents best capture mammalian navigation behaviour. Here, we
  3120. report a comparison of rats, humans and reinforcement learning
  3121. (RL) agents in a novel open-field navigation task (Tartarus Maze)
  3122. requiring dynamic adaptation (shortcuts and detours) to changing
  3123. obstructions in the path to the goal. We find humans and rats are
  3124. remarkably similar in patterns of choice in the task. The
  3125. patterns in their choices, dwell maps and changes over time
  3126. reveal that both species show the greatest similarity to RL
  3127. agents utilising a predictive map: the successor representation.
  3128. Humans also display trajectory features similar to a model-based
  3129. RL agent. Our findings have implications for models seeking to
  3130. explain mammalian navigation in dynamic environments and
  3131. highlight the utility of modelling the behaviour of different
  3132. species in the same frame-work in comparison to RL agents to
  3133. uncover the potential mechanisms used for behaviour. \#\#\#
  3134. Competing Interest Statement The authors have declared no
  3135. competing interest.",
  3136. pages = "2020.09.26.314815",
  3137. month = sep,
  3138. year = 2020,
  3139. language = "en"
  3140. }
  3141. % The entry below contains non-ASCII chars that could not be converted
  3142. % to a LaTeX equivalent.
  3143. @ARTICLE{Walker2018-qp,
  3144. title = "A comparison of two types of running wheel in terms of mouse
  3145. preference, health, and welfare",
  3146. author = "Walker, Michael and Mason, Georgia",
  3147. abstract = "Voluntary wheel running occurs in mice of all strains, sexes, and
  3148. ages. Mice find voluntary wheel running rewarding, and it leads
  3149. to numerous health benefits. For this reason wheels are used both
  3150. to enhance welfare and to create models of exercise. However,
  3151. many designs of running wheel are used. This makes between-study
  3152. comparisons difficult, as this variability could potentially
  3153. affect the amount, pattern, and/or intensity of running
  3154. behaviour, and thence the wheels' effects on welfare and
  3155. exercise-related changes in anatomy and physiology. This study
  3156. therefore evaluated two commercially available models, chosen
  3157. because safe for group-housed mice: Bio Serv\textregistered{}'s
  3158. ``fast-trac'' wheel combo and Ware Manufacturing Inc.'s stainless
  3159. steel mesh 5″ upright wheel. Working with a total of three
  3160. hundred and fifty one female C57BL/6, DBA/2 and BALB/c mice, we
  3161. assessed these wheels' relative utilization by mice when access
  3162. was free; the strength of motivation for each wheel-type when
  3163. access required crossing an electrified grid; and the impact each
  3164. wheel had on mouse well-being (inferred from acoustic startle
  3165. responses and neophobia) and exercise-related anatomical changes
  3166. (BMI; heart and hind limb masses). Mice ran more on the
  3167. ``fast-trac'' wheel regardless of whether both wheel-types were
  3168. available at once, or only if one was present. In terms of
  3169. motivation, subjects required to work to access a single wheel
  3170. worked equally hard for both wheel-types (even if locked and thus
  3171. not useable for running), but if provided with one working wheel
  3172. for free and the other type of wheel (again unlocked) accessible
  3173. via crossing the electrified grid, the ``fast-trac'' wheel
  3174. emerged as more motivating, as the Maximum Price Paid for the
  3175. Ware metal wheel was lower than that paid for the ``fast-trac''
  3176. plastic wheel, at least for C57BL/6s and DBA/2s. No deleterious
  3177. consequences were noted with either wheel in terms of health and
  3178. welfare, but only mice with plastic wheels developed
  3179. significantly larger hearts and hind limbs than control animals
  3180. with locked wheels. Thus, where differences emerged, Bio
  3181. Serv\textregistered{}'s ``fast-trac'' wheel combos appeared to
  3182. better meet the aims of exercise provision than Ware
  3183. Manufacturing's steel upright wheels.",
  3184. journal = "Physiol. Behav.",
  3185. volume = 191,
  3186. pages = "82--90",
  3187. month = jul,
  3188. year = 2018,
  3189. keywords = "Health; Motivation; Mouse; Preference; Welfare; Wheel running",
  3190. language = "en"
  3191. }
  3192. @ARTICLE{Lemieux2016-fx,
  3193. title = "{Speed-Dependent} Modulation of the Locomotor Behavior in Adult
  3194. Mice Reveals Attractor and Transitional Gaits",
  3195. author = "Lemieux, Maxime and Josset, Nicolas and Roussel, Marie and
  3196. Couraud, S{\'e}bastien and Bretzner, Fr{\'e}d{\'e}ric",
  3197. abstract = "Locomotion results from an interplay between biomechanical
  3198. constraints of the muscles attached to the skeleton and the
  3199. neuronal circuits controlling and coordinating muscle activities.
  3200. Quadrupeds exhibit a wide range of locomotor gaits. Given our
  3201. advances in the genetic identification of spinal and supraspinal
  3202. circuits important to locomotion in the mouse, it is now
  3203. important to get a better understanding of the full repertoire of
  3204. gaits in the freely walking mouse. To assess this range, young
  3205. adult C57BL/6J mice were trained to walk and run on a treadmill
  3206. at different locomotor speeds. Instead of using the classical
  3207. paradigm defining gaits according to their footfall pattern, we
  3208. combined the inter-limb coupling and the duty cycle of the stance
  3209. phase, thus identifying several types of gaits: lateral walk,
  3210. trot, out-of-phase walk, rotary gallop, transverse gallop, hop,
  3211. half-bound, and full-bound. Out-of-phase walk, trot, and
  3212. full-bound were robust and appeared to function as attractor
  3213. gaits (i.e., a state to which the network flows and stabilizes)
  3214. at low, intermediate, and high speeds respectively. In contrast,
  3215. lateral walk, hop, transverse gallop, rotary gallop, and
  3216. half-bound were more transient and therefore considered
  3217. transitional gaits (i.e., a labile state of the network from
  3218. which it flows to the attractor state). Surprisingly, lateral
  3219. walk was less frequently observed. Using graph analysis, we
  3220. demonstrated that transitions between gaits were predictable, not
  3221. random. In summary, the wild-type mouse exhibits a wider
  3222. repertoire of locomotor gaits than expected. Future locomotor
  3223. studies should benefit from this paradigm in assessing transgenic
  3224. mice or wild-type mice with neurotraumatic injury or
  3225. neurodegenerative disease affecting gait.",
  3226. journal = "Front. Neurosci.",
  3227. volume = 10,
  3228. pages = "42",
  3229. month = feb,
  3230. year = 2016,
  3231. keywords = "graph analysis; kinematic; locomotor gaits; mouse; speed;
  3232. steady-state;Locomotion",
  3233. language = "en"
  3234. }
  3235. @ARTICLE{Herbin2006-mc,
  3236. title = "How does a mouse increase its velocity? A model for investigation
  3237. in the control of locomotion",
  3238. author = "Herbin, Marc and Gasc, Jean-Pierre and Renous, Sabine",
  3239. abstract = "We analysed treadmill locomotion of the adult SWISS-OF1 mice over
  3240. a large range of velocities. The use of a high-speed video camera
  3241. combined with cinefluoroscopic equipment allowed us to quantify
  3242. in detail the various space and time parameters of limb
  3243. kinematics. We find that velocity adjustments depend upon whether
  3244. animal used a symmetrical or non-symmetrical gait. In symmetrical
  3245. gaits, the increase of velocity generally results equally from an
  3246. increase in the stride frequency and the stride length. On the
  3247. other hand, in non-symmetrical gaits, the increase in velocity is
  3248. achieved differently according to the level of velocity used. As
  3249. speed increases, velocity increases first as a consequence of
  3250. increased stride frequency, then as in symmetrical gaits, by an
  3251. equal increase in both variables, and finally at high speed,
  3252. velocity increases through increased stride length. In both
  3253. symmetrical and non-symmetrical gaits, stance and swing-time
  3254. shortening contributed to the increase of the stride frequency,
  3255. with stance time decrease being the major contributor. The
  3256. pattern of locomotion obtained in the present study may be used
  3257. as a model mouse system for studying locomotor deficits resulting
  3258. from specific mutations in the nervous system. To cite this
  3259. article: M. Herbin et al., C. R. Palevol 5 (2006). R{\'e}sum{\'e}
  3260. Comment la souris augmente-elle sa vitesse ? Un mod{\`e}le pour
  3261. la recherche sur le contr{\^o}le moteur de la locomotion. La
  3262. locomotion sur tapis roulant de la souche de souris SWISS-OF1 a
  3263. {\'e}t{\'e} analys{\'e}e {\`a} travers une large gamme de
  3264. vitesses. L'utilisation de la vid{\'e}oradiographie {\`a} grande
  3265. vitesse a permis de quantifier de fa{\c c}on tr{\`e}s
  3266. d{\'e}taill{\'e}e tous les param{\`e}tres de la cin{\'e}matique
  3267. du membre de r{\'e}f{\'e}rence. Les r{\'e}sultats ainsi obtenus
  3268. montrent que la fr{\'e}quence et l'enjamb{\'e}e n'interviennent
  3269. pas de la m{\^e}me fa{\c c}on dans l'augmentation de la vitesse,
  3270. selon l'allure utilis{\'e}e. Lorsque l'animal est en allure
  3271. sym{\'e}trique, l'augmentation de la vitesse est
  3272. g{\'e}n{\'e}ralement obtenue par une {\'e}gale augmentation de la
  3273. fr{\'e}quence et de l'enjamb{\'e}e. En revanche, si la souris
  3274. utilise une allure non sym{\'e}trique, l'augmentation de la
  3275. vitesse est obtenue diff{\'e}remment selon la valeur de cette
  3276. derni{\`e}re. L'augmentation de la vitesse est d'abord surtout
  3277. assur{\'e}e par une augmentation de la fr{\'e}quence, puis par
  3278. l'augmentation {\'e}gale des deux variables et enfin surtout par
  3279. l'augmentation de l'enjamb{\'e}e. L'augmentation de la
  3280. fr{\'e}quence est, en revanche, surtout assur{\'e}e par une
  3281. diminution de la dur{\'e}e du pos{\'e} et cela, quelle que soit
  3282. l'allure utilis{\'e}e. Cette mod{\'e}lisation de la locomotion
  3283. normale de la souris pourra {\^e}tre utilis{\'e}e comme
  3284. r{\'e}f{\'e}rentiel pour les {\'e}tudes portant sur les
  3285. d{\'e}ficits moteurs de certaines souches de souris mutantes ou
  3286. transg{\'e}niques. Pour citer cet article : M. Herbin et al., C.
  3287. R. Palevol 5 (2006).",
  3288. journal = "C. R. Palevol",
  3289. volume = 5,
  3290. number = 3,
  3291. pages = "531--540",
  3292. month = mar,
  3293. year = 2006,
  3294. keywords = "Stride frequency; Stride length; Treadmill; Locomotion;
  3295. SWISS-OF1; Fr{\'e}quence; Enjamb{\'e}e; Tapis roulant;
  3296. Locomotion; SWISS-OF1;Locomotion"
  3297. }
  3298. @ARTICLE{Walter2003-pb,
  3299. title = "Kinematics of 90 degrees running turns in wild mice",
  3300. author = "Walter, Rebecca M",
  3301. abstract = "Turning is a requirement for locomotion on the variable terrain
  3302. that most terrestrial animals inhabit and is a deciding factor in
  3303. many predator-prey interactions. Despite this, the kinematics and
  3304. mechanics of quadrupedal turns are not well understood. To gain
  3305. insight to the turning kinematics of small quadrupedal mammals,
  3306. six adult wild mice were videotaped at 250 Hz from below as they
  3307. performed 90 degrees running turns. Four markers placed along the
  3308. sagittal axis were digitized to allow observation of lateral
  3309. bending and body rotation throughout the turn. Ground contact
  3310. periods of the fore- and hindlimbs were also noted for each
  3311. frame. During turning, mice increased their ground contact time,
  3312. but did not change their stride frequency relative to straight
  3313. running at maximum speed. Postcranial body rotation preceded
  3314. deflection in heading, and did not occur in one continuous
  3315. motion, but rather in bouts of 15-53 degrees. These bouts were
  3316. synchronized with the stride cycle, such that the majority of
  3317. rotation occurred during the second half of forelimb support and
  3318. the first half of hindlimb support. In this phase of the stride
  3319. cycle, the trunk was sagittally flexed and rotational inertia was
  3320. 65\% of that during maximal extension. By synchronizing body
  3321. rotation with this portion of the stride cycle, mice can achieve
  3322. a given angular acceleration with much lower applied torque.
  3323. Compared with humans running along curved trajectories, mice
  3324. maintained relatively higher speeds at proportionately smaller
  3325. radii. A possible explanation for this difference lies in the
  3326. more crouched limb posture of mice, which increases the
  3327. mechanical advantage for horizontal ground force production. The
  3328. occurrence of body rotation prior to deflection in heading may
  3329. facilitate acceleration in the new direction by making use of the
  3330. relatively greater force production inherent in the parasagittal
  3331. limb posture of mice.",
  3332. journal = "J. Exp. Biol.",
  3333. volume = 206,
  3334. number = "Pt 10",
  3335. pages = "1739--1749",
  3336. month = may,
  3337. year = 2003,
  3338. language = "en"
  3339. }
  3340. @ARTICLE{Herbin2004-ma,
  3341. title = "Symmetrical and asymmetrical gaits in the mouse: patterns to
  3342. increase velocity",
  3343. author = "Herbin, Marc and Gasc, Jean-Pierre and Renous, Sabine",
  3344. abstract = "The gaits of the adult SWISS mice during treadmill locomotion at
  3345. velocities ranging from 15 to 85 cm s(-1) have been analysed
  3346. using a high-speed video camera combined with cinefluoroscopic
  3347. equipment. The sequences of locomotion were analysed to determine
  3348. the various space and time parameters of limb kinematics. We
  3349. found that velocity adjustments are accounted for differently by
  3350. the stride frequency and the stride length if the animal showed a
  3351. symmetrical or an asymmetrical gait. In symmetrical gaits, the
  3352. increase of velocity is provided by an equal increase in the
  3353. stride length and the stride frequency. In asymmetrical gaits,
  3354. the increase in velocity is mainly assured by an increase in the
  3355. stride frequency in velocities ranging from 15 to 29 cm s(-1).
  3356. Above 68 cm s(-1), velocity increase is achieved by stride length
  3357. increase. In velocities ranging from 29 to 68 cm s(-1), the
  3358. contribution of both variables is equal as in symmetrical gaits.
  3359. Both stance time and swing time shortening contributed to the
  3360. increase of the stride frequency in both gaits, though with a
  3361. major contribution from stance time decrease. The pattern of
  3362. locomotion obtained in a normal mouse should be used as a
  3363. template for studying locomotor control deficits after lesions or
  3364. in different mutations affecting the nervous system.",
  3365. journal = "J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol.",
  3366. volume = 190,
  3367. number = 11,
  3368. pages = "895--906",
  3369. month = nov,
  3370. year = 2004,
  3371. keywords = "Locomotion",
  3372. language = "en"
  3373. }
  3374. @ARTICLE{Josset2018-js,
  3375. title = "Distinct Contributions of Mesencephalic Locomotor Region Nuclei
  3376. to Locomotor Control in the Freely Behaving Mouse",
  3377. author = "Josset, Nicolas and Roussel, Marie and Lemieux, Maxime and
  3378. Lafrance-Zoubga, David and Rastqar, Ali and Bretzner, Frederic",
  3379. abstract = "The mesencephalic locomotor region (MLR) has been initially
  3380. identified as a supraspinal center capable of initiating and
  3381. modulating locomotion. Whereas its functional contribution to
  3382. locomotion has been widely documented throughout the phylogeny
  3383. from the lamprey to humans, there is still debate about its exact
  3384. organization. Combining kinematic and electrophysiological
  3385. recordings in mouse genetics, our study reveals that
  3386. glutamatergic neurons of the cuneiform nucleus initiate
  3387. locomotion and induce running gaits, whereas glutamatergic and
  3388. cholinergic neurons of the pedunculopontine nucleus modulate
  3389. locomotor pattern and rhythm, contributing to slow-walking gaits.
  3390. By initiating, modulating, and accelerating locomotion, our study
  3391. identifies and characterizes distinct neuronal populations of
  3392. this functional region important to locomotor command.",
  3393. journal = "Curr. Biol.",
  3394. volume = 28,
  3395. number = 6,
  3396. pages = "884--901.e3",
  3397. month = mar,
  3398. year = 2018,
  3399. keywords = "cuneiform nucleus; electrophysiology; glutamatergic and
  3400. cholinergic neurons; kinematic analysis; locomotor command;
  3401. locomotor pattern rhythm and gait; mesencephalic locomotor
  3402. region; optogenetic tools; pedunculopontine nucleus;Locomotion",
  3403. language = "en"
  3404. }
  3405. @ARTICLE{Herbin2007-us,
  3406. title = "Gait parameters of treadmill versus overground locomotion in
  3407. mouse",
  3408. author = "Herbin, Marc and Hackert, R{\'e}mi and Gasc, Jean-Pierre and
  3409. Renous, Sabine",
  3410. abstract = "Many studies of interest in motor behaviour and motor impairment
  3411. in mice use equally treadmill or track as a routine test.
  3412. However, the literature in mammals shows a wide difference of
  3413. results between the kinematics of treadmill and overground
  3414. locomotion. To study these discrepancies, we analyzed the
  3415. locomotion of adult SWISS-OF1 mice over a large range of
  3416. velocities using treadmill and overground track. The use of a
  3417. high-speed video camera combined with cinefluoroscopic equipment
  3418. allowed us to quantify in detail the various space and time
  3419. parameters of limb kinematics. The results show that mice
  3420. maintain the same gait pattern in both conditions. However, they
  3421. also demonstrate that during treadmill exercise mice always
  3422. exhibit higher stride frequency and consequently lower stride
  3423. length. The relationship of the stance time and the swing time
  3424. against the stride frequency are still the same in both
  3425. conditions. We conclude that the conflict related to the
  3426. discrepancy between the proprioceptive, vestibular, and visual
  3427. inputs contribute to an increase in the stride frequency during
  3428. the treadmill locomotion.",
  3429. journal = "Behav. Brain Res.",
  3430. volume = 181,
  3431. number = 2,
  3432. pages = "173--179",
  3433. month = aug,
  3434. year = 2007,
  3435. keywords = "Locomotion",
  3436. language = "en"
  3437. }
  3438. @ARTICLE{Ahmad_Abu_Hatab2013-st,
  3439. title = "Dynamic Modelling of {Differential-Drive} Mobile Robots using
  3440. Lagrange and {Newton-Euler} Methodologies: A Unified Framework",
  3441. author = "Ahmad Abu Hatab, Rached Dhaouadi",
  3442. journal = "Adv Robot Autom",
  3443. volume = 02,
  3444. number = 02,
  3445. year = 2013
  3446. }
  3447. @MISC{noauthor_undated-dh,
  3448. title = "Two wheeled robot thesis chapter from some gui"
  3449. }
  3450. @ARTICLE{Muryy_undated-jw,
  3451. title = "Route selection in non-Euclidean virtual environments",
  3452. author = "Muryy, Alexander and Glennerster, Andrew"
  3453. }
  3454. @ARTICLE{Bermudez-Contreras2020-wk,
  3455. title = "The Neuroscience of Spatial Navigation and the Relationship to
  3456. Artificial Intelligence",
  3457. author = "Bermudez-Contreras, Edgar and Clark, Benjamin J and Wilber, Aaron",
  3458. abstract = "Recent advances in artificial intelligence (AI) and neuroscience
  3459. are impressive. In AI, this includes the development of computer
  3460. programs that can beat a grandmaster at GO or outperform human
  3461. radiologists at cancer detection. A great deal of these
  3462. technological developments are directly related to progress in
  3463. artificial neural networks---initially inspired by our knowledge
  3464. about how the brain carries out computation. In parallel,
  3465. neuroscience has also experienced significant advances in
  3466. understanding the brain. For example, in the field of spatial
  3467. navigation, knowledge about the mechanisms and brain regions
  3468. involved in neural computations of cognitive maps---an internal
  3469. representation of space---recently received the Nobel Prize in
  3470. medicine. Much of the recent progress in neuroscience has partly
  3471. been due to the development of technology used to record from
  3472. very large populations of neurons in multiple regions of the
  3473. brain with exquisite temporal and spatial resolution in behaving
  3474. animals. With the advent of the vast quantities of data that
  3475. these techniques allow us to collect there has been an increased
  3476. interest in the intersection between AI and neuroscience, many of
  3477. these intersections involve using AI as a novel tool to explore
  3478. and analyze these large data sets. However, given the common
  3479. initial motivation point---to understand the brain---these
  3480. disciplines could be more strongly linked. Currently much of this
  3481. potential synergy is not being realized. We propose that spatial
  3482. navigation is an excellent area in which these two disciplines
  3483. can converge to help advance what we know about the brain. In
  3484. this review, we first summarize progress in the neuroscience of
  3485. spatial navigation and reinforcement learning. We then turn our
  3486. attention to discuss how spatial navigation has been modeled
  3487. using descriptive, mechanistic, and normative approaches and the
  3488. use of AI in such models. Next, we discuss how AI can advance
  3489. neuroscience, how neuroscience can advance AI, and the
  3490. limitations of these approaches. We finally conclude by
  3491. highlighting promising lines of research in which spatial
  3492. navigation can be the point of intersection between neuroscience
  3493. and AI and how this can contribute to the advancement of the
  3494. understanding of intelligent behavior.",
  3495. journal = "Front. Comput. Neurosci.",
  3496. volume = 14,
  3497. pages = "63",
  3498. year = 2020
  3499. }
  3500. @ARTICLE{Chapuis1987-yd,
  3501. title = "The role of exploratory experience in a shortcut task by golden
  3502. hamsters (Mesocricetus auratus)",
  3503. author = "Chapuis, N and Durup, M and Thinus-Blanc, C",
  3504. abstract = "The aim of this experiment was to examine the role of exploratory
  3505. experience on the ability to take a shortcut. In the first phase,
  3506. two subspaces, X and Y, each consisting of two baited tables
  3507. related by a runway, were separately explored by hamsters. In the
  3508. second phase, the experimental group explored a connecting
  3509. pathway between X and Y. The animals were finally submitted to a
  3510. shortcut test during 2 days: in this test, in order to go from X
  3511. to Y, they could choose between the longer familiar pathway and
  3512. two shorter new pathways. In comparison with a control group,
  3513. which did not undergo the second phase, the experimental group
  3514. displayed a significant preference for the shortcut that did not
  3515. cross the linking path with which they had had experience or
  3516. either of the two distant portions whose linkage the animals had
  3517. experienced. These results suggest that, in this simple
  3518. situation, additional experience of a linking element between two
  3519. separated subspaces has a beneficial effect on the setting up of
  3520. spatial relationships between them, and perhaps on the
  3521. representation of the whole situation.",
  3522. journal = "Anim. Learn. Behav.",
  3523. volume = 15,
  3524. number = 2,
  3525. pages = "174--178",
  3526. month = jun,
  3527. year = 1987
  3528. }
  3529. @UNPUBLISHED{Arshadi2020-my,
  3530. title = "{SNT}: A Unifying Toolbox for Quantification of Neuronal Anatomy",
  3531. author = "Arshadi, Cameron and Eddison, Mark and Gunther, Ulrik A and
  3532. Harrington, Kyle I and Ferreira, Tiago A",
  3533. abstract = "Quantification of neuronal morphology is essential for
  3534. understanding neuronal connectivity and many software tools have
  3535. been developed for neuronal reconstruction and morphometry.
  3536. However, such tools remain domain-specific, tethered to specific
  3537. imaging modalities, and were not designed to accommodate the rich
  3538. metadata generated by recent whole-brain cellular connectomics.
  3539. To address these limitations, we created SNT: a unifying
  3540. framework for neuronal morphometry and analysis of single-cell
  3541. connectomics for the widely used Fiji and ImageJ platforms. We
  3542. demonstrate that SNT -that replaces the popular Simple Neurite
  3543. Tracer software- can be used to tackle important problems in
  3544. contemporary neuroscience, validate its utility, and illustrate
  3545. how it establishes an end-to-end platform for tracing,
  3546. proof-editing, visualization, quantification, and modeling of
  3547. neuroanatomy. With an open and scriptable architecture, a large
  3548. user base, and thorough community-based documentation, SNT is an
  3549. accessible and scalable resource for the broad neuroscience
  3550. community that synergizes well with existing software. \#\#\#
  3551. Competing Interest Statement The authors have declared no
  3552. competing interest.",
  3553. journal = "bioRxiv",
  3554. pages = "2020.07.13.179325",
  3555. month = jul,
  3556. year = 2020,
  3557. language = "en"
  3558. }
  3559. @UNPUBLISHED{Chen2020-uy,
  3560. title = "Between-subject prediction reveals a shared representational
  3561. geometry in the rodent hippocampus",
  3562. author = "Chen, Hung-Tu and Manning, Jeremy R and van der Meer, Matthijs A
  3563. A",
  3564. abstract = "Summary How a memory system encodes related experiences has
  3565. consequences for what operations the system supports. For
  3566. instance, independent coding enables retention of potentially
  3567. important idiosyncratic details by reducing interference, but
  3568. makes it difficult to generalize across experiences. Strikingly,
  3569. the rodent hippocampus constructs statistically independent
  3570. representations across environments (``global remapping'') and
  3571. assigns individual neuron firing fields to locations within an
  3572. environment in an apparently random fashion, processes thought to
  3573. contribute to the role of the hippocampus in episodic memory.
  3574. This random mapping implies that it should be challenging to
  3575. predict hippocampal encoding of a given experience in a one
  3576. subject based on the encoding of that same experience in another
  3577. subject. Contrary to this prediction, we find that by
  3578. constructing a common representational space across rats
  3579. (``hyperalignment''), we can consistently predict data of
  3580. ``right'' trials (R) on a T-maze in a target rat based on 1) the
  3581. ``left'' trials (L) of the target rat, and 2) the relationship
  3582. between L and R trials from a different source rat. These
  3583. cross-subject predictions outperformed a number of control
  3584. mappings, such as those based on permuted data that broke the
  3585. relationship between L and R activity for individual neurons, and
  3586. those based solely on within-subject prediction. This work
  3587. constitutes proof-of-principle for successful cross-subject
  3588. prediction of ensemble activity patterns in the hippocampus. This
  3589. novel approach provides new insights in understanding how
  3590. different experiences are structured, and suggests further work
  3591. identifying what aspects of experience encoding are shared vs.
  3592. unique to an individual.",
  3593. journal = "bioRxiv",
  3594. pages = "2020.01.27.922062",
  3595. month = jan,
  3596. year = 2020,
  3597. language = "en"
  3598. }
  3599. @ARTICLE{Cheng2005-pv,
  3600. title = "Is there a geometric module for spatial orientation? Squaring
  3601. theory and evidence",
  3602. author = "Cheng, Ken and Newcombe, Nora S",
  3603. abstract = "There is evidence, beginning with Cheng (1986), that mobile
  3604. animals may use the geometry of surrounding areas to reorient
  3605. following disorientation. Gallistel (1990) proposed that geometry
  3606. is used to compute the major or minor axes of space and suggested
  3607. that such information might form an encapsulated cognitive
  3608. module. Research reviewed here, conducted on a wide variety of
  3609. species since the initial discovery of the use of geometry and
  3610. the formulation of the modularity claim, has supported some
  3611. aspects of the approach, while casting doubt on others. Three
  3612. possible processing models are presented that vary in the way in
  3613. which (and the extent to which) they instantiate the modularity
  3614. claim. The extant data do not permit us to discriminate among
  3615. them. We propose a modified concept of modularity for which an
  3616. empirical program of research is more tractable.",
  3617. journal = "Psychon. Bull. Rev.",
  3618. volume = 12,
  3619. number = 1,
  3620. pages = "1--23",
  3621. month = feb,
  3622. year = 2005,
  3623. language = "en"
  3624. }
  3625. @ARTICLE{Toledo2020-he,
  3626. title = "Cognitive map--based navigation in wild bats revealed by a new
  3627. high-throughput tracking system",
  3628. author = "Toledo, Sivan and Shohami, David and Schiffner, Ingo and Lourie,
  3629. Emmanuel and Orchan, Yotam and Bartan, Yoav and Nathan, Ran",
  3630. abstract = "The presence of a cognitive map is essential to our ability to
  3631. navigate through areas we know because it facilitates the use of
  3632. spatial knowledge to derive new routes. Whether such maps exist
  3633. in nonhuman animals has been debated, largely because of the
  3634. difficulty of demonstrating qualifying components of the map
  3635. outside of a laboratory. In two studies on Egyptian fruit bats,
  3636. Harten et al. and Toledo et al. together show that this
  3637. species's navigational strategies meet the requirements for the
  3638. use of a cognitive map of their environment, confirming that
  3639. this skill occurs outside of humans (see the Perspective by
  3640. Fenton). Science , this issue p. [194][1], p. [188][2]; see also
  3641. p. [142][3] Seven decades of research on the ``cognitive map,''
  3642. the allocentric representation of space, have yielded key
  3643. neurobiological insights, yet field evidence from free-ranging
  3644. wild animals is still lacking. Using a system capable of
  3645. tracking dozens of animals simultaneously at high accuracy and
  3646. resolution, we assembled a large dataset of 172 foraging
  3647. Egyptian fruit bats comprising >18 million localizations
  3648. collected over 3449 bat-nights across 4 years. Detailed track
  3649. analysis, combined with translocation experiments and exhaustive
  3650. mapping of fruit trees, revealed that wild bats seldom exhibit
  3651. random search but instead repeatedly forage in goal-directed,
  3652. long, and straight flights that include frequent shortcuts.
  3653. Alternative, non--map-based strategies were ruled out by
  3654. simulations, time-lag embedding, and other trajectory analyses.
  3655. Our results are consistent with expectations from cognitive
  3656. map--like navigation and support previous neurobiological
  3657. evidence from captive bats. [1]:
  3658. /lookup/doi/10.1126/science.aay3354 [2]:
  3659. /lookup/doi/10.1126/science.aax6904 [3]:
  3660. /lookup/doi/10.1126/science.abd1213",
  3661. journal = "Science",
  3662. publisher = "American Association for the Advancement of Science",
  3663. volume = 369,
  3664. number = 6500,
  3665. pages = "188--193",
  3666. month = jul,
  3667. year = 2020,
  3668. language = "en"
  3669. }
  3670. @ARTICLE{Mugan2020-bu,
  3671. title = "Spatial planning with long visual range benefits escape from
  3672. visual predators in complex naturalistic environments",
  3673. author = "Mugan, Ugurcan and MacIver, Malcolm A",
  3674. abstract = "It is uncontroversial that land animals have more elaborated
  3675. cognitive abilities than their aquatic counterparts such as fish.
  3676. Yet there is no apparent a-priori reason for this. A key
  3677. cognitive faculty is planning. We show that in visually guided
  3678. predator-prey interactions, planning provides a significant
  3679. advantage, but only on land. During animal evolution, the
  3680. water-to-land transition resulted in a massive increase in visual
  3681. range. Simulations of behavior identify a specific type of
  3682. terrestrial habitat, clustered open and closed areas
  3683. (savanna-like), where the advantage of planning peaks. Our
  3684. computational experiments demonstrate how this patchy terrestrial
  3685. structure, in combination with enhanced visual range, can reveal
  3686. and hide agents as a function of their movement and create a
  3687. selective benefit for imagining, evaluating, and selecting among
  3688. possible future scenarios-in short, for planning. The vertebrate
  3689. invasion of land may have been an important step in their
  3690. cognitive evolution.",
  3691. journal = "Nat. Commun.",
  3692. volume = 11,
  3693. number = 1,
  3694. pages = "3057",
  3695. month = jun,
  3696. year = 2020,
  3697. language = "en"
  3698. }
  3699. @ARTICLE{Dodd2000-uu,
  3700. title = "Use of cues by Lipophrys pholis L. (Teleostei, Blenniidae) in
  3701. learning the position of a refuge",
  3702. author = "Dodd, J and Gibson, R N and Hughes, R N",
  3703. abstract = "The ability of Lipophrys pholis to remember the position of a
  3704. refuge was tested in an artificial habitat under the influence
  3705. of different visual clues. L. pholis learned the position of the
  3706. refuge in the presence of a clue consisting only of a small
  3707. black screen. They responded to this clue by moving towards it
  3708. and pressing themselves up against it. Lego towers and a white
  3709. screen clue did not provoke such a response. In a further
  3710. experiment L. pholis continued to respond to the black screen in
  3711. this way when the screen was moved to another location further
  3712. from the refuge. After 12 days L. pholis learned to use the
  3713. black screen in its new position as an indirect clue and
  3714. navigate to the refuge directly without first approaching the
  3715. black screen. These results suggested that when placed in a
  3716. novel habitat the immediate reaction of L. pholis is to move
  3717. quickly towards the first dark area they see but, with
  3718. experience, they can use the position of large objects around
  3719. them to navigate quickly and efficiently to a refuge.",
  3720. journal = "Behav. Processes",
  3721. publisher = "Elsevier",
  3722. volume = 49,
  3723. number = 2,
  3724. pages = "69--75",
  3725. month = apr,
  3726. year = 2000,
  3727. language = "en"
  3728. }
  3729. % The entry below contains non-ASCII chars that could not be converted
  3730. % to a LaTeX equivalent.
  3731. @ARTICLE{Markel1994-dh,
  3732. title = "An adaptive value of spatial learning and memory in the blackeye
  3733. goby, Coryphopterus nicholsi",
  3734. author = "Markel, Russell W",
  3735. abstract = "The adaptive value of spatial learning and memory has been
  3736. demonstrated in birds and mammals (eg Shettleworth \& Krebs
  3737. 1983; Clarke et al. 1993), but few investigations have found
  3738. evidence of similar learning abilities in fish. The goa{\l} of
  3739. this project was to demonstrate spatial learning and memory in
  3740. blackeye gobies, Coryphopterus nicholsi in a context in which it
  3741. may be adaptive, namely predator evasion. Fish can learn spatial
  3742. relationships among rel-evant features of their environments
  3743. (Aronson 1951, 1971; Roitblat …",
  3744. journal = "Anim. Behav.",
  3745. publisher = "Elsevier",
  3746. volume = 47,
  3747. number = 6,
  3748. pages = "1462--1464",
  3749. month = jun,
  3750. year = 1994
  3751. }
  3752. @ARTICLE{Burt_de_Perera2008-ki,
  3753. title = "Rapid learning of shelter position in an intertidal fish, the
  3754. shanny Lipophrys pholis {L}",
  3755. author = "Burt de Perera, T and Guilford, T C",
  3756. abstract = "The homing ability of an intertidal fish, the shanny Lipophrys
  3757. pholis, was investigated using two experiments that were based
  3758. on the shanny?s natural propensity to home to a refuge. A
  3759. displacement experiment demonstrated that the fish were able to
  3760. accurately locate the previous position of a refuge once the
  3761. shelter itself had been removed so that it could not be used as
  3762. a cue to directly signal the goal location. This shows that the
  3763. shanny can encode information about its familiar surroundings
  3764. into a spatial map and use this information to home. A second
  3765. experiment in which the cues internal and external to the
  3766. experimental tank were put in conflict with one another
  3767. suggested that the shanny can encode cues that are both intra-
  3768. and external-tank cues in its representation of space, but that
  3769. there is individual variation in the type of cues that are used,
  3770. or memorized.",
  3771. journal = "J. Fish Biol.",
  3772. publisher = "Wiley Online Library",
  3773. volume = 72,
  3774. number = 6,
  3775. pages = "1386--1392",
  3776. month = apr,
  3777. year = 2008
  3778. }
  3779. @INCOLLECTION{Chapuis1987-dz,
  3780. title = "Detour and Shortcut Abilities in Several Species of Mammals",
  3781. booktitle = "Cognitive Processes and Spatial Orientation in Animal and Man:
  3782. Volume {I} Experimental Animal Psychology and Ethology",
  3783. author = "Chapuis, Nicole",
  3784. editor = "Ellen, Paul and Thinus-Blanc, Catherine",
  3785. abstract = "This study is concerned with two properties of cognitive
  3786. mapping. The first is plasticity, by which I mean the ability of
  3787. an animal to reorganize its previous experience of a given
  3788. situation. Thus, for example, when modifications are introduced
  3789. into a familiar spatial task, some animals can find original
  3790. solutions; they do not become lost and they can still reach the
  3791. goal. The second property is optimalization. It entails the
  3792. choice and the planning of the best adapted solution: for
  3793. example, taking the most direct of several possible ways to
  3794. reach a goal.",
  3795. publisher = "Springer Netherlands",
  3796. pages = "97--106",
  3797. year = 1987,
  3798. address = "Dordrecht"
  3799. }
  3800. @ARTICLE{Collett1982-hv,
  3801. title = "Do Toads Plan Routes ? A Study of the Detour Behaviour of Bufo
  3802. viridis",
  3803. author = "Collett, T S",
  3804. journal = "J. Comp. Physiol.",
  3805. volume = 146,
  3806. pages = "261--271",
  3807. year = 1982
  3808. }
  3809. @UNPUBLISHED{Metz2017-xq,
  3810. title = "Evolution and genetics of precocious burrowing behavior in
  3811. Peromyscus mice",
  3812. author = "Metz, Hillery C and Bedford, Nicole L and Pan, Linda and
  3813. Hoekstra, Hopi E",
  3814. abstract = "Summary A central challenge in biology is to understand how
  3815. innate behaviors evolve between closely related species. One way
  3816. to elucidate how differences arise is to compare the development
  3817. of behavior in species with distinct adult traits. Here, we
  3818. report that Peromyscus polionotus is strikingly precocious with
  3819. regard to burrowing behavior, but not other behaviors, compared
  3820. to its sister species P. maniculatus. In P. polionotus, burrows
  3821. were excavated as early as 17 days of age, while P. maniculatus
  3822. did not build burrows until 10 days later. Moreover, the
  3823. well-known differences in burrow architecture between adults of
  3824. these species---P. polionotus adults excavate long burrows with
  3825. an escape tunnel, while P. maniculatus dig short, single-tunnel
  3826. burrows---were intact in juvenile burrowers. To test whether this
  3827. juvenile behavior is influenced by early-life environment, pups
  3828. of both species were reciprocally cross-fostered. Fostering did
  3829. not alter the characteristic burrowing behavior of either
  3830. species, suggesting these differences are genetic. In backcross
  3831. F2 hybrids, we show that precocious burrowing and adult tunnel
  3832. length are genetically correlated, and that a single P.
  3833. polionotus allele in a genomic region linked to adult tunnel
  3834. length is predictive of precocious burrow construction. The
  3835. co-inheritance of developmental and adult traits indicates the
  3836. same genetic region---either a single gene with pleiotropic
  3837. effects, or closely linked genes--- acts on distinct aspects of
  3838. the same behavior across life stages. Such genetic variants
  3839. likely affect behavioral drive (i.e. motivation) to burrow, and
  3840. thereby affect both the development and adult expression of
  3841. burrowing behavior.Highlights Juvenile P. polionotus construct
  3842. burrows precociously compared to its sister species P.
  3843. maniculatusCross-fostering does not alter species-specific
  3844. burrowing behaviorA QTL linked to adult tunnel length predicts
  3845. developmental onset of burrow construction in hybridsPleiotropic
  3846. genetic variant(s) may affect behavioral drive across life stages",
  3847. journal = "bioRxiv",
  3848. pages = "150243",
  3849. month = jun,
  3850. year = 2017,
  3851. language = "en"
  3852. }
  3853. @ARTICLE{Jackson2020-te,
  3854. title = "Many Paths to the Same Goal: Balancing Exploration and
  3855. Exploitation during Probabilistic Route Planning",
  3856. author = "Jackson, Brian J and Fatima, Gusti Lulu and Oh, Sujean and Gire,
  3857. David H",
  3858. abstract = "During self-guided behaviors, animals identify constraints of the
  3859. problems they face and adaptively employ appropriate strategies
  3860. (Marsh, 2002). In the case of foraging, animals must balance
  3861. sensory-guided exploration of an environment with memory-guided
  3862. exploitation of known resource locations. Here, we show that
  3863. animals adaptively shift cognitive resources between sensory and
  3864. memory systems during foraging to optimize route planning under
  3865. uncertainty. We demonstrate this using a new, laboratory-based
  3866. discovery method to define the strategies used to solve a
  3867. difficult route optimization scenario, the probabilistic
  3868. ``traveling salesman'' problem (Raman and Gill, 2017; Fuentes et
  3869. al., 2018; Mukherjee et al., 2019). Using this system, we
  3870. precisely manipulated the strength of prior information as well
  3871. as the complexity of the problem. We find that rats are capable
  3872. of efficiently solving this route-planning problem, even under
  3873. conditions with unreliable prior information and a large space of
  3874. possible solutions. Through analysis of animals' trajectories, we
  3875. show that they shift the balance between exploiting known
  3876. locations and searching for new locations of rewards based on the
  3877. predictability of reward locations. When compared with a Bayesian
  3878. search, we found that animal performance is consistent with an
  3879. approach that adaptively allocates cognitive resources between
  3880. sensory processing and memory, enhancing sensory acuity and
  3881. reducing memory load under conditions in which prior information
  3882. is unreliable. Our findings establish new approaches to
  3883. understand neural substrates of natural behavior as well as the
  3884. rational development of biologically inspired approaches for
  3885. complex real-world optimization.",
  3886. journal = "eNeuro",
  3887. volume = 7,
  3888. number = 3,
  3889. month = jun,
  3890. year = 2020,
  3891. keywords = "Bayesian; foraging; navigation",
  3892. language = "en"
  3893. }
  3894. @MISC{Juszczak2016-fk,
  3895. title = "Detour Behavior of Mice Trained with Transparent, Semitransparent
  3896. and Opaque Barriers",
  3897. author = "Juszczak, Grzegorz R and Miller, Michal",
  3898. editor = "Burne, Thomas H J",
  3899. abstract = "Detour tasks are commonly used to study problem solving skills
  3900. and inhibitory control in canids and primates. However, there is
  3901. no comparable detour test designed for rodents despite its
  3902. significance for studying the development of executive skills.
  3903. Furthermore, mice offer research opportunities that are not
  3904. currently possible to achieve when primates are used. Therefore,
  3905. the aim of the study was to translate the classic detour task to
  3906. mice and to compare obtained data with key findings obtained
  3907. previously in other mammals. The experiment was performed with
  3908. V-shaped barriers and was based on the water escape paradigm. The
  3909. study showed that an apparently simple task requiring mice to
  3910. move around a small barrier constituted in fact a challenge that
  3911. was strongly affected by the visibility of the target. The most
  3912. difficult task involved a completely transparent barrier, which
  3913. forced the mice to resolve a conflict between vision and tactile
  3914. perception. The performance depended both on the inhibitory
  3915. skills and on previous experiences. Additionally, all mice
  3916. displayed a preference for one side of the barrier and most of
  3917. them relied on the egocentric strategy. Obtained results show for
  3918. the first time that the behavior of mice subjected to the detour
  3919. task is comparable to the behavior of other mammals tested
  3920. previously with free-standing barriers. This detailed
  3921. characterization of the detour behavior of mice constitutes the
  3922. first step toward the substitution of rodents for primates in
  3923. laboratory experiments employing the detour task.",
  3924. month = sep,
  3925. year = 2016
  3926. }
  3927. @ARTICLE{Uribe-Marino2012-pw,
  3928. title = "Anti-aversive effects of cannabidiol on innate fear-induced
  3929. behaviors evoked by an ethological model of panic attacks based
  3930. on a prey vs the wild snake Epicrates cenchria crassus
  3931. confrontation paradigm",
  3932. author = "Uribe-Mari{\~n}o, Andr{\'e}s and Francisco, Audrey and
  3933. Castiblanco-Urbina, Maria Ang{\'e}lica and Twardowschy, Andr{\'e}
  3934. and Salgado-Rohner, Carlos Jos{\'e} and Crippa, Jos{\'e}
  3935. Alexandre S and Hallak, Jaime Eduardo Cec{\'\i}lio and Zuardi,
  3936. Ant{\^o}nio Waldo and Coimbra, Norberto Cysne",
  3937. abstract = "Several pharmacological targets have been proposed as modulators
  3938. of panic-like reactions. However, interest should be given to
  3939. other potential therapeutic neurochemical agents. Recent
  3940. attention has been given to the potential anxiolytic properties
  3941. of cannabidiol, because of its complex actions on the
  3942. endocannabinoid system together with its effects on other
  3943. neurotransmitter systems. The aim of this study was to
  3944. investigate the effects of cannabidiol on innate fear-related
  3945. behaviors evoked by a prey vs predator paradigm. Male Swiss mice
  3946. were submitted to habituation in an arena containing a burrow and
  3947. subsequently pre-treated with intraperitoneal administrations of
  3948. vehicle or cannabidiol. A constrictor snake was placed inside the
  3949. arena, and defensive and non-defensive behaviors were recorded.
  3950. Cannabidiol caused a clear anti-aversive effect, decreasing
  3951. explosive escape and defensive immobility behaviors outside and
  3952. inside the burrow. These results show that cannabidiol modulates
  3953. defensive behaviors evoked by the presence of threatening
  3954. stimuli, even in a potentially safe environment following a fear
  3955. response, suggesting a panicolytic effect.",
  3956. journal = "Neuropsychopharmacology",
  3957. volume = 37,
  3958. number = 2,
  3959. pages = "412--421",
  3960. month = jan,
  3961. year = 2012,
  3962. language = "en"
  3963. }
  3964. @ARTICLE{Kabadayi2018-pq,
  3965. title = "The detour paradigm in animal cognition",
  3966. author = "Kabadayi, Can and Bobrowicz, Katarzyna and Osvath, Mathias",
  3967. abstract = "In this paper, we review one of the oldest paradigms used in
  3968. animal cognition: the detour paradigm. The paradigm presents the
  3969. subject with a situation where a direct route to the goal is
  3970. blocked and a detour must be made to reach it. Often being an
  3971. ecologically valid and a versatile tool, the detour paradigm has
  3972. been used to study diverse cognitive skills like insight, social
  3973. learning, inhibitory control and route planning. Due to the
  3974. relative ease of administrating detour tasks, the paradigm has
  3975. lately been used in large-scale comparative studies in order to
  3976. investigate the evolution of inhibitory control. Here we review
  3977. the detour paradigm and some of its cognitive requirements, we
  3978. identify various ecological and contextual factors that might
  3979. affect detour performance, we also discuss developmental and
  3980. neurological underpinnings of detour behaviors, and we suggest
  3981. some methodological approaches to make species comparisons more
  3982. robust.",
  3983. journal = "Anim. Cogn.",
  3984. volume = 21,
  3985. number = 1,
  3986. pages = "21--35",
  3987. month = jan,
  3988. year = 2018,
  3989. keywords = "Comparative psychology; Detour behavior; Inhibitory control;
  3990. Route planning",
  3991. language = "en"
  3992. }
  3993. @UNPUBLISHED{Alonso2020-is,
  3994. title = "The {HexMaze}: A previous knowledge and schema task for mice",
  3995. author = "Alonso, Alejandra and Bokeria, Levan and van der Meij, Jacqueline
  3996. and Samanta, Anumita and Eichler, Ronny and Spooner, Patrick and
  3997. Lobato, Irene Navarro and Genzel, Lisa",
  3998. abstract = "Abstract New information is rarely learned in isolation, instead
  3999. most of what we experience can be incorporated into or uses
  4000. previous knowledge networks in some form. However, most rodent
  4001. laboratory tasks assume the animal to be na{\"\i}ve with no
  4002. previous experience influencing the results. Previous knowledge
  4003. in form of a schema can facilitate knowledge acquisition and
  4004. accelerate systems consolidation: memories become more rapidly
  4005. hippocampal independent and instead rely more on the prefrontal
  4006. cortex. Here, we developed a new spatial navigation task where
  4007. food locations are learned in a large, gangway maze -- the
  4008. HexMaze. Analysing performance across sessions as well as on
  4009. specific trials, we can show simple memory effects as well as
  4010. multiple effects of previous knowledge accelerating both online
  4011. learning and performance increases over offline periods.
  4012. Importantly, we are the first to show that schema build-up is
  4013. dependent on how much time passes, not how often the animal is
  4014. trained.",
  4015. journal = "bioRxiv",
  4016. pages = "441048",
  4017. month = mar,
  4018. year = 2020,
  4019. language = "en"
  4020. }
  4021. @ARTICLE{Hein2018-el,
  4022. title = "Conserved behavioral circuits govern high-speed decision-making
  4023. in wild fish shoals",
  4024. author = "Hein, Andrew M and Gil, Michael A and Twomey, Colin R and Couzin,
  4025. Iain D and Levin, Simon A",
  4026. abstract = "To evade their predators, animals must quickly detect potential
  4027. threats, gauge risk, and mount a response. Putative neural
  4028. circuits responsible for these tasks have been isolated in
  4029. laboratory studies. However, it is unclear whether and how these
  4030. circuits combine to generate the flexible, dynamic sequences of
  4031. evasion behavior exhibited by wild, freely moving animals. Here,
  4032. we report that evasion behavior of wild fish on a coral reef is
  4033. generated through a sequence of well-defined decision rules that
  4034. convert visual sensory input into behavioral actions. Using an
  4035. automated system to present visual threat stimuli to fish in
  4036. situ, we show that individuals initiate escape maneuvers in
  4037. response to the perceived size and expansion rate of an oncoming
  4038. threat using a decision rule that matches dynamics of known
  4039. loom-sensitive neural circuits. After initiating an evasion
  4040. maneuver, fish adjust their trajectories using a control rule
  4041. based on visual feedback to steer away from the threat and toward
  4042. shelter. These decision rules accurately describe evasion
  4043. behavior of fish from phylogenetically distant families,
  4044. illustrating the conserved nature of escape decision-making. Our
  4045. results reveal how the flexible behavioral responses required for
  4046. survival can emerge from relatively simple, conserved
  4047. decision-making mechanisms.",
  4048. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  4049. volume = 115,
  4050. number = 48,
  4051. pages = "12224--12228",
  4052. month = nov,
  4053. year = 2018,
  4054. keywords = "decision-making; evasion; neural circuit; neuroethology;
  4055. predator--prey interactions",
  4056. language = "en"
  4057. }
  4058. @ARTICLE{Carandini2012-xx,
  4059. title = "From circuits to behavior: a bridge too far?",
  4060. author = "Carandini, Matteo",
  4061. abstract = "Neuroscience seeks to understand how neural circuits lead to
  4062. behavior. However, the gap between circuits and behavior is too
  4063. wide. An intermediate level is one of neural computations, which
  4064. occur in individual neurons and populations of neurons. Some
  4065. computations seem to be canonical: repeated and combined in
  4066. different ways across the brain. To understand neural
  4067. computations, we must record from a myriad of neurons in multiple
  4068. brain regions. Understanding computation guides research in the
  4069. underlying circuits and provides a language for theories of
  4070. behavior.",
  4071. journal = "Nat. Neurosci.",
  4072. volume = 15,
  4073. number = 4,
  4074. pages = "507--509",
  4075. month = mar,
  4076. year = 2012,
  4077. language = "en"
  4078. }
  4079. % The entry below contains non-ASCII chars that could not be converted
  4080. % to a LaTeX equivalent.
  4081. @ARTICLE{Fiker2020-kd,
  4082. title = "Visual Gait Lab: A user-friendly approach to gait analysis",
  4083. author = "Fiker, Robert and Kim, Linda H and Molina, Leonardo A and
  4084. Chomiak, Taylor and Whelan, Patrick J",
  4085. abstract = "BACKGROUND: Gait analysis forms a critical part of many lab
  4086. workflows, ranging from those interested in preclinical
  4087. neurological models to others who use locomotion as part of a
  4088. standard battery of tests. Unfortunately, while paw detection can
  4089. be semi-automated, it becomes generally a time-consuming process
  4090. with error corrections. Improvement in paw tracking would aid in
  4091. better gait analysis performance and experience. NEW METHOD: Here
  4092. we show the use of Visual Gait Lab (VGL), a high-level software
  4093. with an intuitive, easy to use interface, that is built on
  4094. DeepLabCut™. VGL is optimized to generate gait metrics and allows
  4095. for quick manual error corrections. VGL comes with a single
  4096. executable, streamlining setup on Windows systems. We demonstrate
  4097. the use of VGL to analyze gait. RESULTS: Training and evaluation
  4098. of VGL were conducted using 200 frames (80/20 train-test split)
  4099. of video from mice walking on a treadmill. The trained network
  4100. was then used to visually track paw placements to compute gait
  4101. metrics. These are processed and presented on the screen where
  4102. the user can rapidly identify and correct errors. COMPARISON WITH
  4103. EXISTING METHODS: Gait analysis remains cumbersome, even with
  4104. commercial software due to paw detection errors. DeepLabCut™ is
  4105. an alternative that can improve visual tracking but is not
  4106. optimized for gait analysis functionality. CONCLUSIONS: VGL
  4107. allows for gait analysis to be performed in a rapid, unbiased
  4108. manner, with a set-up that can be easily implemented and executed
  4109. by those without a background in computer programming.",
  4110. journal = "J. Neurosci. Methods",
  4111. volume = 341,
  4112. pages = "108775",
  4113. month = may,
  4114. year = 2020,
  4115. keywords = "DeepLabCut™; Gait analysis; Gait tracking system; Motor control;
  4116. Mouse locomotion",
  4117. language = "en"
  4118. }
  4119. @ARTICLE{Lecca2020-xk,
  4120. title = "Heterogeneous Habenular Neuronal Ensembles during Selection of
  4121. Defensive Behaviors",
  4122. author = "Lecca, Salvatore and Namboodiri, Vijay M K and Restivo, Leonardo
  4123. and Gervasi, Nicolas and Pillolla, Giuliano and Stuber, Garret D
  4124. and Mameli, Manuel",
  4125. journal = "Cell Rep.",
  4126. volume = 31,
  4127. number = 10,
  4128. pages = "107752",
  4129. month = jun,
  4130. year = 2020
  4131. }
  4132. @ARTICLE{De_Oca2007-jy,
  4133. title = "Brief flight to a familiar enclosure in response to a conditional
  4134. stimulus in rats",
  4135. author = "de Oca, Beatrice M and Minor, Thomas R and Fanselow, Michael S",
  4136. abstract = "The authors observed brief, directed movement to a familiar
  4137. enclosure in rats to determine whether this behavior is part of a
  4138. rat's defensive repertoire when exposed to a conditional-fear
  4139. stimulus. In Experiment 1, upon exposure to the compound
  4140. conditional-fear stimulus of tone and light, only rats that
  4141. received paired presentations of the conditional stimuli and
  4142. shock fled into a small, familiar enclosure where they then
  4143. froze. Rats that had received unpaired presentations did not
  4144. enter the enclosure in significant amounts when later tested. In
  4145. Experiment 2, the authors observed rats' freezing and use of
  4146. either a familiar or an unfamiliar enclosure when tested with a
  4147. conditional-fear stimulus. Rats tested with a familiar enclosure
  4148. entered it more quickly than did rats without prior exposure to
  4149. the enclosure. Freezing was greatest when both training and
  4150. testing environments were similar with respect to access to the
  4151. enclosure. The results of these 2 experiments support the idea
  4152. that brief, directed flight in rats is a component of the
  4153. postencounter stage of predatory imminence (M. S. Fanselow \& L.
  4154. S. Lester, 1988) and is compatible with freezing.",
  4155. journal = "J. Gen. Psychol.",
  4156. volume = 134,
  4157. number = 2,
  4158. pages = "153--172",
  4159. month = apr,
  4160. year = 2007,
  4161. language = "en"
  4162. }
  4163. @ARTICLE{Hein2020-gg,
  4164. title = "An Algorithmic Approach to Natural Behavior",
  4165. author = "Hein, Andrew M and Altshuler, Douglas L and Cade, David E and
  4166. Liao, James C and Martin, Benjamin T and Taylor, Graham K",
  4167. abstract = "SummaryUncovering the mechanisms and implications of natural
  4168. behavior is a goal that unites many fields of biology. Yet, the
  4169. diversity, flexibility, and multi-scale nature of these
  4170. behaviors often make understanding elusive. Here, we review
  4171. studies of animal pursuit and evasion --- two special classes of
  4172. behavior where theory-driven experiments and new modeling
  4173. techniques are beginning to uncover the general control
  4174. principles underlying natural behavior. A key finding of these
  4175. studies is that intricate sequences of pursuit and evasion
  4176. behavior can often be constructed through simple, repeatable
  4177. rules that link sensory input to motor output: we refer to these
  4178. rules as behavioral algorithms. Identifying and mathematically
  4179. characterizing these algorithms has led to important insights,
  4180. including the discovery of guidance rules that attacking
  4181. predators use to intercept mobile prey, and coordinated neural
  4182. and biomechanical mechanisms that animals use to avoid impending
  4183. collisions. Here, we argue that algorithms provide a good
  4184. starting point for studies of natural behavior more generally.
  4185. Rather than beginning at the neural or ecological levels of
  4186. organization, we advocate starting in the middle, where the
  4187. algorithms that link sensory input to behavioral output can
  4188. provide a solid foundation from which to explore both the
  4189. implementation and the ecological outcomes of behavior. We
  4190. review insights that have been gained through such an
  4191. algorithmic approach to pursuit and evasion behaviors. From
  4192. these, we synthesize theoretical principles and lay out key
  4193. modeling tools needed to apply an algorithmic approach to the
  4194. study of other complex natural behaviors.",
  4195. journal = "Curr. Biol.",
  4196. publisher = "Elsevier",
  4197. volume = 30,
  4198. number = 11,
  4199. pages = "R663--R675",
  4200. month = jun,
  4201. year = 2020,
  4202. language = "en"
  4203. }
  4204. @ARTICLE{Kafkafi2005-es,
  4205. title = "Texture of locomotor path: a replicable characterization of a
  4206. complex behavioral phenotype",
  4207. author = "Kafkafi, N and Elmer, G I",
  4208. abstract = "A database of mouse locomotor path in spatial tests can be used
  4209. to search in silico for behavioral measures that better
  4210. discriminate between genotypes and are more replicable across
  4211. laboratories. In this study, software for the exploration of
  4212. exploration (SEE) was used to search a large database for a novel
  4213. behavioral measure that would characterize complex movement
  4214. paths. The database included mouse open-field behavior assessed
  4215. in 3 laboratories, 7 inbred strains, several pharmacological
  4216. treatments and hundreds of animals. The new behavioral measure,
  4217. ``path texture'', was characterized using the local curvature of
  4218. the path (the change of direction per unit distance, in
  4219. degrees/cm) across several spatial scales, starting from scales
  4220. smaller than the animal's body length and up to the scale of the
  4221. arena size. Path texture analysis differs from fractal dimension
  4222. analysis in that it does not assume self-similarity across
  4223. scales. Path texture was found to discriminate inbred strains
  4224. with relatively high broad-sense heritability (43\%-71\%) and
  4225. high replicability across laboratories. Even genotypes that had
  4226. similar path curvatures in some scales usually differed in other
  4227. scales, and self-similarity across scales was not displayed by
  4228. all genotypes. Amphetamine decreased the path curvature of
  4229. C57BL/6 mice in small and medium scales, while having no effect
  4230. on DBA/2J mice. Diazepam dose-dependently decreased the curvature
  4231. of C57BL/6 mice across all scales, while 2 anxiogenic drugs,
  4232. FG-7142 and pentylenetetrazole, increased it. Path texture thus
  4233. has high potential for behavioral phenotyping and the study of
  4234. drug effects in the mouse.",
  4235. journal = "Genes Brain Behav.",
  4236. volume = 4,
  4237. number = 7,
  4238. pages = "431--443",
  4239. month = oct,
  4240. year = 2005,
  4241. language = "en"
  4242. }
  4243. @MISC{Dvorkin2010-js,
  4244. title = "Knots: Attractive Places with High Path Tortuosity in Mouse Open
  4245. Field Exploration",
  4246. author = "Dvorkin, Anna and Szechtman, Henry and Golani, Ilan",
  4247. editor = "Bourne, Philip E",
  4248. abstract = "When introduced into a novel environment, mammals establish in it
  4249. a preferred place marked by the highest number of visits and
  4250. highest cumulative time spent in it. Examination of exploratory
  4251. behavior in reference to this ``home base'' highlights important
  4252. features of its organization. It might therefore be fruitful to
  4253. search for other types of marked places in mouse exploratory
  4254. behavior and examine their influence on overall
  4255. behavior.Examination of path curvatures of mice exploring a large
  4256. empty arena revealed the presence of circumscribed locales marked
  4257. by the performance of tortuous paths full of twists and turns. We
  4258. term these places knots, and the behavior performed in
  4259. them-knot-scribbling. There is typically no more than one knot
  4260. per session; it has distinct boundaries and it is maintained both
  4261. within and across sessions. Knots are mostly situated in the
  4262. place of introduction into the arena, here away from walls. Knots
  4263. are not characterized by the features of a home base, except for
  4264. a high speed during inbound and a low speed during outbound
  4265. paths. The establishment of knots is enhanced by injecting the
  4266. mouse with saline and placing it in an exposed portion of the
  4267. arena, suggesting that stress and the arousal associated with it
  4268. consolidate a long-term contingency between a particular locale
  4269. and knot-scribbling.In an environment devoid of proximal cues
  4270. mice mark a locale associated with arousal by twisting and
  4271. turning in it. This creates a self-generated, often centrally
  4272. located landmark. The tortuosity of the path traced during the
  4273. behavior implies almost concurrent multiple views of the
  4274. environment. Knot-scribbling could therefore function as a way to
  4275. obtain an overview of the entire environment, allowing
  4276. re-calibration of the mouse's locale map and compass directions.
  4277. The rich vestibular input generated by scribbling could improve
  4278. the interpretation of the visual scene.",
  4279. month = jan,
  4280. year = 2010
  4281. }
  4282. @MISC{Valente2007-qx,
  4283. title = "Analysis of the Trajectory of Drosophila melanogaster in a
  4284. Circular Open Field Arena",
  4285. author = "Valente, Dan and Golani, Ilan and Mitra, Partha P",
  4286. editor = "Scalas, Enrico",
  4287. abstract = "BACKGROUND Obtaining a complete phenotypic characterization of a
  4288. freely moving organism is a difficult task, yet such a
  4289. description is desired in many neuroethological studies. Many
  4290. metrics currently used in the literature to describe locomotor
  4291. and exploratory behavior are typically based on average
  4292. quantities or subjectively chosen spatial and temporal
  4293. thresholds. All of these measures are relatively coarse-grained
  4294. in the time domain. It is advantageous, however, to employ
  4295. metrics based on the entire trajectory that an organism takes
  4296. while exploring its environment. METHODOLOGY/PRINCIPAL FINDINGS
  4297. To characterize the locomotor behavior of Drosophila
  4298. melanogaster, we used a video tracking system to record the
  4299. trajectory of a single fly walking in a circular open field
  4300. arena. The fly was tracked for two hours. Here, we present
  4301. techniques with which to analyze the motion of the fly in this
  4302. paradigm, and we discuss the methods of calculation. The measures
  4303. we introduce are based on spatial and temporal probability
  4304. distributions and utilize the entire time-series trajectory of
  4305. the fly, thus emphasizing the dynamic nature of locomotor
  4306. behavior. Marginal and joint probability distributions of speed,
  4307. position, segment duration, path curvature, and reorientation
  4308. angle are examined and related to the observed behavior.
  4309. CONCLUSIONS/SIGNIFICANCE The measures discussed in this paper
  4310. provide a detailed profile of the behavior of a single fly and
  4311. highlight the interaction of the fly with the environment. Such
  4312. measures may serve as useful tools in any behavioral study in
  4313. which the movement of a fly is an important variable and can be
  4314. incorporated easily into many setups, facilitating
  4315. high-throughput phenotypic characterization.",
  4316. month = oct,
  4317. year = 2007
  4318. }
  4319. @ARTICLE{Benjamini2010-cs,
  4320. title = "Ten ways to improve the quality of descriptions of whole-animal
  4321. movement",
  4322. author = "Benjamini, Yoav and Lipkind, Dina and Horev, Guy and Fonio, Ehud
  4323. and Kafkafi, Neri and Golani, Ilan",
  4324. abstract = "The demand for replicability of behavioral results across
  4325. laboratories is viewed as a burden in behavior genetics. We
  4326. demonstrate how it can become an asset offering a quantitative
  4327. criterion that guides the design of better ways to describe
  4328. behavior. Passing the high benchmark dictated by the
  4329. replicability demand requires less stressful and less restraining
  4330. experimental setups, less noisy data, individually customized
  4331. cutoff points between the building blocks of movement, and less
  4332. variable yet discriminative dynamic representations that would
  4333. capture more faithfully the nature of the behavior, unmasking
  4334. similarities and differences and revealing novel animal-centered
  4335. measures. Here we review ten tools that enhance replicability
  4336. without compromising discrimination. While we demonstrate the
  4337. usefulness of these tools in the context of inbred mouse
  4338. exploratory behavior they can readily be used in any study
  4339. involving a high-resolution analysis of spatial behavior. Viewing
  4340. replicability as a design concept and using the ten
  4341. methodological improvements may prove useful in many fields not
  4342. necessarily related to spatial behavior.",
  4343. journal = "Neurosci. Biobehav. Rev.",
  4344. volume = 34,
  4345. number = 8,
  4346. pages = "1351--1365",
  4347. month = jul,
  4348. year = 2010,
  4349. language = "en"
  4350. }
  4351. @ARTICLE{Eilam1989-uk,
  4352. title = "Home base behavior of rats (Rattus norvegicus) exploring a novel
  4353. environment",
  4354. author = "Eilam, D and Golani, I",
  4355. abstract = "When rats are placed in a novel environment, they alternate
  4356. between progression and stopping: in the course of a session
  4357. they stop briefly in many places, but in one or two places they
  4358. also stop for very long periods. The place in which they stay
  4359. for the longest cumulative time is defined as the rat's home
  4360. base. In this place the incidences of grooming and of rearing
  4361. are high and often the highest. In addition, the number of
  4362. visits to the home base is typically the highest. Some rats
  4363. establish a secondary base with similar properties to those of
  4364. the main home base. The location of the base influences the mode
  4365. of progression throughout the environment: progression away from
  4366. base is slower and includes more stops than progression back. It
  4367. is suggested that this paradigm may be used for the analysis of
  4368. the spatial organization of locomotor behavior in neuroscience
  4369. research.",
  4370. journal = "Behav. Brain Res.",
  4371. publisher = "tau.ac.il",
  4372. volume = 34,
  4373. number = 3,
  4374. pages = "199--211",
  4375. month = sep,
  4376. year = 1989,
  4377. language = "en"
  4378. }
  4379. @ARTICLE{Kimchi2001-oj,
  4380. title = "Spatial learning and memory in the blind mole-rat in comparison
  4381. with the laboratory rat and Levant vole",
  4382. author = "Kimchi, Tali and Terkel, Joseph",
  4383. abstract = "Studies dealing with spatial orientation in mammals have mostly
  4384. dealt with surface-dwelling species. We studied the ability of a
  4385. subterranean rodent to orient in space and compared it with two
  4386. species of rodents that spend most of their lives above ground.
  4387. The solitary blind mole-rat, Spalax ehrenbergi, inhabits an
  4388. extensive, branching tunnel system that it digs itself and in
  4389. which it spends its entire life. We examined its ability to
  4390. learn and remember a winding path towards a goal in a multiple
  4391. labyrinth and compared it with Levant voles, Microtus guentheri,
  4392. and laboratory rats, Rattus norvegicus. The mole-rats learned
  4393. significantly faster than the rats and voles. Furthermore, their
  4394. ability to remember the maze was significantly better than that
  4395. of the rats after 2, 7, 30 and 60 days from the end of the
  4396. learning experiment and significantly better than the voles
  4397. after 120 days. The mole-rats still retained ca. 45\% of their
  4398. optimal performance at the end of the learning experiment after
  4399. 4 months compared with 20\% for the voles after 4 months and
  4400. less than 20\% for the rats after 2 months. Despite having lost
  4401. its vision, the mole-rat was thus more able to orient in a
  4402. complex maze than the surface-dwelling vole and laboratory rat.
  4403. We suggest that the mole-rat compensates for the sensory
  4404. limitations imposed by the subterranean niche and for its loss
  4405. of vision by relying on the Earth's magnetic field and internal
  4406. cues to steer its course efficiently. We discuss the possible
  4407. mechanisms of orientation. Copyright 2001 The Association for
  4408. the Study of Animal Behaviour.",
  4409. journal = "Anim. Behav.",
  4410. publisher = "Elsevier",
  4411. volume = 61,
  4412. number = 1,
  4413. pages = "171--180",
  4414. month = jan,
  4415. year = 2001,
  4416. language = "en"
  4417. }
  4418. @ARTICLE{Kimchi2003-xz,
  4419. title = "Detours by the blind mole-rat follow assessment of location and
  4420. physical properties of underground obstacles",
  4421. author = "Kimchi, Tali and Terkel, Joseph",
  4422. abstract = "Orientation by an animal inhabiting an underground environment
  4423. must be extremely efficient if it is to contend effectively with
  4424. the high energetic costs of excavating soil for a tunnel system.
  4425. We examined, in the field, the ability of a fossorial rodent,
  4426. the blind mole-rat, Spalax ehrenbergi, to detour different types
  4427. of obstacles blocking its tunnel and rejoin the disconnected
  4428. tunnel section. To create obstacles, we dug ditches, which we
  4429. either left open or filled with stone or wood. Most (77\%)
  4430. mole-rats reconnected the two parts of their tunnel and
  4431. accurately returned to their orginal path by digging a parallel
  4432. bypass tunnel around the obstacle at a distance of 10--20cm from
  4433. the open ditch boundaries or 3--8cm from the filled ditch
  4434. boundaries. When the ditch was placed asymmetrically across the
  4435. tunnel, the mole-rats detoured around the shorter side. These
  4436. findings demonstrate that mole-rats seem to be able to assess
  4437. the nature of an obstacle ahead and their own distance from the
  4438. obstacle boundaries, as well as the relative location of the far
  4439. section of disconnected tunnel. We suggest that mole-rats mainly
  4440. use reverberating self-produced seismic vibrations as a
  4441. mechanism to determine the size, nature and location of the
  4442. obstacle, as well as internal self-generated references to
  4443. determine their location relative to the disconnected tunnel
  4444. section. Copyright 2003 The Association for the Study of Animal
  4445. Behaviour. Published by Elsevier Ltd. All rights reserved.",
  4446. journal = "Anim. Behav.",
  4447. publisher = "Elsevier",
  4448. volume = 66,
  4449. number = 5,
  4450. pages = "885--891",
  4451. month = nov,
  4452. year = 2003
  4453. }
  4454. @ARTICLE{Hennig1976-es,
  4455. title = "The Effect of Distance Between Predator and Prey and the
  4456. Opportunity to Escape on Tonic Immobility in Anolis Carolinensis",
  4457. author = "Hennig, Charles W and Dunlap, William P and Gallup, Gordon G",
  4458. abstract = "The idea that tonic immobility (TI) may be a reaction to
  4459. predation has received increasing support in recent years. It
  4460. follows, from this view, that distance between predator and prey
  4461. and opportunity for escape should have predictable effects on
  4462. immobility. The first experiment showed that the presence of
  4463. large bushes, as an explicit escape manipulation, reduced
  4464. immobility durations in anoles (Anolis carolinensis) in
  4465. comparison to what occurred when they were immobilized in an open
  4466. area, with the effect being most evident the closer the predator
  4467. was to the prey. In the second experiment it was shown that close
  4468. proximity between anoles and the experimenter produced longer
  4469. durations of immobility in an open area, while a third experiment
  4470. showed that with bushes nearby this relationship was reversed;
  4471. that is, shorter durations of TI with anoles in close proximity
  4472. to the experimenter.",
  4473. journal = "Psychol. Rec.",
  4474. volume = 26,
  4475. number = 3,
  4476. pages = "312--320",
  4477. month = jul,
  4478. year = 1976
  4479. }
  4480. @ARTICLE{Eaton1991-cg,
  4481. title = "How stimulus direction determines the trajectory of the
  4482. Mauthner-initiated escape response in a teleost fish",
  4483. author = "Eaton, R C and Emberley, D S",
  4484. abstract = "Fishes use the Mauthner-initiated C-start for short-latency
  4485. evasion of predators. C-starts consist of a sudden turn (stage
  4486. 1) and a rapid acceleration (stage 2). We analyzed high-speed
  4487. cin{\'e} films of goldfish C-starts elicited by dropping a ball
  4488. into the water. It was previously thought that stage 1 angle
  4489. does not vary concomitantly with the angle of the threatening
  4490. stimulus relative to the position of the fish. We found,
  4491. however, a significant inverse relationship between the
  4492. direction of the impact of the ball and the angle turned by the
  4493. end of stage 1. When starting near a wall, or when its usual
  4494. trajectory was blocked by a wall, the fish used an escape route
  4495. that was not predictable from the stimulus angle. The fish did
  4496. not appear to correct its trajectory if it began to turn towards
  4497. the ball. This behavioral evidence supports the previous notion
  4498. that the underlying neural command is ballistic and does not use
  4499. sensory information from the stimulus once the movement begins.
  4500. If this is so, the fish probably utilizes information on
  4501. obstacle location in the interval leading up to the trigger
  4502. stimulus.",
  4503. journal = "J. Exp. Biol.",
  4504. publisher = "jeb.biologists.org",
  4505. volume = 161,
  4506. pages = "469--487",
  4507. month = nov,
  4508. year = 1991,
  4509. language = "en"
  4510. }
  4511. @ARTICLE{Herberholz2012-ib,
  4512. title = "Decision Making and Behavioral Choice during Predator Avoidance",
  4513. author = "Herberholz, Jens and Marquart, Gregory D",
  4514. abstract = "One of the most important decisions animals have to make is how
  4515. to respond to an attack from a potential predator. The response
  4516. must be prompt and appropriate to ensure survival. Invertebrates
  4517. have been important models in studying the underlying
  4518. neurobiology of the escape response due to their accessible
  4519. nervous systems and easily quantifiable behavioral output.
  4520. Moreover, invertebrates provide opportunities for investigating
  4521. these processes at a level of analysis not available in most
  4522. other organisms. Recently, there has been a renewed focus in
  4523. understanding how value-based calculations are made on the level
  4524. of the nervous system, i.e., when decisions are made under
  4525. conflicting circumstances, and the most desirable choice must be
  4526. selected by weighing the costs and benefits for each behavioral
  4527. choice. This article reviews samples from the current literature
  4528. on anti-predator decision making in invertebrates, from single
  4529. neurons to complex behaviors. Recent progress in understanding
  4530. the mechanisms underlying value-based behavioral decisions is
  4531. also discussed.",
  4532. journal = "Front. Neurosci.",
  4533. volume = 6,
  4534. pages = "125",
  4535. month = aug,
  4536. year = 2012,
  4537. keywords = "behavioral choice; decision making; escape; neural circuits;
  4538. predation",
  4539. language = "en"
  4540. }
  4541. % The entry below contains non-ASCII chars that could not be converted
  4542. % to a LaTeX equivalent.
  4543. @MISC{Eason2019-sm,
  4544. title = "Squirrels Do the Math: Flight Trajectories in Eastern Gray
  4545. Squirrels (Sciurus carolinensis)",
  4546. author = "Eason, Perri K and Nason, Lindsay D and Alexander, Jr., James E",
  4547. abstract = "Animals are under strong selective pressures to make correct
  4548. decisions when attempting to escape an approaching predator, and
  4549. not surprisingly many studies have shown that animals adjust
  4550. their flight initiation behavior in response to risk. However, we
  4551. have a poor understanding of animals' capability to select an
  4552. appropriate flight trajectory. We investigated whether eastern
  4553. gray squirrels would adjust their flight trajectory based on the
  4554. relative locations of the squirrel, the approaching threat, and
  4555. potential refuges. We used a person running toward a focal
  4556. squirrel (N = 122) as the threat and considered the three trees
  4557. nearest the squirrel and taller than 8 m to be potential refuges.
  4558. Squirrels were strongly affected by the angle () formed by the
  4559. locations of person, squirrel, and the three nearest trees. A
  4560. squirrel was less likely to run to the nearest tree (Tree 1) when
  4561. 1 was relatively acute, but also less likely to run to Tree 1
  4562. when 2 was obtuse, making Tree 2 a more attractive refuge. A
  4563. squirrel was more likely to run to Tree 1 if it was close and if
  4564. Tree 2 was relatively far. Subtle differences in the effects of
  4565. 1 versus 2 on squirrel refuge choice support the idea that
  4566. squirrels prefer a nearby refuge. Squirrels were more likely to
  4567. select Trees 2 and 3 rather than Tree 1 only when 2 was obtuse
  4568. (105). In contrast, most squirrels chose to run to Tree 1 when
  4569. 1 was greater than 65; thus squirrels were more likely to
  4570. choose Tree 1 even when doing so required running at least partly
  4571. toward the approaching threat. The decisions made by focal
  4572. squirrels provide evidence that this species' assessment of risk
  4573. is highly nuanced. A great deal of variation has been reported in
  4574. responses to predators within species. While part of the
  4575. variation may be due to strategic unpredictability on the part of
  4576. the prey, part of it may also be due to differences in flight
  4577. trajectory and refuge preferences that have not been well
  4578. studied.",
  4579. month = mar,
  4580. year = 2019
  4581. }
  4582. @ARTICLE{Fellini2011-wd,
  4583. title = "Geometric information is required for allothetic navigation in
  4584. mice",
  4585. author = "Fellini, Laetitia and Morellini, Fabio",
  4586. abstract = "In tasks for allothetic navigation, animals should orientate by
  4587. means of distal cues. We have previously shown that mice use
  4588. several forms of information to navigate, among which geometry,
  4589. i.e. the shape of the environment, seems to play an important
  4590. role. Here we investigated whether geometric features of the
  4591. environment are necessary for allothetic navigation in mice.
  4592. Mice were trained to navigate in a circular water maze by means
  4593. of four distal landmarks distributed either symmetrically
  4594. (symmetry group) or asymmetrically (asymmetry group) around the
  4595. maze. Thus, mice could locate a hidden platform by either
  4596. differentiating the landmarks based on their intrinsic features
  4597. (symmetry group) or in addition by geometric information, i.e.
  4598. based on the relative distances between landmarks (asymmetry
  4599. group). Data indicated that place learning occurred only in the
  4600. asymmetry group. The results support the idea that mice navigate
  4601. by using the relational properties between distal landmarks and
  4602. that geometric information is required for proper allothetic
  4603. navigation in this species.",
  4604. journal = "Behav. Brain Res.",
  4605. publisher = "Elsevier",
  4606. volume = 222,
  4607. number = 2,
  4608. pages = "380--384",
  4609. month = sep,
  4610. year = 2011,
  4611. language = "en"
  4612. }
  4613. @ARTICLE{Roberts2007-ec,
  4614. title = "Rats take correct novel routes and shortcuts in an enclosed maze",
  4615. author = "Roberts, William A and Cruz, Catherine and Tremblay, Joseph",
  4616. abstract = "In 3 experiments, rats were allowed to travel selected routes
  4617. along the internal alleys of a cross-maze that led from one
  4618. distinctive end box to another. The maze and procedures used were
  4619. designed to control the rats' ability to use intrinsic and
  4620. extrinsic cues to their location in the maze; thus, only the
  4621. internal geometry of the maze could be learned and used to travel
  4622. between one end box and another. After an initial exploration
  4623. phase, rats were given novel routes and shortcut tests that
  4624. involved peripheral alleys not before traveled. Rats chose the
  4625. correct novel path or shortcut significantly above chance on some
  4626. tests in Experiments 1 and 2 and significantly better than a
  4627. control group in Experiment 3. The findings suggest that rats
  4628. were able to compute novel routes and shortcuts within the maze
  4629. on the basis of limited experience with the internal geometry of
  4630. the maze.",
  4631. journal = "J. Exp. Psychol. Anim. Behav. Process.",
  4632. volume = 33,
  4633. number = 2,
  4634. pages = "79--91",
  4635. month = apr,
  4636. year = 2007,
  4637. language = "en"
  4638. }
  4639. @ARTICLE{Grieves2013-ji,
  4640. title = "Cognitive maps and spatial inference in animals: Rats fail to
  4641. take a novel shortcut, but can take a previously experienced one",
  4642. author = "Grieves, Roderick M and Dudchenko, Paul A",
  4643. abstract = "Previous work has shown that children are able to make a spatial
  4644. inference about adjacent locations that have only been
  4645. experienced indirectly (Hazen, Lockman, \& Pick, 1978). We
  4646. sought to replicate this finding in rats, on a conceptually
  4647. analogous task. In a first experiment, rats (n=8) were given 110
  4648. training trials on a task in which they entered a series of four
  4649. square environments via connecting alleyways. Following
  4650. training, we conducted a probe session in which the original
  4651. training route was blocked and three novel routes were
  4652. introduced, one of which led directly to the food reward.
  4653. Surprisingly, rats failed to choose this shortcut route over the
  4654. alternative routes. In a second experiment, following additional
  4655. training with a series of platforms that were visible from one
  4656. another, rats again failed to take a shortcut when given the
  4657. opportunity to do so. In a third experiment with naive rats
  4658. (n=11), a shortcut was chosen, but only by rats that were given
  4659. unrewarded preexposure to the shortcut route. These tests
  4660. suggest that, despite their dedicated neural representations of
  4661. location and direction, rats lack the capacity for a novel
  4662. spatial inference. For rats, the use of a shortcut requires
  4663. learning.",
  4664. journal = "Learn. Motiv.",
  4665. publisher = "Elsevier",
  4666. volume = 44,
  4667. number = 2,
  4668. pages = "81--92",
  4669. month = may,
  4670. year = 2013,
  4671. keywords = "Tolman; Maier; Cognitive map; Shortcutting; Intramaze landmarks;
  4672. Extramaze landmarks; Spatial inference"
  4673. }
  4674. @ARTICLE{Morellini2013-qr,
  4675. title = "Spatial memory tasks in rodents: what do they model?",
  4676. author = "Morellini, Fabio",
  4677. abstract = "The analysis of spatial learning and memory in rodents is
  4678. commonly used to investigate the mechanisms underlying certain
  4679. forms of human cognition and to model their dysfunction in
  4680. neuropsychiatric and neurodegenerative diseases. Proper
  4681. interpretation of rodent behavior in terms of spatial memory and
  4682. as a model of human cognitive functions is only possible if
  4683. various navigation strategies and factors controlling the
  4684. performance of the animal in a spatial task are taken into
  4685. consideration. The aim of this review is to describe the
  4686. experimental approaches that are being used for the study of
  4687. spatial memory in rats and mice and the way that they can be
  4688. interpreted in terms of general memory functions. After an
  4689. introduction to the classification of memory into various
  4690. categories and respective underlying neuroanatomical substrates,
  4691. I explain the concept of spatial memory and its measurement in
  4692. rats and mice by analysis of their navigation strategies.
  4693. Subsequently, I describe the most common paradigms for spatial
  4694. memory assessment with specific focus on methodological issues
  4695. relevant for the correct interpretation of the results in terms
  4696. of cognitive function. Finally, I present recent advances in the
  4697. use of spatial memory tasks to investigate episodic-like memory
  4698. in mice.",
  4699. journal = "Cell Tissue Res.",
  4700. volume = 354,
  4701. number = 1,
  4702. pages = "273--286",
  4703. month = oct,
  4704. year = 2013,
  4705. language = "en"
  4706. }
  4707. @MISC{noauthor_undated-ot,
  4708. title = "The travelling sales Rat",
  4709. howpublished = "\url{https://iopscience.iop.org/article/10.1088/1741-2560/8/6/065010/pdf?casa_token=9NbJ5AY4fAcAAAAA:wtKCXaD6lR9nIMBH4wCuthuuSp_UhnchIGvU7UldVIyT_Za-XIUFdiascZx3rgBaslD4Py_dQw}",
  4710. note = "Accessed: 2020-6-4"
  4711. }
  4712. @ARTICLE{Munteanu2016-dx,
  4713. title = "Take the long way home: Behaviour of a neotropical frog,
  4714. Allobates femoralis, in a detour task",
  4715. author = "Munteanu, Alexandru Marian and Starnberger, Iris and Pa{\v
  4716. s}ukonis, Andrius and Bugnyar, Thomas and H{\"o}dl, Walter and
  4717. Fitch, William Tecumseh",
  4718. abstract = "Detour behaviour, an individual's ability to reach its goal by
  4719. taking an indirect route, has been used to test spatial cognitive
  4720. abilities across a variety of taxa. Although many amphibians show
  4721. a strong homing ability, there is currently little evidence of
  4722. amphibian spatial cognitive flexibility. We tested whether a
  4723. territorial frog, Allobates femoralis, can flexibly adjust its
  4724. homing path when faced with an obstacle. We displaced male frogs
  4725. from their calling sites into the centre of circular arenas and
  4726. recorded their escape routes. In the first experiment we provided
  4727. an arena with equally high walls. In the second experiment we
  4728. doubled the height of the homeward facing wall. Finally, we
  4729. provided a tube as a shortcut through the high wall. In the
  4730. equal-height arena, most frogs chose to escape via the quadrant
  4731. facing their former calling site. However, when challenged with
  4732. different heights, nearly all frogs chose the low wall, directing
  4733. their movements away from the calling site. In the ``escape
  4734. tunnel'' experiment most frogs still chose the low wall. Our
  4735. results show that displaced A. femoralis males can flexibly
  4736. adjust their homing path and avoid (presumably energetically
  4737. costly) obstacles, providing experimental evidence of spatial
  4738. cognitive flexibility in an amphibian.",
  4739. journal = "Behav. Processes",
  4740. volume = 126,
  4741. pages = "71--75",
  4742. month = may,
  4743. year = 2016,
  4744. keywords = "Amphibian behaviour; Dendrobatidae; Homing; Obstacle avoidance;
  4745. Spatial cognitive flexibility",
  4746. language = "en"
  4747. }
  4748. @ARTICLE{Nesterova2009-em,
  4749. title = "Simple and integrated detours: field tests with Columbian ground
  4750. squirrels",
  4751. author = "Nesterova, Anna Pavlovna and Hansen, Frank",
  4752. abstract = "An internal representation of space offers flexibility to animals
  4753. during orientation and allows execution of short cuts and
  4754. detours. We tested the ability of 19 free-ranging Columbian
  4755. ground squirrels (Spermophilus columbianus) to perform integrated
  4756. detours that required travelling under- and aboveground.
  4757. Squirrels were individually tested on their territories (2 tests)
  4758. and in an arena (7 tests). During tests, animals could reach food
  4759. by running aboveground and then through tunnels. For the
  4760. territory tests, natural tunnels were available. For the arena
  4761. tests, animals used artificial tunnels within a fenced-in part of
  4762. the meadow. For the last arena test, tubes were placed
  4763. aboveground replicating the underground structure. In this test
  4764. animals were asked to make a simple detour, when the full path to
  4765. the goal was visible. On their territories, 41\% of squirrels
  4766. performed detours. All animals reached the food in the arena.
  4767. When choosing an arena detour, squirrels based their decision on
  4768. the proximity of the burrow as well as on whether it led to food.
  4769. On the last arena test, more squirrels performed correct detours
  4770. on the first attempt compared to other tests. The results suggest
  4771. that ground squirrels can perform simple and integrated detours,
  4772. but animals perform better if the full path is visible.",
  4773. journal = "Anim. Cogn.",
  4774. volume = 12,
  4775. number = 5,
  4776. pages = "655--670",
  4777. month = sep,
  4778. year = 2009,
  4779. language = "en"
  4780. }
  4781. @ARTICLE{Floreano2010-wg,
  4782. title = "Evolution of adaptive behaviour in robots by means of Darwinian
  4783. selection",
  4784. author = "Floreano, Dario and Keller, Laurent",
  4785. journal = "PLoS Biol.",
  4786. volume = 8,
  4787. number = 1,
  4788. pages = "e1000292",
  4789. month = jan,
  4790. year = 2010,
  4791. language = "en"
  4792. }
  4793. @ARTICLE{Cooper1999-nv,
  4794. title = "Escape behavior by prey blocked from entering the nearest refuge",
  4795. author = "{Cooper} and {Jr.} and William, E",
  4796. abstract = "Current models of optimal antipredation behavior do not apply to
  4797. prey blocked by a predator from access to the primary refuge
  4798. because the predator is closer than the optimal approach
  4799. distance and flight toward the refuge would increase risk. If
  4800. other alternative refuges are available, the prey should flee
  4801. toward the best alternative one. I studied the effect of an
  4802. approaching human simulated predator interposed between prey and
  4803. refuge on the use of alternative refuges and on
  4804. flight-initiation distance in the keeled earless lizard,
  4805. Holbrookia propinqua. When the predator approached on a line
  4806. between a lizard and its closest refuge, the lizard invariably
  4807. fled to or toward an alternative refuge. Lizards were
  4808. significantly more likely to use alternative refuges than
  4809. lizards approached on a line connecting the closest refuge,
  4810. prey, and predator, but with the lizard between the predator and
  4811. the refuge. Flight-initiation distance was significantly greater
  4812. for lizards having free access to the closest refuge than for
  4813. those blocked from it, perhaps because of the time required to
  4814. assess the new risk posed by blockage of the closest refuge, to
  4815. select the best alternative refuge, or to wait for the predator
  4816. to commit to a closing pattern before choosing the best flight
  4817. option.",
  4818. journal = "Can. J. Zool.",
  4819. publisher = "NRC Research Press",
  4820. volume = 77,
  4821. number = 4,
  4822. pages = "671--674",
  4823. month = sep,
  4824. year = 1999
  4825. }
  4826. @ARTICLE{Kopena2015-jl,
  4827. title = "Escape strategy of Schreiber's green lizards (Lacerta
  4828. schreiberi) is determined by environment but not season or sex",
  4829. author = "Kopena, Ren{\'a}ta and Herczeg, G{\'a}bor and L{\'o}pez, Pilar
  4830. and Mart{\'\i}n, Jos{\'e}",
  4831. abstract = "Antipredator escape behaviour varies with several
  4832. well-established sources of variation ranging from the physical
  4833. environment to reproductive status. However, the relative roles
  4834. of these sources are rarely assessed together. We measured (i)
  4835. the distance to the nearest refuge that Schreiber's green
  4836. lizards, Lacerta schreiberi, maintained before an attack (refuge
  4837. distance) and (ii) the distance lizards allowed a simulated
  4838. predator to approach before fleeing (flight initiation distance,
  4839. FID). Refuge distance was unaffected by studied variables.
  4840. However, FID was positively related to refuge distance on
  4841. grassy, but not on rocky substrates. Furthermore, refuge
  4842. distance and escape angle interacted in a substrate-independent
  4843. manner: lizards allowed predators close when refuges were close
  4844. or when lizards had to flee towards the predator. In contrast,
  4845. neither mating season nor sex affected FID. We suggest that the
  4846. escape strategy of L. schreiberi is determined more by the
  4847. physical environment than by sex or reproductive condition.",
  4848. journal = "Behaviour",
  4849. publisher = "Brill",
  4850. volume = 152,
  4851. number = 11,
  4852. pages = "1527--1542",
  4853. month = jan,
  4854. year = 2015,
  4855. keywords = "Biology \& Environmental Sciences; Biology; Journal",
  4856. language = "en"
  4857. }
  4858. @ARTICLE{Mattingly2005-qc,
  4859. title = "The choice of arboreal escape paths and its consequences for the
  4860. locomotor behaviour of four species of Anolis lizards",
  4861. author = "Mattingly, W Brett and Jayne, Bruce C",
  4862. abstract = "The direction and speed of escape locomotion can affect the
  4863. ability of an animal to evade a predator, and variation in
  4864. habitat structure often affects speed. Consequently, the escape
  4865. paths chosen by animals may affect their performance and
  4866. subsequent survival. Arboreal locomotion is well suited for
  4867. gaining insight into the choice of escape routes because of the
  4868. discrete paths formed by branches. Decreased branch diameter and
  4869. increased angles between branches can significantly decrease
  4870. locomotor speeds, but no previous study has determined whether
  4871. arboreal lizards selectively choose alternative paths. We
  4872. quantified choice of escape paths and locomotor performance of
  4873. four syntopic species of arboreal Anolis lizards in their natural
  4874. habitat and in the laboratory. In the field, species with shorter
  4875. limbs occurred more commonly on narrow perches than did
  4876. long-limbed species, but all species favoured escape paths with
  4877. larger-diameter perches and straighter interperch angles. Thus,
  4878. short-limbed species used narrower perches than long-limbed
  4879. species merely as a result of what they encountered, rather than
  4880. as a result of a biased choice at branching points. In natural
  4881. vegetation, choosing branches with the largest diameter often
  4882. results in the straightest path. However, in the laboratory, most
  4883. lizards preferred large-diameter perches with a sharp turn to
  4884. continuing a straight path onto a small-diameter perch. Although
  4885. an overriding preference for larger perch diameter may optimize
  4886. escape speed within a single perch, a maladaptive side-effect
  4887. could be a compromise of the overall rate of gaining distance
  4888. from starting points in paths with turns.",
  4889. journal = "Anim. Behav.",
  4890. volume = 70,
  4891. number = 6,
  4892. pages = "1239--1250",
  4893. month = dec,
  4894. year = 2005
  4895. }
  4896. @ARTICLE{Tajima2019-xa,
  4897. title = "Optimal policy for multi-alternative decisions",
  4898. author = "Tajima, Satohiro and Drugowitsch, Jan and Patel, Nisheet and
  4899. Pouget, Alexandre",
  4900. abstract = "Everyday decisions frequently require choosing among multiple
  4901. alternatives. Yet the optimal policy for such decisions is
  4902. unknown. Here we derive the normative policy for general
  4903. multi-alternative decisions. This strategy requires evidence
  4904. accumulation to nonlinear, time-dependent bounds that trigger
  4905. choices. A geometric symmetry in those boundaries allows the
  4906. optimal strategy to be implemented by a simple neural circuit
  4907. involving normalization with fixed decision bounds and an
  4908. urgency signal. The model captures several key features of the
  4909. response of decision-making neurons as well as the increase in
  4910. reaction time as a function of the number of alternatives, known
  4911. as Hick's law. In addition, we show that in the presence of
  4912. divisive normalization and internal variability, our model can
  4913. account for several so-called 'irrational' behaviors, such as
  4914. the similarity effect as well as the violation of both the
  4915. independence of irrelevant alternatives principle and the
  4916. regularity principle.",
  4917. journal = "Nat. Neurosci.",
  4918. publisher = "nature.com",
  4919. volume = 22,
  4920. number = 9,
  4921. pages = "1503--1511",
  4922. month = sep,
  4923. year = 2019,
  4924. language = "en"
  4925. }
  4926. @ARTICLE{Cooper1997-wy,
  4927. title = "Escape by a refuging prey, the broad-headed skink (Eumeces
  4928. laticeps)",
  4929. author = "Cooper, Jr., William E",
  4930. abstract = "Factors influencing escape to refuge by the broad-headed skink
  4931. (Eumeces laticeps) were examined by multiple regression and
  4932. correlation of quantitative escape variables and distance and
  4933. direction to refuge. I simulated a predator by walking toward a
  4934. lizard and recorded aspects of escape. Approach distance
  4935. (distance from me when escape began) increased with distance and
  4936. angle to refuge, suggesting that the skinks assessed that risk
  4937. increased with relative times required for prey and predator to
  4938. reach the refuge. Distance fled was affected jointly by distance
  4939. from the predator when escape began and distance to refuge; it
  4940. increased with distance to refuge. It also increased with the
  4941. angle between the predator's path and refuge due to declining
  4942. distance from the predator per unit distance fled. Direction to
  4943. the nearest refuge and direction fled were nearly identical.
  4944. Distance and direction to refuge should strongly affect escape
  4945. behaviour in prey that are active some distance from refuges but
  4946. rely on them to avoid predation. These relationships may be
  4947. weaker or absent in anachoric species (those nearly continuously
  4948. occupying refuges) and those remaining close to refuges, as well
  4949. as in species relying more on speed and fleeing for long
  4950. distances than on refuges.",
  4951. journal = "Can. J. Zool.",
  4952. publisher = "NRC Research Press",
  4953. volume = 75,
  4954. number = 6,
  4955. pages = "943--947",
  4956. month = jun,
  4957. year = 1997
  4958. }
  4959. % The entry below contains non-ASCII chars that could not be converted
  4960. % to a LaTeX equivalent.
  4961. @ARTICLE{Domenici2010-ma,
  4962. title = "Context-dependent variability in the components of fish escape
  4963. response: integrating locomotor performance and behavior",
  4964. author = "Domenici, Paolo",
  4965. abstract = "Escape responses are used by most fish species in order to avoid
  4966. predation. Escape responses include a number of behavioral and
  4967. kinematic components, such as responsiveness, reaction distance,
  4968. escape latency, directionality, and distance‐derived
  4969. performance. All of these components can contribute to escape
  4970. success. Work on the context‐dependent variability has focused
  4971. on reaction distance, and suggests that this component is
  4972. largely determined by the relative cost and benefits of escaping
  4973. (economic …",
  4974. journal = "J. Exp. Zool. A Ecol. Genet. Physiol.",
  4975. publisher = "Wiley Online Library",
  4976. volume = 313,
  4977. number = 2,
  4978. pages = "59--79",
  4979. year = 2010
  4980. }
  4981. @ARTICLE{Husak2006-yq,
  4982. title = "Does survival depend on how fast you can run or how fast you do
  4983. run?",
  4984. author = "Husak, J F",
  4985. abstract = "Summary 1 Natural selection is generally thought to operate on
  4986. organisms? maximal abilities to perform ecological tasks in
  4987. nature (i.e. whole-animal performance). However, selection may
  4988. instead operate on the manner in which that performance trait is
  4989. used (i.e. ?ecological performance?). 2 I tested whether
  4990. survival of adult Collared Lizards (Crotaphytus collaris)
  4991. depended on maximal sprint speed capacity or on the speed at
  4992. which they actually performed two important ecological tasks:
  4993. chasing a prey item and escaping a predator. 3 Maximal sprint
  4994. speed did not significantly predict annual survival as
  4995. determined by daily censuses of the site the following season,
  4996. nor did speed while foraging, but speed while escaping a
  4997. predator did. Survival also was positively related to the
  4998. proportion of maximal capacity used while escaping. 4 These
  4999. results suggest that selection may operate on ecological
  5000. performance that is constrained, but not necessarily determined,
  5001. by maximal performance capacity, suggesting that researchers
  5002. should consider how organisms utilize maximal performance in
  5003. nature when testing for a performance?survival relationship.",
  5004. journal = "Funct. Ecol.",
  5005. publisher = "Wiley Online Library",
  5006. volume = 20,
  5007. number = 6,
  5008. pages = "1080--1086",
  5009. month = dec,
  5010. year = 2006
  5011. }
  5012. % The entry below contains non-ASCII chars that could not be converted
  5013. % to a LaTeX equivalent.
  5014. @ARTICLE{Barthas2017-zs,
  5015. title = "Secondary motor cortex: where `sensory'meets `motor'in the
  5016. rodent frontal cortex",
  5017. author = "Barthas, Florent and Kwan, Alex C",
  5018. abstract = "In rodents, the medial aspect of the secondary motor cortex (M2)
  5019. is known by other names, including medial agranular cortex
  5020. (AGm), medial precentral cortex (PrCm), and frontal orienting
  5021. field (FOF). As a subdivision of the medial prefrontal cortex
  5022. (mPFC), M2 can be defined by a distinct set of afferent and
  5023. efferent connections, microstimulation responses, and lesion
  5024. outcomes. However, the behavioral role of M2 remains mysterious.
  5025. Here, we focus on evidence from rodent studies, highlighting
  5026. recent findings of early and context …",
  5027. journal = "Trends Neurosci.",
  5028. publisher = "Elsevier",
  5029. volume = 40,
  5030. number = 3,
  5031. pages = "181--193",
  5032. year = 2017,
  5033. keywords = "To Read"
  5034. }
  5035. @ARTICLE{Tchernichovski1995-qv,
  5036. title = "A phase plane representation of rat exploratory behavior",
  5037. author = "Tchernichovski, O and Golani, I",
  5038. abstract = "Rat spontaneous spatial behavior is considered to be stochastic
  5039. and is therefore commonly analyzed in terms of cumulative
  5040. measures. Here, we suggest a method which generates a
  5041. moment-to-moment representation of this behavior. It has been
  5042. proposed earlier that rat spatial behavior can be partitioned
  5043. into natural units termed excursions (round trips) performed from
  5044. a reference place termed the rat's home base. We offer a phase
  5045. plane representation of excursions (plotting the rat's momentary
  5046. location against its momentary velocity). The results reveal a
  5047. geometrical pattern, typical of young age and early exposure. It
  5048. consists of low velocity and intermittent progression while
  5049. moving away from the home base (upstream segment), and high
  5050. velocity while moving back to it (downstream segment). The
  5051. asymmetry between the two segments defines a field of
  5052. significance in the rat's operational world. This field undergoes
  5053. regular transformations, revealing thereby the rat's strategy of
  5054. occupancy of the environment. The presented dynamics could
  5055. provide a framework for the interpretation of concurrent neural
  5056. events associated with navigation and spatial memory.",
  5057. journal = "J. Neurosci. Methods",
  5058. volume = 62,
  5059. number = "1-2",
  5060. pages = "21--27",
  5061. month = nov,
  5062. year = 1995,
  5063. language = "en"
  5064. }
  5065. @MISC{noauthor_undated-wh,
  5066. title = "{PIC88.pdf}"
  5067. }
  5068. @ARTICLE{Egnor2017-sl,
  5069. title = "Spatial Memory: Mice Quickly Learn a Safe Haven",
  5070. author = "Egnor, S E Roian",
  5071. abstract = "New work on innate escape behavior shows that mice spontaneously
  5072. form a spatially precise memory of the location of shelter, which
  5073. is laid down quickly and updated continuously.",
  5074. journal = "Curr. Biol.",
  5075. volume = 27,
  5076. number = 10,
  5077. pages = "R388--R390",
  5078. month = may,
  5079. year = 2017,
  5080. language = "en"
  5081. }
  5082. @ARTICLE{Li2018-zt,
  5083. title = "Modulation of Innate Defensive Responses by Locus
  5084. {Coeruleus-Superior} Colliculus Circuit",
  5085. author = "Li, Lei and Wang, Liping",
  5086. abstract = "Among key survival circuits, defensive response circuits are one
  5087. of the most intensively studied. A consensus is emerging that
  5088. multiple, independent circuitries are involved in different
  5089. conditioned and unconditioned defensive responses. Investigating
  5090. these well-conserved defensive responses would help us to
  5091. decipher the basic working mechanism of the brain at a circuitry
  5092. level and thus shed light on new diagnoses and treatments for
  5093. neural diseases and disorders. We showed that the visually evoked
  5094. innate defensive response was modulated by a locus
  5095. coeruleus-superior colliculus (LC-SC) projection. Our work
  5096. demonstrates that as conserved and instinctive as the survival
  5097. circuits are, they are flexible and subject to fine-tuned
  5098. modulation by experience or internal states of the animals. Here,
  5099. we provide more data to further discuss the possible downstream
  5100. mechanisms of the LC-SC pathway for this important modulation of
  5101. the defensive response, the wide range of flight latency between
  5102. individual flight responses, and the interpretations of our data
  5103. with additional statistical analysis.",
  5104. journal = "J. Exp. Neurosci.",
  5105. volume = 12,
  5106. pages = "1179069518792035",
  5107. month = aug,
  5108. year = 2018,
  5109. keywords = "Locus coeruleus; defensive circuitry; looming; modulation;
  5110. norepinephrine; stress; superior colliculus",
  5111. language = "en"
  5112. }
  5113. @ARTICLE{Gentry1964-sw,
  5114. title = "Homing in the {Old-Field} Mouse",
  5115. author = "Gentry, John B",
  5116. abstract = "Abstract. Homing was successful in 31 of 39 old-field mice
  5117. (Peromyscus polionotus) released from the center of a 9-acre
  5118. plowed field 340 to 640 feet from trap",
  5119. journal = "J. Mammal.",
  5120. publisher = "Oxford Academic",
  5121. volume = 45,
  5122. number = 2,
  5123. pages = "276--283",
  5124. month = may,
  5125. year = 1964
  5126. }
  5127. @UNPUBLISHED{Seidenbecher2019-im,
  5128. title = "Foraging fruit flies mix navigational and learning-based
  5129. decision-making strategies",
  5130. author = "Seidenbecher, Sophie E and Sanders, Joshua I and von Philipsborn,
  5131. Anne C and Kvitsiani, Duda",
  5132. abstract = "Abstract Animals often navigate environments that are uncertain,
  5133. volatile and complex, making it challenging to locate reliable
  5134. food sources. Therefore, it is not surprising that many species
  5135. evolved multiple, parallel and complementary foraging strategies
  5136. to survive. Current research on animal behavior is largely driven
  5137. by a reductionist approach and attempts to study one particular
  5138. aspect of behavior in isolation. This is justified by the huge
  5139. success of past and current research in understanding neural
  5140. circuit mechanisms of behaviors. But focusing on only one aspect
  5141. of behaviors obscures their inherent multidimensional nature. To
  5142. fill this gap we aimed to identify and characterize distinct
  5143. behavioral modules using a simple reward foraging assay. For this
  5144. we developed a single-animal, trial-based probabilistic foraging
  5145. task, where freely walking fruit flies experience optogenetic
  5146. sugar-receptor neuron stimulation. By carefully analyzing the
  5147. walking trajectories of flies, we were able to dissect the
  5148. animals foraging decisions into multiple underlying systems. We
  5149. show that flies perform local searches, cue-based navigation and
  5150. learn task relevant contingencies. Using probabilistic reward
  5151. delivery allowed us to bid several competing reinforcement
  5152. learning (RL) models against each other. We discover that flies
  5153. accumulate chosen option values, forget unchosen option values
  5154. and seek novelty. We further show that distinct behavioral
  5155. modules -learning and navigation-based systems-cooperate,
  5156. suggesting that reinforcement learning in flies operates on
  5157. dimensionality reduced representations. We therefore argue that
  5158. animals will apply combinations of multiple behavioral strategies
  5159. to generate foraging decisions.",
  5160. journal = "bioRxiv",
  5161. pages = "842096",
  5162. month = nov,
  5163. year = 2019,
  5164. language = "en"
  5165. }
  5166. @UNPUBLISHED{Kaplan2017-ez,
  5167. title = "Planning and navigation as active inference",
  5168. author = "Kaplan, Raphael and Friston, Karl J",
  5169. abstract = "Abstract This paper introduces an active inference formulation of
  5170. planning and navigation. It illustrates how the
  5171. exploitation--exploration dilemma is dissolved by acting to
  5172. minimise uncertainty (i.e., expected surprise or free energy). We
  5173. use simulations of a maze problem to illustrate how agents can
  5174. solve quite complicated problems using context sensitive prior
  5175. preferences to form subgoals. Our focus is on how epistemic
  5176. behaviour -- driven by novelty and the imperative to reduce
  5177. uncertainty about the world -- contextualises pragmatic or
  5178. goal-directed behaviour. Using simulations, we illustrate the
  5179. underlying process theory with synthetic behavioural and
  5180. electrophysiological responses during exploration of a maze and
  5181. subsequent navigation to a target location. An interesting
  5182. phenomenon that emerged from the simulations was a putative
  5183. distinction between `place cells' -- that fire when a subgoal is
  5184. reached -- and `path cells' -- that fire until a subgoal is
  5185. reached.",
  5186. journal = "bioRxiv",
  5187. pages = "230599",
  5188. month = dec,
  5189. year = 2017,
  5190. language = "en"
  5191. }
  5192. @UNPUBLISHED{Sutton2018-vu,
  5193. title = "Born to run? Quantifying the balance of prior bias and new
  5194. information in prey escape decisions",
  5195. author = "Sutton, Nicholas M and O'Dwyer, James P",
  5196. abstract = "Abstract Animal behaviors can often be challenging to model and
  5197. predict, though optimality theory has improved our ability to do
  5198. so. While many qualitative predictions of behavior exist,
  5199. accurate quantitative models, tested by empirical data, are often
  5200. lacking. This is likely due to variation in biases across
  5201. individuals and variation in the way new information is gathered
  5202. and used. We propose a modeling framework based on a novel
  5203. interpretation of Bayes' theorem to integrate optimization of
  5204. energetic constraints with both prior biases and specific sources
  5205. of new information gathered by individuals. We present methods
  5206. for inferring distributions of prior biases within populations
  5207. rather than assuming known priors, as is common in Bayesian
  5208. approaches to modelling behavior, and for evaluating the goodness
  5209. of fit of overall model descriptions. We apply this framework to
  5210. predict optimal escape during predator-prey encounters, based on
  5211. prior biases and variation in what information prey use. Using
  5212. this approach we collected and analyzed data characterizing
  5213. white-tailed deer (Odocoileus virginianus) escape behavior in
  5214. response to human approaches. We found that distance to predator
  5215. alone was not sufficient for predicting deer flight response, and
  5216. have shown that the inclusion of additional information is
  5217. necessary. Additionally, we compared differences in the inferred
  5218. distributions of prior biases across different populations and
  5219. discuss the possible role of human activity in influencing these
  5220. distributions.",
  5221. journal = "bioRxiv",
  5222. pages = "297218",
  5223. month = apr,
  5224. year = 2018,
  5225. language = "en"
  5226. }
  5227. @UNPUBLISHED{Vale2020-yy,
  5228. title = "A cortico-collicular circuit for accurate orientation to shelter
  5229. during escape",
  5230. author = "Vale, Ruben and Campagner, Dario and Iordanidou, Panagiota and
  5231. Arocas, Oriol Pav{\'o}n and Tan, Yu Lin and Vanessa Stempel, A
  5232. and Keshavarzi, Sepideh and Petersen, Rasmus S and Margrie, Troy
  5233. W and Branco, Tiago",
  5234. abstract = "When faced with predatorial threats, escaping towards shelter is
  5235. an adaptive action that offers long-term protection against the
  5236. attacker. From crustaceans to mammals, animals rely on knowledge
  5237. of safe locations in the environment to rapidly execute
  5238. shelter-directed escape actions[1][1]--[3][2]. While previous
  5239. work has identified neural mechanisms of instinctive
  5240. escape[4][3]--[9][4], it is not known how the escape circuit
  5241. incorporates spatial information to execute rapid and accurate
  5242. flights to safety. Here we show that mouse retrosplenial cortex
  5243. (RSP) and superior colliculus (SC) form a monosynaptic circuit
  5244. that continuously encodes the shelter direction. Inactivation of
  5245. SC-projecting RSP neurons decreases SC shelter-direction tuning
  5246. while preserving SC motor function. Moreover, specific
  5247. inactivation of RSP input onto SC neurons disrupts orientation
  5248. and subsequent escapes to shelter, but not orientation accuracy
  5249. to a sensory cue. We conclude that the RSC-SC circuit supports an
  5250. egocentric representation of shelter direction and is necessary
  5251. for optimal shelter-directed escapes. This cortical-subcortical
  5252. interface may be a general blueprint for increasing the
  5253. sophistication and flexibility of instinctive behaviours. \#\#\#
  5254. Competing Interest Statement The authors have declared no
  5255. competing interest. [1]: \#ref-1 [2]: \#ref-3 [3]: \#ref-4 [4]:
  5256. \#ref-9",
  5257. journal = "bioRxiv",
  5258. pages = "2020.05.26.117598",
  5259. month = may,
  5260. year = 2020,
  5261. language = "en"
  5262. }
  5263. % The entry below contains non-ASCII chars that could not be converted
  5264. % to a LaTeX equivalent.
  5265. @UNPUBLISHED{Kawabata2020-yt,
  5266. title = "Geometrical model explains multiple preferred escape trajectories
  5267. of fish",
  5268. author = "Kawabata, Yuuki and Akada, Hideyuki and Shimatani, Ken-Ichiro and
  5269. Nishihara, Gregory N and Kimura, Hibiki and Nozomi, Nishiumi and
  5270. Domenici, Paolo",
  5271. abstract = "Abstract To evade predators, many prey perform rapid escape
  5272. movements. The resulting escape trajectory (ET) -- measured as
  5273. the angle of escape direction relative to the predator's approach
  5274. path -- plays a major role in avoiding predation. Previous
  5275. geometrical models predict a single ET; however, many animals
  5276. (fish and other animal taxa) show highly variable ETs with
  5277. multiple preferred directions. Although such a high ET
  5278. variability may confer unpredictability, preventing predators
  5279. from adopting counter-strategies, the reasons why animals prefer
  5280. specific multiple ETs remain unclear. Here, we constructed a
  5281. novel geometrical model in which Tdiff (the time difference
  5282. between the prey entering the safety zone and the predator
  5283. reaching that entry point) is expected to be maximized. We tested
  5284. this prediction by analyzing the escape responses of Pagrus major
  5285. attacked by a dummy predator. At each initial body orientation of
  5286. the prey relative to the predator, our model predicts a
  5287. multimodal ET with an optimal ET at the maximum Tdiff (Tdiff,1)
  5288. and a suboptimal ET at a second local maximum of Tdiff (Tdiff,2).
  5289. Our experiments show that when Tdiff, 1--Tdiff, 2 is negligible,
  5290. the prey uses optimal or suboptimal ETs to a similar extent, in
  5291. line with the idea of unpredictability. The experimentally
  5292. observed ET distribution is consistent with the model, showing
  5293. two large peaks at 110--130° and 170--180° away from the
  5294. predator. Because various animal taxa show multiple preferred ETs
  5295. similar to those observed here, this behavioral phenotype may
  5296. result from convergent evolution that combines maximal Tdiff with
  5297. a high level of unpredictability.Significance Statement Animals
  5298. from many taxa escape from suddenly approaching threats, such as
  5299. ambush predators, by using multiple preferred escape
  5300. trajectories. However, the reason why these multiple preferred
  5301. escape trajectories are used is still unknown. By fitting a newly
  5302. constructed model to the empirical escape response data, we show
  5303. that the seemingly complex multiple preferred escape trajectories
  5304. can arise from a simple geometrical rule which maximizes the time
  5305. difference between when the prey enters the safety zone and when
  5306. the predator reaches that entry point. Our results open new
  5307. avenues of investigation for understanding how animals choose
  5308. their escape trajectories from behavioral and neurosensory
  5309. perspectives.",
  5310. journal = "bioRxiv",
  5311. pages = "2020.04.27.049833",
  5312. month = apr,
  5313. year = 2020,
  5314. language = "en"
  5315. }
  5316. @ARTICLE{Olson2020-lm,
  5317. title = "Secondary Motor Cortex Transforms Spatial Information into
  5318. Planned Action during Navigation",
  5319. author = "Olson, Jacob M and Li, Jamie K and Montgomery, Sarah E and Nitz,
  5320. Douglas A",
  5321. abstract = "Fluid navigation requires constant updating of planned movements
  5322. to adapt to evolving obstacles and goals. For that reason, a
  5323. neural substrate for navigation demands spatial and
  5324. environmental information and the ability to effect actions
  5325. through efferents. The secondary motor cortex (M2) is a prime
  5326. candidate for this role given its interconnectivity with
  5327. association cortices that encode spatial relationships and its
  5328. projection to the primary motor cortex. Here, we report that M2
  5329. neurons robustly encode both planned and current left/right
  5330. turning actions across multiple turn locations in a multi-route
  5331. navigational task. Comparisons within a common statistical
  5332. framework reveal that M2 neurons differentiate contextual
  5333. factors, including environmental position, route, action
  5334. sequence, orientation, and choice availability. Despite
  5335. significant modulation by environmental factors, action
  5336. planning, and execution are the dominant output signals of M2
  5337. neurons. These results identify the M2 as a structure
  5338. integrating spatial information toward the updating of planned
  5339. movements.",
  5340. journal = "Curr. Biol.",
  5341. publisher = "Elsevier",
  5342. volume = 30,
  5343. number = 10,
  5344. pages = "1845--1854.e4",
  5345. month = may,
  5346. year = 2020,
  5347. keywords = "M2; action; allocentric; cortical circuits; decision making;
  5348. egocentric; in vivo electrophysiology; navigation; parietal
  5349. cortex; retrosplenial cortex; systems neuroscience;Locomotion",
  5350. language = "en"
  5351. }
  5352. @ARTICLE{Jordan2008-dx,
  5353. title = "Descending command systems for the initiation of locomotion in
  5354. mammals",
  5355. author = "Jordan, Larry M and Liu, Jun and Hedlund, Peter B and Akay,
  5356. Turgay and Pearson, Keir G",
  5357. abstract = "Neurons in the brainstem implicated in the initiation of
  5358. locomotion include glutamatergic, noradrenergic (NA),
  5359. dopaminergic (DA), and serotonergic (5-HT) neurons giving rise
  5360. to descending tracts. Glutamate antagonists block mesencephalic
  5361. locomotor region-induced and spontaneous locomotion, and
  5362. glutamatergic agonists induce locomotion in spinal animals. NA
  5363. and 5-HT inputs to the spinal cord originate in the brainstem,
  5364. while the descending dopaminergic pathway originates in the
  5365. hypothalamus. Agonists acting at NA, DA or 5-HT receptors
  5366. facilitate or induce locomotion in spinal animals. 5-HT neurons
  5367. located in the parapyramidal region (PPR) produce locomotion
  5368. when stimulated in the isolated neonatal rat brainstem-spinal
  5369. cord preparation, and they constitute the first anatomically
  5370. discrete group of spinally-projecting neurons demonstrated to be
  5371. involved in the initiation of locomotion in mammals. Neurons in
  5372. the PPR are activated during treadmill locomotion in adult rats.
  5373. Locomotion evoked from the PPR is mediated by 5-HT(7) and
  5374. 5-HT(2A) receptors, and 5-HT(7) antagonists block locomotion in
  5375. cat, rat and mouse preparations, but have little effect in mice
  5376. lacking 5-HT(7) receptors. 5-HT induced activity in 5-HT(7)
  5377. knockout mice is rhythmic, but coordination among flexor and
  5378. extensor motor nuclei and left and right sides of the spinal
  5379. cord is disrupted. In the adult wild-type mouse, 5-HT(7)
  5380. receptor antagonists impair locomotion, producing patterns of
  5381. activity resembling those induced by 5-HT in 5-HT(7) knockout
  5382. mice. 5-HT(7) receptor antagonists have a reduced effect on
  5383. locomotion in adult 5-HT(7) receptor knockout mice. We conclude
  5384. that the PPR is the source of a descending 5-HT command pathway
  5385. that activates the CPG via 5-HT(7) and 5-HT(2A) receptors.
  5386. Further experiments are necessary to define the putative
  5387. glutamatergic, DA, and NA command pathways.",
  5388. journal = "Brain Res. Rev.",
  5389. publisher = "Elsevier",
  5390. volume = 57,
  5391. number = 1,
  5392. pages = "183--191",
  5393. month = jan,
  5394. year = 2008,
  5395. language = "en"
  5396. }
  5397. @ARTICLE{Kaneshige2018-cg,
  5398. title = "A Descending Circuit Derived From the Superior Colliculus
  5399. Modulates Vibrissal Movements",
  5400. author = "Kaneshige, Miki and Shibata, Ken-Ichi and Matsubayashi, Jun and
  5401. Mitani, Akira and Furuta, Takahiro",
  5402. abstract = "The superior colliculus (SC) is an essential structure for the
  5403. control of eye movements. In rodents, the SC is also considered
  5404. to play an important role in whisking behavior, in which animals
  5405. actively move their vibrissae (mechanosensors) to gather tactile
  5406. information about the space around them during exploration. We
  5407. investigated how the SC contributes to vibrissal movement
  5408. control. We found that when the SC was unilaterally lesioned,
  5409. the resting position of the vibrissae shifted backward on the
  5410. side contralateral to the lesion. The unilateral SC lesion also
  5411. induced an increase in the whisking amplitude on the
  5412. contralateral side. To explore the anatomical basis for SC
  5413. involvement in vibrissal movement control, we then
  5414. quantitatively evaluated axonal projections from the SC to the
  5415. brainstem using neuronal labeling with a virus vector. Neurons
  5416. of the SC mainly sent axons to the contralateral side in the
  5417. lower brainstem. We found that the facial nucleus received input
  5418. directly from the SC, and that the descending projections from
  5419. the SC also reached the intermediate reticular formation and
  5420. pre-B{\"o}tzinger complex, which are both considered to contain
  5421. neural oscillators generating rhythmic movements of the
  5422. vibrissae. Together, these results indicate the existence of a
  5423. neural circuit in which the SC modulates vibrissal movements
  5424. mainly on the contralateral side, via direct connections to
  5425. motoneurons, and via indirect connections including the central
  5426. pattern generators.",
  5427. journal = "Front. Neural Circuits",
  5428. publisher = "frontiersin.org",
  5429. volume = 12,
  5430. pages = "100",
  5431. month = nov,
  5432. year = 2018,
  5433. keywords = "CPGs; anterograde tracing; kinematic analysis; premotor neurons;
  5434. rat; whisker",
  5435. language = "en"
  5436. }
  5437. @ARTICLE{Ryczko2013-pw,
  5438. title = "The multifunctional mesencephalic locomotor region",
  5439. author = "Ryczko, Dimitri and Dubuc, R{\'e}jean",
  5440. abstract = "In 1966, Shik, Severin and Orlovskii discovered that electrical
  5441. stimulation of a region at the junction between the midbrain and
  5442. hindbrain elicited controlled walking and running in the cat.
  5443. The region was named Mesencephalic Locomotor Region (MLR). Since
  5444. then, this locomotor center was shown to control locomotion in
  5445. various vertebrate species, including the lamprey, salamander,
  5446. stingray, rat, guinea-pig, rabbit or monkey. In human subjects
  5447. asked to imagine they are walking, there is an increased
  5448. activity in brainstem nuclei corresponding to the MLR (i.e.
  5449. pedunculopontine, cuneiform and subcuneiform nuclei). Clinicians
  5450. are now stimulating (deep brain stimulation) structures
  5451. considered to be part of the MLR to alleviate locomotor symptoms
  5452. of patients with Parkinson's disease. However, the anatomical
  5453. constituents of the MLR still remain a matter of debate,
  5454. especially relative to the pedunculopontine, cuneiform and
  5455. subcuneiform nuclei. Furthermore, recent studies in lampreys
  5456. have revealed that the MLR is more complex than a simple relay
  5457. in a serial descending pathway activating the spinal locomotor
  5458. circuits. It has multiple functions. Our goal is to review the
  5459. current knowledge relative to the anatomical constituents of the
  5460. MLR, and its physiological role, from lamprey to man. We will
  5461. discuss these results in the context of the recent clinical
  5462. studies involving stimulation of the MLR in patients with
  5463. Parkinson's disease.",
  5464. journal = "Curr. Pharm. Des.",
  5465. publisher = "ingentaconnect.com",
  5466. volume = 19,
  5467. number = 24,
  5468. pages = "4448--4470",
  5469. year = 2013,
  5470. keywords = "Locomotion",
  5471. language = "en"
  5472. }
  5473. @ARTICLE{Basso2017-nd,
  5474. title = "Circuits for Action and Cognition: A View from the Superior
  5475. Colliculus",
  5476. author = "Basso, Michele A and May, Paul J",
  5477. abstract = "The superior colliculus is one of the most well-studied
  5478. structures in the brain, and with each new report, its proposed
  5479. role in behavior seems to increase in complexity. Forty years of
  5480. evidence show that the colliculus is critical for reorienting an
  5481. organism toward objects of interest. In monkeys, this involves
  5482. saccadic eye movements. Recent work in the monkey colliculus and
  5483. in the homologous optic tectum of the bird extends our
  5484. understanding of the role of the colliculus in higher mental
  5485. functions, such as attention and decision making. In this
  5486. review, we highlight some of these recent results, as well as
  5487. those capitalizing on circuit-based methodologies using
  5488. transgenic mice models, to understand the contribution of the
  5489. colliculus to attention and decision making. The wealth of
  5490. information we have about the colliculus, together with new
  5491. tools, provides a unique opportunity to obtain a detailed
  5492. accounting of the neurons, circuits, and computations that
  5493. underlie complex behavior.",
  5494. journal = "Annu Rev Vis Sci",
  5495. publisher = "annualreviews.org",
  5496. volume = 3,
  5497. pages = "197--226",
  5498. month = sep,
  5499. year = 2017,
  5500. keywords = "attention; decision making; movement; normalization; orienting;
  5501. population coding; saccades; vision",
  5502. language = "en"
  5503. }
  5504. @ARTICLE{Wolf2015-dq,
  5505. title = "An integrative role for the superior colliculus in selecting
  5506. targets for movements",
  5507. author = "Wolf, Andrew B and Lintz, Mario J and Costabile, Jamie D and
  5508. Thompson, John A and Stubblefield, Elizabeth A and Felsen, Gidon",
  5509. abstract = "A fundamental goal of systems neuroscience is to understand the
  5510. neural mechanisms underlying decision making. The midbrain
  5511. superior colliculus (SC) is known to be central to the selection
  5512. of one among many potential spatial targets for movements, which
  5513. represents an important form of decision making that is
  5514. tractable to rigorous experimental investigation. In this
  5515. review, we first discuss data from mammalian models-including
  5516. primates, cats, and rodents-that inform our understanding of how
  5517. neural activity in the SC underlies the selection of targets for
  5518. movements. We then examine the anatomy and physiology of inputs
  5519. to the SC from three key regions that are themselves implicated
  5520. in motor decisions-the basal ganglia, parabrachial region, and
  5521. neocortex-and discuss how they may influence SC activity related
  5522. to target selection. Finally, we discuss the potential for
  5523. methodological advances to further our understanding of the
  5524. neural bases of target selection. Our overarching goal is to
  5525. synthesize what is known about how the SC and its inputs act
  5526. together to mediate the selection of targets for movements, to
  5527. highlight open questions about this process, and to spur future
  5528. studies addressing these questions.",
  5529. journal = "J. Neurophysiol.",
  5530. publisher = "physiology.org",
  5531. volume = 114,
  5532. number = 4,
  5533. pages = "2118--2131",
  5534. month = oct,
  5535. year = 2015,
  5536. keywords = "decision making; laterodorsal tegmental nucleus; motor planning;
  5537. pedunculopontine tegmental nucleus; substantia nigra",
  5538. language = "en"
  5539. }
  5540. @ARTICLE{Grillner2008-ev,
  5541. title = "Neural bases of goal-directed locomotion in vertebrates---An
  5542. overview",
  5543. author = "Grillner, Sten and Wall{\'e}n, Peter and Saitoh, Kazuya and
  5544. Kozlov, Alexander and Robertson, Brita",
  5545. abstract = "The different neural control systems involved in goal-directed
  5546. vertebrate locomotion are reviewed. They include not only the
  5547. central pattern generator networks in the spinal cord that
  5548. generate the basic locomotor synergy and the brainstem command
  5549. systems for locomotion but also the control systems for steering
  5550. and control of body orientation (posture) and finally the neural
  5551. structures responsible for determining which motor programs
  5552. should be turned on in a given instant. The role of the basal
  5553. ganglia is considered in this context. The review summarizes the
  5554. available information from a general vertebrate perspective, but
  5555. specific examples are often derived from the lamprey, which
  5556. provides the most detailed information when considering cellular
  5557. and network perspectives.",
  5558. journal = "Brain Res. Rev.",
  5559. publisher = "Elsevier",
  5560. volume = 57,
  5561. number = 1,
  5562. pages = "2--12",
  5563. month = jan,
  5564. year = 2008,
  5565. keywords = "Basal ganglia; Lamprey; Central pattern generator; Tectum; Brain
  5566. stem--spinal cord; Modeling"
  5567. }
  5568. @ARTICLE{Drew2004-kl,
  5569. title = "Cortical and brainstem control of locomotion",
  5570. author = "Drew, Trevor and Prentice, Stephen and Schepens,
  5571. B{\'e}n{\'e}dicte",
  5572. abstract = "While a basic locomotor rhythm is centrally generated by spinal
  5573. circuits, descending pathways are critical for ensuring
  5574. appropriate anticipatory modifications of gait to accommodate
  5575. uneven terrain. Neurons in the motor cortex command the changes
  5576. in muscle activity required to modify limb trajectory when
  5577. stepping over obstacles. Simultaneously, neurons in the
  5578. brainstem reticular formation ensure that these modifications
  5579. are superimposed on an appropriate base of postural support.
  5580. Recent experiments suggest that the same neurons in the same
  5581. structures also provide similar information during reaching
  5582. movements. It is suggested that, during both locomotion and
  5583. reaching movements, the final expression of descending signals
  5584. is influenced by the state and excitability of the spinal
  5585. circuits upon which they impinge.",
  5586. journal = "Prog. Brain Res.",
  5587. publisher = "Elsevier",
  5588. volume = 143,
  5589. pages = "251--261",
  5590. year = 2004,
  5591. keywords = "Locomotion",
  5592. language = "en"
  5593. }
  5594. @ARTICLE{Warren2019-lv,
  5595. title = "{Non-Euclidean} navigation",
  5596. author = "Warren, William H",
  5597. abstract = "A basic set of navigation strategies supports navigational tasks
  5598. ranging from homing to novel detours and shortcuts. To perform
  5599. these last two tasks, it is generally thought that humans,
  5600. mammals and perhaps some insects possess Euclidean cognitive
  5601. maps, constructed on the basis of input from the path integration
  5602. system. In this article, I review the rationale and behavioral
  5603. evidence for this metric cognitive map hypothesis, and find it
  5604. unpersuasive: in practice, there is little evidence for truly
  5605. novel shortcuts in animals, and human performance is highly
  5606. unreliable and biased by environmental features. I develop the
  5607. alternative hypothesis that spatial knowledge is better
  5608. characterized as a labeled graph: a network of paths between
  5609. places augmented with local metric information. What
  5610. distinguishes such a cognitive graph from a metric cognitive map
  5611. is that this local information is not embedded in a global
  5612. coordinate system, so spatial knowledge is often geometrically
  5613. inconsistent. Human path integration appears to be better suited
  5614. to piecewise measurements of path lengths and turn angles than to
  5615. building a consistent map. In a series of experiments in
  5616. immersive virtual reality, we tested human navigation in
  5617. non-Euclidean environments and found that shortcuts manifest
  5618. large violations of the metric postulates. The results are
  5619. contrary to the Euclidean map hypothesis and support the
  5620. cognitive graph hypothesis. Apparently Euclidean behavior, such
  5621. as taking novel detours and approximate shortcuts, can be
  5622. explained by the adaptive use of non-Euclidean strategies.",
  5623. journal = "J. Exp. Biol.",
  5624. volume = 222,
  5625. number = "Pt Suppl 1",
  5626. month = feb,
  5627. year = 2019,
  5628. keywords = "Cognitive graph; Cognitive map; Path integration; Spatial
  5629. cognition; Wayfinding",
  5630. language = "en"
  5631. }
  5632. @ARTICLE{Ericson2020-rz,
  5633. title = "Probing the invariant structure of spatial knowledge: Support for
  5634. the cognitive graph hypothesis",
  5635. author = "Ericson, Jonathan D and Warren, William H",
  5636. abstract = "We tested four hypotheses about the structure of spatial
  5637. knowledge used for navigation: (1) the Euclidean hypothesis, a
  5638. geometrically consistent map; (2) the Neighborhood hypothesis,
  5639. adjacency relations between spatial regions, based on visible
  5640. boundaries; (3) the Cognitive Graph hypothesis, a network of
  5641. paths between places, labeled with approximate local distances
  5642. and angles; and (4) the Constancy hypothesis, whatever geometric
  5643. properties are invariant during learning. In two experiments,
  5644. different groups of participants learned three virtual hedge
  5645. mazes, which varied specific geometric properties (Euclidean
  5646. Control Maze, Elastic Maze with stretching paths, Swap Maze with
  5647. alternating paths to the same place). Spatial knowledge was then
  5648. tested using three navigation tasks (metric shortcuts on empty
  5649. ground plane, neighborhood shortcuts with visible boundaries,
  5650. route task in corridors). They yielded the following results: (a)
  5651. Metric shortcuts were insensitive to detectable shifts in target
  5652. location, inconsistent with the Euclidean hypothesis. (b)
  5653. Neighborhood shortcuts were constrained by visible boundaries in
  5654. the Elastic Maze, but not in the Swap Maze, contrary to the
  5655. Neighborhood and Constancy hypotheses. (c) The route task
  5656. indicated that a graph of the maze was acquired in all
  5657. environments, including knowledge of local path lengths. We
  5658. conclude that primary spatial knowledge is consistent with the
  5659. Cognitive Graph hypothesis. Neighborhoods are derived from the
  5660. graph, and local distance and angle information is not embedded
  5661. in a geometrically consistent map.",
  5662. journal = "Cognition",
  5663. volume = 200,
  5664. pages = "104276",
  5665. month = may,
  5666. year = 2020,
  5667. keywords = "Cognitive graph; Cognitive map; Human navigation; Spatial
  5668. cognition",
  5669. language = "en"
  5670. }
  5671. @UNPUBLISHED{Duan2019-cb,
  5672. title = "A cortico-collicular pathway for motor planning in a
  5673. memory-dependent perceptual decision task",
  5674. author = "Duan, Chunyu A and Pan, Yuxin and Ma, Guofen and Zhou, Taotao and
  5675. Zhang, Siyu and Xu, Ning-Long",
  5676. abstract = "ABSTRACT Survival in a dynamic environment requires animals to
  5677. plan future actions based on past sensory evidence. However, the
  5678. neural circuit mechanism underlying this crucial brain function,
  5679. referred to as motor planning, remains unclear. Here, we employ
  5680. projection-specific imaging and perturbation methods to
  5681. investigate the direct pathway linking two key nodes in the motor
  5682. planning network, the secondary motor cortex (M2) and the
  5683. midbrain superior colliculus (SC), in mice performing a
  5684. memory-dependent perceptual decision task. We find dynamic coding
  5685. of choice information in SC-projecting M2 neurons during motor
  5686. planning and execution, and disruption of this information by
  5687. inhibiting M2 terminals in SC selectively impaired decision
  5688. maintenance. Furthermore, cell-type-specific optogenetic circuit
  5689. mapping shows that M2 terminals modulate both excitatory and
  5690. inhibitory SC neurons with balanced synaptic strength. Together,
  5691. our results reveal the dynamic recruitment of the
  5692. premotor-collicular pathway as a circuit mechanism for motor
  5693. planning.",
  5694. journal = "bioRxiv",
  5695. pages = "709170",
  5696. month = jul,
  5697. year = 2019,
  5698. language = "en"
  5699. }
  5700. @UNPUBLISHED{Adam2020-zd,
  5701. title = "Cortico-subthalamic projections send brief stop signals to halt
  5702. visually-guided locomotion",
  5703. author = "Adam, Elie M and Johns, Taylor and Sur, Mriganka",
  5704. abstract = "Summary Goal-directed locomotion necessitates control signals
  5705. that propagate from higher-order areas to regulate spinal
  5706. mechanisms. The cortico-subthalamic hyperdirect pathway offers a
  5707. short route for cortical information to reach locomotor centers
  5708. in the brainstem. We developed a task where head-fixed mice run
  5709. to a visual landmark, then stop and wait to collect reward, and
  5710. examined the role of secondary motor cortex (M2) projections to
  5711. the subthalamic nucleus (STN) in controlling locomotion. Our
  5712. modeled behavioral strategy indicates a switching point in
  5713. behavior, suggesting a critical neuronal control signal at stop
  5714. locations. Optogenetic activation of M2 axons in STN leads the
  5715. animal to stop prematurely. By imaging M2 neurons projecting to
  5716. STN, we find neurons that are active at the onset of stops, when
  5717. executed at the landmark but not spontaneously elsewhere. Our
  5718. results suggest that the M2-STN pathway can be recruited during
  5719. visually-guided locomotion to rapidly and precisely control the
  5720. mesencephalic locomotor region through the basal ganglia.",
  5721. journal = "bioRxiv",
  5722. pages = "2020.02.05.936443",
  5723. month = feb,
  5724. year = 2020,
  5725. language = "en"
  5726. }
  5727. @UNPUBLISHED{Storchi2020-em,
  5728. title = "Beyond locomotion: in the mouse the mapping between sensations
  5729. and behaviours unfolds in a higher dimensional space",
  5730. author = "Storchi, Riccardo and Milosavljevic, Nina and Allen, Annette E
  5731. and Cootes, Timothy F and Lucas, Robert J",
  5732. abstract = "Abstract The ability of specific sensory stimuli to evoke
  5733. spontaneous behavioural responses in the mouse represents a
  5734. powerful approach to study how the mammalian brain processes
  5735. sensory information and selects appropriate motor actions. For
  5736. visually and auditory guided behaviours the relevant action has
  5737. been empirically identified as a change in locomotion state.
  5738. However, the extent to which locomotion alone captures the
  5739. diversity of those behaviours and their sensory specificity is
  5740. unknown.To tackle this problem we developed a method to obtain a
  5741. faithful 3D reconstruction of the mouse body that enabled us to
  5742. quantify a wide variety of movements and changes in postures.
  5743. This higher dimensional description of behaviour revealed that
  5744. responses to different sensory inputs is more stimulus-specific
  5745. than indicated by locomotion data alone. Thus, equivalent
  5746. locomotion patterns evoked by different stimuli (e.g. looming and
  5747. sound evoking locomotion arrest) could be well separated along
  5748. other dimensions. The enhanced stimulus-specificity was explained
  5749. by a surprising diversity of behavioural responses. A clustering
  5750. analysis revealed that distinct combinations of motor actions and
  5751. postures, giving rise to at least 7 different behaviours, were
  5752. required to account for stimulus-specificity. Moreover, each
  5753. stimulus evoked more than one behaviour revealing a robust
  5754. one-to-many mapping between sensations and behaviours that could
  5755. not be detected from locomotion data.Our results challenge the
  5756. current view of visually and auditory guided behaviours as purely
  5757. locomotion-based actions (e.g. freeze, escape) and indicate that
  5758. behavioural diversity and sensory specificity unfold in a higher
  5759. dimensional space spanning multiple motor actions.",
  5760. journal = "bioRxiv",
  5761. pages = "2020.02.24.961565",
  5762. month = mar,
  5763. year = 2020,
  5764. keywords = "Locomotion",
  5765. language = "en"
  5766. }
  5767. @ARTICLE{Cregg2020-ie,
  5768. title = "Brainstem neurons that command mammalian locomotor asymmetries",
  5769. author = "Cregg, Jared M and Leiras, Roberto and Montalant, Alexia and
  5770. Wanken, Paulina and Wickersham, Ian R and Kiehn, Ole",
  5771. abstract = "Descending command neurons instruct spinal networks to execute
  5772. basic locomotor functions, such as gait and speed. The command
  5773. functions for gait and speed are symmetric, implying that a
  5774. separate unknown system directs asymmetric movements, including
  5775. the ability to move left or right. In the present study, we
  5776. report that Chx10-lineage reticulospinal neurons act to control
  5777. the direction of locomotor movements in mammals. Chx10 neurons
  5778. exhibit mainly ipsilateral projection, and their selective
  5779. unilateral activation causes ipsilateral turning movements in
  5780. freely moving mice. Unilateral inhibition of Chx10 neurons causes
  5781. contralateral turning movements. Paired left--right motor
  5782. recordings identified distinct mechanisms for directional
  5783. movements mediated via limb and axial spinal circuits. Finally,
  5784. we identify sensorimotor brain regions that project on to Chx10
  5785. reticulospinal neurons, and demonstrate that their unilateral
  5786. activation can impart left--right directional commands. Together
  5787. these data identify the descending motor system that commands
  5788. left--right locomotor asymmetries in mammals.",
  5789. journal = "Nat. Neurosci.",
  5790. month = may,
  5791. year = 2020
  5792. }
  5793. @ARTICLE{Wang2020-lz,
  5794. title = "The Allen Mouse Brain Common Coordinate Framework: A {3D}
  5795. Reference Atlas",
  5796. author = "Wang, Quanxin and Ding, Song-Lin and Li, Yang and Royall, Josh
  5797. and Feng, David and Lesnar, Phil and Graddis, Nile and Naeemi,
  5798. Maitham and Facer, Benjamin and Ho, Anh and Dolbeare, Tim and
  5799. Blanchard, Brandon and Dee, Nick and Wakeman, Wayne and
  5800. Hirokawa, Karla E and Szafer, Aaron and Sunkin, Susan M and Oh,
  5801. Seung Wook and Bernard, Amy and Phillips, John W and Hawrylycz,
  5802. Michael and Koch, Christof and Zeng, Hongkui and Harris, Julie A
  5803. and Ng, Lydia",
  5804. abstract = "SummaryRecent large-scale collaborations are generating major
  5805. surveys of cell types and connections in the mouse brain,
  5806. collecting large amounts of data across modalities, spatial
  5807. scales, and brain areas. Successful integration of these data
  5808. requires a standard 3D reference atlas. Here, we present the
  5809. Allen Mouse Brain Common Coordinate Framework (CCFv3) as such a
  5810. resource. We constructed an average template brain at 10 $\mu$m
  5811. voxel resolution by interpolating high resolution in-plane
  5812. serial two-photon tomography images with 100 $\mu$m z-sampling
  5813. from 1,675 young adult C57BL/6J mice. Then, using multimodal
  5814. reference data, we parcellated the entire brain directly in 3D,
  5815. labeling every voxel with a brain structure spanning 43
  5816. isocortical areas and their layers, 329 subcortical gray matter
  5817. structures, 81 fiber tracts, and 8 ventricular structures. CCFv3
  5818. can be used to analyze, visualize, and integrate multimodal and
  5819. multiscale datasets in 3D and is openly accessible
  5820. (https://atlas.brain-map.org/).",
  5821. journal = "Cell",
  5822. publisher = "Elsevier",
  5823. volume = 0,
  5824. number = 0,
  5825. month = may,
  5826. year = 2020,
  5827. keywords = "average mouse brain; reference atlas; 3D brain atlas; brain
  5828. parcellation; brain anatomy; mouse cortex; common coordinate
  5829. framework; CCFv3; fiber tracts; transgenic mice",
  5830. language = "en"
  5831. }
  5832. @ARTICLE{Taube2007-if,
  5833. title = "The head direction signal: origins and sensory-motor integration",
  5834. author = "Taube, Jeffrey S",
  5835. abstract = "Navigation first requires accurate perception of one's spatial
  5836. orientation within the environment, which consists of knowledge
  5837. about location and directional heading. Cells within several
  5838. limbic system areas of the mammalian brain discharge
  5839. allocentrically as a function of the animal's directional
  5840. heading, independent of the animal's location and ongoing
  5841. behavior. These cells are referred to as head direction (HD)
  5842. cells and are believed to encode the animal's perceived
  5843. directional heading with respect to its environment. Although HD
  5844. cells are found in several areas, the principal circuit for
  5845. generating this signal originates in the dorsal tegmental
  5846. nucleus and projects serially, with some reciprocal connections,
  5847. to the lateral mammillary nucleus --> anterodorsal thalamus -->
  5848. PoS, and terminates in the entorhinal cortex. HD cells receive
  5849. multimodal information about landmarks and self-generated
  5850. movements. Vestibular information appears critical for
  5851. generating the directional signal, but motor/proprioceptive and
  5852. landmark information are important for updating it.",
  5853. journal = "Annu. Rev. Neurosci.",
  5854. publisher = "annualreviews.org",
  5855. volume = 30,
  5856. pages = "181--207",
  5857. year = 2007,
  5858. language = "en"
  5859. }
  5860. @ARTICLE{Crane_undated-dr,
  5861. title = "{D} {ISCRETE} {D} {IFFERENTIAL} {G} {EOMETRY}: A {N} A {PPLIED} {I}
  5862. {NTRODUCTION}",
  5863. author = "Crane, Keenan"
  5864. }
  5865. @ARTICLE{lx_undated-hq,
  5866. title = "{'$HWK\ODWLRQ$WODVRIWKHRXVH\%UDLQDWLQJOH\&HOOHVROXWLRQ}",
  5867. author = "/lx, +dqtlqj and =krx, -Lqjwldq and /xr, Krqj\textbackslashxdq and
  5868. \%duwohww, \$qqd and \$ogulgjh, \$qguhz"
  5869. }
  5870. @INCOLLECTION{Sorscher2019-xx,
  5871. title = "A unified theory for the origin of grid cells through the lens
  5872. of pattern formation",
  5873. booktitle = "Advances in Neural Information Processing Systems 32",
  5874. author = "Sorscher, Ben and Mel, Gabriel and Ganguli, Surya and Ocko,
  5875. Samuel",
  5876. editor = "Wallach, H and Larochelle, H and Beygelzimer, A and
  5877. d\textbackslashtextquotesingle Alch{\'e}-Buc, F and Fox, E and
  5878. Garnett, R",
  5879. publisher = "Curran Associates, Inc.",
  5880. pages = "10003--10013",
  5881. year = 2019
  5882. }
  5883. @ARTICLE{Pnevmatikakis2017-qi,
  5884. title = "{NoRMCorre}: An online algorithm for piecewise rigid motion
  5885. correction of calcium imaging data",
  5886. author = "Pnevmatikakis, Eftychios A and Giovannucci, Andrea",
  5887. abstract = "BACKGROUND: Motion correction is a challenging pre-processing
  5888. problem that arises early in the analysis pipeline of calcium
  5889. imaging data sequences. The motion artifacts in two-photon
  5890. microscopy recordings can be non-rigid, arising from the finite
  5891. time of raster scanning and non-uniform deformations of the brain
  5892. medium. NEW METHOD: We introduce an algorithm for fast Non-Rigid
  5893. Motion Correction (NoRMCorre) based on template matching.
  5894. NoRMCorre operates by splitting the field of view (FOV) into
  5895. overlapping spatial patches along all directions. The patches are
  5896. registered at a sub-pixel resolution for rigid translation
  5897. against a regularly updated template. The estimated alignments
  5898. are subsequently up-sampled to create a smooth motion field for
  5899. each frame that can efficiently approximate non-rigid artifacts
  5900. in a piecewise-rigid manner. EXISTING METHODS: Existing
  5901. approaches either do not scale well in terms of computational
  5902. performance or are targeted to non-rigid artifacts arising just
  5903. from the finite speed of raster scanning, and thus cannot correct
  5904. for non-rigid motion observable in datasets from a large FOV.
  5905. RESULTS: NoRMCorre can be run in an online mode resulting in
  5906. comparable to or even faster than real time motion registration
  5907. of streaming data. We evaluate its performance with simple yet
  5908. intuitive metrics and compare against other non-rigid
  5909. registration methods on simulated data and in vivo two-photon
  5910. calcium imaging datasets. Open source Matlab and Python code is
  5911. also made available. CONCLUSIONS: The proposed method and
  5912. accompanying code can be useful for solving large scale image
  5913. registration problems in calcium imaging, especially in the
  5914. presence of non-rigid deformations.",
  5915. journal = "J. Neurosci. Methods",
  5916. volume = 291,
  5917. pages = "83--94",
  5918. month = nov,
  5919. year = 2017,
  5920. keywords = "Calcium imaging; Image registration; Motion correction",
  5921. language = "en"
  5922. }
  5923. @UNPUBLISHED{Tran2020-yg,
  5924. title = "Automated curation of {CNMF-E-extracted} {ROI} spatial footprints
  5925. and calcium traces using open-source {AutoML} tools",
  5926. author = "Tran, L M and Mocle, A J and Ramsaran, A I and Jacob, A D and
  5927. Frankland, P W and Josselyn, S A",
  5928. abstract = "In vivo 1-photon calcium imaging is an increasingly prevalent
  5929. method in behavioural neuroscience. Numerous analysis pipelines
  5930. have been developed to improve the reliability and scalability of
  5931. pre-processing and ROI extraction for these large calcium imaging
  5932. datasets. Despite these advancements in pre-processing methods,
  5933. manual curation of the extracted spatial footprints and calcium
  5934. traces of neurons remains important for quality control. Here, we
  5935. propose an additional semi-automated curation step for sorting
  5936. spatial footprints and calcium traces from putative neurons
  5937. extracted using the popular CNMF-E algorithm. We used the
  5938. automated machine learning tools TPOT and AutoSklearn to generate
  5939. classifiers to curate the extracted ROIs trained on a subset of
  5940. human-labeled data. AutoSklearn produced the best performing
  5941. classifier, achieving an F1 score > 92\% on the ground truth test
  5942. dataset. This automated approach is a useful strategy for
  5943. filtering ROIs with relatively few labeled data points, and can
  5944. be easily added to pre-existing pipelines currently using CNMF-E
  5945. for ROI extraction.",
  5946. journal = "bioRxiv",
  5947. pages = "2020.03.13.991216",
  5948. month = mar,
  5949. year = 2020,
  5950. language = "en"
  5951. }
  5952. @ARTICLE{Zhou2018-fo,
  5953. title = "Efficient and accurate extraction of in vivo calcium signals from
  5954. microendoscopic video data",
  5955. author = "Zhou, Pengcheng and Resendez, Shanna L and Rodriguez-Romaguera,
  5956. Jose and Jimenez, Jessica C and Neufeld, Shay Q and Giovannucci,
  5957. Andrea and Friedrich, Johannes and Pnevmatikakis, Eftychios A and
  5958. Stuber, Garret D and Hen, Rene and Kheirbek, Mazen A and
  5959. Sabatini, Bernardo L and Kass, Robert E and Paninski, Liam",
  5960. abstract = "In vivo calcium imaging through microendoscopic lenses enables
  5961. imaging of previously inaccessible neuronal populations deep
  5962. within the brains of freely moving animals. However, it is
  5963. computationally challenging to extract single-neuronal activity
  5964. from microendoscopic data, because of the very large background
  5965. fluctuations and high spatial overlaps intrinsic to this
  5966. recording modality. Here, we describe a new constrained matrix
  5967. factorization approach to accurately separate the background and
  5968. then demix and denoise the neuronal signals of interest. We
  5969. compared the proposed method against previous independent
  5970. components analysis and constrained nonnegative matrix
  5971. factorization approaches. On both simulated and experimental data
  5972. recorded from mice, our method substantially improved the quality
  5973. of extracted cellular signals and detected more well-isolated
  5974. neural signals, especially in noisy data regimes. These advances
  5975. can in turn significantly enhance the statistical power of
  5976. downstream analyses, and ultimately improve scientific
  5977. conclusions derived from microendoscopic data.",
  5978. journal = "Elife",
  5979. volume = 7,
  5980. month = feb,
  5981. year = 2018,
  5982. keywords = "calcium imaging; microendoscope; mouse; neuroscience; source
  5983. extraction",
  5984. language = "en"
  5985. }
  5986. @ARTICLE{Weir_undated-oc,
  5987. title = "A molecular filter for the cnidarian stinging response",
  5988. author = "Weir, Keiko and Dupre, Christophe and van Giesen, Lena and Lee, Amy
  5989. S Y and Bellono, Nicholas W"
  5990. }
  5991. @ARTICLE{Dolensek2020-dn,
  5992. title = "Facial expressions of emotion states and their neuronal
  5993. correlates in mice",
  5994. author = "Dolensek, Nejc and Gehrlach, Daniel A and Klein, Alexandra S and
  5995. Gogolla, Nadine",
  5996. abstract = "Understanding the neurobiological underpinnings of emotion relies
  5997. on objective readouts of the emotional state of an individual,
  5998. which remains a major challenge especially in animal models. We
  5999. found that mice exhibit stereotyped facial expressions in
  6000. response to emotionally salient events, as well as upon targeted
  6001. manipulations in emotion-relevant neuronal circuits. Facial
  6002. expressions were classified into distinct categories using
  6003. machine learning and reflected the changing intrinsic value of
  6004. the same sensory stimulus encountered under different homeostatic
  6005. or affective conditions. Facial expressions revealed emotion
  6006. features such as intensity, valence, and persistence. Two-photon
  6007. imaging uncovered insular cortical neuron activity that
  6008. correlated with specific facial expressions and may encode
  6009. distinct emotions. Facial expressions thus provide a means to
  6010. infer emotion states and their neuronal correlates in mice.",
  6011. journal = "Science",
  6012. volume = 368,
  6013. number = 6486,
  6014. pages = "89--94",
  6015. month = apr,
  6016. year = 2020,
  6017. language = "en"
  6018. }
  6019. @UNPUBLISHED{Rayshubskiy2020-ad,
  6020. title = "Neural control of steering in walking Drosophila",
  6021. author = "Rayshubskiy, Aleksandr and Holtz, Stephen L and D'Alessandro,
  6022. Isabel and Li, Anna A and Vanderbeck, Quinn X and Haber, Isabel S
  6023. and Gibb, Peter W and Wilson, Rachel I",
  6024. abstract = "During navigation, the brain must continuously integrate external
  6025. guidance cues with internal spatial maps to update steering
  6026. commands. However, it has been difficult to link spatial maps
  6027. with motor control. Here we identify 9descending steering9
  6028. neurons in the Drosophila brain that lie two synapses downstream
  6029. from the brain9s heading direction map in the central complex.
  6030. These steering neurons predict behavioral turns caused by
  6031. microstimulation of the spatial map. Moreover, these neurons
  6032. receive 9direct9 sensory input that bypasses the central complex,
  6033. and they predict steering evoked by multimodal stimuli.
  6034. Unilateral activation of these neurons can promote turning, while
  6035. bilateral silencing interferes with body and leg movements. In
  6036. short, these neurons combine internal maps with external cues to
  6037. predict and influence steering. They represent a key link between
  6038. cognitive maps, which use an abstract coordinate frame, and motor
  6039. commands, which use a body-centric coordinate frame.",
  6040. journal = "bioRxiv",
  6041. pages = "2020.04.04.024703",
  6042. month = apr,
  6043. year = 2020,
  6044. language = "en"
  6045. }
  6046. @ARTICLE{Branco_undated-ez,
  6047. title = "The Neural Basis of Escape Behavior in Vertebrates",
  6048. author = "Branco, Tiago and Redgrave, Peter"
  6049. }
  6050. % The entry below contains non-ASCII chars that could not be converted
  6051. % to a LaTeX equivalent.
  6052. @ARTICLE{Hsu2019-jc,
  6053. title = "{B-SOiD}: An Open Source Unsupervised Algorithm for Discovery of
  6054. Spontaneous Behaviors",
  6055. author = "Hsu, A I and Yttri, E A",
  6056. abstract = "The motivation, control, and selection of actions comprising
  6057. naturalistic behaviors remains a tantalizing but difficult field
  6058. of study. Detailed and unbiased quantification is critical.
  6059. Interpreting the positions of animals and their limbs can be
  6060. useful in studying behavior, and significant recent advances
  6061. have made this step straightforward. However, body position
  6062. alone does not provide a grasp of the dynamic range of
  6063. naturalistic behaviors. Behavioral Segmentation of Open-field In
  6064. DeepLabCut, or B-SOiD (`` B-side''), is an unsupervised …",
  6065. journal = "bioRxiv",
  6066. publisher = "biorxiv.org",
  6067. year = 2019
  6068. }
  6069. @ARTICLE{Andalman2019-zi,
  6070. title = "Neuronal Dynamics Regulating Brain and Behavioral State
  6071. Transitions",
  6072. author = "Andalman, Aaron S and Burns, Vanessa M and Lovett-Barron,
  6073. Matthew and Broxton, Michael and Poole, Ben and Yang, Samuel J
  6074. and Grosenick, Logan and Lerner, Talia N and Chen, Ritchie and
  6075. Benster, Tyler and Mourrain, Philippe and Levoy, Marc and Rajan,
  6076. Kanaka and Deisseroth, Karl",
  6077. abstract = "Prolonged behavioral challenges can cause animals to switch from
  6078. active to passive coping strategies to manage effort-expenditure
  6079. during stress; such normally adaptive behavioral state
  6080. transitions can become maladaptive in psychiatric disorders such
  6081. as depression. The underlying neuronal dynamics and brainwide
  6082. interactions important for passive coping have remained unclear.
  6083. Here, we develop a paradigm to study these behavioral state
  6084. transitions at cellular-resolution across the entire vertebrate
  6085. brain. Using brainwide imaging in zebrafish, we observed that
  6086. the transition to passive coping is manifested by progressive
  6087. activation of neurons in the ventral (lateral) habenula.
  6088. Activation of these ventral-habenula neurons suppressed
  6089. downstream neurons in the serotonergic raphe nucleus and caused
  6090. behavioral passivity, whereas inhibition of these neurons
  6091. prevented passivity. Data-driven recurrent neural network
  6092. modeling pointed to altered intra-habenula interactions as a
  6093. contributory mechanism. These results demonstrate ongoing
  6094. encoding of experience features in the habenula, which guides
  6095. recruitment of downstream networks and imposes a passive coping
  6096. behavioral strategy.",
  6097. journal = "Cell",
  6098. publisher = "Elsevier",
  6099. volume = 177,
  6100. number = 4,
  6101. pages = "970--985.e20",
  6102. month = may,
  6103. year = 2019,
  6104. language = "en"
  6105. }
  6106. @ARTICLE{Tombaz2020-tl,
  6107. title = "Action representation in the mouse parieto-frontal network",
  6108. author = "Tombaz, Tuce and Dunn, Benjamin A and Hovde, Karoline and Cubero,
  6109. Ryan John and Mimica, Bartul and Mamidanna, Pranav and Roudi,
  6110. Yasser and Whitlock, Jonathan R",
  6111. abstract = "The posterior parietal cortex (PPC) and frontal motor areas
  6112. comprise a cortical network supporting goal-directed behaviour,
  6113. with functions including sensorimotor transformations and
  6114. decision making. In primates, this network links performed and
  6115. observed actions via mirror neurons, which fire both when
  6116. individuals perform an action and when they observe the same
  6117. action performed by a conspecific. Mirror neurons are believed to
  6118. be important for social learning, but it is not known whether
  6119. mirror-like neurons occur in similar networks in other social
  6120. species, such as rodents, or if they can be measured in such
  6121. models using paradigms where observers passively view a
  6122. demonstrator. Therefore, we imaged Ca2+ responses in PPC and
  6123. secondary motor cortex (M2) while mice performed and observed
  6124. pellet-reaching and wheel-running tasks, and found that cell
  6125. populations in both areas robustly encoded several naturalistic
  6126. behaviours. However, neural responses to the same set of observed
  6127. actions were absent, although we verified that observer mice were
  6128. attentive to performers and that PPC neurons responded reliably
  6129. to visual cues. Statistical modelling also indicated that
  6130. executed actions outperformed observed actions in predicting
  6131. neural responses. These results raise the possibility that
  6132. sensorimotor action recognition in rodents could take place
  6133. outside of the parieto-frontal circuit, and underscore that
  6134. detecting socially-driven neural coding depends critically on the
  6135. species and behavioural paradigm used.",
  6136. journal = "Sci. Rep.",
  6137. volume = 10,
  6138. number = 1,
  6139. pages = "5559",
  6140. month = mar,
  6141. year = 2020,
  6142. language = "en"
  6143. }
  6144. @UNPUBLISHED{Benavidez2020-oh,
  6145. title = "The mouse cortico-tectal projectome",
  6146. author = "Benavidez, Nora L and Bienkowski, Michael S and Khanjani, Neda
  6147. and Bowman, Ian and Fayzullina, Marina and Garcia, Luis and Gao,
  6148. Lei and Korobkova, Laura and Gou, Lin and Cotter, Kaelan and
  6149. Becerra, Marlene and Aquino, Sarvia and Cao, Chunru and Foster,
  6150. Nicholas N and Song, Monica Y and Zhang, Bin and Yamashita, Seita
  6151. and Zhu, Muye and Lo, Darrick and Boesen, Tyler and Zingg, Brian
  6152. and Santarelli, Anthony and Wickersham, Ian R and Ascoli, Giorgio
  6153. A and Hintiryan, Houri and Dong, Hong-Wei",
  6154. abstract = "The superior colliculus (SC) is a midbrain structure that
  6155. receives diverse and robust cortical inputs to drive a range of
  6156. cognitive and sensorimotor behaviors. However, it remains unclear
  6157. how descending cortical inputs arising from higher-order
  6158. associative areas coordinate with SC sensorimotor networks to
  6159. influence its outputs. In this study, we constructed a
  6160. comprehensive map of all cortico-tectal projections and
  6161. identified four collicular zones with differential cortical
  6162. inputs: medial (SC.m), centromedial (SC.cm), centrolateral
  6163. (SC.cl) and lateral (SC.l). Computational analyses revealed that
  6164. cortico-tectal projections are organized as multiple subnetworks
  6165. that are consistent with previously identified cortico-cortical
  6166. and cortico-striatal subnetworks. Furthermore, we delineated the
  6167. brain-wide input/output organization of each collicular zone and
  6168. described a subset of their constituent neuronal cell types based
  6169. on distinct connectional and morphological features. Altogether,
  6170. this work provides a novel structural foundation for the
  6171. integrative role of the SC in controlling cognition, orientation,
  6172. and other sensorimotor behaviors.",
  6173. journal = "bioRxiv",
  6174. pages = "2020.03.24.006775",
  6175. month = mar,
  6176. year = 2020,
  6177. language = "en"
  6178. }
  6179. @INPROCEEDINGS{Pascanu2013-zg,
  6180. title = "On the difficulty of training recurrent neural networks",
  6181. booktitle = "International Conference on Machine Learning",
  6182. author = "Pascanu, Razvan and Mikolov, Tomas and Bengio, Yoshua",
  6183. abstract = "There are two widely known issues with properly training
  6184. recurrent neural networks, the vanishing and the exploding
  6185. gradient problems detailed in Bengio et al. (1994). In this
  6186. paper we attempt to i...",
  6187. publisher = "jmlr.org",
  6188. pages = "1310--1318",
  6189. month = feb,
  6190. year = 2013,
  6191. language = "en",
  6192. conference = "International Conference on Machine Learning"
  6193. }
  6194. @ARTICLE{Remington2018-aw,
  6195. title = "A Dynamical Systems Perspective on Flexible Motor Timing",
  6196. author = "Remington, Evan D and Egger, Seth W and Narain, Devika and Wang,
  6197. Jing and Jazayeri, Mehrdad",
  6198. abstract = "A hallmark of higher brain function is the ability to rapidly and
  6199. flexibly adjust behavioral responses based on internal and
  6200. external cues. Here, we examine the computational principles that
  6201. allow decisions and actions to unfold flexibly in time. We adopt
  6202. a dynamical systems perspective and outline how temporal
  6203. flexibility in such a system can be achieved through
  6204. manipulations of inputs and initial conditions. We then review
  6205. evidence from experiments in nonhuman primates that support this
  6206. interpretation. Finally, we explore the broader utility and
  6207. limitations of the dynamical systems perspective as a general
  6208. framework for addressing open questions related to the temporal
  6209. control of movements, as well as in the domains of learning and
  6210. sequence generation.",
  6211. journal = "Trends Cogn. Sci.",
  6212. volume = 22,
  6213. number = 10,
  6214. pages = "938--952",
  6215. month = oct,
  6216. year = 2018,
  6217. keywords = "dynamical systems; flexible timing; learning; movement planning;
  6218. movement sequences; sensorimotor control",
  6219. language = "en"
  6220. }
  6221. @ARTICLE{Remington2018-tc,
  6222. title = "Flexible Sensorimotor Computations through Rapid Reconfiguration
  6223. of Cortical Dynamics",
  6224. author = "Remington, Evan D and Narain, Devika and Hosseini, Eghbal A and
  6225. Jazayeri, Mehrdad",
  6226. abstract = "Neural mechanisms that support flexible sensorimotor computations
  6227. are not well understood. In a dynamical system whose state is
  6228. determined by interactions among neurons, computations can be
  6229. rapidly reconfigured by controlling the system's inputs and
  6230. initial conditions. To investigate whether the brain employs such
  6231. control mechanisms, we recorded from the dorsomedial frontal
  6232. cortex of monkeys trained to measure and produce time intervals
  6233. in two sensorimotor contexts. The geometry of neural trajectories
  6234. during the production epoch was consistent with a mechanism
  6235. wherein the measured interval and sensorimotor context exerted
  6236. control over cortical dynamics by adjusting the system's initial
  6237. condition and input, respectively. These adjustments, in turn,
  6238. set the speed at which activity evolved in the production epoch,
  6239. allowing the animal to flexibly produce different time intervals.
  6240. These results provide evidence that the language of dynamical
  6241. systems can be used to parsimoniously link brain activity to
  6242. sensorimotor computations.",
  6243. journal = "Neuron",
  6244. volume = 98,
  6245. number = 5,
  6246. pages = "1005--1019.e5",
  6247. month = jun,
  6248. year = 2018,
  6249. keywords = "Dynamical Systems; cognitive flexibility; electrophysiology;
  6250. frontal cortex; motor planning; population coding; recurrent
  6251. neural networks; sensorimotor coordination; timing",
  6252. language = "en"
  6253. }
  6254. @ARTICLE{Wang2018-gi,
  6255. title = "Flexible timing by temporal scaling of cortical responses",
  6256. author = "Wang, Jing and Narain, Devika and Hosseini, Eghbal A and
  6257. Jazayeri, Mehrdad",
  6258. abstract = "Musicians can perform at different tempos, speakers can control
  6259. the cadence of their speech, and children can flexibly vary their
  6260. temporal expectations of events. To understand the neural basis
  6261. of such flexibility, we recorded from the medial frontal cortex
  6262. of nonhuman primates trained to produce different time intervals
  6263. with different effectors. Neural responses were heterogeneous,
  6264. nonlinear, and complex, and they exhibited a remarkable form of
  6265. temporal invariance: firing rate profiles were temporally scaled
  6266. to match the produced intervals. Recording from downstream
  6267. neurons in the caudate and from thalamic neurons projecting to
  6268. the medial frontal cortex indicated that this phenomenon
  6269. originates within cortical networks. Recurrent neural network
  6270. models trained to perform the task revealed that temporal scaling
  6271. emerges from nonlinearities in the network and that the degree of
  6272. scaling is controlled by the strength of external input. These
  6273. findings demonstrate a simple and general mechanism for
  6274. conferring temporal flexibility upon sensorimotor and cognitive
  6275. functions.",
  6276. journal = "Nat. Neurosci.",
  6277. volume = 21,
  6278. number = 1,
  6279. pages = "102--110",
  6280. month = jan,
  6281. year = 2018,
  6282. keywords = "RNN To read;RNN",
  6283. language = "en"
  6284. }
  6285. @ARTICLE{Song2016-cr,
  6286. title = "Training {Excitatory-Inhibitory} Recurrent Neural Networks for
  6287. Cognitive Tasks: A Simple and Flexible Framework",
  6288. author = "Song, H Francis and Yang, Guangyu R and Wang, Xiao-Jing",
  6289. abstract = "The ability to simultaneously record from large numbers of
  6290. neurons in behaving animals has ushered in a new era for the
  6291. study of the neural circuit mechanisms underlying cognitive
  6292. functions. One promising approach to uncovering the dynamical and
  6293. computational principles governing population responses is to
  6294. analyze model recurrent neural networks (RNNs) that have been
  6295. optimized to perform the same tasks as behaving animals. Because
  6296. the optimization of network parameters specifies the desired
  6297. output but not the manner in which to achieve this output,
  6298. ``trained'' networks serve as a source of mechanistic hypotheses
  6299. and a testing ground for data analyses that link neural
  6300. computation to behavior. Complete access to the activity and
  6301. connectivity of the circuit, and the ability to manipulate them
  6302. arbitrarily, make trained networks a convenient proxy for
  6303. biological circuits and a valuable platform for theoretical
  6304. investigation. However, existing RNNs lack basic biological
  6305. features such as the distinction between excitatory and
  6306. inhibitory units (Dale's principle), which are essential if RNNs
  6307. are to provide insights into the operation of biological
  6308. circuits. Moreover, trained networks can achieve the same
  6309. behavioral performance but differ substantially in their
  6310. structure and dynamics, highlighting the need for a simple and
  6311. flexible framework for the exploratory training of RNNs. Here, we
  6312. describe a framework for gradient descent-based training of
  6313. excitatory-inhibitory RNNs that can incorporate a variety of
  6314. biological knowledge. We provide an implementation based on the
  6315. machine learning library Theano, whose automatic differentiation
  6316. capabilities facilitate modifications and extensions. We validate
  6317. this framework by applying it to well-known experimental
  6318. paradigms such as perceptual decision-making, context-dependent
  6319. integration, multisensory integration, parametric working memory,
  6320. and motor sequence generation. Our results demonstrate the wide
  6321. range of neural activity patterns and behavior that can be
  6322. modeled, and suggest a unified setting in which diverse cognitive
  6323. computations and mechanisms can be studied.",
  6324. journal = "PLoS Comput. Biol.",
  6325. volume = 12,
  6326. number = 2,
  6327. pages = "e1004792",
  6328. month = feb,
  6329. year = 2016,
  6330. language = "en"
  6331. }
  6332. @ARTICLE{Stubblefield2013-dj,
  6333. title = "Optogenetic investigation of the role of the superior colliculus
  6334. in orienting movements",
  6335. author = "Stubblefield, Elizabeth A and Costabile, Jamie D and Felsen,
  6336. Gidon",
  6337. abstract = "In vivo studies have demonstrated that the superior colliculus
  6338. (SC) integrates sensory information and plays a role in
  6339. controlling orienting motor output. However, how the complex
  6340. microcircuitry within the SC, as documented by slice studies,
  6341. subserves these functions is unclear. Optogenetics affords the
  6342. potential to examine, in behaving animals, the functional roles
  6343. of specific neuron types that comprise heterogeneous nuclei. As
  6344. a first step toward understanding how SC microcircuitry
  6345. underlies motor output, we applied optogenetics to mice
  6346. performing an odor discrimination task in which sensory
  6347. decisions are reported by either a leftward or rightward
  6348. SC-dependent orienting movement. We unilaterally expressed
  6349. either channelrhodopsin-2 or halorhodopsin in the SC and
  6350. delivered light in order to excite or inhibit motor-related SC
  6351. activity as the movement was planned. We found that manipulating
  6352. SC activity predictably affected the direction of the selected
  6353. movement in a manner that depended on the difficulty of the odor
  6354. discrimination. This study demonstrates that the SC plays a
  6355. similar role in directional orienting movements in mice as it
  6356. does in other species, and provides a framework for future
  6357. investigations into how specific SC cell types contribute to
  6358. motor control.",
  6359. journal = "Behav. Brain Res.",
  6360. publisher = "Elsevier",
  6361. volume = 255,
  6362. pages = "55--63",
  6363. month = oct,
  6364. year = 2013,
  6365. keywords = "ChR2; Channelrhodopsin-2; Decision making; Halorhodopsin;
  6366. Midbrain; Motor planning; Mouse behavior; NpHR; SC;
  6367. channelrhodopsin-2; contra.; contraversive; halorhodopsin;
  6368. ipsi.; ipsiversive; mW/mm(2); milliwatts per millimeter squared
  6369. (power output measured at optic fiber tip); superior colliculus",
  6370. language = "en"
  6371. }
  6372. @UNPUBLISHED{Coletta2020-gb,
  6373. title = "Network structure of the mouse brain connectome with voxel
  6374. resolution",
  6375. author = "Coletta, Ludovico and Pagani, Marco and Whitesell, Jennifer D and
  6376. Harris, Julie A and Bernhardt, Boris and Gozzi, Alessandro",
  6377. abstract = "Fine-grained descriptions of brain connectivity are fundamental
  6378. for understanding how neural information is processed and relayed
  6379. across spatial scales. Prior investigations of the mouse brain
  6380. connectome have employed discrete anatomical parcellations,
  6381. limiting spatial resolution and potentially concealing network
  6382. attributes critical to the organization of the mammalian
  6383. connectome. Here we provide a voxel-level description of the
  6384. network and hierarchical structure of the directed mouse
  6385. connectome, unconstrained by regional partitioning. We show that
  6386. integrative hub regions can be directionally segregated into
  6387. neural sinks and sources, defining a hierarchical axis. We
  6388. describe a set of structural communities that spatially
  6389. reconstitute previously described fMRI networks of the mouse
  6390. brain, and document that neuromodulatory nuclei are strategically
  6391. wired as critical orchestrators of inter-modular and network
  6392. communicability. Notably, like in primates, the directed mouse
  6393. connectome is organized along two superimposed cortical gradients
  6394. reflecting unimodal-transmodal functional processing and a
  6395. modality-specific sensorimotor axis. These structural features
  6396. can be related to patterns of intralaminar connectivity and to
  6397. the spatial topography of dynamic fMRI brain states,
  6398. respectively. Together, our results reveal a high-resolution
  6399. structural scaffold linking mesoscale connectome topography to
  6400. its macroscale functional organization, and create opportunities
  6401. for identifying targets of interventions to modulate brain
  6402. function in a physiologically-accessible species.",
  6403. journal = "bioRxiv",
  6404. pages = "2020.03.06.973164",
  6405. month = mar,
  6406. year = 2020,
  6407. language = "en"
  6408. }
  6409. @ARTICLE{Doykos2020-or,
  6410. title = "Monosynaptic inputs to specific cell types of the intermediate
  6411. and deep layers of the superior colliculus",
  6412. author = "Doykos, Ted K and Gilmer, Jesse I and Person, Abigail L and
  6413. Felsen, Gidon",
  6414. abstract = "The intermediate and deep layers of the midbrain superior
  6415. colliculus (SC) are a key locus for several critical functions,
  6416. including spatial attention, multisensory integration, and
  6417. behavioral responses. While the SC is known to integrate input
  6418. from a variety of brain regions, progress in understanding how
  6419. these inputs contribute to SC-dependent functions has been
  6420. hindered by the paucity of data on innervation patterns to
  6421. specific types of SC neurons. Here, we use G-deleted rabies
  6422. virus-mediated monosynaptic tracing to identify inputs to
  6423. excitatory and inhibitory neurons of the intermediate and deep
  6424. SC. We observed stronger and more numerous projections to
  6425. excitatory than inhibitory SC neurons. However, a subpopulation
  6426. of excitatory neurons thought to mediate behavioral output
  6427. received weaker inputs, from far fewer brain regions, than the
  6428. overall population of excitatory neurons. Additionally,
  6429. extrinsic inputs tended to target rostral excitatory and
  6430. inhibitory SC neurons more strongly than their caudal
  6431. counterparts, and commissural SC neurons tended to project to
  6432. similar rostrocaudal positions in the other SC. Our findings
  6433. support the view that active intrinsic processes are critical to
  6434. SC-dependent functions, and will enable the examination of how
  6435. specific inputs contribute to these functions.",
  6436. journal = "J. Comp. Neurol.",
  6437. publisher = "Wiley Online Library",
  6438. month = feb,
  6439. year = 2020,
  6440. keywords = "RRIDs: FIJI software: SCR\_002285; RRIDs: ImageJ software:
  6441. SCR\_003070; RRIDs: Jackson labs heterozygous Gad2-Cre mice:
  6442. IMSR\_JAX:010802; RRIDs: Jackson labs homozygous Vglut2-Cre
  6443. mice: IMSR\_JAX:028863; RRIDs: MATLAB Computer Vision System
  6444. Toolbox software: SCR\_017581; RRIDs: MATLAB software:
  6445. SCR\_001622; RRIDs: Thermo Fisher Scientific Nissl: AB\_2572212;
  6446. RRIDs: $\mu$Manager software: SCR\_016865; excitatory;
  6447. inhibitory; monosynaptic; neuroanatomy; rabies; sensorimotor;
  6448. superior colliculus",
  6449. language = "en"
  6450. }
  6451. % The entry below contains non-ASCII chars that could not be converted
  6452. % to a LaTeX equivalent.
  6453. @INCOLLECTION{Belkin2002-ei,
  6454. title = "Laplacian Eigenmaps and Spectral Techniques for Embedding and
  6455. Clustering",
  6456. booktitle = "Advances in Neural Information Processing Systems 14",
  6457. author = "Belkin, Mikhail and Niyogi, Partha",
  6458. editor = "Dietterich, T G and Becker, S and Ghahramani, Z",
  6459. abstract = "Drawing on the correspondence between the graph Laplacian, the
  6460. Laplace-Beltrami operator on a manifold, and the connections to
  6461. the heat equation, we propose a geometrically motivated
  6462. algorithm for constructing a representation for data sampled
  6463. from a low dimensional manifold embedded in a higher dimensional
  6464. space. The algorithm provides a computationally efficient
  6465. approach to nonlinear dimensionality reduction that has locality
  6466. preserving properties and a natural connection to clustering.
  6467. Several applications are …",
  6468. publisher = "MIT Press",
  6469. pages = "585--591",
  6470. year = 2002
  6471. }
  6472. % The entry below contains non-ASCII chars that could not be converted
  6473. % to a LaTeX equivalent.
  6474. @ARTICLE{Maaten2008-tv,
  6475. title = "Visualizing Data using {t-SNE}",
  6476. author = "Maaten, Laurens van der and Hinton, Geoffrey",
  6477. abstract = "We present a new technique called`` t-SNE'' that visualizes
  6478. high-dimensional data by giving each datapoint a location in a
  6479. two or three-dimensional map. The technique is a variation of
  6480. Stochastic Neighbor Embedding (Hinton and Roweis, 2002) that is
  6481. much easier to optimize …",
  6482. journal = "J. Mach. Learn. Res.",
  6483. publisher = "jmlr.org",
  6484. volume = 9,
  6485. number = "Nov",
  6486. pages = "2579--2605",
  6487. year = 2008
  6488. }
  6489. % The entry below contains non-ASCII chars that could not be converted
  6490. % to a LaTeX equivalent.
  6491. @ARTICLE{Balasubramanian2002-ly,
  6492. title = "The isomap algorithm and topological stability",
  6493. author = "Balasubramanian, Mukund and Schwartz, Eric L",
  6494. abstract = "Tenenbaum et al.(1) presented an algorithm, Isomap , for
  6495. computing a quasi-isometric, low- dimensional embedding of a set
  6496. of high-dimensional data points. Two issues need to be raised
  6497. concerning this work. First, the basic approach presented by
  6498. Tenenbaum et al. is not …",
  6499. journal = "Science",
  6500. publisher = "science.sciencemag.org",
  6501. volume = 295,
  6502. number = 5552,
  6503. pages = "7",
  6504. month = jan,
  6505. year = 2002,
  6506. language = "en"
  6507. }
  6508. @UNPUBLISHED{Essig2020-br,
  6509. title = "Inhibitory midbrain neurons mediate decision making",
  6510. author = "Essig, Jaclyn and Hunt, Joshua B and Felsen, Gidon",
  6511. abstract = "Decision making is critical for survival but its neural basis is
  6512. unclear. Here we examine how functional neural circuitry in the
  6513. output layers of the midbrain superior colliculus (SC) mediates
  6514. spatial choice, an SC-dependent tractable form of decision
  6515. making. We focus on the role of inhibitory SC neurons, using
  6516. optogenetics to record and manipulate their activity in behaving
  6517. mice. Based on data from SC slice experiments and on a canonical
  6518. role of inhibitory neurons in cortical microcircuits, we
  6519. hypothesized that inhibitory SC neurons locally inhibit premotor
  6520. output neurons that represent contralateral targets. However, our
  6521. experimental results refuted this hypothesis. An attractor model
  6522. revealed that our results were instead consistent with inhibitory
  6523. neurons providing long-range inhibition between the two SCs, and
  6524. terminal activation experiments supported this architecture. Our
  6525. study provides mechanistic evidence for competitive inhibition
  6526. between populations representing discrete choices, a common motif
  6527. in theoretical models of decision making.",
  6528. journal = "bioRxiv",
  6529. pages = "2020.02.25.965699",
  6530. month = feb,
  6531. year = 2020,
  6532. language = "en"
  6533. }
  6534. @UNPUBLISHED{Liu2020-pe,
  6535. title = "Accurate localization of linear probe electrodes across multiple
  6536. brains",
  6537. author = "Liu, Liu D and Chen, Susu and Economo, Michael N and Li, Nuo and
  6538. Svoboda, Karel",
  6539. abstract = "Recently developed silicon probes have large numbers of recording
  6540. electrodes on long linear shanks. Specifically, Neuropixels
  6541. probes have 960 recording electrodes distributed over 9.6 mm
  6542. shanks. Because of their length, Neuropixels probe recordings in
  6543. rodents naturally span multiple brain areas. Typical studies
  6544. collate recordings across several recording sessions and animals.
  6545. Neurons recorded in different sessions and animals have to be
  6546. aligned to each other and to a standardized brain coordinate
  6547. system. Here we report a workflow for accurate localization of
  6548. individual electrodes in standardized coordinates and aligned
  6549. across individual brains. This workflow relies on imaging brains
  6550. with fluorescent probe tracks and warping 3-dimensional image
  6551. stacks to standardized brain atlases. Electrophysiological
  6552. features are then used to anchor particular electrodes along the
  6553. reconstructed tracks to specific locations in the brain atlas and
  6554. therefore to specific brain structures. We performed ground-truth
  6555. experiments, in which motor cortex outputs are labelled with ChR2
  6556. and a fluorescence protein. Recording from brain regions targeted
  6557. by these outputs reveals better than 100 $\mu$m accuracy for
  6558. electrode localization.",
  6559. journal = "bioRxiv",
  6560. pages = "2020.02.25.965210",
  6561. month = feb,
  6562. year = 2020,
  6563. language = "en"
  6564. }
  6565. @ARTICLE{Becht2018-hp,
  6566. title = "Dimensionality reduction for visualizing single-cell data using
  6567. {UMAP}",
  6568. author = "Becht, Etienne and McInnes, Leland and Healy, John and Dutertre,
  6569. Charles-Antoine and Kwok, Immanuel W H and Ng, Lai Guan and
  6570. Ginhoux, Florent and Newell, Evan W",
  6571. abstract = "Advances in single-cell technologies have enabled
  6572. high-resolution dissection of tissue composition. Several tools
  6573. for dimensionality reduction are available to analyze the large
  6574. number of parameters generated in single-cell studies. Recently,
  6575. a nonlinear dimensionality-reduction technique, uniform manifold
  6576. approximation and projection (UMAP), was developed for the
  6577. analysis of any type of high-dimensional data. Here we apply it
  6578. to biological data, using three well-characterized mass
  6579. cytometry and single-cell RNA sequencing datasets. Comparing the
  6580. performance of UMAP with five other tools, we find that UMAP
  6581. provides the fastest run times, highest reproducibility and the
  6582. most meaningful organization of cell clusters. The work
  6583. highlights the use of UMAP for improved visualization and
  6584. interpretation of single-cell data.",
  6585. journal = "Nat. Biotechnol.",
  6586. publisher = "nature.com",
  6587. month = dec,
  6588. year = 2018,
  6589. language = "en"
  6590. }
  6591. @ARTICLE{McInnes2018-dg,
  6592. title = "{UMAP}: Uniform Manifold Approximation and Projection for
  6593. Dimension Reduction",
  6594. author = "McInnes, Leland and Healy, John and Melville, James",
  6595. abstract = "UMAP (Uniform Manifold Approximation and Projection) is a
  6596. novel manifold learning technique for dimension reduction.
  6597. UMAP is constructed from a theoretical framework based in
  6598. Riemannian geometry and algebraic topology. The result is a
  6599. practical scalable algorithm that applies to real world
  6600. data. The UMAP algorithm is competitive with t-SNE for
  6601. visualization quality, and arguably preserves more of the
  6602. global structure with superior run time performance.
  6603. Furthermore, UMAP has no computational restrictions on
  6604. embedding dimension, making it viable as a general purpose
  6605. dimension reduction technique for machine learning.",
  6606. month = feb,
  6607. year = 2018,
  6608. archivePrefix = "arXiv",
  6609. primaryClass = "stat.ML",
  6610. eprint = "1802.03426"
  6611. }
  6612. @ARTICLE{Kobak2019-se,
  6613. title = "The art of using {t-SNE} for single-cell transcriptomics",
  6614. author = "Kobak, Dmitry and Berens, Philipp",
  6615. abstract = "Single-cell transcriptomics yields ever growing data sets
  6616. containing RNA expression levels for thousands of genes from up
  6617. to millions of cells. Common data analysis pipelines include a
  6618. dimensionality reduction step for visualising the data in two
  6619. dimensions, most frequently performed using t-distributed
  6620. stochastic neighbour embedding (t-SNE). It excels at revealing
  6621. local structure in high-dimensional data, but naive applications
  6622. often suffer from severe shortcomings, e.g. the global structure
  6623. of the data is not represented accurately. Here we describe how
  6624. to circumvent such pitfalls, and develop a protocol for creating
  6625. more faithful t-SNE visualisations. It includes PCA
  6626. initialisation, a high learning rate, and multi-scale similarity
  6627. kernels; for very large data sets, we additionally use
  6628. exaggeration and downsampling-based initialisation. We use
  6629. published single-cell RNA-seq data sets to demonstrate that this
  6630. protocol yields superior results compared to the naive
  6631. application of t-SNE.",
  6632. journal = "Nat. Commun.",
  6633. publisher = "nature.com",
  6634. volume = 10,
  6635. number = 1,
  6636. pages = "5416",
  6637. month = nov,
  6638. year = 2019,
  6639. language = "en"
  6640. }
  6641. @ARTICLE{Kuwabara2020-qm,
  6642. title = "Neural mechanisms of economic choices in mice",
  6643. author = "Kuwabara, Masaru and Kang, Ningdong and Holy, Timothy E and
  6644. Padoa-Schioppa, Camillo",
  6645. abstract = "Economic choices entail computing and comparing subjective
  6646. values. Evidence from primates indicates that this behavior
  6647. relies on the orbitofrontal cortex. Conversely, previous work in
  6648. rodents provided conflicting results. Here we present a mouse
  6649. model of economic choice behavior, and we show that the lateral
  6650. orbital (LO) area is intimately related to the decision process.
  6651. In the experiments, mice chose between different juices offered
  6652. in variable amounts. Choice patterns closely resembled those
  6653. measured in primates. Optogenetic inactivation of LO dramatically
  6654. disrupted choices by inducing erratic changes of relative value
  6655. and by increasing choice variability. Neuronal recordings
  6656. revealed that different groups of cells encoded the values of
  6657. individual options, the binary choice outcome and the chosen
  6658. value. These groups match those previously identified in
  6659. primates, except that the neuronal representation in mice is
  6660. spatial (in monkeys it is good-based). Our results lay the
  6661. foundations for a circuit-level analysis of economic decisions.",
  6662. journal = "Elife",
  6663. volume = 9,
  6664. month = feb,
  6665. year = 2020,
  6666. keywords = "mouse; neuroscience",
  6667. language = "en"
  6668. }
  6669. % The entry below contains non-ASCII chars that could not be converted
  6670. % to a LaTeX equivalent.
  6671. @ARTICLE{Ascoli2007-hp,
  6672. title = "{NeuroMorpho.Org}: a central resource for neuronal morphologies",
  6673. author = "Ascoli, Giorgio A and Donohue, Duncan E and Halavi, Maryam",
  6674. abstract = "The structure of dendrites and axons plays fundamental roles in
  6675. synaptic integration and network connectivity. Synergistic
  6676. advances in neurobiology (eg, intracellular injections,
  6677. fluorescent protein expression), microscopy (eg, multiphoton
  6678. laser scanning, computer controllers), and imaging software (eg,
  6679. Neurolucida tracing, blind deconvolution) are rapidly
  6680. transforming the three-dimensional (3D) reconstruction of
  6681. neuronal morphology into a mainstream technique. In the course
  6682. of electrophysiological, pharmacological, or …",
  6683. journal = "J. Neurosci.",
  6684. publisher = "Soc Neuroscience",
  6685. volume = 27,
  6686. number = 35,
  6687. pages = "9247--9251",
  6688. month = aug,
  6689. year = 2007,
  6690. language = "en"
  6691. }
  6692. @ARTICLE{Muller2015-ws,
  6693. title = "Python in neuroscience",
  6694. author = "Muller, Eilif and Bednar, James A and Diesmann, Markus and
  6695. Gewaltig, Marc-Oliver and Hines, Michael and Davison, Andrew P",
  6696. journal = "Front. Neuroinform.",
  6697. volume = 9,
  6698. pages = "11",
  6699. month = apr,
  6700. year = 2015,
  6701. keywords = "collaboration; interoperability; python language; scientific
  6702. computing; software development",
  6703. language = "en"
  6704. }
  6705. @ARTICLE{Niedworok2016-wj,
  6706. title = "{aMAP} is a validated pipeline for registration and segmentation
  6707. of high-resolution mouse brain data",
  6708. author = "Niedworok, Christian J and Brown, Alexander P Y and Jorge
  6709. Cardoso, M and Osten, Pavel and Ourselin, Sebastien and Modat,
  6710. Marc and Margrie, Troy W",
  6711. abstract = "The validation of automated image registration and segmentation
  6712. is crucial for accurate and reliable mapping of brain
  6713. connectivity and function in three-dimensional (3D) data sets.
  6714. While validation standards are necessarily high and routinely met
  6715. in the clinical arena, they have to date been lacking for
  6716. high-resolution microscopy data sets obtained from the rodent
  6717. brain. Here we present a tool for optimized automated mouse atlas
  6718. propagation (aMAP) based on clinical registration software
  6719. (NiftyReg) for anatomical segmentation of high-resolution 3D
  6720. fluorescence images of the adult mouse brain. We empirically
  6721. evaluate aMAP as a method for registration and subsequent
  6722. segmentation by validating it against the performance of expert
  6723. human raters. This study therefore establishes a benchmark
  6724. standard for mapping the molecular function and cellular
  6725. connectivity of the rodent brain.",
  6726. journal = "Nat. Commun.",
  6727. volume = 7,
  6728. pages = "11879",
  6729. month = jul,
  6730. year = 2016,
  6731. language = "en"
  6732. }
  6733. @UNPUBLISHED{Guitchounts2020-xj,
  6734. title = "Encoding of {3D} Head Orienting Movements in Primary Visual
  6735. Cortex",
  6736. author = "Guitchounts, Grigori and Masis, Javier and Wolff, Steffen B E and
  6737. Cox, David",
  6738. abstract = "Animals actively sample from the sensory world by generating
  6739. complex patterns of movement that evolve in three dimensions. At
  6740. least some of these movements have been shown to influence neural
  6741. codes in sensory areas. For example, in primary visual cortex
  6742. (V1), locomotion-related neural activity influences sensory gain,
  6743. encodes running speed, and predicts the direction of visual flow.
  6744. As most experiments exploring movement-related modulation of V1
  6745. have been performed in head-fixed animals, it remains unclear
  6746. whether or how the naturalistic movements used to interact with
  6747. sensory stimuli--like head orienting--influence visual
  6748. processing. Here we show that 3D head orienting movements
  6749. modulate V1 neuronal activity in a direction-specific manner that
  6750. also depends on the presence or absence of light. We identify two
  6751. largely independent populations of movement-direction-tuned
  6752. neurons that support this modulation, one of which is
  6753. direction-tuned in the dark and the other in the light. Finally,
  6754. we demonstrate that V1 gains access to a motor efference copy
  6755. related to orientation from secondary motor cortex, which has
  6756. been shown to control head orienting movements. These results
  6757. suggest a mechanism through which sensory signals generated by
  6758. purposeful movement can be distinguished from those arising in
  6759. the outside world, and reveal a pervasive role of 3D movement in
  6760. shaping sensory cortical dynamics.",
  6761. journal = "bioRxiv",
  6762. pages = "2020.01.16.909473",
  6763. month = jan,
  6764. year = 2020,
  6765. language = "en"
  6766. }
  6767. @ARTICLE{Mobbs2020-vp,
  6768. title = "Space, Time, and Fear: Survival Computations along Defensive
  6769. Circuits",
  6770. author = "Mobbs, Dean and Headley, Drew B and Ding, Weilun and Dayan,
  6771. Peter",
  6772. abstract = "Naturalistic observations show that decisions to avoid or escape
  6773. predators occur at different spatiotemporal scales and that they
  6774. are supported by different computations and neural circuits. At
  6775. their extremes, proximal threats are addressed by a limited
  6776. repertoire of reflexive and myopic actions, reflecting reduced
  6777. decision and state spaces and model-free (MF) architectures.
  6778. Conversely, distal threats allow increased information
  6779. processing supported by model-based (MB) operations, including
  6780. affective prospection, replay, and planning. However, MF and MB
  6781. computations are often intertwined, and under conditions of
  6782. safety the foundations for future effective reactive execution
  6783. can be laid through MB instruction of MF control. Together,
  6784. these computations are associated with distinct population codes
  6785. embedded within a distributed defensive circuitry whose goal is
  6786. to determine and realize the best policy.",
  6787. journal = "Trends Cogn. Sci.",
  6788. publisher = "Elsevier",
  6789. volume = 0,
  6790. number = 0,
  6791. month = feb,
  6792. year = 2020,
  6793. keywords = "fear; anxiety; threat imminence continuum; periaqueductal gray;
  6794. hippocampus; prefrontal cortex; model free; model based; Dyna",
  6795. language = "en"
  6796. }
  6797. @ARTICLE{Hasson2020-zv,
  6798. title = "Direct Fit to Nature: An Evolutionary Perspective on Biological
  6799. and Artificial Neural Networks",
  6800. author = "Hasson, Uri and Nastase, Samuel A and Goldstein, Ariel",
  6801. abstract = "SummaryEvolution is a blind fitting process by which organisms
  6802. become adapted to their environment. Does the brain use similar
  6803. brute-force fitting processes to learn how to perceive and act
  6804. upon the world? Recent advances in artificial neural networks
  6805. have exposed the power of optimizing millions of synaptic
  6806. weights over millions of observations to operate robustly in
  6807. real-world contexts. These models do not learn simple,
  6808. human-interpretable rules or representations of the world;
  6809. rather, they use local computations to interpolate over
  6810. task-relevant manifolds in a high-dimensional parameter space.
  6811. Counterintuitively, similar to evolutionary processes,
  6812. over-parameterized models can be simple and parsimonious, as
  6813. they provide a versatile, robust solution for learning a diverse
  6814. set of functions. This new family of direct-fit models present a
  6815. radical challenge to many of the theoretical assumptions in
  6816. psychology and neuroscience. At the same time, this shift in
  6817. perspective establishes unexpected links with developmental and
  6818. ecological psychology.",
  6819. journal = "Neuron",
  6820. publisher = "Elsevier",
  6821. volume = 105,
  6822. number = 3,
  6823. pages = "416--434",
  6824. month = feb,
  6825. year = 2020,
  6826. keywords = "evolution; experimental design; interpolation; learning; neural
  6827. networks",
  6828. language = "en"
  6829. }
  6830. @ARTICLE{Barabasi2019-ou,
  6831. title = "A Genetic Model of the Connectome",
  6832. author = "Barab{\'a}si, D{\'a}niel L and Barab{\'a}si,
  6833. Albert-L{\'a}szl{\'o}",
  6834. abstract = "The connectomes of organisms of the same species show remarkable
  6835. architectural and often local wiring similarity, raising the
  6836. question: where and how is neuronal connectivity encoded? Here,
  6837. we start from the hypothesis that the genetic identity of neurons
  6838. guides synapse and gap-junction formation and show that such
  6839. genetically driven wiring predicts the existence of specific
  6840. biclique motifs in the connectome. We identify a family of large,
  6841. statistically significant biclique subgraphs in the connectomes
  6842. of three species and show that within many of the observed
  6843. bicliques the neurons share statistically significant expression
  6844. patterns and morphological characteristics, supporting our
  6845. expectation of common genetic factors that drive the synapse
  6846. formation within these subgraphs. The proposed connectome model
  6847. offers a self-consistent framework to link the genetics of an
  6848. organism to the reproducible architecture of its connectome,
  6849. offering experimentally falsifiable predictions on the genetic
  6850. factors that drive the formation of individual neuronal circuits.",
  6851. journal = "Neuron",
  6852. month = nov,
  6853. year = 2019,
  6854. keywords = "Brain Networks; C. elegans; Connectomics; Generative Model;
  6855. Theory; development; encoding",
  6856. language = "en"
  6857. }
  6858. @ARTICLE{Kaplan2019-zb,
  6859. title = "Nested Neuronal Dynamics Orchestrate a Behavioral Hierarchy
  6860. across Timescales",
  6861. author = "Kaplan, Harris S and Salazar Thula, Oriana and Khoss, Niklas and
  6862. Zimmer, Manuel",
  6863. abstract = "Classical and modern ethological studies suggest that animal
  6864. behavior is organized hierarchically across timescales, such that
  6865. longer-timescale behaviors are composed of specific
  6866. shorter-timescale actions. Despite progress relating neuronal
  6867. dynamics to single-timescale behavior, it remains unclear how
  6868. different timescale dynamics interact to give rise to such
  6869. higher-order behavioral organization. Here, we show, in the
  6870. nematode Caenorhabditis elegans, that a behavioral hierarchy
  6871. spanning three timescales is implemented by nested neuronal
  6872. dynamics. At the uppermost hierarchical level, slow neuronal
  6873. population dynamics spanning brain and motor periphery control
  6874. two faster motor neuron oscillations, toggling them between
  6875. different activity states and functional roles. At lower
  6876. hierarchical levels, these faster oscillations are further nested
  6877. in a manner that enables flexible behavioral control in an
  6878. otherwise rigid hierarchical framework. Our findings establish
  6879. nested neuronal activity patterns as a repeated dynamical motif
  6880. of the C. elegans nervous system, which together implement a
  6881. controllable hierarchical organization of behavior.",
  6882. journal = "Neuron",
  6883. month = nov,
  6884. year = 2019,
  6885. keywords = "C. elegans neuroscience; behavior organization; behavioral
  6886. hierarchy; ethology; hierarchical organization; motor control;
  6887. neuronal dynamics; neuronal oscillations; quantitative behavior;
  6888. whole-brain imaging",
  6889. language = "en"
  6890. }
  6891. @ARTICLE{Mendes-Gomes2020-de,
  6892. title = "Defensive behaviors and brain regional activation changes in rats
  6893. confronting a snake",
  6894. author = "Mendes-Gomes, Joyce and Motta, Simone Cristina and Passoni Bindi,
  6895. Ricardo and de Oliveira, Amanda Ribeiro and Ullah, Farhad and
  6896. Baldo, Marcus Vinicius C and Coimbra, Norberto Cysne and
  6897. Canteras, Newton Sabino and Blanchard, D Caroline",
  6898. abstract = "In the present study, we examined behavioral and brain regional
  6899. activation changes of rats). To a nonmammalian predator, a wild
  6900. rattler snake (Crotalus durissus terrificus). Accordingly, during
  6901. snake threat, rat subjects showed a striking and highly
  6902. significant behavioral response of freezing, stretch attend, and,
  6903. especially, spatial avoidance of this threat. The brain regional
  6904. activation patterns for these rats were in broad outline similar
  6905. to those of rats encountering other predator threats, showing Fos
  6906. activation of sites in the amygdala, hypothalamus, and
  6907. periaqueductal gray matter. In the amygdala, only the lateral
  6908. nucleus showed significant activation, although the medial
  6909. nucleus, highly responsive to olfaction, also showed higher
  6910. activation. Importantly, the hypothalamus, in particular, was
  6911. somewhat different, with significant Fos increases in the
  6912. anterior and central parts of the ventromedial hypothalamic
  6913. nucleus (VMH), in contrast to patterns of enhanced Fos expression
  6914. in the dorsomedial VMH to cat predators, and in the ventrolateral
  6915. VMH to an attacking conspecific. In addition, the
  6916. juxtodorsalmedial region of the lateral hypothalamus showed
  6917. enhanced Fos activation, where inputs from the septo-hippocampal
  6918. system may suggest the potential involvement of hippocampal
  6919. boundary cells in the very strong spatial avoidance of the snake
  6920. and the area it occupied. Notably, these two hypothalamic paths
  6921. appear to merge into the dorsomedial part of the dorsal
  6922. premammillary nucleus and dorsomedial and lateral parts of the
  6923. periaqueductal gray, all of which present significant increases
  6924. in Fos expression and are likely to be critical for the
  6925. expression of defensive behaviors in responses to the snake
  6926. threat.",
  6927. journal = "Behav. Brain Res.",
  6928. volume = 381,
  6929. pages = "112469",
  6930. month = mar,
  6931. year = 2020,
  6932. keywords = "Amygdala; Antipredatory defense; Hippocampus; Hypothalamus;
  6933. Periaqueductal gray; Prey versus rattlesnake confrontation
  6934. paradigm",
  6935. language = "en"
  6936. }
  6937. @ARTICLE{Grillner2002-si,
  6938. title = "Cellular bases of a vertebrate locomotor system-steering,
  6939. intersegmental and segmental co-ordination and sensory control",
  6940. author = "Grillner, Sten and Wall{\'e}n, Peter",
  6941. abstract = "The isolated brainstem-spinal cord of the lamprey is used as an
  6942. experimental model in the analysis of the cellular bases of
  6943. vertebrate locomotor behaviour. In this article we review the
  6944. neural mechanisms involved in the control of steering,
  6945. intersegmental co-ordination, as well as the segmental burst
  6946. generation and the sensory contribution to motor pattern
  6947. generation. Within these four components of the control system
  6948. for locomotion, we now have good knowledge of not only the
  6949. neurones that take part and their synaptic interactions, but also
  6950. the membrane properties of these neurones, including ion channel
  6951. subtypes, and their contribution to motor pattern generation.",
  6952. journal = "Brain Res. Brain Res. Rev.",
  6953. volume = 40,
  6954. number = "1-3",
  6955. pages = "92--106",
  6956. month = oct,
  6957. year = 2002,
  6958. language = "en"
  6959. }
  6960. @ARTICLE{Krumin2018-yz,
  6961. title = "Decision and navigation in mouse parietal cortex",
  6962. author = "Krumin, Michael and Lee, Julie J and Harris, Kenneth D and
  6963. Carandini, Matteo",
  6964. abstract = "Posterior parietal cortex (PPC) has been implicated in
  6965. navigation, in the control of movement, and in visually-guided
  6966. decisions. To relate these views, we measured activity in PPC
  6967. while mice performed a virtual navigation task driven by visual
  6968. decisions. PPC neurons were selective for specific combinations
  6969. of the animal's spatial position and heading angle. This
  6970. selectivity closely predicted both the activity of individual PPC
  6971. neurons, and the arrangement of their collective firing patterns
  6972. in choice-selective sequences. These sequences reflected PPC
  6973. encoding of the animal's navigation trajectory. Using decision as
  6974. a predictor instead of heading yielded worse fits, and using it
  6975. in addition to heading only slightly improved the fits.
  6976. Alternative models based on visual or motor variables were
  6977. inferior. We conclude that when mice use vision to choose their
  6978. trajectories, a large fraction of parietal cortex activity can be
  6979. predicted from simple attributes such as spatial position and
  6980. heading.",
  6981. journal = "Elife",
  6982. volume = 7,
  6983. month = nov,
  6984. year = 2018,
  6985. keywords = "cortex; decision; mouse; navigation; neuroscience; visual
  6986. processing",
  6987. language = "en"
  6988. }
  6989. @ARTICLE{Stringer2019-fm,
  6990. title = "Spontaneous behaviors drive multidimensional, brainwide activity",
  6991. author = "Stringer, Carsen and Pachitariu, Marius and Steinmetz, Nicholas
  6992. and Reddy, Charu Bai and Carandini, Matteo and Harris, Kenneth D",
  6993. abstract = "Neuronal populations in sensory cortex produce variable responses
  6994. to sensory stimuli and exhibit intricate spontaneous activity
  6995. even without external sensory input. Cortical variability and
  6996. spontaneous activity have been variously proposed to represent
  6997. random noise, recall of prior experience, or encoding of ongoing
  6998. behavioral and cognitive variables. Recording more than 10,000
  6999. neurons in mouse visual cortex, we observed that spontaneous
  7000. activity reliably encoded a high-dimensional latent state, which
  7001. was partially related to the mouse's ongoing behavior and was
  7002. represented not just in visual cortex but also across the
  7003. forebrain. Sensory inputs did not interrupt this ongoing signal
  7004. but added onto it a representation of external stimuli in
  7005. orthogonal dimensions. Thus, visual cortical population activity,
  7006. despite its apparently noisy structure, reliably encodes an
  7007. orthogonal fusion of sensory and multidimensional behavioral
  7008. information.",
  7009. journal = "Science",
  7010. volume = 364,
  7011. number = 6437,
  7012. pages = "255",
  7013. month = apr,
  7014. year = 2019,
  7015. language = "en"
  7016. }
  7017. @ARTICLE{Stringer2019-nc,
  7018. title = "High-dimensional geometry of population responses in visual
  7019. cortex",
  7020. author = "Stringer, Carsen and Pachitariu, Marius and Steinmetz, Nicholas
  7021. and Carandini, Matteo and Harris, Kenneth D",
  7022. abstract = "A neuronal population encodes information most efficiently when
  7023. its stimulus responses are high-dimensional and uncorrelated, and
  7024. most robustly when they are lower-dimensional and correlated.
  7025. Here we analysed the dimensionality of the encoding of natural
  7026. images by large populations of neurons in the visual cortex of
  7027. awake mice. The evoked population activity was high-dimensional,
  7028. and correlations obeyed an unexpected power law: the nth
  7029. principal component variance scaled as 1/n. This scaling was not
  7030. inherited from the power law spectrum of natural images, because
  7031. it persisted after stimulus whitening. We proved mathematically
  7032. that if the variance spectrum was to decay more slowly then the
  7033. population code could not be smooth, allowing small changes in
  7034. input to dominate population activity. The theory also predicts
  7035. larger power-law exponents for lower-dimensional stimulus
  7036. ensembles, which we validated experimentally. These results
  7037. suggest that coding smoothness may represent a fundamental
  7038. constraint that determines correlations in neural population
  7039. codes.",
  7040. journal = "Nature",
  7041. volume = 571,
  7042. number = 7765,
  7043. pages = "361--365",
  7044. month = jul,
  7045. year = 2019,
  7046. language = "en"
  7047. }
  7048. % The entry below contains non-ASCII chars that could not be converted
  7049. % to a LaTeX equivalent.
  7050. @ARTICLE{Spellman2015-qp,
  7051. title = "Hippocampal--prefrontal input supports spatial encoding in
  7052. working memory",
  7053. author = "Spellman, T and Rigotti, M and Ahmari, S E and Fusi, S and
  7054. Gogos, J A and {others}",
  7055. abstract = "Spatial working memory, the caching of behaviourally relevant
  7056. spatial cues on a timescale of seconds, is a fundamental
  7057. constituent of cognition. Although the prefrontal cortex and
  7058. hippocampus are known to contribute jointly to successful
  7059. spatial working memory, the …",
  7060. journal = "Nature",
  7061. publisher = "nature.com",
  7062. year = 2015
  7063. }
  7064. @ARTICLE{Rigotti2013-ls,
  7065. title = "The importance of mixed selectivity in complex cognitive tasks",
  7066. author = "Rigotti, Mattia and Barak, Omri and Warden, Melissa R and Wang,
  7067. Xiao-Jing and Daw, Nathaniel D and Miller, Earl K and Fusi,
  7068. Stefano",
  7069. abstract = "Single-neuron activity in the prefrontal cortex (PFC) is tuned to
  7070. mixtures of multiple task-related aspects. Such mixed selectivity
  7071. is highly heterogeneous, seemingly disordered and therefore
  7072. difficult to interpret. We analysed the neural activity recorded
  7073. in monkeys during an object sequence memory task to identify a
  7074. role of mixed selectivity in subserving the cognitive functions
  7075. ascribed to the PFC. We show that mixed selectivity neurons
  7076. encode distributed information about all task-relevant aspects.
  7077. Each aspect can be decoded from the population of neurons even
  7078. when single-cell selectivity to that aspect is eliminated.
  7079. Moreover, mixed selectivity offers a significant computational
  7080. advantage over specialized responses in terms of the repertoire
  7081. of input-output functions implementable by readout neurons. This
  7082. advantage originates from the highly diverse nonlinear
  7083. selectivity to mixtures of task-relevant variables, a signature
  7084. of high-dimensional neural representations. Crucially, this
  7085. dimensionality is predictive of animal behaviour as it collapses
  7086. in error trials. Our findings recommend a shift of focus for
  7087. future studies from neurons that have easily interpretable
  7088. response tuning to the widely observed, but rarely analysed,
  7089. mixed selectivity neurons.",
  7090. journal = "Nature",
  7091. volume = 497,
  7092. number = 7451,
  7093. pages = "585--590",
  7094. month = may,
  7095. year = 2013,
  7096. language = "en"
  7097. }
  7098. @ARTICLE{Shang2019-nm,
  7099. title = "A subcortical excitatory circuit for sensory-triggered predatory
  7100. hunting in mice",
  7101. author = "Shang, Congping and Liu, Aixue and Li, Dapeng and Xie, Zhiyong
  7102. and Chen, Zijun and Huang, Meizhu and Li, Yang and Wang, Yi and
  7103. Shen, Wei L and Cao, Peng",
  7104. abstract = "Predatory hunting plays a fundamental role in animal survival.
  7105. Little is known about the neural circuits that convert sensory
  7106. cues into neural signals to drive this behavior. Here we
  7107. identified an excitatory subcortical neural circuit from the
  7108. superior colliculus to the zona incerta that triggers predatory
  7109. hunting. The superior colliculus neurons that form this pathway
  7110. integrate motion-related visual and vibrissal somatosensory cues
  7111. of prey. During hunting, these neurons send out neural signals
  7112. that are temporally correlated with predatory attacks, but not
  7113. with feeding after prey capture. Synaptic inactivation of this
  7114. pathway selectively blocks hunting for prey without impairing
  7115. other sensory-triggered behaviors. These data reveal a
  7116. subcortical neural circuit that is specifically engaged in
  7117. translating sensory cues into neural signals to provoke predatory
  7118. hunting.",
  7119. journal = "Nat. Neurosci.",
  7120. volume = 22,
  7121. number = 6,
  7122. pages = "909--920",
  7123. month = jun,
  7124. year = 2019,
  7125. language = "en"
  7126. }
  7127. @UNPUBLISHED{Cadena2019-ly,
  7128. title = "How well do deep neural networks trained on object recognition
  7129. characterize the mouse visual system?",
  7130. author = "Cadena, Santiago A and Sinz, Fabian H and Muhammad, Taliah and
  7131. Froudarakis, Emmanouil and Cobos, Erick and Walker, Edgar Y and
  7132. Reimer, Jake and Bethge, Matthias and Tolias, Andreas and Ecker,
  7133. Alexander S",
  7134. abstract = "Recent work on modeling neural responses in the primate visual
  7135. system has benefited from deep neural networks trained on
  7136. large-scale object recognition, and found a hierarchical
  7137. correspondence between layers of the artificial neural network
  7138. and brain areas along the ventral visual stream. However, we
  7139. neither know whether such task-optimized networks enable equally
  7140. good models of the rodent visual system, nor if a similar
  7141. hierarchical correspondence exists. Here, we address these
  7142. questions in the mouse visual system by extracting features at
  7143. several layers of a convolutional neural network (CNN) trained on
  7144. ImageNet to predict the responses of thousands of neurons in four
  7145. visual areas (V1, LM, AL, RL) to natural images. We found that
  7146. the CNN features outperform classical subunit energy models, but
  7147. found no evidence for an order of the areas we recorded via a
  7148. correspondence to the hierarchy of CNN layers. Moreover, the same
  7149. CNN but with random weights provided an equivalently useful
  7150. feature space for predicting neural responses. Our results
  7151. suggest that object recognition as a high-level task does not
  7152. provide more discriminative features to characterize the mouse
  7153. visual system than a random network. Unlike in the primate,
  7154. training on ethologically relevant visually guided behaviors --
  7155. beyond static object recognition -- may be needed to unveil the
  7156. functional organization of the mouse visual cortex.",
  7157. month = sep,
  7158. year = 2019
  7159. }
  7160. @ARTICLE{Dean1986-hf,
  7161. title = "Head and body movements produced by electrical stimulation of
  7162. superior colliculus in rats: effects of interruption of crossed
  7163. tectoreticulospinal pathway",
  7164. author = "Dean, P and Redgrave, P and Sahibzada, N and Tsuji, K",
  7165. abstract = "Stimulation of the superior colliculus in rats produces movements
  7166. of the head and body that resemble either orientation and
  7167. approach towards a contralateral stimulus, or avoidance of, or
  7168. escape from, such a stimulus. A variety of evidence indicates
  7169. that the crossed descending pathway, which runs in the
  7170. contralateral predorsal bundle to the pontomedullary reticular
  7171. formation and the spinal cord, is involved in orienting
  7172. movements. The nature of this involvement was investigated, by
  7173. assessing the effects on tectally-elicited movements of midbrain
  7174. knife-cuts intended to section the pathway as it crosses midline
  7175. in the dorsal tegmental decussation. As expected, ipsilateral
  7176. movements resembling avoidance or escape were little affected by
  7177. dorsal tegmental decussation section, whereas contralateral
  7178. circling movements of the body were almost abolished. However,
  7179. contralateral movements of the head in response to electrical
  7180. stimulation were not eliminated, nor were orienting head
  7181. movements to visual or tactile stimuli. There was some suggestion
  7182. that section of the dorsal tegmental decussation increased the
  7183. latency of head movements from electrical stimulation at lateral
  7184. sites, and decreased the accuracy of orienting movements to
  7185. sensory stimuli. These results support the view that the crossed
  7186. tectoreticulospinal system is concerned with approach rather than
  7187. avoidance movements. However, it appears that other, as yet
  7188. unidentified, tectal efferent systems are also involved in
  7189. orienting head movements. It is possible that this division of
  7190. labour may reflect functional differences between various kinds
  7191. of apparently similar orienting responses. One suggestion is that
  7192. the tectoreticulospinal system is concerned less in open-loop
  7193. orienting responses (that are initiated but not subsequently
  7194. guided by sensory stimuli), than in following or pursuit
  7195. movements.",
  7196. journal = "Neuroscience",
  7197. volume = 19,
  7198. number = 2,
  7199. pages = "367--380",
  7200. month = oct,
  7201. year = 1986,
  7202. language = "en"
  7203. }
  7204. @ARTICLE{Saitoh2007-yr,
  7205. title = "Tectal control of locomotion, steering, and eye movements in
  7206. lamprey",
  7207. author = "Saitoh, Kazuya and M{\'e}nard, Ariane and Grillner, Sten",
  7208. abstract = "The intrinsic function of the brain stem-spinal cord networks
  7209. eliciting the locomotor synergy is well described in the
  7210. lamprey-a vertebrate model system. This study addresses the role
  7211. of tectum in integrating eye, body orientation, and locomotor
  7212. movements as in steering and goal-directed behavior. Electrical
  7213. stimuli were applied to different areas within the optic tectum
  7214. in head-restrained semi-intact lampreys (n = 40). Motions of the
  7215. eyes and body were recorded simultaneously (videotaped). Brief
  7216. pulse trains (0.5 s) lateral bending movements of the body
  7217. (orientation movements) were added, and with even longer stimuli
  7218. locomotor movements were initiated. Depending on the tectal area
  7219. stimulated, four characteristic response patterns were observed.
  7220. In a lateral area conjugate horizontal eye movements combined
  7221. with lateral bending movements of the body and locomotor
  7222. movements were elicited, depending on stimulus duration. The
  7223. amplitude of the eye movement and bending movements was site
  7224. specific within this region. In a rostromedial area, bilateral
  7225. downward vertical eye movements occurred. In a caudomedial tectal
  7226. area, large-amplitude undulatory body movements akin to
  7227. struggling behavior were elicited, combined with large-amplitude
  7228. eye movements that were antiphasic to the body movements. The
  7229. alternating eye movements were not dependent on vestibuloocular
  7230. reflexes. Finally, in a caudolateral area locomotor movements
  7231. without eye or bending movements could be elicited. These results
  7232. show that tectum can provide integrated motor responses of eye,
  7233. body orientation, and locomotion of the type that would be
  7234. required in goal-directed locomotion.",
  7235. journal = "J. Neurophysiol.",
  7236. volume = 97,
  7237. number = 4,
  7238. pages = "3093--3108",
  7239. month = apr,
  7240. year = 2007,
  7241. language = "en"
  7242. }
  7243. @ARTICLE{Sussillo2013-ey,
  7244. title = "Opening the black box: low-dimensional dynamics in
  7245. high-dimensional recurrent neural networks",
  7246. author = "Sussillo, David and Barak, Omri",
  7247. abstract = "Recurrent neural networks (RNNs) are useful tools for learning
  7248. nonlinear relationships between time-varying inputs and outputs
  7249. with complex temporal dependencies. Recently developed
  7250. algorithms have been successful at training RNNs to perform a
  7251. wide variety of tasks, but the resulting networks have been
  7252. treated as black boxes: their mechanism of operation remains
  7253. unknown. Here we explore the hypothesis that fixed points, both
  7254. stable and unstable, and the linearized dynamics around them,
  7255. can reveal crucial aspects of how RNNs implement their
  7256. computations. Further, we explore the utility of linearization
  7257. in areas of phase space that are not true fixed points but
  7258. merely points of very slow movement. We present a simple
  7259. optimization technique that is applied to trained RNNs to find
  7260. the fixed and slow points of their dynamics. Linearization
  7261. around these slow regions can be used to explore, or
  7262. reverse-engineer, the behavior of the RNN. We describe the
  7263. technique, illustrate it using simple examples, and finally
  7264. showcase it on three high-dimensional RNN examples: a 3-bit
  7265. flip-flop device, an input-dependent sine wave generator, and a
  7266. two-point moving average. In all cases, the mechanisms of
  7267. trained networks could be inferred from the sets of fixed and
  7268. slow points and the linearized dynamics around them.",
  7269. journal = "Neural Comput.",
  7270. publisher = "MIT Press",
  7271. volume = 25,
  7272. number = 3,
  7273. pages = "626--649",
  7274. month = mar,
  7275. year = 2013,
  7276. keywords = "RNN",
  7277. language = "en"
  7278. }
  7279. @ARTICLE{Werbos1990-qo,
  7280. title = "Backpropagation through time: what it does and how to do it",
  7281. author = "Werbos, P J",
  7282. abstract = "Basic backpropagation, which is a simple method now being widely
  7283. used in areas like pattern recognition and fault diagnosis, is
  7284. reviewed. The basic equations for backpropagation through time,
  7285. and applications to areas like pattern recognition involving
  7286. dynamic systems, systems identification, and control are
  7287. discussed. Further extensions of this method, to deal with
  7288. systems other than neural networks, systems involving
  7289. simultaneous equations, or true recurrent networks, and other
  7290. practical issues arising with the method are described.
  7291. Pseudocode is provided to clarify the algorithms. The chain rule
  7292. for ordered derivatives-the theorem which underlies
  7293. backpropagation-is briefly discussed. The focus is on designing a
  7294. simpler version of backpropagation which can be translated into
  7295. computer code and applied directly by neutral network users.",
  7296. journal = "Proc. IEEE",
  7297. volume = 78,
  7298. number = 10,
  7299. pages = "1550--1560",
  7300. month = oct,
  7301. year = 1990,
  7302. keywords = "identification;neural nets;pattern recognition;pseudocode;pattern
  7303. recognition;fault diagnosis;backpropagation;systems
  7304. identification;neural networks;Backpropagation;Artificial neural
  7305. networks;Supervised learning;Pattern recognition;Neural
  7306. networks;Power system modeling;Equations;Control systems;Fluid
  7307. dynamics;Books"
  7308. }
  7309. @ARTICLE{Mante2013-yl,
  7310. title = "Context-dependent computation by recurrent dynamics in prefrontal
  7311. cortex",
  7312. author = "Mante, Valerio and Sussillo, David and Shenoy, Krishna V and
  7313. Newsome, William T",
  7314. abstract = "Prefrontal cortex is thought to have a fundamental role in
  7315. flexible, context-dependent behaviour, but the exact nature of
  7316. the computations underlying this role remains largely unknown. In
  7317. particular, individual prefrontal neurons often generate
  7318. remarkably complex responses that defy deep understanding of
  7319. their contribution to behaviour. Here we study prefrontal cortex
  7320. activity in macaque monkeys trained to flexibly select and
  7321. integrate noisy sensory inputs towards a choice. We find that the
  7322. observed complexity and functional roles of single neurons are
  7323. readily understood in the framework of a dynamical process
  7324. unfolding at the level of the population. The population dynamics
  7325. can be reproduced by a trained recurrent neural network, which
  7326. suggests a previously unknown mechanism for selection and
  7327. integration of task-relevant inputs. This mechanism indicates
  7328. that selection and integration are two aspects of a single
  7329. dynamical process unfolding within the same prefrontal circuits,
  7330. and potentially provides a novel, general framework for
  7331. understanding context-dependent computations.",
  7332. journal = "Nature",
  7333. volume = 503,
  7334. number = 7474,
  7335. pages = "78--84",
  7336. month = nov,
  7337. year = 2013,
  7338. language = "en"
  7339. }
  7340. @ARTICLE{Athalye2019-gr,
  7341. title = "Neural reinforcement: re-entering and refining neural dynamics
  7342. leading to desirable outcomes",
  7343. author = "Athalye, Vivek R and Carmena, Jose M and Costa, Rui M",
  7344. abstract = "How do organisms learn to do again, on-demand, a behavior that
  7345. led to a desirable outcome? Dopamine-dependent cortico-striatal
  7346. plasticity provides a framework for learning behavior's value,
  7347. but it is less clear how it enables the brain to re-enter desired
  7348. behaviors and refine them over time. Reinforcing behavior is
  7349. achieved by re-entering and refining the neural patterns that
  7350. produce it. We review studies using brain-machine interfaces
  7351. which reveal that reinforcing cortical population activity
  7352. requires cortico-basal ganglia circuits. Then, we propose a
  7353. formal framework for how reinforcement in cortico-basal ganglia
  7354. circuits acts on the neural dynamics of cortical populations. We
  7355. propose two parallel mechanisms: i) fast reinforcement which
  7356. selects the inputs that permit the re-entrance of the particular
  7357. cortical population dynamics which naturally produced the desired
  7358. behavior, and ii) slower reinforcement which leads to refinement
  7359. of cortical population dynamics and more reliable production of
  7360. neural trajectories driving skillful behavior on-demand.",
  7361. journal = "Curr. Opin. Neurobiol.",
  7362. volume = 60,
  7363. pages = "145--154",
  7364. month = dec,
  7365. year = 2019,
  7366. language = "en"
  7367. }
  7368. @ARTICLE{Sussillo2014-mo,
  7369. title = "Neural circuits as computational dynamical systems",
  7370. author = "Sussillo, David",
  7371. abstract = "Many recent studies of neurons recorded from cortex reveal
  7372. complex temporal dynamics. How such dynamics embody the
  7373. computations that ultimately lead to behavior remains a mystery.
  7374. Approaching this issue requires developing plausible hypotheses
  7375. couched in terms of neural dynamics. A tool ideally suited to aid
  7376. in this question is the recurrent neural network (RNN). RNNs
  7377. straddle the fields of nonlinear dynamical systems and machine
  7378. learning and have recently seen great advances in both theory and
  7379. application. I summarize recent theoretical and technological
  7380. advances and highlight an example of how RNNs helped to explain
  7381. perplexing high-dimensional neurophysiological data in the
  7382. prefrontal cortex.",
  7383. journal = "Curr. Opin. Neurobiol.",
  7384. volume = 25,
  7385. pages = "156--163",
  7386. month = apr,
  7387. year = 2014,
  7388. keywords = "RNN To read;RNN",
  7389. language = "en"
  7390. }
  7391. @ARTICLE{DePasquale2018-on,
  7392. title = "{full-FORCE}: A target-based method for training recurrent
  7393. networks",
  7394. author = "DePasquale, Brian and Cueva, Christopher J and Rajan, Kanaka and
  7395. Escola, G Sean and Abbott, L F",
  7396. abstract = "Trained recurrent networks are powerful tools for modeling
  7397. dynamic neural computations. We present a target-based method for
  7398. modifying the full connectivity matrix of a recurrent network to
  7399. train it to perform tasks involving temporally complex
  7400. input/output transformations. The method introduces a second
  7401. network during training to provide suitable ``target'' dynamics
  7402. useful for performing the task. Because it exploits the full
  7403. recurrent connectivity, the method produces networks that perform
  7404. tasks with fewer neurons and greater noise robustness than
  7405. traditional least-squares (FORCE) approaches. In addition, we
  7406. show how introducing additional input signals into the
  7407. target-generating network, which act as task hints, greatly
  7408. extends the range of tasks that can be learned and provides
  7409. control over the complexity and nature of the dynamics of the
  7410. trained, task-performing network.",
  7411. journal = "PLoS One",
  7412. volume = 13,
  7413. number = 2,
  7414. pages = "e0191527",
  7415. month = feb,
  7416. year = 2018,
  7417. language = "en"
  7418. }
  7419. @ARTICLE{Sussillo2009-tf,
  7420. title = "Generating coherent patterns of activity from chaotic neural
  7421. networks",
  7422. author = "Sussillo, David and Abbott, L F",
  7423. abstract = "Neural circuits display complex activity patterns both
  7424. spontaneously and when responding to a stimulus or generating a
  7425. motor output. How are these two forms of activity related? We
  7426. develop a procedure called FORCE learning for modifying synaptic
  7427. strengths either external to or within a model neural network to
  7428. change chaotic spontaneous activity into a wide variety of
  7429. desired activity patterns. FORCE learning works even though the
  7430. networks we train are spontaneously chaotic and we leave feedback
  7431. loops intact and unclamped during learning. Using this approach,
  7432. we construct networks that produce a wide variety of complex
  7433. output patterns, input-output transformations that require
  7434. memory, multiple outputs that can be switched by control inputs,
  7435. and motor patterns matching human motion capture data. Our
  7436. results reproduce data on premovement activity in motor and
  7437. premotor cortex, and suggest that synaptic plasticity may be a
  7438. more rapid and powerful modulator of network activity than
  7439. generally appreciated.",
  7440. journal = "Neuron",
  7441. volume = 63,
  7442. number = 4,
  7443. pages = "544--557",
  7444. month = aug,
  7445. year = 2009,
  7446. language = "en"
  7447. }
  7448. @ARTICLE{Helmbrecht2018-ux,
  7449. title = "Topography of a Visuomotor Transformation",
  7450. author = "Helmbrecht, Thomas O and Dal Maschio, Marco and Donovan, Joseph C
  7451. and Koutsouli, Styliani and Baier, Herwig",
  7452. abstract = "The brain converts perceptual information into appropriate
  7453. patterns of muscle activity depending on the categorization and
  7454. localization of sensory cues. Sensorimotor information might
  7455. either be encoded by distributed networks or by ``labeled lines''
  7456. connecting sensory channels to dedicated behavioral pathways.
  7457. Here we investigate, in the context of natural behavior, how the
  7458. tectum of larval zebrafish can inform downstream premotor areas.
  7459. Optogenetic mapping revealed a tectal motor map underlying
  7460. locomotor maneuvers for escape and approach. Single-cell
  7461. reconstructions and high-resolution functional imaging showed
  7462. that two spatially segregated and uncrossed descending axon
  7463. tracts selectively transmit approach and escape signals to the
  7464. hindbrain. Moreover, the approach pathway conveys information
  7465. about retinotopic target coordinates to specific premotor
  7466. ensembles via spatially ordered axonal projections. This
  7467. topographic organization supports a tectum-generated space code
  7468. sufficient to steer orienting movements. We conclude that
  7469. specific labeled lines guide object-directed behavior in the
  7470. larval zebrafish brain.",
  7471. journal = "Neuron",
  7472. volume = 100,
  7473. number = 6,
  7474. pages = "1429--1445.e4",
  7475. month = dec,
  7476. year = 2018,
  7477. keywords = "hindbrain; motor map; optic tectum; optogenetics; reticular
  7478. formation; space code; superior colliculus; tectal projectome;
  7479. visuomotor transformation; zebrafish",
  7480. language = "en"
  7481. }
  7482. @ARTICLE{Kardamakis2015-yc,
  7483. title = "Tectal microcircuit generating visual selection commands on
  7484. gaze-controlling neurons",
  7485. author = "Kardamakis, Andreas A and Saitoh, Kazuya and Grillner, Sten",
  7486. abstract = "The optic tectum (called superior colliculus in mammals) is
  7487. critical for eye-head gaze shifts as we navigate in the terrain
  7488. and need to adapt our movements to the visual scene. The neuronal
  7489. mechanisms underlying the tectal contribution to stimulus
  7490. selection and gaze reorientation remains, however, unclear at the
  7491. microcircuit level. To analyze this complex--yet phylogenetically
  7492. conserved--sensorimotor system, we developed a novel in vitro
  7493. preparation in the lamprey that maintains the eye and midbrain
  7494. intact and allows for whole-cell recordings from prelabeled
  7495. tectal gaze-controlling cells in the deep layer, while visual
  7496. stimuli are delivered. We found that receptive field activation
  7497. of these cells provide monosynaptic retinal excitation followed
  7498. by local GABAergic inhibition (feedforward). The entire remaining
  7499. retina, on the other hand, elicits only inhibition (surround
  7500. inhibition). If two stimuli are delivered simultaneously, one
  7501. inside and one outside the receptive field, the former excitatory
  7502. response is suppressed. When local inhibition is
  7503. pharmacologically blocked, the suppression induced by competing
  7504. stimuli is canceled. We suggest that this rivalry between visual
  7505. areas across the tectal map is triggered through long-range
  7506. inhibitory tectal connections. Selection commands conveyed via
  7507. gaze-controlling neurons in the optic tectum are, thus, formed
  7508. through synaptic integration of local retinotopic excitation and
  7509. global tectal inhibition. We anticipate that this mechanism not
  7510. only exists in lamprey but is also conserved throughout
  7511. vertebrate evolution.",
  7512. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  7513. volume = 112,
  7514. number = 15,
  7515. pages = "E1956--65",
  7516. month = apr,
  7517. year = 2015,
  7518. keywords = "GABAergic inhibition; evolution; gaze control; optic tectum;
  7519. superior colliculus",
  7520. language = "en"
  7521. }
  7522. @ARTICLE{Weldon1983-oz,
  7523. title = "Rotational behavior following cholinergic stimulation of the
  7524. superior colliculus in rats",
  7525. author = "Weldon, D A and Calabrese, L C and Nicklaus, K J",
  7526. abstract = "Rate which received microinjections of carbachol into the
  7527. superior colliculus exhibited pronounced dose-dependent
  7528. rotational behavior contralateral to the site of injection
  7529. (Experiment 1). Wet dog shakes were also observed in some
  7530. animals. Similar injections in the midbrain reticular formation
  7531. produced immobility with slight contralateral flexion of the
  7532. neck. Convulsions were observed in some rats after injections
  7533. into either anatomical location. In Experiment 2, circling
  7534. induced by carbachol in the superior colliculus was blocked by
  7535. prior injection of either the muscarinic receptor antagonist
  7536. scopolamine or the nicotinic receptor antagonist mecamylamine,
  7537. suggesting that both nicotinic and muscarinic receptors are
  7538. involved in the effect. In Experiment 3 contralateral rotational
  7539. behavior was induced by intracollicular microinjections of the
  7540. combination of acetylcholine chloride and physostigmine. The
  7541. results suggest that collicular mediation of contralateral
  7542. rotational behavior, and perhaps orientation, might involve
  7543. cholinergic receptors.",
  7544. journal = "Pharmacol. Biochem. Behav.",
  7545. publisher = "Elsevier",
  7546. volume = 19,
  7547. number = 5,
  7548. pages = "813--820",
  7549. month = nov,
  7550. year = 1983,
  7551. language = "en"
  7552. }
  7553. @ARTICLE{Geula1984-sq,
  7554. title = "Circling and bodily asymmetry induced by injection of {GABA}
  7555. agonists and antagonists into the superior colliculus",
  7556. author = "Geula, C and Asdourian, D",
  7557. abstract = "The observation that ipsiversive circling follows unilateral
  7558. lesions of the deep layers of the superior colliculus (DLSC),
  7559. combined with the recent demonstration of an ipsilateral
  7560. inhibitory GABAergic projection from substantia nigra pars
  7561. reticulata (SNr) to the DLSC suggests a role for tectal GABA in
  7562. circling behavior. In the present experiment, GABA, the GABA
  7563. agonist muscimol, and the GABA antagonists picrotoxin and
  7564. bicuculline were injected into the DLSC through chronic
  7565. cannulae. GABA and muscimol produced significantly higher
  7566. ipsiversive circling and bodily asymmetry than saline
  7567. injections. Picrotoxin and bicuculline resulted in significantly
  7568. higher contraversive circling and asymmetry than saline
  7569. injections. All drugs except bicuculline produced dose-dependent
  7570. circling. GABA injections were also made into the mesencephalic
  7571. reticular formation (MRF) and the periaqueductal gray (PAG). The
  7572. MRF injections produced the same degree of circling and
  7573. asymmetry as the DLSC injections. The PAG injections resulted in
  7574. significantly lower amounts of circling than the DLSC GABA
  7575. injections, but they resulted in equivalent measures of
  7576. asymmetry. These results demonstrate that DLSC GABA produces
  7577. circling and asymmetry, and suggest that the DLSC as well as the
  7578. MRF serve as output stations for the expression of circling
  7579. behavior initiated at the striatum.",
  7580. journal = "Pharmacol. Biochem. Behav.",
  7581. publisher = "Elsevier",
  7582. volume = 21,
  7583. number = 6,
  7584. pages = "853--858",
  7585. month = dec,
  7586. year = 1984,
  7587. language = "en"
  7588. }
  7589. @ARTICLE{Ito2017-gi,
  7590. title = "Segregation of Visual Response Properties in the Mouse Superior
  7591. Colliculus and Their Modulation during Locomotion",
  7592. author = "Ito, Shinya and Feldheim, David A and Litke, Alan M",
  7593. abstract = "The superior colliculus (SC) receives direct input from the
  7594. retina and integrates it with information about sound, touch,
  7595. and state of the animal that is relayed from other parts of the
  7596. brain to initiate specific behavioral outcomes. The superficial
  7597. SC layers (sSC) contain cells that respond to visual stimuli,
  7598. whereas the deep SC layers (dSC) contain cells that also respond
  7599. to auditory and somatosensory stimuli. Here, we used a
  7600. large-scale silicon probe recording system to examine the visual
  7601. response properties of SC cells of head-fixed and alert male
  7602. mice. We found cells with diverse response properties including:
  7603. (1) orientation/direction-selective (OS/DS) cells with a firing
  7604. rate that is suppressed by drifting sinusoidal gratings
  7605. (negative OS/DS cells); (2) suppressed-by-contrast cells; (3)
  7606. cells with complex-like spatial summation nonlinearity; and (4)
  7607. cells with Y-like spatial summation nonlinearity. We also found
  7608. specific response properties that are enriched in different
  7609. depths of the SC. The sSC is enriched with cells with small RFs,
  7610. high evoked firing rates (FRs), and sustained temporal
  7611. responses, whereas the dSC is enriched with the negative OS/DS
  7612. cells and with cells with large RFs, low evoked FRs, and
  7613. transient temporal responses. Locomotion modulates the activity
  7614. of the SC cells both additively and multiplicatively and changes
  7615. the preferred spatial frequency of some SC cells. These results
  7616. provide the first description of the negative OS/DS cells and
  7617. demonstrate that the SC segregates cells with different response
  7618. properties and that the behavioral state of a mouse affects SC
  7619. activity.SIGNIFICANCE STATEMENT The superior colliculus (SC)
  7620. receives visual input from the retina in its superficial layers
  7621. (sSC) and induces eye/head-orientating movements and innate
  7622. defensive responses in its deeper layers (dSC). Despite their
  7623. importance, very little is known about the visual response
  7624. properties of dSC neurons. Using high-density electrode
  7625. recordings and novel model-based analysis, we found several
  7626. novel visual response properties of the SC cells, including
  7627. encoding of a cell's preferred orientation or direction by
  7628. suppression of the firing rate. The sSC and the dSC are enriched
  7629. with cells with different visual response properties. Locomotion
  7630. modulates the cells in the SC. These findings contribute to our
  7631. understanding of how the SC processes visual inputs, a critical
  7632. step in comprehending visually guided behaviors.",
  7633. journal = "J. Neurosci.",
  7634. publisher = "Soc Neuroscience",
  7635. volume = 37,
  7636. number = 35,
  7637. pages = "8428--8443",
  7638. month = aug,
  7639. year = 2017,
  7640. keywords = "mouse; silicon probe; superior colliculus; vision",
  7641. language = "en"
  7642. }
  7643. @ARTICLE{Bretzner2013-si,
  7644. title = "{Lhx3-Chx10} reticulospinal neurons in locomotor circuits",
  7645. author = "Bretzner, Fr{\'e}d{\'e}ric and Brownstone, Robert M",
  7646. abstract = "Motor behaviors result from the interplay between the brain and
  7647. the spinal cord. Reticulospinal neurons, situated between the
  7648. supraspinal structures that initiate motor movements and the
  7649. spinal cord that executes them, play key integrative roles in
  7650. these behaviors. However, the molecular identities of mammalian
  7651. reticular formation neurons that mediate motor behaviors have
  7652. not yet been determined, thus limiting their study in health and
  7653. disease. In the medullary reticular formation of the mouse, we
  7654. identified neurons that express the transcription factors Lhx3
  7655. and/or Chx10, and demonstrate that these neurons form a
  7656. significant component of glutamatergic reticulospinal pathways.
  7657. Lhx3-positive medullary reticular formation neurons express Fos
  7658. following a locomotor task in the adult, indicating that they
  7659. are active during walking. Furthermore, they receive functional
  7660. inputs from the mesencephalic locomotor region and have
  7661. electrophysiological properties to support tonic repetitive
  7662. firing, both of which are necessary for neurons that mediate the
  7663. descending command for locomotion. Together, these results
  7664. suggest that Lhx3/Chx10 medullary reticular formation neurons
  7665. are involved in locomotion.",
  7666. journal = "J. Neurosci.",
  7667. publisher = "Soc Neuroscience",
  7668. volume = 33,
  7669. number = 37,
  7670. pages = "14681--14692",
  7671. month = sep,
  7672. year = 2013,
  7673. keywords = "Locomotion",
  7674. language = "en"
  7675. }
  7676. @ARTICLE{Grillner2020-uq,
  7677. title = "Current Principles of Motor Control, with Special Reference to
  7678. Vertebrate Locomotion",
  7679. author = "Grillner, Sten and El Manira, Abdeljabbar",
  7680. abstract = "The vertebrate control of locomotion involves all levels of the
  7681. nervous system from cortex to the spinal cord. Here, we aim to
  7682. cover all main aspects of this complex behavior, from the
  7683. operation of the microcircuits in the spinal cord to the systems
  7684. and behavioral levels and extend from mammalian locomotion to
  7685. the basic undulatory movements of lamprey and fish. The cellular
  7686. basis of propulsion represents the core of the control system,
  7687. and it involves the spinal central pattern generator networks
  7688. (CPGs) controlling the timing of different muscles, the sensory
  7689. compensation for perturbations, and the brain stem command
  7690. systems controlling the level of activity of the CPGs and the
  7691. speed of locomotion. The forebrain and in particular the basal
  7692. ganglia are involved in determining which motor programs should
  7693. be recruited at a given point of time and can both initiate and
  7694. stop locomotor activity. The propulsive control system needs to
  7695. be integrated with the postural control system to maintain body
  7696. orientation. Moreover, the locomotor movements need to be
  7697. steered so that the subject approaches the goal of the locomotor
  7698. episode, or avoids colliding with elements in the environment or
  7699. simply escapes at high speed. These different aspects will all
  7700. be covered in the review.",
  7701. journal = "Physiol. Rev.",
  7702. publisher = "physiology.org",
  7703. volume = 100,
  7704. number = 1,
  7705. pages = "271--320",
  7706. month = jan,
  7707. year = 2020,
  7708. keywords = "basal ganglia; central pattern generators; cerebellum; spinal
  7709. cord; vestibular; visuomotor",
  7710. language = "en"
  7711. }
  7712. % The entry below contains non-ASCII chars that could not be converted
  7713. % to a LaTeX equivalent.
  7714. @ARTICLE{Kleinman2019-ks,
  7715. title = "Recurrent neural network models of multi-area computation
  7716. underlying decision-making",
  7717. author = "Kleinman, M and Chandrasekaran, C and Kao, J C",
  7718. abstract = "Cognition emerges from the coordination of computations in
  7719. multiple brain areas. However, elucidating these coordinated
  7720. computations within and across brain regions is challenging
  7721. because intra-and inter-areal connectivity are typically
  7722. unknown. Testable hypotheses about …",
  7723. journal = "bioRxiv",
  7724. publisher = "biorxiv.org",
  7725. year = 2019
  7726. }
  7727. @UNPUBLISHED{Shamash2018-ky,
  7728. title = "A tool for analyzing electrode tracks from slice histology",
  7729. author = "Shamash, Philip and Carandini, Matteo and Harris, Kenneth and
  7730. Steinmetz, Nick",
  7731. abstract = "It is now possible to record from hundreds of neurons across
  7732. multiple brain regions in a single electrophysiology experiment.
  7733. An essential step in the ensuing data analysis is to assign
  7734. recorded neurons to the correct brain regions. Brain regions are
  7735. typically identified after the recordings by comparing images of
  7736. brain slices to a reference atlas by eye. This introduces error,
  7737. in particular when slices are not cut at a perfectly coronal
  7738. angle or when electrode tracks span multiple slices. Here we
  7739. introduce SHARP-Track, a tool to localize regions of interest and
  7740. plot the brain regions they pass through. SHARP-Track offers a
  7741. MATLAB user interface to explore the Allen Mouse Brain Atlas,
  7742. register asymmetric slice images to the atlas using manual input,
  7743. and interactively analyze electrode tracks. We find that it
  7744. reduces error compared to localizing electrodes in a reference
  7745. atlas by eye. See [github.com/cortex-lab/allenCCF][1] for the
  7746. software and wiki. [1]: http://github.com/cortex-lab/allenCCF",
  7747. journal = "bioRxiv",
  7748. pages = "447995",
  7749. month = oct,
  7750. year = 2018,
  7751. language = "en"
  7752. }
  7753. @ARTICLE{Savage2017-un,
  7754. title = "Segregated fronto-cortical and midbrain connections in the mouse
  7755. and their relation to approach and avoidance orienting behaviors",
  7756. author = "Savage, Michael Anthony and McQuade, Richard and Thiele,
  7757. Alexander",
  7758. abstract = "The orchestration of orienting behaviors requires the interaction
  7759. of many cortical and subcortical areas, for example the superior
  7760. colliculus (SC), as well as prefrontal areas responsible for
  7761. top-down control. Orienting involves different behaviors, such as
  7762. approach and avoidance. In the rat, these behaviors are at least
  7763. partially mapped onto different SC subdomains, the lateral (SCl)
  7764. and medial (SCm), respectively. To delineate the circuitry
  7765. involved in the two types of orienting behavior in mice, we
  7766. injected retrograde tracer into the intermediate and deep layers
  7767. of the SCm and SCl, and thereby determined the main input
  7768. structures to these subdomains. Overall the SCm receives larger
  7769. numbers of afferents compared to the SCl. The prefrontal
  7770. cingulate area (Cg), visual, oculomotor, and auditory areas
  7771. provide strong input to the SCm, while prefrontal motor area 2
  7772. (M2), and somatosensory areas provide strong input to the SCl.
  7773. The prefrontal areas Cg and M2 in turn connect to different
  7774. cortical and subcortical areas, as determined by anterograde
  7775. tract tracing. Even though connectivity pattern often overlap,
  7776. our labeling approaches identified segregated neural circuits
  7777. involving SCm, Cg, secondary visual cortices, auditory areas, and
  7778. the dysgranular retrospenial cortex likely to be involved in
  7779. avoidance behaviors. Conversely, SCl, M2, somatosensory cortex,
  7780. and the granular retrospenial cortex comprise a network likely
  7781. involved in approach/appetitive behaviors.",
  7782. journal = "J. Comp. Neurol.",
  7783. volume = 525,
  7784. number = 8,
  7785. pages = "1980--1999",
  7786. month = jun,
  7787. year = 2017,
  7788. keywords = "RRID:SCR\_013672; approach behaviors; avoidance behaviors;
  7789. cingulate area; motor cortex area 2; superior colliculus",
  7790. language = "en"
  7791. }
  7792. @ARTICLE{Lomber2001-zl,
  7793. title = "Role of the superior colliculus in analyses of space: superficial
  7794. and intermediate layer contributions to visual orienting,
  7795. auditory orienting, and visuospatial discriminations during
  7796. unilateral and bilateral deactivations",
  7797. author = "Lomber, S G and Payne, B R and Cornwell, P",
  7798. abstract = "The superior colliculus (SC) has been implicated in spatial
  7799. analyses of the environment, although few behavioral studies have
  7800. explicitly tested this role. To test its imputed role in spatial
  7801. analyses, we used a battery of four spatial tasks combined with
  7802. unilateral and bilateral cooling deactivation of the upper and
  7803. intermediate layers of the superior colliculus. We tested the
  7804. abilities of cats to orient to three different stimuli: (1)
  7805. moving visual, (2) stationary visual, (3) stationary white-noise
  7806. aural. Furthermore, we tested the ability of the cats to
  7807. discriminate the relative spatial position of a landmark.
  7808. Unilateral cooling deactivation of the superficial layers of the
  7809. SC induced a profound neglect of both moving and stationary
  7810. visual stimuli presented in, and landmark objects located within,
  7811. the contralateral hemifield. However, responses to auditory
  7812. stimuli were unimpaired. Unilateral cooling deactivation of both
  7813. the superficial and intermediate layers induced a profound
  7814. contralateral neglect of the auditory stimulus. Additional and
  7815. equivalent deactivation of the opposite SC largely restored
  7816. orienting to either moving visual or auditory stimuli, and
  7817. restored landmark position reporting to normal levels. However,
  7818. during bilateral SC deactivation, orienting to the static visual
  7819. stimulus was abolished throughout the entire visual field.
  7820. Overall, unilateral SC deactivation results show that the upper
  7821. and intermediate layers of the SC contribute in different ways to
  7822. guiding behavioral responses to visual and auditory stimuli cues.
  7823. Finally, bilateral superior colliculus deactivations reveal that
  7824. other structures are sufficient to support spatial analyses and
  7825. guide visual behaviors in the absence of neural operations in the
  7826. superior colliculus, but only under certain circumstances.",
  7827. journal = "J. Comp. Neurol.",
  7828. volume = 441,
  7829. number = 1,
  7830. pages = "44--57",
  7831. month = dec,
  7832. year = 2001,
  7833. language = "en"
  7834. }
  7835. @ARTICLE{Angelaki2019-ky,
  7836. title = "The head direction cell network: attractor dynamics, integration
  7837. within the navigation system, and three-dimensional properties",
  7838. author = "Angelaki, Dora E and Laurens, Jean",
  7839. abstract = "Knowledge of head direction cell function has progressed
  7840. remarkably in recent years. The predominant theory that they form
  7841. an attractor has been confirmed by several experiments. Candidate
  7842. pathways that may convey visual input have been identified. The
  7843. pre-subicular circuitry that conveys head direction signals to
  7844. the medial entorhinal cortex, potentially sustaining path
  7845. integration by grid cells, has been resolved. Although the
  7846. neuronal substrate of the attractor remains unknown in mammals, a
  7847. simple head direction network, whose structure is astoundingly
  7848. similar to neuronal models theorized decades earlier, has been
  7849. identified in insects. Finally, recent experiments have revealed
  7850. that these cells do not encode head direction in the horizontal
  7851. plane only, but also in vertical planes, thus providing a 3D
  7852. orientation signal.",
  7853. journal = "Curr. Opin. Neurobiol.",
  7854. volume = 60,
  7855. pages = "136--144",
  7856. month = dec,
  7857. year = 2019,
  7858. language = "en"
  7859. }
  7860. @ARTICLE{Sauerbrei2019-uc,
  7861. title = "Cortical pattern generation during dexterous movement is
  7862. input-driven",
  7863. author = "Sauerbrei, Britton A and Guo, Jian-Zhong and Cohen, Jeremy D and
  7864. Mischiati, Matteo and Guo, Wendy and Kabra, Mayank and Verma,
  7865. Nakul and Mensh, Brett and Branson, Kristin and Hantman, Adam W",
  7866. abstract = "The motor cortex controls skilled arm movement by sending
  7867. temporal patterns of activity to lower motor centres1. Local
  7868. cortical dynamics are thought to shape these patterns throughout
  7869. movement execution2-4. External inputs have been implicated in
  7870. setting the initial state of the motor cortex5,6, but they may
  7871. also have a pattern-generating role. Here we dissect the
  7872. contribution of local dynamics and inputs to cortical pattern
  7873. generation during a prehension task in mice. Perturbing cortex to
  7874. an aberrant state prevented movement initiation, but after the
  7875. perturbation was released, cortex either bypassed the normal
  7876. initial state and immediately generated the pattern that controls
  7877. reaching or failed to generate this pattern. The difference in
  7878. these two outcomes was probably a result of external inputs. We
  7879. directly investigated the role of inputs by inactivating the
  7880. thalamus; this perturbed cortical activity and disrupted limb
  7881. kinematics at any stage of the movement. Activation of
  7882. thalamocortical axon terminals at different frequencies disrupted
  7883. cortical activity and arm movement in a graded manner.
  7884. Simultaneous recordings revealed that both thalamic activity and
  7885. the current state of cortex predicted changes in cortical
  7886. activity. Thus, the pattern generator for dexterous arm movement
  7887. is distributed across multiple, strongly interacting brain
  7888. regions.",
  7889. journal = "Nature",
  7890. month = dec,
  7891. year = 2019,
  7892. language = "en"
  7893. }
  7894. @UNPUBLISHED{Sabatini2019-zw,
  7895. title = "The impact of reporter kinetics on the interpretation of data
  7896. gathered with fluorescent reporters",
  7897. author = "Sabatini, Bernardo L",
  7898. abstract = "Abstract Fluorescent reporters of biological functions are used
  7899. to monitor biochemical events and signals in cells and tissue.
  7900. For neurobiology, these have been particularly useful for
  7901. monitoring signals in the brains of behaving animals. In order to
  7902. enhance signal-to-noise, fluorescent reporters typically have
  7903. kinetics that are slower than that of the underlying biological
  7904. process. This low-pass filtering by the reporter renders the
  7905. fluorescence transient a leaking integrated version of the
  7906. biological signal. Here I discuss the effects that low-pass
  7907. filtering, or more precisely of integrating by convolving with an
  7908. exponentially decaying kernel, has on the interpretation of the
  7909. relationship between the reporter fluorescence transient and the
  7910. events that underlie it. Unfortunately, when the biological
  7911. events being monitored are impulse-like, such as the firing of an
  7912. action potential or the release of neurotransmitter, filtering
  7913. greatly reduces the maximum correlation coefficient that can be
  7914. found between the events and the fluorescence signal. This can
  7915. erroneously support the conclusion that the fluorescence
  7916. transient and the biological signal that it reports are only
  7917. weakly related. Furthermore, when examining the encoding of
  7918. behavioral state variables by nervous system, filtering by the
  7919. reporter kinetics will favor the interpretation that fluorescence
  7920. transients encode integrals of measured variables as opposed to
  7921. the variables themselves. For these reasons, it is necessary to
  7922. take into account the filtering effects of the indicator by
  7923. deconvolving with the convolution kernel and recovering the
  7924. underlying biological events before making conclusions about what
  7925. is encoded in the signals emitted by fluorescent reporters.",
  7926. journal = "bioRxiv",
  7927. pages = "834895",
  7928. month = nov,
  7929. year = 2019,
  7930. language = "en"
  7931. }
  7932. @ARTICLE{Reinhard2019-zc,
  7933. title = "A projection specific logic to sampling visual inputs in mouse
  7934. superior colliculus",
  7935. author = "Reinhard, Katja and Li, Chen and Do, Quan and Burke, Emily G and
  7936. Heynderickx, Steven and Farrow, Karl",
  7937. abstract = "Using sensory information to trigger different behaviors relies
  7938. on circuits that pass through brain regions. The rules by which
  7939. parallel inputs are routed to downstream targets are poorly
  7940. understood. The superior colliculus mediates a set of innate
  7941. behaviors, receiving input from >30 retinal ganglion cell types
  7942. and projecting to behaviorally important targets including the
  7943. pulvinar and parabigeminal nucleus. Combining transsynaptic
  7944. circuit tracing with in vivo and ex vivo electrophysiological
  7945. recordings, we observed a projection-specific logic where each
  7946. collicular output pathway sampled a distinct set of retinal
  7947. inputs. Neurons projecting to the pulvinar or the parabigeminal
  7948. nucleus showed strongly biased sampling from four cell types
  7949. each, while six others innervated both pathways. The visual
  7950. response properties of retinal ganglion cells correlated well
  7951. with those of their disynaptic targets. These findings open the
  7952. possibility that projection-specific sampling of retinal inputs
  7953. forms a basis for the selective triggering of behaviors by the
  7954. superior colliculus.",
  7955. journal = "Elife",
  7956. volume = 8,
  7957. month = nov,
  7958. year = 2019,
  7959. keywords = "mouse; neuroscience; parabigeminal nucleus; pulvinar; retina;
  7960. superior colliculus; visual circuits",
  7961. language = "en"
  7962. }
  7963. @UNPUBLISHED{Kietzmann2018-ou,
  7964. title = "Deep Neural Networks in Computational Neuroscience",
  7965. author = "Kietzmann, Tim C and McClure, Patrick and Kriegeskorte, Nikolaus",
  7966. abstract = "Summary The goal of computational neuroscience is to find
  7967. mechanistic explanations of how the nervous system processes
  7968. information to give rise to cognitive function and behaviour. At
  7969. the heart of the field are its models, i.e. mathematical and
  7970. computational descriptions of the system being studied, which map
  7971. sensory stimuli to neural responses and/or neural to behavioural
  7972. responses. These models range from simple to complex. Recently,
  7973. deep neural networks (DNNs) have come to dominate several domains
  7974. of artificial intelligence (AI). As the term ``neural network''
  7975. suggests, these models are inspired by biological brains.
  7976. However, current DNNs neglect many details of biological neural
  7977. networks. These simplifications contribute to their computational
  7978. efficiency, enabling them to perform complex feats of
  7979. intelligence, ranging from perceptual (e.g. visual object and
  7980. auditory speech recognition) to cognitive tasks (e.g. machine
  7981. translation), and on to motor control (e.g. playing computer
  7982. games or controlling a robot arm). In addition to their ability
  7983. to model complex intelligent behaviours, DNNs excel at predicting
  7984. neural responses to novel sensory stimuli with accuracies well
  7985. beyond any other currently available model type. DNNs can have
  7986. millions of parameters, which are required to capture the domain
  7987. knowledge needed for successful task performance. Contrary to the
  7988. intuition that this renders them into impenetrable black boxes,
  7989. the computational properties of the network units are the result
  7990. of four directly manipulable elements: input statistics, network
  7991. structure, functional objective, and learning algorithm. With
  7992. full access to the activity and connectivity of all units,
  7993. advanced visualization techniques, and analytic tools to map
  7994. network representations to neural data, DNNs represent a powerful
  7995. framework for building task-performing models and will drive
  7996. substantial insights in computational neuroscience.",
  7997. journal = "bioRxiv",
  7998. pages = "133504",
  7999. month = jun,
  8000. year = 2018,
  8001. language = "en"
  8002. }
  8003. @UNPUBLISHED{Nallapu2019-wq,
  8004. title = "Interacting roles of lateral and medial Orbitofrontal cortex in
  8005. decision-making and learning : A system-level computational model",
  8006. author = "Nallapu, Bhargav Teja and Alexandre, Fr{\'e}d{\'e}ric",
  8007. abstract = "In the context of flexible and adaptive animal behavior, the
  8008. orbitofrontal cortex (OFC) is found to be one of the crucial
  8009. regions in the prefrontal cortex (PFC) influencing the downstream
  8010. processes of decision-making and learning in the sub-cortical
  8011. regions. Although OFC has been implicated to be important in a
  8012. variety of related behavioral processes, the exact mechanisms are
  8013. unclear, through which the OFC encodes or processes information
  8014. related to decision-making and learning. Here, we propose a
  8015. systems-level view of the OFC, positioning it at the nexus of
  8016. sub-cortical systems and other prefrontal regions. Particularly
  8017. we focus on one of the most recent implications of
  8018. neuroscientific evidences regarding the OFC - possible functional
  8019. dissociation between two of its sub-regions : lateral and medial.
  8020. We present a system-level computational model of decision-making
  8021. and learning involving the two sub-regions taking into account
  8022. their individual roles as commonly implicated in neuroscientific
  8023. studies. We emphasize on the role of the interactions between the
  8024. sub-regions within the OFC as well as the role of other
  8025. sub-cortical structures which form a network with them. We
  8026. leverage well-known computational architecture of
  8027. thalamo-cortical basal ganglia loops, accounting for recent
  8028. experimental findings on monkeys with lateral and medial OFC
  8029. lesions, performing a 3-arm bandit task. First we replicate the
  8030. seemingly dissociate effects of lesions to lateral and medial OFC
  8031. during decision-making as a function of value-difference of the
  8032. presented options. Further we demonstrate and argue that such an
  8033. effect is not necessarily due to the dissociate roles of both the
  8034. subregions, but rather a result of complex temporal dynamics
  8035. between the interacting networks in which they are involved.
  8036. Author summary We first highlight the role of the Orbitofrontal
  8037. Cortex (OFC) in value-based decision making and goal-directed
  8038. behavior in primates. We establish the position of OFC at the
  8039. intersection of cortical mechanisms and thalamo-basal ganglial
  8040. circuits. In order to understand possible mechanisms through
  8041. which the OFC exerts emotional control over behavior, among
  8042. several other possibilities, we consider the case of dissociate
  8043. roles of two of its topographical subregions - lateral and medial
  8044. parts of OFC. We gather predominant roles of each of these
  8045. sub-regions as suggested by numerous experimental evidences in
  8046. the form of a system-level computational model that is based on
  8047. existing neuronal architectures. We argue that besides possible
  8048. dissociation, there could be possible interaction of these
  8049. sub-regions within themselves and through other sub-cortical
  8050. structures, in distinct mechanisms of choice and learning. The
  8051. computational framework described accounts for experimental data
  8052. and can be extended to more comprehensive detail of
  8053. representations required to understand the processes of
  8054. decision-making, learning and the role of OFC and subsequently
  8055. the regions of prefrontal cortex in general.",
  8056. journal = "bioRxiv",
  8057. pages = "867515",
  8058. month = dec,
  8059. year = 2019,
  8060. language = "en"
  8061. }
  8062. @ARTICLE{Yoo2019-jz,
  8063. title = "The Transition from Evaluation to Selection Involves Neural
  8064. Subspace Reorganization in Core Reward Regions",
  8065. author = "Yoo, Seng Bum Michael and Hayden, Benjamin Y",
  8066. abstract = "Economic choice proceeds from evaluation, in which we contemplate
  8067. options, to selection, in which we weigh options and choose one.
  8068. These stages must be differentiated so that decision makers do
  8069. not proceed to selection before evaluation is complete. We
  8070. examined responses of neurons in two core reward regions,
  8071. orbitofrontal (OFC) and ventromedial prefrontal cortex (vmPFC),
  8072. during two-option choice with asynchronous offer presentation.
  8073. Our data suggest that neurons selective during the first
  8074. (presumed evaluation) and second (presumed comparison and
  8075. selection) offer epochs come from a single pool. Stage transition
  8076. is accompanied by a shift toward orthogonality in the
  8077. low-dimensional population response manifold. Nonetheless, the
  8078. relative position of each option in driving responses in the
  8079. population subspace is preserved. The orthogonalization we
  8080. observe supports the hypothesis that the transition from
  8081. evaluation to selection leads to reorganization of response
  8082. subspace and suggests a mechanism by which value-related signals
  8083. are prevented from prematurely driving choice.",
  8084. journal = "Neuron",
  8085. month = nov,
  8086. year = 2019,
  8087. keywords = "comparison; covariance matrix; neuroeconomics; orbitofrontal
  8088. cortex; orthogonalization; valuation; ventromedial prefrontal
  8089. cortex",
  8090. language = "en"
  8091. }
  8092. @ARTICLE{Woon2019-lj,
  8093. title = "Involvement of the rodent prelimbic and medial orbitofrontal
  8094. cortices in goal-directed action: A brief review",
  8095. author = "Woon, Ellen P and Sequeira, Michelle K and Barbee, Britton R and
  8096. Gourley, Shannon L",
  8097. abstract = "Goal-directed action refers to selecting behaviors based on the
  8098. expectation that they will be reinforced with desirable outcomes.
  8099. It is typically conceptualized as opposing habit-based behaviors,
  8100. which are instead supported by stimulus-response associations and
  8101. insensitive to consequences. The prelimbic prefrontal cortex (PL)
  8102. is positioned along the medial wall of the rodent prefrontal
  8103. cortex. It is indispensable for action-outcome-driven
  8104. (goal-directed) behavior, consolidating action-outcome
  8105. relationships and linking contextual information with
  8106. instrumental behavior. In this brief review, we will discuss the
  8107. growing list of molecular factors involved in PL function.
  8108. Ventral to the PL is the medial orbitofrontal cortex (mOFC). We
  8109. will also summarize emerging evidence from rodents (complementing
  8110. existing literature describing humans) that it too is involved in
  8111. action-outcome conditioning. We describe experiments using
  8112. procedures that quantify responding based on reward value, the
  8113. likelihood of reinforcement, or effort requirements, touching
  8114. also on experiments assessing food consumption more generally. We
  8115. synthesize these findings with the argument that the mOFC is
  8116. essential to goal-directed action when outcome value information
  8117. is not immediately observable and must be recalled and inferred.",
  8118. journal = "J. Neurosci. Res.",
  8119. month = dec,
  8120. year = 2019,
  8121. keywords = "action-outcome; contingency degradation; devaluation; habit;
  8122. mouse; rat; response-outcome; review; reward",
  8123. language = "en"
  8124. }
  8125. @ARTICLE{Kao2019-hv,
  8126. title = "Considerations in using recurrent neural networks to probe neural
  8127. dynamics",
  8128. author = "Kao, Jonathan C",
  8129. abstract = "Recurrent neural networks (RNNs) are increasingly being used to
  8130. model complex cognitive and motor tasks performed by behaving
  8131. animals. RNNs are trained to reproduce animal behavior while also
  8132. capturing key statistics of empirically recorded neural activity.
  8133. In this manner, the RNN can be viewed as an in silico circuit
  8134. whose computational elements share similar motifs with the
  8135. cortical area it is modeling. Furthermore, because the RNN's
  8136. governing equations and parameters are fully known, they can be
  8137. analyzed to propose hypotheses for how neural populations
  8138. compute. In this context, we present important considerations
  8139. when using RNNs to model motor behavior in a delayed reach task.
  8140. First, by varying the network's nonlinear activation and rate
  8141. regularization, we show that RNNs reproducing single-neuron
  8142. firing rate motifs may not adequately capture important
  8143. population motifs. Second, we find that even when RNNs reproduce
  8144. key neurophysiological features on both the single neuron and
  8145. population levels, they can do so through distinctly different
  8146. dynamical mechanisms. To distinguish between these mechanisms, we
  8147. show that an RNN consistent with a previously proposed dynamical
  8148. mechanism is more robust to input noise. Finally, we show that
  8149. these dynamics are sufficient for the RNN to generalize to tasks
  8150. it was not trained on. Together, these results emphasize
  8151. important considerations when using RNN models to probe neural
  8152. dynamics.NEW \& NOTEWORTHY Artificial neurons in a recurrent
  8153. neural network (RNN) may resemble empirical single-unit activity
  8154. but not adequately capture important features on the neural
  8155. population level. Dynamics of RNNs can be visualized in
  8156. low-dimensional projections to provide insight into the RNN's
  8157. dynamical mechanism. RNNs trained in different ways may reproduce
  8158. neurophysiological motifs but do so with distinctly different
  8159. mechanisms. RNNs trained to only perform a delayed reach task can
  8160. generalize to perform tasks where the target is switched or the
  8161. target location is changed.",
  8162. journal = "J. Neurophysiol.",
  8163. volume = 122,
  8164. number = 6,
  8165. pages = "2504--2521",
  8166. month = dec,
  8167. year = 2019,
  8168. keywords = "artificial neural network; motor cortex; neural computation;
  8169. neural dynamics; recurrent neural network;RNN To read;RNN",
  8170. language = "en"
  8171. }
  8172. @ARTICLE{Wang2019-ot,
  8173. title = "Zona Incerta: An Integrative Node for Global Behavioral
  8174. Modulation",
  8175. author = "Wang, Xiyue and Chou, Xiao-Lin and Zhang, Li I and Tao, Huizhong
  8176. Whit",
  8177. abstract = "Zona incerta (ZI) is a largely inhibitory subthalamic region
  8178. connecting with many brain areas. Early studies have suggested
  8179. involvement of ZI in various functions such as visceral
  8180. activities, arousal, attention, and locomotion, but the specific
  8181. roles of different ZI subdomains or cell types have not been well
  8182. examined. Recent studies combining optogenetics, behavioral
  8183. assays, neural tracing, and neural activity-recording reveal
  8184. novel functional roles of ZI depending on specific input-output
  8185. connectivity patterns. Here, we review these studies and
  8186. summarize functions of ZI into four categories: sensory
  8187. integration, behavioral output control, motivational drive, and
  8188. neural plasticity. In view of these new findings, we propose that
  8189. ZI serves as an integrative node for global modulation of
  8190. behaviors and physiological states.",
  8191. journal = "Trends Neurosci.",
  8192. month = dec,
  8193. year = 2019,
  8194. keywords = "animal behavior; gain modulation; inhibitory nucleus;
  8195. input/output pattern; motivation; neural plasticity;
  8196. physiological state; processing node; subthalamic region",
  8197. language = "en"
  8198. }
  8199. @ARTICLE{Marques2019-dw,
  8200. title = "Internal state dynamics shape brainwide activity and foraging
  8201. behaviour",
  8202. author = "Marques, Jo{\~a}o C and Li, Meng and Schaak, Diane and Robson,
  8203. Drew N and Li, Jennifer M",
  8204. abstract = "The brain has persistent internal states that can modulate every
  8205. aspect of an animal's mental experience1-4. In complex tasks such
  8206. as foraging, the internal state is dynamic5-8. Caenorhabditis
  8207. elegans alternate between local search and global dispersal5.
  8208. Rodents and primates exhibit trade-offs between exploitation and
  8209. exploration6,7. However, fundamental questions remain about how
  8210. persistent states are maintained in the brain, which upstream
  8211. networks drive state transitions and how state-encoding neurons
  8212. exert neuromodulatory effects on sensory perception and
  8213. decision-making to govern appropriate behaviour. Here, using
  8214. tracking microscopy to monitor whole-brain neuronal activity at
  8215. cellular resolution in freely moving zebrafish larvae9, we show
  8216. that zebrafish spontaneously alternate between two persistent
  8217. internal states during foraging for live prey (Paramecia). In the
  8218. exploitation state, the animal inhibits locomotion and promotes
  8219. hunting, generating small, localized trajectories. In the
  8220. exploration state, the animal promotes locomotion and suppresses
  8221. hunting, generating long-ranging trajectories that enhance
  8222. spatial dispersion. We uncover a dorsal raphe subpopulation with
  8223. persistent activity that robustly encodes the exploitation state.
  8224. The exploitation-state-encoding neurons, together with a
  8225. multimodal trigger network that is associated with state
  8226. transitions, form a stochastically activated nonlinear dynamical
  8227. system. The activity of this oscillatory network correlates with
  8228. a global retuning of sensorimotor transformations during foraging
  8229. that leads to marked changes in both the motivation to hunt for
  8230. prey and the accuracy of motor sequences during hunting. This
  8231. work reveals an important hidden variable that shapes the
  8232. temporal structure of motivation and decision-making.",
  8233. journal = "Nature",
  8234. month = dec,
  8235. year = 2019,
  8236. language = "en"
  8237. }
  8238. @UNPUBLISHED{Sanchez-Bellot2019-me,
  8239. title = "Push-pull regulation of exploratory behavior by two opposing
  8240. hippocampal to prefrontal cortex pathways",
  8241. author = "S{\'a}nchez-Bellot, Candela and MacAskill, Andrew F",
  8242. abstract = "We found that the hippocampal projection to prefrontal cortex is
  8243. composed of two parallel circuits located in the superficial or
  8244. deep hippocampal pyramidal layers. These circuits have unique
  8245. upstream and downstream connectivity, and are differentially
  8246. active during exploration of a potentially threatening
  8247. environment. Artificial activation of the superficial circuit
  8248. promotes exploration via preferential recruitment of PFC
  8249. inhibition, while activation of the deep circuit promotes
  8250. avoidance via direct excitation.",
  8251. journal = "bioRxiv",
  8252. pages = "2019.12.18.880831",
  8253. month = dec,
  8254. year = 2019,
  8255. language = "en"
  8256. }
  8257. @ARTICLE{Pollock_undated-sq,
  8258. title = "Engineering recurrent neural networks from task-relevant manifolds
  8259. and dynamics",
  8260. author = "Pollock, Eli and Jazayeri, Mehrdad"
  8261. }
  8262. @ARTICLE{Mearns2019-md,
  8263. title = "Deconstructing Hunting Behavior Reveals a Tightly Coupled
  8264. {Stimulus-Response} Loop",
  8265. author = "Mearns, Duncan S and Donovan, Joseph C and Fernandes,
  8266. Ant{\'o}nio M and Semmelhack, Julia L and Baier, Herwig",
  8267. abstract = "SummaryAnimal behavior often forms sequences, built from simple
  8268. stereotyped actions and shaped by environmental cues. A
  8269. comprehensive characterization of the interplay between an
  8270. animal's movements and its environment is necessary to
  8271. understand the sensorimotor transformations performed by the
  8272. brain. Here, we use unsupervised methods to study behavioral
  8273. sequences in zebrafish larvae. We generate a map of swim bouts,
  8274. revealing that fish modulate their tail movements along a
  8275. continuum. During prey capture, larvae produce stereotyped
  8276. sequences using a subset of bouts from a broader behavioral
  8277. repertoire. These sequences exhibit low-order transition
  8278. dynamics and immediately respond to changes in visual cues.
  8279. Chaining of prey capture bouts is disrupted in visually impaired
  8280. (lakritz and blumenkohl) mutants, and removing the prey stimulus
  8281. during ongoing behavior in closed-loop virtual reality causes
  8282. larvae to immediately abort the hunting sequence. These results
  8283. suggest that the continuous integration of sensory information
  8284. is necessary to structure the behavior. This stimulus-response
  8285. loop serves to bring prey into the anterior dorsal visual field
  8286. of the larvae. Fish then release a capture strike maneuver
  8287. comprising a stereotyped jaw movement and tail movements
  8288. fine-tuned to the distance of the prey. Fish with only one
  8289. intact eye fail to correctly position the prey in the strike
  8290. zone, but are able to produce the strike itself. Our analysis
  8291. shows that short-term integration of binocular visual cues
  8292. shapes the behavioral dynamics of hunting, thus uncovering the
  8293. temporal organization of a goal-directed behavior in a
  8294. vertebrate.",
  8295. journal = "Curr. Biol.",
  8296. publisher = "Elsevier",
  8297. volume = 0,
  8298. number = 0,
  8299. month = dec,
  8300. year = 2019,
  8301. keywords = "zebrafish; ethology; prey capture; unsupervised machine
  8302. learning; behavioral sequences",
  8303. language = "en"
  8304. }
  8305. @ARTICLE{Lefler2019-ke,
  8306. title = "The role of the periaqueductal gray in escape behavior",
  8307. author = "Lefler, Yaara and Campagner, Dario and Branco, Tiago",
  8308. abstract = "Escape behavior is a defensive action deployed by animals in
  8309. response to imminent threats. In mammalian species, a variety of
  8310. different brain circuits are known to participate in this crucial
  8311. survival behavior. One of these circuits is the periaqueductal
  8312. gray, a midbrain structure that can command a variety of
  8313. instinctive behaviors. Recent experiments using modern systems
  8314. neuroscience techniques have begun to elucidate the specific role
  8315. of the periaqueductal gray in controlling escape. These have
  8316. shown that periaqueductal gray neurons are crucial units for
  8317. gating and commanding the initiation of escape, specifically
  8318. activated in situations of imminent, escapable threat. In
  8319. addition, it is becoming clear that the periaqueductal gray
  8320. integrates brain-wide information that can modulate escape
  8321. initiation to generate flexible defensive behaviors.",
  8322. journal = "Curr. Opin. Neurobiol.",
  8323. volume = 60,
  8324. pages = "115--121",
  8325. month = dec,
  8326. year = 2019,
  8327. language = "en"
  8328. }
  8329. @ARTICLE{Perreault2019-da,
  8330. title = "Diversity of reticulospinal systems in mammals",
  8331. author = "Perreault, Marie-Claude and Giorgi, Andrea",
  8332. abstract = "Reticulospinal (RS) neurons provide the spinal cord with the
  8333. executive signals for a large repertoire of motor and autonomic
  8334. functions, ensuring at the same time that these functions are
  8335. adapted to the different behavioral contexts. This requires the
  8336. coordinated action of many RS neurons. In this mini-review, we
  8337. examine how the RS neurons that carry out specific functions
  8338. distribute across the three parts of the brain stem. Extensive
  8339. overlap between populations suggests a need to explore
  8340. multi-functionality at the single cell-level. We next contrast
  8341. functional diversity and homogeneity in transmitter phenotype.
  8342. Then, we examine the molecular genetic mechanisms that specify
  8343. brain stem development and likely contribute to RS neurons
  8344. identities. We advocate that a better knowledge of the
  8345. developmental lineage of the RS neurons and a better knowledge of
  8346. RS neuron activity across multiple behaviors will help uncover
  8347. the fundamental principles behind the diversity of RS systems in
  8348. mammals.",
  8349. journal = "Curr Opin Physiol",
  8350. volume = 8,
  8351. pages = "161--169",
  8352. month = apr,
  8353. year = 2019,
  8354. keywords = "Brain Stem; Control of Movement and Bodily Functions; Reticular
  8355. Formation; Reticulospinal",
  8356. language = "en"
  8357. }
  8358. @UNPUBLISHED{Finkelstein2019-ne,
  8359. title = "Attractor dynamics gate cortical information flow during
  8360. decision-making",
  8361. author = "Finkelstein, Arseny and Fontolan, Lorenzo and Economo, Michael N
  8362. and Li, Nuo and Romani, Sandro and Svoboda, Karel",
  8363. abstract = "Decisions about future actions are held in memory until enacted,
  8364. making them vulnerable to distractors. The neural mechanisms
  8365. controlling decision robustness to distractors remain unknown. We
  8366. trained mice to report optogenetic stimulation of somatosensory
  8367. cortex, with a delay separating sensation and action. Distracting
  8368. stimuli influenced behavior less when delivered later during
  8369. delay --- demonstrating temporal gating of sensory information
  8370. flow. Gating occurred even though distractor-evoked activity
  8371. percolated through the cortex without attenuation. Instead,
  8372. choice-related dynamics in frontal cortex became progressively
  8373. robust to distractors as time passed. Reverse-engineering of
  8374. neural networks trained to reproduce frontal-cortex activity
  8375. revealed that chosen actions were stabilized via attractor
  8376. dynamics, which gated out distracting stimuli. Our results reveal
  8377. a dynamic gating mechanism that operates by controlling the
  8378. degree of commitment to a chosen course of action.",
  8379. journal = "bioRxiv",
  8380. pages = "2019.12.14.876425",
  8381. month = dec,
  8382. year = 2019,
  8383. language = "en"
  8384. }
  8385. @UNPUBLISHED{Isa2019-ww,
  8386. title = "Difference in context-dependency between orienting and
  8387. defense-like responses induced by the superior colliculus",
  8388. author = "Isa, Kaoru and Sooksawate, Thongchai and Kobayashi, Kenta and
  8389. Kobayashi, Kazuto and Redgrave, Peter and Isa, Tadashi",
  8390. abstract = "Abstract Previous electrical stimulation and lesion experiments
  8391. have suggested that the crossed descending output pathway from
  8392. the deeper layers (SCd) of superior colliculus (SC) controls
  8393. orienting responses, while the uncrossed pathway mediates
  8394. defense-like behavior. Here we extended these investigations by
  8395. using selective optogenetic activation of each pathway in mice
  8396. with channelrhodopsin 2 expression by double viral vector
  8397. techniques. Brief photo-stimulation of the crossed pathway evoked
  8398. short latency contraversive orienting-like head turns, while
  8399. extended stimulation induced contraversive circling responses. In
  8400. contrast, stimulation of uncrossed pathway induced short-latency
  8401. upward head movements followed by longer-latency defense-like
  8402. behaviors including retreat and flight. The novel discovery was
  8403. that the evoked defense-like responses varied depending on the
  8404. environment, suggesting that uncrossed output can be influenced
  8405. by top-down modification of the SC or its downstream. This
  8406. further suggests that the SCd-defense system can be profoundly
  8407. modulated by non-motor, affective and cognitive components, in
  8408. addition to direct sensory inputs.",
  8409. journal = "bioRxiv",
  8410. pages = "729772",
  8411. month = aug,
  8412. year = 2019,
  8413. language = "en"
  8414. }
  8415. @ARTICLE{Stone2017-ix,
  8416. title = "An Anatomically Constrained Model for Path Integration in the Bee
  8417. Brain",
  8418. author = "Stone, Thomas and Webb, Barbara and Adden, Andrea and Weddig,
  8419. Nicolai Ben and Honkanen, Anna and Templin, Rachel and Wcislo,
  8420. William and Scimeca, Luca and Warrant, Eric and Heinze, Stanley",
  8421. abstract = "Path integration is a widespread navigational strategy in which
  8422. directional changes and distance covered are continuously
  8423. integrated on an outward journey, enabling a straight-line return
  8424. to home. Bees use vision for this task-a celestial-cue-based
  8425. visual compass and an optic-flow-based visual odometer-but the
  8426. underlying neural integration mechanisms are unknown. Using
  8427. intracellular electrophysiology, we show that
  8428. polarized-light-based compass neurons and optic-flow-based
  8429. speed-encoding neurons converge in the central complex of the bee
  8430. brain, and through block-face electron microscopy, we identify
  8431. potential integrator cells. Based on plausible output targets for
  8432. these cells, we propose a complete circuit for path integration
  8433. and steering in the central complex, with anatomically identified
  8434. neurons suggested for each processing step. The resulting model
  8435. circuit is thus fully constrained biologically and provides a
  8436. functional interpretation for many previously unexplained
  8437. architectural features of the central complex. Moreover, we show
  8438. that the receptive fields of the newly discovered speed neurons
  8439. can support path integration for the holonomic motion (i.e., a
  8440. ground velocity that is not precisely aligned with body
  8441. orientation) typical of bee flight, a feature not captured in any
  8442. previously proposed model of path integration. In a broader
  8443. context, the model circuit presented provides a general mechanism
  8444. for producing steering signals by comparing current and desired
  8445. headings-suggesting a more basic function for central complex
  8446. connectivity, from which path integration may have evolved.",
  8447. journal = "Curr. Biol.",
  8448. volume = 27,
  8449. number = 20,
  8450. pages = "3069--3085.e11",
  8451. month = oct,
  8452. year = 2017,
  8453. keywords = "central complex; circuit modeling; compass orientation; insect
  8454. brain; navigation; neuroanatomy; optic flow; path integration;
  8455. polarized light; robotics",
  8456. language = "en"
  8457. }
  8458. @UNPUBLISHED{Plitt2019-hy,
  8459. title = "Experience dependent contextual codes in the hippocampus",
  8460. author = "Plitt, Mark H and Giocomo, Lisa M",
  8461. abstract = "The hippocampus is a medial temporal lobe brain structure that
  8462. contains circuitry and neural representations capable of
  8463. supporting declarative memory. Hippocampal place cells fire in
  8464. one or few restricted spatial locations in a given environment.
  8465. Between environmental contexts, place cell firing fields remap
  8466. (turning on/off or moving to a new spatial location), providing a
  8467. unique population-wide neural code for context specificity.
  8468. However, the manner by which features associated with a given
  8469. context combine to drive place cell remapping remains a matter of
  8470. debate. Here we show that remapping of neural representations in
  8471. region CA1 of the hippocampus is strongly driven by prior beliefs
  8472. about the frequency of certain contexts, and that remapping is
  8473. equivalent to an optimal estimate of the identity of the current
  8474. context under that prior. This prior-driven remapping is learned
  8475. early in training and remains robust to changes in behavioral
  8476. task-demands. Furthermore, a simple associative learning
  8477. mechanism is sufficient to reproduce these results. Our findings
  8478. demonstrate that place cell remapping is a generalization of
  8479. representing an animal9s location. Rather than simply
  8480. representing location in physical space, the hippocampus
  8481. represents an optimal estimate of location in a multi-dimensional
  8482. stimulus space.",
  8483. journal = "bioRxiv",
  8484. pages = "864090",
  8485. month = dec,
  8486. year = 2019,
  8487. language = "en"
  8488. }
  8489. @ARTICLE{Wang2020-ee,
  8490. title = "Egocentric and allocentric representations of space in the rodent
  8491. brain",
  8492. author = "Wang, Cheng and Chen, Xiaojing and Knierim, James J",
  8493. abstract = "Spatial signals are prevalent within the hippocampus and its
  8494. neighboring regions. It is generally accepted that these signals
  8495. are defined with respect to the external world (i.e., a
  8496. world-centered, or allocentric, frame of reference). Recently,
  8497. evidence of egocentric processing (i.e., self-centered, defined
  8498. relative to the subject) in the extended hippocampal system has
  8499. accumulated. These results support the idea that egocentric
  8500. sensory information, derived from primary sensory cortical areas,
  8501. may be transformed to allocentric representations that interact
  8502. with the allocentric hippocampal system. We propose a framework
  8503. to explain the implications of the egocentric-allocentric
  8504. transformations to the functions of the medial temporal lobe
  8505. memory system.",
  8506. journal = "Curr. Opin. Neurobiol.",
  8507. volume = 60,
  8508. pages = "12--20",
  8509. month = feb,
  8510. year = 2020
  8511. }
  8512. @ARTICLE{Hoy2019-jh,
  8513. title = "Defined Cell Types in Superior Colliculus Make Distinct
  8514. Contributions to Prey Capture Behavior in the Mouse",
  8515. author = "Hoy, Jennifer L and Bishop, Hannah I and Niell, Cristopher M",
  8516. abstract = "The superior colliculus (SC) plays a highly conserved role in
  8517. visual processing and mediates visual orienting behaviors across
  8518. species, including both overt motor orienting [1, 2] and
  8519. orienting of attention [3, 4]. To determine the specific circuits
  8520. within the superficial superior colliculus (sSC) that drive
  8521. orienting and approach behavior toward appetitive stimuli, we
  8522. explored the role of three genetically defined cell types in
  8523. mediating prey capture in mice. Chemogenetic inactivation of two
  8524. classically defined cell types, the wide-field (WF) and
  8525. narrow-field (NF) vertical neurons, revealed that they are
  8526. involved in distinct aspects of prey capture. WF neurons were
  8527. required for rapid prey detection and distant approach
  8528. initiation, whereas NF neurons were required for accurate
  8529. orienting during pursuit as well as approach initiation and
  8530. continuity. In contrast, prey capture did not require
  8531. parvalbumin-expressing (PV) neurons that have previously been
  8532. implicated in fear responses. The visual coding and projection
  8533. targets of WF and NF cells were consistent with their roles in
  8534. prey detection versus pursuit, respectively. Thus, our studies
  8535. link specific neural circuit connectivity and function with
  8536. stimulus detection and orienting behavior, providing insight into
  8537. visuomotor and attentional mechanisms mediated by superior
  8538. colliculus.",
  8539. journal = "Curr. Biol.",
  8540. month = nov,
  8541. year = 2019,
  8542. keywords = "mouse vision; prey capture; receptive fields; superior colliculus",
  8543. language = "en"
  8544. }
  8545. @ARTICLE{Winnubst2019-eo,
  8546. title = "Reconstruction of 1,000 Projection Neurons Reveals New Cell Types
  8547. and Organization of {Long-Range} Connectivity in the Mouse Brain",
  8548. author = "Winnubst, Johan and Bas, Erhan and Ferreira, Tiago A and Wu,
  8549. Zhuhao and Economo, Michael N and Edson, Patrick and Arthur, Ben
  8550. J and Bruns, Christopher and Rokicki, Konrad and Schauder, David
  8551. and Olbris, Donald J and Murphy, Sean D and Ackerman, David G and
  8552. Arshadi, Cameron and Baldwin, Perry and Blake, Regina and
  8553. Elsayed, Ahmad and Hasan, Mashtura and Ramirez, Daniel and Dos
  8554. Santos, Bruno and Weldon, Monet and Zafar, Amina and Dudman,
  8555. Joshua T and Gerfen, Charles R and Hantman, Adam W and Korff,
  8556. Wyatt and Sternson, Scott M and Spruston, Nelson and Svoboda,
  8557. Karel and Chandrashekar, Jayaram",
  8558. abstract = "Neuronal cell types are the nodes of neural circuits that
  8559. determine the flow of information within the brain. Neuronal
  8560. morphology, especially the shape of the axonal arbor, provides an
  8561. essential descriptor of cell type and reveals how individual
  8562. neurons route their output across the brain. Despite the
  8563. importance of morphology, few projection neurons in the mouse
  8564. brain have been reconstructed in their entirety. Here we present
  8565. a robust and efficient platform for imaging and reconstructing
  8566. complete neuronal morphologies, including axonal arbors that span
  8567. substantial portions of the brain. We used this platform to
  8568. reconstruct more than 1,000 projection neurons in the motor
  8569. cortex, thalamus, subiculum, and hypothalamus. Together, the
  8570. reconstructed neurons constitute more than 85 meters of axonal
  8571. length and are available in a searchable online database. Axonal
  8572. shapes revealed previously unknown subtypes of projection neurons
  8573. and suggest organizational principles of long-range connectivity.",
  8574. journal = "Cell",
  8575. volume = 179,
  8576. number = 1,
  8577. pages = "268--281.e13",
  8578. month = sep,
  8579. year = 2019,
  8580. keywords = "automated reconstruction; axonal morphology; long-range
  8581. projections; morphology database; neuronal cell types; neuronal
  8582. connectivity; projection neurons; single-cell reconstruction;
  8583. whole brain",
  8584. language = "en"
  8585. }
  8586. @ARTICLE{Ragan2012-yo,
  8587. title = "Serial two-photon tomography for automated ex vivo mouse brain
  8588. imaging",
  8589. author = "Ragan, Timothy and Kadiri, Lolahon R and Venkataraju, Kannan
  8590. Umadevi and Bahlmann, Karsten and Sutin, Jason and Taranda,
  8591. Julian and Arganda-Carreras, Ignacio and Kim, Yongsoo and Seung,
  8592. H Sebastian and Osten, Pavel",
  8593. abstract = "Here we describe an automated method, named serial two-photon
  8594. (STP) tomography, that achieves high-throughput fluorescence
  8595. imaging of mouse brains by integrating two-photon microscopy and
  8596. tissue sectioning. STP tomography generates high-resolution
  8597. datasets that are free of distortions and can be readily warped
  8598. in three dimensions, for example, for comparing multiple
  8599. anatomical tracings. This method opens the door to routine
  8600. systematic studies of neuroanatomy in mouse models of human
  8601. brain disorders.",
  8602. journal = "Nat. Methods",
  8603. publisher = "nature.com",
  8604. volume = 9,
  8605. number = 3,
  8606. pages = "255--258",
  8607. month = jan,
  8608. year = 2012,
  8609. language = "en"
  8610. }
  8611. @ARTICLE{Kuan2015-mt,
  8612. title = "Neuroinformatics of the Allen Mouse Brain Connectivity Atlas",
  8613. author = "Kuan, Leonard and Li, Yang and Lau, Chris and Feng, David and
  8614. Bernard, Amy and Sunkin, Susan M and Zeng, Hongkui and Dang,
  8615. Chinh and Hawrylycz, Michael and Ng, Lydia",
  8616. abstract = "The Allen Mouse Brain Connectivity Atlas is a mesoscale whole
  8617. brain axonal projection atlas of the C57Bl/6J mouse brain.
  8618. Anatomical trajectories throughout the brain were mapped into a
  8619. common 3D space using a standardized platform to generate a
  8620. comprehensive and quantitative database of inter-areal and
  8621. cell-type-specific projections. This connectivity atlas has
  8622. several desirable features, including brain-wide coverage,
  8623. validated and versatile experimental techniques, a single
  8624. standardized data format, a quantifiable and integrated
  8625. neuroinformatics resource, and an open-access public online
  8626. database (http://connectivity.brain-map.org/). Meaningful
  8627. informatics data quantification and comparison is key to
  8628. effective use and interpretation of connectome data. This relies
  8629. on successful definition of a high fidelity atlas template and
  8630. framework, mapping precision of raw data sets into the 3D
  8631. reference framework, accurate signal detection and quantitative
  8632. connection strength algorithms, and effective presentation in an
  8633. integrated online application. Here we describe key informatics
  8634. pipeline steps in the creation of the Allen Mouse Brain
  8635. Connectivity Atlas and include basic application use cases.",
  8636. journal = "Methods",
  8637. publisher = "Elsevier",
  8638. volume = 73,
  8639. pages = "4--17",
  8640. month = feb,
  8641. year = 2015,
  8642. keywords = "Digital atlas; Image registration; Mouse connectivity atlas;
  8643. Neuronal projection; Signal detection",
  8644. language = "en"
  8645. }
  8646. @ARTICLE{Sunkin2013-ap,
  8647. title = "Allen Brain Atlas: an integrated spatio-temporal portal for
  8648. exploring the central nervous system",
  8649. author = "Sunkin, Susan M and Ng, Lydia and Lau, Chris and Dolbeare, Tim
  8650. and Gilbert, Terri L and Thompson, Carol L and Hawrylycz,
  8651. Michael and Dang, Chinh",
  8652. abstract = "The Allen Brain Atlas (http://www.brain-map.org) provides a
  8653. unique online public resource integrating extensive gene
  8654. expression data, connectivity data and neuroanatomical
  8655. information with powerful search and viewing tools for the adult
  8656. and developing brain in mouse, human and non-human primate.
  8657. Here, we review the resources available at the Allen Brain
  8658. Atlas, describing each product and data type [such as in situ
  8659. hybridization (ISH) and supporting histology, microarray, RNA
  8660. sequencing, reference atlases, projection mapping and magnetic
  8661. resonance imaging]. In addition, standardized and unique
  8662. features in the web applications are described that enable users
  8663. to search and mine the various data sets. Features include both
  8664. simple and sophisticated methods for gene searches, colorimetric
  8665. and fluorescent ISH image viewers, graphical displays of ISH,
  8666. microarray and RNA sequencing data, Brain Explorer software for
  8667. 3D navigation of anatomy and gene expression, and an interactive
  8668. reference atlas viewer. In addition, cross data set searches
  8669. enable users to query multiple Allen Brain Atlas data sets
  8670. simultaneously. All of the Allen Brain Atlas resources can be
  8671. accessed through the Allen Brain Atlas data portal.",
  8672. journal = "Nucleic Acids Res.",
  8673. publisher = "academic.oup.com",
  8674. volume = 41,
  8675. number = "Database issue",
  8676. pages = "D996--D1008",
  8677. month = jan,
  8678. year = 2013,
  8679. language = "en"
  8680. }
  8681. @ARTICLE{Lein2007-di,
  8682. title = "Genome-wide atlas of gene expression in the adult mouse brain",
  8683. author = "Lein, Ed S and Hawrylycz, Michael J and Ao, Nancy and Ayres,
  8684. Mikael and Bensinger, Amy and Bernard, Amy and Boe, Andrew F and
  8685. Boguski, Mark S and Brockway, Kevin S and Byrnes, Emi J and
  8686. Chen, Lin and Chen, Li and Chen, Tsuey-Ming and Chin, Mei Chi
  8687. and Chong, Jimmy and Crook, Brian E and Czaplinska, Aneta and
  8688. Dang, Chinh N and Datta, Suvro and Dee, Nick R and Desaki, Aimee
  8689. L and Desta, Tsega and Diep, Ellen and Dolbeare, Tim A and
  8690. Donelan, Matthew J and Dong, Hong-Wei and Dougherty, Jennifer G
  8691. and Duncan, Ben J and Ebbert, Amanda J and Eichele, Gregor and
  8692. Estin, Lili K and Faber, Casey and Facer, Benjamin A and Fields,
  8693. Rick and Fischer, Shanna R and Fliss, Tim P and Frensley, Cliff
  8694. and Gates, Sabrina N and Glattfelder, Katie J and Halverson,
  8695. Kevin R and Hart, Matthew R and Hohmann, John G and Howell,
  8696. Maureen P and Jeung, Darren P and Johnson, Rebecca A and Karr,
  8697. Patrick T and Kawal, Reena and Kidney, Jolene M and Knapik,
  8698. Rachel H and Kuan, Chihchau L and Lake, James H and Laramee,
  8699. Annabel R and Larsen, Kirk D and Lau, Christopher and Lemon,
  8700. Tracy A and Liang, Agnes J and Liu, Ying and Luong, Lon T and
  8701. Michaels, Jesse and Morgan, Judith J and Morgan, Rebecca J and
  8702. Mortrud, Marty T and Mosqueda, Nerick F and Ng, Lydia L and Ng,
  8703. Randy and Orta, Geralyn J and Overly, Caroline C and Pak, Tu H
  8704. and Parry, Sheana E and Pathak, Sayan D and Pearson, Owen C and
  8705. Puchalski, Ralph B and Riley, Zackery L and Rockett, Hannah R
  8706. and Rowland, Stephen A and Royall, Joshua J and Ruiz, Marcos J
  8707. and Sarno, Nadia R and Schaffnit, Katherine and Shapovalova,
  8708. Nadiya V and Sivisay, Taz and Slaughterbeck, Clifford R and
  8709. Smith, Simon C and Smith, Kimberly A and Smith, Bryan I and
  8710. Sodt, Andy J and Stewart, Nick N and Stumpf, Kenda-Ruth and
  8711. Sunkin, Susan M and Sutram, Madhavi and Tam, Angelene and
  8712. Teemer, Carey D and Thaller, Christina and Thompson, Carol L and
  8713. Varnam, Lee R and Visel, Axel and Whitlock, Ray M and Wohnoutka,
  8714. Paul E and Wolkey, Crissa K and Wong, Victoria Y and Wood,
  8715. Matthew and Yaylaoglu, Murat B and Young, Rob C and Youngstrom,
  8716. Brian L and Yuan, Xu Feng and Zhang, Bin and Zwingman, Theresa A
  8717. and Jones, Allan R",
  8718. abstract = "Molecular approaches to understanding the functional circuitry
  8719. of the nervous system promise new insights into the relationship
  8720. between genes, brain and behaviour. The cellular diversity of
  8721. the brain necessitates a cellular resolution approach towards
  8722. understanding the functional genomics of the nervous system. We
  8723. describe here an anatomically comprehensive digital atlas
  8724. containing the expression patterns of approximately 20,000 genes
  8725. in the adult mouse brain. Data were generated using automated
  8726. high-throughput procedures for in situ hybridization and data
  8727. acquisition, and are publicly accessible online. Newly developed
  8728. image-based informatics tools allow global genome-scale
  8729. structural analysis and cross-correlation, as well as
  8730. identification of regionally enriched genes. Unbiased
  8731. fine-resolution analysis has identified highly specific cellular
  8732. markers as well as extensive evidence of cellular heterogeneity
  8733. not evident in classical neuroanatomical atlases. This highly
  8734. standardized atlas provides an open, primary data resource for a
  8735. wide variety of further studies concerning brain organization
  8736. and function.",
  8737. journal = "Nature",
  8738. publisher = "nature.com",
  8739. volume = 445,
  8740. number = 7124,
  8741. pages = "168--176",
  8742. month = jan,
  8743. year = 2007,
  8744. language = "en"
  8745. }
  8746. @ARTICLE{Osten2013-ze,
  8747. title = "Mapping brain circuitry with a light microscope",
  8748. author = "Osten, Pavel and Margrie, Troy W",
  8749. abstract = "The beginning of the 21st century has seen a renaissance in light
  8750. microscopy and anatomical tract tracing that together are rapidly
  8751. advancing our understanding of the form and function of neuronal
  8752. circuits. The introduction of instruments for automated imaging
  8753. of whole mouse brains, new cell type--specific and trans-synaptic
  8754. tracers, and computational methods for handling the whole-brain
  8755. data sets has opened the door to neuroanatomical studies at an
  8756. unprecedented scale. We present an overview of the present state
  8757. and future opportunities in charting long-range and local
  8758. connectivity in the entire mouse brain and in linking brain
  8759. circuits to function.",
  8760. journal = "Nat. Methods",
  8761. volume = 10,
  8762. number = 6,
  8763. pages = "515--523",
  8764. month = jun,
  8765. year = 2013,
  8766. language = "en"
  8767. }
  8768. @ARTICLE{Steinmetz2019-av,
  8769. title = "Distributed coding of choice, action and engagement across the
  8770. mouse brain",
  8771. author = "Steinmetz, Nicholas A and Zatka-Haas, Peter and Carandini,
  8772. Matteo and Harris, Kenneth D",
  8773. abstract = "Vision, choice, action and behavioural engagement arise from
  8774. neuronal activity that may be distributed across brain regions.
  8775. Here we delineate the spatial distribution of neurons underlying
  8776. these processes. We used Neuropixels probes1,2 to record from
  8777. approximately 30,000 neurons in 42 brain regions of mice
  8778. performing a visual discrimination task3. Neurons in nearly all
  8779. regions responded non-specifically when the mouse initiated an
  8780. action. By contrast, neurons encoding visual stimuli and
  8781. upcoming choices occupied restricted regions in the neocortex,
  8782. basal ganglia and midbrain. Choice signals were rare and emerged
  8783. with indistinguishable timing across regions. Midbrain neurons
  8784. were activated before contralateral choices and were suppressed
  8785. before ipsilateral choices, whereas forebrain neurons could
  8786. prefer either side. Brain-wide pre-stimulus activity predicted
  8787. engagement in individual trials and in the overall task, with
  8788. enhanced subcortical but suppressed neocortical activity during
  8789. engagement. These results reveal organizing principles for the
  8790. distribution of neurons encoding behaviourally relevant
  8791. variables across the mouse brain.",
  8792. journal = "Nature",
  8793. publisher = "Nature Publishing Group",
  8794. pages = "1--8",
  8795. month = nov,
  8796. year = 2019,
  8797. language = "en"
  8798. }
  8799. @ARTICLE{Economo2018-pj,
  8800. title = "Distinct descending motor cortex pathways and their roles in
  8801. movement",
  8802. author = "Economo, Michael N and Viswanathan, Sarada and Tasic, Bosiljka
  8803. and Bas, Erhan and Winnubst, Johan and Menon, Vilas and Graybuck,
  8804. Lucas T and Nguyen, Thuc Nghi and Smith, Kimberly A and Yao,
  8805. Zizhen and Wang, Lihua and Gerfen, Charles R and Chandrashekar,
  8806. Jayaram and Zeng, Hongkui and Looger, Loren L and Svoboda, Karel",
  8807. abstract = "Activity in the motor cortex predicts movements, seconds before
  8808. they are initiated. This preparatory activity has been observed
  8809. across cortical layers, including in descending pyramidal tract
  8810. neurons in layer 5. A key question is how preparatory activity is
  8811. maintained without causing movement, and is ultimately converted
  8812. to a motor command to trigger appropriate movements. Here, using
  8813. single-cell transcriptional profiling and axonal reconstructions,
  8814. we identify two types of pyramidal tract neuron. Both types
  8815. project to several targets in the basal ganglia and brainstem.
  8816. One type projects to thalamic regions that connect back to motor
  8817. cortex; populations of these neurons produced early preparatory
  8818. activity that persisted until the movement was initiated. The
  8819. second type projects to motor centres in the medulla and mainly
  8820. produced late preparatory activity and motor commands. These
  8821. results indicate that two types of motor cortex output neurons
  8822. have specialized roles in motor control.",
  8823. journal = "Nature",
  8824. volume = 563,
  8825. number = 7729,
  8826. pages = "79--84",
  8827. month = nov,
  8828. year = 2018,
  8829. language = "en"
  8830. }
  8831. @ARTICLE{Cunningham2014-aw,
  8832. title = "Dimensionality reduction for large-scale neural recordings",
  8833. author = "Cunningham, John P and Yu, Byron M",
  8834. abstract = "Most sensory, cognitive and motor functions depend on the
  8835. interactions of many neurons. In recent years, there has been
  8836. rapid development and increasing use of technologies for
  8837. recording from large numbers of neurons, either sequentially or
  8838. simultaneously. A key question is what scientific insight can be
  8839. gained by studying a population of recorded neurons beyond
  8840. studying each neuron individually. Here, we examine three
  8841. important motivations for population studies: single-trial
  8842. hypotheses requiring statistical power, hypotheses of population
  8843. response structure and exploratory analyses of large data sets.
  8844. Many recent studies have adopted dimensionality reduction to
  8845. analyze these populations and to find features that are not
  8846. apparent at the level of individual neurons. We describe the
  8847. dimensionality reduction methods commonly applied to population
  8848. activity and offer practical advice about selecting methods and
  8849. interpreting their outputs. This review is intended for
  8850. experimental and computational researchers who seek to understand
  8851. the role dimensionality reduction has had and can have in systems
  8852. neuroscience, and who seek to apply these methods to their own
  8853. data.",
  8854. journal = "Nat. Neurosci.",
  8855. volume = 17,
  8856. number = 11,
  8857. pages = "1500--1509",
  8858. month = nov,
  8859. year = 2014,
  8860. language = "en"
  8861. }
  8862. @ARTICLE{Bassett2017-qa,
  8863. title = "Network neuroscience",
  8864. author = "Bassett, Danielle S and Sporns, Olaf",
  8865. abstract = "Despite substantial recent progress, our understanding of the
  8866. principles and mechanisms underlying complex brain function and
  8867. cognition remains incomplete. Network neuroscience proposes to
  8868. tackle these enduring challenges. Approaching brain structure and
  8869. function from an explicitly integrative perspective, network
  8870. neuroscience pursues new ways to map, record, analyze and model
  8871. the elements and interactions of neurobiological systems. Two
  8872. parallel trends drive the approach: the availability of new
  8873. empirical tools to create comprehensive maps and record dynamic
  8874. patterns among molecules, neurons, brain areas and social
  8875. systems; and the theoretical framework and computational tools of
  8876. modern network science. The convergence of empirical and
  8877. computational advances opens new frontiers of scientific inquiry,
  8878. including network dynamics, manipulation and control of brain
  8879. networks, and integration of network processes across
  8880. spatiotemporal domains. We review emerging trends in network
  8881. neuroscience and attempt to chart a path toward a better
  8882. understanding of the brain as a multiscale networked system.",
  8883. journal = "Nat. Neurosci.",
  8884. volume = 20,
  8885. number = 3,
  8886. pages = "353--364",
  8887. month = feb,
  8888. year = 2017,
  8889. language = "en"
  8890. }
  8891. @ARTICLE{Engel2019-qn,
  8892. title = "New perspectives on dimensionality and variability from
  8893. large-scale cortical dynamics",
  8894. author = "Engel, Tatiana A and Steinmetz, Nicholas A",
  8895. abstract = "The neocortex is a multi-scale network, with intricate local
  8896. circuitry interwoven into a global mesh of long-range
  8897. connections. Neural activity propagates within this network on a
  8898. wide range of temporal and spatial scales. At the micro scale,
  8899. neurophysiological recordings reveal coordinated dynamics in
  8900. local neural populations, which support behaviorally relevant
  8901. computations. At the macro scale, neuroimaging modalities measure
  8902. global activity fluctuations organized into spatiotemporal
  8903. patterns across the entire brain. Here we review recent advances
  8904. linking the local and global scales of cortical dynamics and
  8905. their relationship to behavior. We argue that diverse
  8906. experimental observations on the dimensionality and variability
  8907. of neural activity can be reconciled by considering how activity
  8908. propagates in space and time on multiple spatial scales.",
  8909. journal = "Curr. Opin. Neurobiol.",
  8910. volume = 58,
  8911. pages = "181--190",
  8912. month = oct,
  8913. year = 2019,
  8914. language = "en"
  8915. }
  8916. @ARTICLE{Kao2019-wk,
  8917. title = "Neuroscience out of control: control-theoretic perspectives on
  8918. neural circuit dynamics",
  8919. author = "Kao, Ta-Chu and Hennequin, Guillaume",
  8920. abstract = "A major challenge in systems neuroscience is to understand how
  8921. the dynamics of neural circuits give rise to behaviour. Analysis
  8922. of complex dynamical systems is also at the heart of control
  8923. engineering, where it is central to the design of robust control
  8924. strategies. Although a rich engineering literature has grown over
  8925. decades to facilitate the analysis of such systems, little of it
  8926. has percolated into neuroscience so far. Here, we give a brief
  8927. introduction to a number of core control-theoretic concepts that
  8928. provide useful perspectives on neural circuit dynamics. We
  8929. introduce important mathematical tools related to these concepts,
  8930. and establish connections to neural circuit analysis, focusing on
  8931. a number of themes that have arisen from the modern 'state-space'
  8932. view on neural population dynamics.",
  8933. journal = "Curr. Opin. Neurobiol.",
  8934. volume = 58,
  8935. pages = "122--129",
  8936. month = oct,
  8937. year = 2019,
  8938. language = "en"
  8939. }
  8940. @ARTICLE{Wilting2019-lp,
  8941. title = "25 years of criticality in neuroscience - established results,
  8942. open controversies, novel concepts",
  8943. author = "Wilting, J and Priesemann, V",
  8944. abstract = "Twenty-five years ago, Dunkelmann and Radons (1994) showed that
  8945. neural networks can self-organize to a critical state. In models,
  8946. the critical state offers a number of computational advantages.
  8947. Thus this hypothesis, and in particular the experimental work by
  8948. Beggs and Plenz (2003), has triggered an avalanche of research,
  8949. with thousands of studies referring to it. Nonetheless,
  8950. experimental results are still contradictory. How is it possible,
  8951. that a hypothesis has attracted active research for decades, but
  8952. nonetheless remains controversial? We discuss the experimental
  8953. and conceptual controversy, and then present a parsimonious
  8954. solution that (i) unifies the contradictory experimental results,
  8955. (ii) avoids disadvantages of a critical state, and (iii) enables
  8956. rapid, adaptive tuning of network properties to task
  8957. requirements.",
  8958. journal = "Curr. Opin. Neurobiol.",
  8959. volume = 58,
  8960. pages = "105--111",
  8961. month = oct,
  8962. year = 2019,
  8963. language = "en"
  8964. }
  8965. @ARTICLE{Sharpee2019-zx,
  8966. title = "An argument for hyperbolic geometry in neural circuits",
  8967. author = "Sharpee, Tatyana O",
  8968. abstract = "This review connects several lines of research to argue that
  8969. hyperbolic geometry should be broadly applicable to neural
  8970. circuits as well as other biological circuits. The reason for
  8971. this is that networks that conform to hyperbolic geometry are
  8972. maximally responsive to external and internal perturbations.
  8973. These networks also allow for efficient communication under
  8974. conditions where nodes are added or removed. We will argue that
  8975. one of the signatures of hyperbolic geometry is the celebrated
  8976. Zipf's law (also sometimes known as the Pareto distribution) that
  8977. states that the probability to observe a given pattern is
  8978. inversely related to its rank. Zipf's law is observed in a
  8979. variety of biological systems - from protein sequences, neural
  8980. networks to economics. These observations provide further
  8981. evidence for the ubiquity of networks with an underlying
  8982. hyperbolic metric structure. Recent studies in neuroscience
  8983. specifically point to the relevance of a three-dimensional
  8984. hyperbolic space for neural signaling. The three-dimensional
  8985. hyperbolic space may confer additional robustness compared to
  8986. other dimensions. We illustrate how the use of hyperbolic
  8987. coordinates revealed a novel topographic organization within the
  8988. olfactory system. The use of such coordinates may facilitate
  8989. representation of relevant signals elsewhere in the brain.",
  8990. journal = "Curr. Opin. Neurobiol.",
  8991. volume = 58,
  8992. pages = "101--104",
  8993. month = oct,
  8994. year = 2019,
  8995. language = "en"
  8996. }
  8997. @ARTICLE{Whiteway2019-fp,
  8998. title = "The quest for interpretable models of neural population activity",
  8999. author = "Whiteway, Matthew R and Butts, Daniel A",
  9000. abstract = "Many aspects of brain function arise from the coordinated
  9001. activity of large populations of neurons. Recent developments in
  9002. neural recording technologies are providing unprecedented access
  9003. to the activity of such populations during increasingly complex
  9004. experimental contexts; however, extracting scientific insights
  9005. from such recordings requires the concurrent development of
  9006. analytical tools that relate this population activity to
  9007. system-level function. This is a primary motivation for latent
  9008. variable models, which seek to provide a low-dimensional
  9009. description of population activity that can be related to
  9010. experimentally controlled variables, as well as uncontrolled
  9011. variables such as internal states (e.g. attention and arousal)
  9012. and elements of behavior. While deriving an understanding of
  9013. function from traditional latent variable methods relies on
  9014. low-dimensional visualizations, new approaches are targeting more
  9015. interpretable descriptions of the components underlying
  9016. system-level function.",
  9017. journal = "Curr. Opin. Neurobiol.",
  9018. volume = 58,
  9019. pages = "86--93",
  9020. month = oct,
  9021. year = 2019,
  9022. language = "en"
  9023. }
  9024. @ARTICLE{Curto2019-ar,
  9025. title = "Relating network connectivity to dynamics: opportunities and
  9026. challenges for theoretical neuroscience",
  9027. author = "Curto, Carina and Morrison, Katherine",
  9028. abstract = "We review recent work relating network connectivity to the
  9029. dynamics of neural activity. While concepts stemming from network
  9030. science provide a valuable starting point, the interpretation of
  9031. graph-theoretic structures and measures can be highly dependent
  9032. on the dynamics associated to the network. Properties that are
  9033. quite meaningful for linear dynamics, such as random walk and
  9034. network flow models, may be of limited relevance in the
  9035. neuroscience setting. Theoretical and computational neuroscience
  9036. are playing a vital role in understanding the relationship
  9037. between network connectivity and the nonlinear dynamics
  9038. associated to neural networks.",
  9039. journal = "Curr. Opin. Neurobiol.",
  9040. volume = 58,
  9041. pages = "11--20",
  9042. month = oct,
  9043. year = 2019,
  9044. language = "en"
  9045. }
  9046. @ARTICLE{Comoli2012-li,
  9047. title = "Segregated anatomical input to sub-regions of the rodent superior
  9048. colliculus associated with approach and defense",
  9049. author = "Comoli, Eliane and Das Neves Favaro, Pl{\'\i}nio and Vautrelle,
  9050. Nicolas and Leriche, Mariana and Overton, Paul G and Redgrave,
  9051. Peter",
  9052. abstract = "The superior colliculus (SC) is responsible for sensorimotor
  9053. transformations required to direct gaze toward or away from
  9054. unexpected, biologically salient events. Significant changes in
  9055. the external world are signaled to SC through primary
  9056. multisensory afferents, spatially organized according to a
  9057. retinotopic topography. For animals, where an unexpected event
  9058. could indicate the presence of either predator or prey, early
  9059. decisions to approach or avoid are particularly important.
  9060. Rodents' ecology dictates predators are most often detected
  9061. initially as movements in upper visual field (mapped in medial
  9062. SC), while appetitive stimuli are normally found in lower visual
  9063. field (mapped in lateral SC). Our purpose was to exploit this
  9064. functional segregation to reveal neural sites that can bias or
  9065. modulate initial approach or avoidance responses. Small
  9066. injections of Fluoro-Gold were made into medial or lateral
  9067. sub-regions of intermediate and deep layers of SC (SCm/SCl). A
  9068. remarkable segregation of input to these two functionally defined
  9069. areas was found. (i) There were structures that projected only to
  9070. SCm (e.g., specific cortical areas, lateral geniculate and
  9071. suprageniculate thalamic nuclei, ventromedial and premammillary
  9072. hypothalamic nuclei, and several brainstem areas) or SCl (e.g.,
  9073. primary somatosensory cortex representing upper body parts and
  9074. vibrissae and parvicellular reticular nucleus in the brainstem).
  9075. (ii) Other structures projected to both SCm and SCl but from
  9076. topographically segregated populations of neurons (e.g., zona
  9077. incerta and substantia nigra pars reticulata). (iii) There were a
  9078. few brainstem areas in which retrogradely labeled neurons were
  9079. spatially overlapping (e.g., pedunculopontine nucleus and locus
  9080. coeruleus). These results indicate significantly more structures
  9081. across the rat neuraxis are in a position to modulate defense
  9082. responses evoked from SCm, and that neural mechanisms modulating
  9083. SC-mediated defense or appetitive behavior are almost entirely
  9084. segregated.",
  9085. journal = "Front. Neuroanat.",
  9086. volume = 6,
  9087. pages = "9",
  9088. month = apr,
  9089. year = 2012,
  9090. keywords = "approach; defense; segregated anatomical inputs; superior
  9091. colliculus",
  9092. language = "en"
  9093. }
  9094. @ARTICLE{Pandarinath2018-lc,
  9095. title = "Inferring single-trial neural population dynamics using
  9096. sequential auto-encoders",
  9097. author = "Pandarinath, Chethan and O'Shea, Daniel J and Collins, Jasmine
  9098. and Jozefowicz, Rafal and Stavisky, Sergey D and Kao, Jonathan C
  9099. and Trautmann, Eric M and Kaufman, Matthew T and Ryu, Stephen I
  9100. and Hochberg, Leigh R and Henderson, Jaimie M and Shenoy, Krishna
  9101. V and Abbott, L F and Sussillo, David",
  9102. abstract = "Neuroscience is experiencing a revolution in which simultaneous
  9103. recording of thousands of neurons is revealing population
  9104. dynamics that are not apparent from single-neuron responses. This
  9105. structure is typically extracted from data averaged across many
  9106. trials, but deeper understanding requires studying phenomena
  9107. detected in single trials, which is challenging due to incomplete
  9108. sampling of the neural population, trial-to-trial variability,
  9109. and fluctuations in action potential timing. We introduce latent
  9110. factor analysis via dynamical systems, a deep learning method to
  9111. infer latent dynamics from single-trial neural spiking data. When
  9112. applied to a variety of macaque and human motor cortical
  9113. datasets, latent factor analysis via dynamical systems accurately
  9114. predicts observed behavioral variables, extracts precise firing
  9115. rate estimates of neural dynamics on single trials, infers
  9116. perturbations to those dynamics that correlate with behavioral
  9117. choices, and combines data from non-overlapping recording
  9118. sessions spanning months to improve inference of underlying
  9119. dynamics.",
  9120. journal = "Nat. Methods",
  9121. volume = 15,
  9122. number = 10,
  9123. pages = "805--815",
  9124. month = oct,
  9125. year = 2018,
  9126. language = "en"
  9127. }
  9128. @ARTICLE{Yu2009-pm,
  9129. title = "Gaussian-process factor analysis for low-dimensional single-trial
  9130. analysis of neural population activity",
  9131. author = "Yu, Byron M and Cunningham, John P and Santhanam, Gopal and Ryu,
  9132. Stephen I and Shenoy, Krishna V and Sahani, Maneesh",
  9133. abstract = "We consider the problem of extracting smooth, low-dimensional
  9134. neural trajectories that summarize the activity recorded
  9135. simultaneously from many neurons on individual experimental
  9136. trials. Beyond the benefit of visualizing the high-dimensional,
  9137. noisy spiking activity in a compact form, such trajectories can
  9138. offer insight into the dynamics of the neural circuitry
  9139. underlying the recorded activity. Current methods for extracting
  9140. neural trajectories involve a two-stage process: the spike trains
  9141. are first smoothed over time, then a static
  9142. dimensionality-reduction technique is applied. We first describe
  9143. extensions of the two-stage methods that allow the degree of
  9144. smoothing to be chosen in a principled way and that account for
  9145. spiking variability, which may vary both across neurons and
  9146. across time. We then present a novel method for extracting neural
  9147. trajectories-Gaussian-process factor analysis (GPFA)-which
  9148. unifies the smoothing and dimensionality-reduction operations in
  9149. a common probabilistic framework. We applied these methods to the
  9150. activity of 61 neurons recorded simultaneously in macaque
  9151. premotor and motor cortices during reach planning and execution.
  9152. By adopting a goodness-of-fit metric that measures how well the
  9153. activity of each neuron can be predicted by all other recorded
  9154. neurons, we found that the proposed extensions improved the
  9155. predictive ability of the two-stage methods. The predictive
  9156. ability was further improved by going to GPFA. From the extracted
  9157. trajectories, we directly observed a convergence in neural state
  9158. during motor planning, an effect that was shown indirectly by
  9159. previous studies. We then show how such methods can be a powerful
  9160. tool for relating the spiking activity across a neural population
  9161. to the subject's behavior on a single-trial basis. Finally, to
  9162. assess how well the proposed methods characterize neural
  9163. population activity when the underlying time course is known, we
  9164. performed simulations that revealed that GPFA performed tens of
  9165. percent better than the best two-stage method.",
  9166. journal = "J. Neurophysiol.",
  9167. volume = 102,
  9168. number = 1,
  9169. pages = "614--635",
  9170. month = jul,
  9171. year = 2009,
  9172. language = "en"
  9173. }
  9174. @ARTICLE{Kim2019-fo,
  9175. title = "Generation of stable heading representations in diverse visual
  9176. scenes",
  9177. author = "Kim, Sung Soo and Hermundstad, Ann M and Romani, Sandro and
  9178. Abbott, L F and Jayaraman, Vivek",
  9179. abstract = "Many animals rely on an internal heading representation when
  9180. navigating in varied environments1--10. How this representation
  9181. is linked to the sensory cues that define different surroundings
  9182. is unclear. In the fly brain, heading is represented by `compass'
  9183. neurons that innervate a ring-shaped structure known as the
  9184. ellipsoid body3,11,12. Each compass neuron receives inputs from
  9185. `ring' neurons that are selective for particular visual
  9186. features13--16; this combination provides an ideal substrate for
  9187. the extraction of directional information from a visual scene.
  9188. Here we combine two-photon calcium imaging and optogenetics in
  9189. tethered flying flies with circuit modelling, and show how the
  9190. correlated activity of compass and visual neurons drives
  9191. plasticity17--22, which flexibly transforms two-dimensional
  9192. visual cues into a stable heading representation. We also
  9193. describe how this plasticity enables the fly to convert a partial
  9194. heading representation, established from orienting within part of
  9195. a novel setting, into a complete heading representation. Our
  9196. results provide mechanistic insight into the memory-related
  9197. computations that are essential for flexible navigation in varied
  9198. surroundings.",
  9199. journal = "Nature",
  9200. month = nov,
  9201. year = 2019
  9202. }
  9203. @ARTICLE{Dabaghian2012-lc,
  9204. title = "A topological paradigm for hippocampal spatial map formation
  9205. using persistent homology",
  9206. author = "Dabaghian, Y and M{\'e}moli, F and Frank, L and Carlsson, G",
  9207. abstract = "An animal's ability to navigate through space rests on its
  9208. ability to create a mental map of its environment. The
  9209. hippocampus is the brain region centrally responsible for such
  9210. maps, and it has been assumed to encode geometric information
  9211. (distances, angles). Given, however, that hippocampal output
  9212. consists of patterns of spiking across many neurons, and
  9213. downstream regions must be able to translate those patterns into
  9214. accurate information about an animal's spatial environment, we
  9215. hypothesized that 1) the temporal pattern of neuronal firing,
  9216. particularly co-firing, is key to decoding spatial information,
  9217. and 2) since co-firing implies spatial overlap of place fields, a
  9218. map encoded by co-firing will be based on connectivity and
  9219. adjacency, i.e., it will be a topological map. Here we test this
  9220. topological hypothesis with a simple model of hippocampal
  9221. activity, varying three parameters (firing rate, place field
  9222. size, and number of neurons) in computer simulations of rat
  9223. trajectories in three topologically and geometrically distinct
  9224. test environments. Using a computational algorithm based on
  9225. recently developed tools from Persistent Homology theory in the
  9226. field of algebraic topology, we find that the patterns of
  9227. neuronal co-firing can, in fact, convey topological information
  9228. about the environment in a biologically realistic length of time.
  9229. Furthermore, our simulations reveal a ``learning region'' that
  9230. highlights the interplay between the parameters in combining to
  9231. produce hippocampal states that are more or less adept at map
  9232. formation. For example, within the learning region a lower number
  9233. of neurons firing can be compensated by adjustments in firing
  9234. rate or place field size, but beyond a certain point map
  9235. formation begins to fail. We propose that this learning region
  9236. provides a coherent theoretical lens through which to view
  9237. conditions that impair spatial learning by altering place cell
  9238. firing rates or spatial specificity.",
  9239. journal = "PLoS Comput. Biol.",
  9240. volume = 8,
  9241. number = 8,
  9242. pages = "e1002581",
  9243. month = aug,
  9244. year = 2012,
  9245. language = "en"
  9246. }
  9247. @ARTICLE{McNaughton2004-qd,
  9248. title = "A two-dimensional neuropsychology of defense: fear/anxiety and
  9249. defensive distance",
  9250. author = "McNaughton, Neil and Corr, Philip J",
  9251. abstract = "We present in this paper a picture of the neural systems
  9252. controlling defense that updates and simplifies Gray's
  9253. ``Neuropsychology of Anxiety''. It is based on two behavioural
  9254. dimensions: 'defensive distance' as defined by the Blanchards
  9255. and 'defensive direction'. Defensive direction is a categorical
  9256. dimension with avoidance of threat corresponding to fear and
  9257. approach to threat corresponding to anxiety. These two
  9258. psychological dimensions are mapped to underlying neural
  9259. dimensions. Defensive distance is mapped to neural level, with
  9260. the shortest defensive distances involving the lowest neural
  9261. level (periaqueductal grey) and the largest defensive distances
  9262. the highest neural level (prefrontal cortex). Defensive
  9263. direction is mapped to separate parallel streams that run across
  9264. these levels. A significant departure from prior models is the
  9265. proposal that both fear and anxiety are represented at all
  9266. levels. The theory is presented in a simplified form that does
  9267. not incorporate the interactions that must occur between
  9268. non-adjacent levels of the system. It also requires expansion to
  9269. include the dimension of escapability of threat. Our current
  9270. development and these proposed future extensions do not change
  9271. the core concepts originally proposed by Gray and, we argue,
  9272. demonstrate their enduring value.",
  9273. journal = "Neurosci. Biobehav. Rev.",
  9274. publisher = "Elsevier",
  9275. volume = 28,
  9276. number = 3,
  9277. pages = "285--305",
  9278. month = may,
  9279. year = 2004,
  9280. language = "en"
  9281. }
  9282. @ARTICLE{Martin2011-fc,
  9283. title = "Molecular and neuroanatomical characterization of single neurons
  9284. in the mouse medullary gigantocellular reticular nucleus",
  9285. author = "Martin, E M and Devidze, N and Shelley, D N and Westberg, L and
  9286. Fontaine, C and Pfaff, D W",
  9287. abstract = "Medullary gigantocellular reticular nucleus (mGi) neurons have
  9288. been ascribed a variety of behaviors, many of which may fall
  9289. under the concepts of either arousal or motivation. Despite this,
  9290. many details of the connectivity of mGi neurons, particularly in
  9291. reference to those neurons with ascending axons, remain unknown.
  9292. To provide a neuroanatomical and molecular characterization of
  9293. these cells, with reference to arousal and level-setting systems,
  9294. large medullary reticular neurons were characterized with
  9295. retrograde dye techniques and with real-time reverse
  9296. transcriptase PCR (RT-PCR) analyses of single-neuron mRNA
  9297. expression in the mouse. We have shown that receptors consistent
  9298. with participation in generalized arousal are expressed by single
  9299. mGi neurons and that receptors from different families of
  9300. arousal-related neurotransmitters are rarely coexpressed. Through
  9301. retrograde labeling, we have shown that neurons with ascending
  9302. axons and neurons with descending axons tend to form
  9303. like-with-like clusters, a finding that is consistent across age
  9304. and gender. In comparing the two groups of retrogradely labeled
  9305. neurons in neonatal animals, those neurons with axons that ascend
  9306. to the midbrain show markers for GABAergic or coincident
  9307. GABAergic and glutamatergic function; in contrast, approximately
  9308. 60\% of the neurons with axons that descend to the spinal cord
  9309. are glutamatergic. We discuss the mGi's relationship to the
  9310. voluntary and emotional motor systems and speculate that neurons
  9311. in the mGi may represent a mammalian analogue to Mauthner cells,
  9312. with a separation of function for neurons with ascending and
  9313. descending axons.",
  9314. journal = "J. Comp. Neurol.",
  9315. volume = 519,
  9316. number = 13,
  9317. pages = "2574--2593",
  9318. month = sep,
  9319. year = 2011,
  9320. language = "en"
  9321. }
  9322. @ARTICLE{Liang2016-po,
  9323. title = "Terminations of reticulospinal fibers originating from the
  9324. gigantocellular reticular formation in the mouse spinal cord",
  9325. author = "Liang, Huazheng and Watson, Charles and Paxinos, George",
  9326. abstract = "The present study investigated the projections of the
  9327. gigantocellular reticular nucleus (Gi) and its neighbors--the
  9328. dorsal paragigantocellular reticular nucleus (DPGi), the
  9329. alpha/ventral part of the gigantocellular reticular nucleus
  9330. (GiA/V), and the lateral paragigantocellular reticular nucleus
  9331. (LPGi)--to the mouse spinal cord by injecting the anterograde
  9332. tracer biotinylated dextran amine (BDA) into the Gi, DPGi,
  9333. GiA/GiV, and LPGi. The Gi projected to the entire spinal cord
  9334. bilaterally with an ipsilateral predominance. Its fibers traveled
  9335. in both the ventral and lateral funiculi with a greater presence
  9336. in the ventral funiculus. As the fibers descended in the spinal
  9337. cord, their density in the lateral funiculus increased. The
  9338. terminals were present mainly in laminae 7-10 with a dorsolateral
  9339. expansion caudally. In the lumbar and sacral cord, a considerable
  9340. number of terminals were also present in laminae 5 and 6.
  9341. Contralateral fibers shared a similar pattern to their
  9342. ipsilateral counterparts and some fibers were seen to cross the
  9343. midline. Fibers arising from the DPGi were similarly distributed
  9344. in the spinal cord except that there was no dorsolateral
  9345. expansion in the lumbar and sacral segments and there were fewer
  9346. fiber terminals. Fibers arising from GiA/V predominantly traveled
  9347. in the ventral and lateral funiculi ipsilaterally. Ipsilaterally,
  9348. the density of fibers in the ventral funiculus decreased along
  9349. the rostrocaudal axis, whereas the density of fibers in the
  9350. lateral funiculus increased. They terminate mainly in the medial
  9351. ventral horn and lamina 10 with a smaller number of fibers in the
  9352. dorsal horn. Fibers arising from the LPGi traveled in both the
  9353. ventral and lateral funiculi and the density of these fibers in
  9354. the ventral and lateral funiculi decreased dramatically in the
  9355. lumbar and sacral segments. Their terminals were present in the
  9356. ventral horn with a large portion of them terminating in the
  9357. motor neuron columns. The present study is the first
  9358. demonstration of the termination pattern of fibers arising from
  9359. the Gi, DPGi, GiA/GiV, and LPGi in the mouse spinal cord. It
  9360. provides an anatomical foundation for those who are conducting
  9361. spinal cord injury and locomotion related research.",
  9362. journal = "Brain Struct. Funct.",
  9363. volume = 221,
  9364. number = 3,
  9365. pages = "1623--1633",
  9366. month = apr,
  9367. year = 2016,
  9368. keywords = "Blood pressure control; Gigantocellular reticular nucleus;
  9369. Locomotion; Medullary reticulospinal tract; Paragigantocellular
  9370. nucleus; Spinal cord",
  9371. language = "en"
  9372. }
  9373. @ARTICLE{Mobbs2018-li,
  9374. title = "Foraging for foundations in decision neuroscience: insights from
  9375. ethology",
  9376. author = "Mobbs, Dean and Trimmer, Pete C and Blumstein, Daniel T and
  9377. Dayan, Peter",
  9378. abstract = "Modern decision neuroscience offers a powerful and broad account
  9379. of human behaviour using computational techniques that link
  9380. psychological and neuroscientific approaches to the ways that
  9381. individuals can generate near-optimal choices in complex
  9382. controlled environments. However, until recently, relatively
  9383. little attention has been paid to the extent to which the
  9384. structure of experimental environments relates to natural
  9385. scenarios, and the survival problems that individuals have
  9386. evolved to solve. This situation not only risks leaving
  9387. decision-theoretic accounts ungrounded but also makes various
  9388. aspects of the solutions, such as hard-wired or Pavlovian
  9389. policies, difficult to interpret in the natural world. Here, we
  9390. suggest importing concepts, paradigms and approaches from the
  9391. fields of ethology and behavioural ecology, which concentrate on
  9392. the contextual and functional correlates of decisions made about
  9393. foraging and escape and address these lacunae.",
  9394. journal = "Nat. Rev. Neurosci.",
  9395. volume = 19,
  9396. number = 7,
  9397. pages = "419--427",
  9398. month = jul,
  9399. year = 2018,
  9400. language = "en"
  9401. }
  9402. % The entry below contains non-ASCII chars that could not be converted
  9403. % to a LaTeX equivalent.
  9404. @ARTICLE{Cooper1998-ii,
  9405. title = "Superior colliculus and active navigation: Role of visual and
  9406. non-visual cues in controlling cellular representations of space",
  9407. author = "Cooper, B G and Miya, D Y and Mizumori, S J Y",
  9408. abstract = "To begin investigation of the contribution of the superior
  9409. colliculus to unrestrained navigation, the nature of behavioral
  9410. representation by individual neurons was identified as rats
  9411. performed a spatial memory task. Similar to what has been
  9412. observed for hippocampus, many superior collicular cells showed
  9413. elevated firing as animals traversed particular locations on the
  9414. maze, and also during directional movement. However, when
  9415. compared to hippocampal place fields, superior collicular
  9416. location fields were found to be more broad …",
  9417. journal = "Hippocampus",
  9418. publisher = "Wiley Online Library",
  9419. volume = 8,
  9420. number = 4,
  9421. pages = "340--372",
  9422. year = 1998
  9423. }
  9424. @ARTICLE{Kraskov2004-de,
  9425. title = "Estimating mutual information",
  9426. author = "Kraskov, Alexander and St{\"o}gbauer, Harald and Grassberger,
  9427. Peter",
  9428. abstract = "We present two classes of improved estimators for mutual
  9429. information M(X,Y), from samples of random points distributed
  9430. according to some joint probability density mu(x,y). In contrast
  9431. to conventional estimators based on binnings, they are based on
  9432. entropy estimates from k -nearest neighbor distances. This means
  9433. that they are data efficient (with k=1 we resolve structures down
  9434. to the smallest possible scales), adaptive (the resolution is
  9435. higher where data are more numerous), and have minimal bias.
  9436. Indeed, the bias of the underlying entropy estimates is mainly
  9437. due to nonuniformity of the density at the smallest resolved
  9438. scale, giving typically systematic errors which scale as
  9439. functions of k/N for N points. Numerically, we find that both
  9440. families become exact for independent distributions, i.e. the
  9441. estimator M(X,Y) vanishes (up to statistical fluctuations) if
  9442. mu(x,y)=mu(x)mu(y). This holds for all tested marginal
  9443. distributions and for all dimensions of x and y. In addition, we
  9444. give estimators for redundancies between more than two random
  9445. variables. We compare our algorithms in detail with existing
  9446. algorithms. Finally, we demonstrate the usefulness of our
  9447. estimators for assessing the actual independence of components
  9448. obtained from independent component analysis (ICA), for improving
  9449. ICA, and for estimating the reliability of blind source
  9450. separation.",
  9451. journal = "Phys. Rev. E Stat. Nonlin. Soft Matter Phys.",
  9452. volume = 69,
  9453. number = "6 Pt 2",
  9454. pages = "066138",
  9455. month = jun,
  9456. year = 2004,
  9457. language = "en"
  9458. }
  9459. @ARTICLE{Sooksawate2013-hx,
  9460. title = "Viral vector-mediated selective and reversible blockade of the
  9461. pathway for visual orienting in mice",
  9462. author = "Sooksawate, Thongchai and Isa, Kaoru and Matsui, Ryosuke and
  9463. Kato, Shigeki and Kinoshita, Masaharu and Kobayashi, Kenta and
  9464. Watanabe, Dai and Kobayashi, Kazuto and Isa, Tadashi",
  9465. abstract = "Recently, by using a combination of two viral vectors, we
  9466. developed a technique for pathway-selective and reversible
  9467. synaptic transmission blockade, and successfully induced a
  9468. behavioral deficit of dexterous hand movements in macaque monkeys
  9469. by affecting a population of spinal interneurons. To explore the
  9470. capacity of this technique to work in other pathways and species,
  9471. and to obtain fundamental methodological information, we tried to
  9472. block the crossed tecto-reticular pathway, which is known to
  9473. control orienting responses to visual targets, in mice. A
  9474. neuron-specific retrograde gene transfer vector with the gene
  9475. encoding enhanced tetanus neurotoxin (eTeNT) tagged with enhanced
  9476. green fluorescent protein (EGFP) under the control of a
  9477. tetracycline responsive element was injected into the left medial
  9478. pontine reticular formation. 7-17 days later, an adeno-associated
  9479. viral vector with a highly efficient Tet-ON sequence, rtTAV16,
  9480. was injected into the right superior colliculus. 5-9 weeks later,
  9481. the daily administration of doxycycline (Dox) was initiated.
  9482. Visual orienting responses toward the left side were impaired 1-4
  9483. days after Dox administration. Anti-GFP immunohistochemistry
  9484. revealed that a number of neurons in the intermediate and deep
  9485. layers of the right superior colliculus were positively stained,
  9486. indicating eTeNT expression. After the termination of Dox
  9487. administration, the anti-GFP staining returned to the baseline
  9488. level within 28 days. A second round of Dox administration,
  9489. starting from 28 days after the termination of the first Dox
  9490. administration, resulted in the reappearance of the behavioral
  9491. impairment. These findings showed that pathway-selective and
  9492. reversible blockade of synaptic transmission also causes
  9493. behavioral effects in rodents, and that the crossed
  9494. tecto-reticular pathway clearly controls visual orienting
  9495. behaviors.",
  9496. journal = "Front. Neural Circuits",
  9497. volume = 7,
  9498. pages = "162",
  9499. month = oct,
  9500. year = 2013,
  9501. keywords = "Tet-ON; mouse; orienting behavior; pontine reticular formation;
  9502. superior colliculus; tetanus neurotoxin; viral vector",
  9503. language = "en"
  9504. }
  9505. @ARTICLE{Felsen2008-nl,
  9506. title = "Neural substrates of sensory-guided locomotor decisions in the
  9507. rat superior colliculus",
  9508. author = "Felsen, Gidon and Mainen, Zachary F",
  9509. abstract = "Deciding in which direction to move is a ubiquitous feature of
  9510. animal behavior, but the neural substrates of locomotor choices
  9511. are not well understood. The superior colliculus (SC) is a
  9512. midbrain structure known to be important for controlling the
  9513. direction of gaze, particularly when guided by visual or auditory
  9514. cues, but which may play a more general role in behavior
  9515. involving spatial orienting. To test this idea, we recorded and
  9516. manipulated activity in the SC of freely moving rats performing
  9517. an odor-guided spatial choice task. In this context, not only did
  9518. a substantial majority of SC neurons encode choice direction
  9519. during goal-directed locomotion, but many also predicted the
  9520. upcoming choice and maintained selectivity for it after movement
  9521. completion. Unilateral inactivation of SC activity profoundly
  9522. altered spatial choices. These results indicate that the SC
  9523. processes information necessary for spatial locomotion,
  9524. suggesting a broad role for this structure in sensory-guided
  9525. orienting and navigation.",
  9526. journal = "Neuron",
  9527. volume = 60,
  9528. number = 1,
  9529. pages = "137--148",
  9530. month = oct,
  9531. year = 2008,
  9532. keywords = "Locomotion",
  9533. language = "en"
  9534. }
  9535. @ARTICLE{Dean1986-of,
  9536. title = "Head and body movements produced by electrical stimulation of
  9537. superior colliculus in rats: effects of interruption of crossed
  9538. tectoreticulospinal pathway",
  9539. author = "Dean, P and Redgrave, P and Sahibzada, N and Tsuji, K",
  9540. abstract = "Stimulation of the superior colliculus in rats produces movements
  9541. of the head and body that resemble either orientation and
  9542. approach towards a contralateral stimulus, or avoidance of, or
  9543. escape from, such a stimulus. A variety of evidence indicates
  9544. that the crossed descending pathway, which runs in the
  9545. contralateral predorsal bundle to the pontomedullary reticular
  9546. formation and the spinal cord, is involved in orienting
  9547. movements. The nature of this involvement was investigated, by
  9548. assessing the effects on tectally-elicited movements of midbrain
  9549. knife-cuts intended to section the pathway as it crosses midline
  9550. in the dorsal tegmental decussation. As expected, ipsilateral
  9551. movements resembling avoidance or escape were little affected by
  9552. dorsal tegmental decussation section, whereas contralateral
  9553. circling movements of the body were almost abolished. However,
  9554. contralateral movements of the head in response to electrical
  9555. stimulation were not eliminated, nor were orienting head
  9556. movements to visual or tactile stimuli. There was some suggestion
  9557. that section of the dorsal tegmental decussation increased the
  9558. latency of head movements from electrical stimulation at lateral
  9559. sites, and decreased the accuracy of orienting movements to
  9560. sensory stimuli. These results support the view that the crossed
  9561. tectoreticulospinal system is concerned with approach rather than
  9562. avoidance movements. However, it appears that other, as yet
  9563. unidentified, tectal efferent systems are also involved in
  9564. orienting head movements. It is possible that this division of
  9565. labour may reflect functional differences between various kinds
  9566. of apparently similar orienting responses. One suggestion is that
  9567. the tectoreticulospinal system is concerned less in open-loop
  9568. orienting responses (that are initiated but not subsequently
  9569. guided by sensory stimuli), than in following or pursuit
  9570. movements.",
  9571. journal = "Neuroscience",
  9572. volume = 19,
  9573. number = 2,
  9574. pages = "367--380",
  9575. month = oct,
  9576. year = 1986,
  9577. language = "en"
  9578. }
  9579. @ARTICLE{Masullo2019-mk,
  9580. title = "Genetically Defined Functional Modules for Spatial Orienting in
  9581. the Mouse Superior Colliculus",
  9582. author = "Masullo, Laura and Mariotti, Letizia and Alexandre, Nicolas and
  9583. Freire-Pritchett, Paula and Boulanger, Jerome and Tripodi, Marco",
  9584. abstract = "Summary In order to explore and interact with their
  9585. surroundings, animals need to orient toward specific positions
  9586. in space. Throughout the animal kingdom, head movements
  9587. represent a primary form of orienting behavior. The superior
  9588. colliculus (SC) is a fundamental structure for the generation of
  9589. orienting responses, but how genetically distinct groups of
  9590. collicular neurons contribute to these spatially tuned behaviors
  9591. remains largely to be defined. Here, through the genetic
  9592. dissection of the murine SC, we identify a functionally and
  9593. genetically homogeneous subclass of glutamatergic neurons
  9594. defined by the expression of the paired-like homeodomain
  9595. transcription factor Pitx2. We show that the optogenetic
  9596. stimulation of Pitx2ON neurons drives three-dimensional head
  9597. displacements characterized by stepwise, saccade-like
  9598. kinematics. Furthermore, during naturalistic foraging behavior,
  9599. the activity of Pitx2ON neurons precedes and predicts the onset
  9600. of spatially tuned head movements. Intriguingly, we reveal that
  9601. Pitx2ON neurons are clustered in an orderly array of anatomical
  9602. modules that tile the entire intermediate layer of the SC. Such
  9603. a modular organization gives origin to a discrete and
  9604. discontinuous representation of the motor space, with each
  9605. Pitx2ON module subtending a defined portion of the animal's
  9606. egocentric space. The modularity of Pitx2ON neurons provides an
  9607. anatomical substrate for the convergence of spatially coherent
  9608. sensory and motor signals of cortical and subcortical origins,
  9609. thereby promoting the recruitment of appropriate movement
  9610. vectors. Overall, these data support the view of the superior
  9611. colliculus as a selectively addressable and modularly organized
  9612. spatial-motor register.",
  9613. journal = "Curr. Biol.",
  9614. publisher = "Elsevier",
  9615. volume = 29,
  9616. number = 17,
  9617. pages = "2892--2904.e8",
  9618. month = sep,
  9619. year = 2019,
  9620. keywords = "superior colliculus; Pitx2; motor control; head movement;
  9621. orienting behaviour"
  9622. }
  9623. @ARTICLE{Wilson2018-yh,
  9624. title = "{Three-Dimensional} Representation of Motor Space in the Mouse
  9625. Superior Colliculus",
  9626. author = "Wilson, Jonathan J and Alexandre, Nicolas and Trentin, Caterina
  9627. and Tripodi, Marco",
  9628. abstract = "From the act of exploring an environment to that of grasping a
  9629. cup of tea, animals must put in register their motor acts with
  9630. their surrounding space. In the motor domain, this is likely to
  9631. be defined by a register of three-dimensional (3D) displacement
  9632. vectors, whose recruitment allows motion in the direction of a
  9633. target. One such spatially targeted action is seen in the head
  9634. reorientation behavior of mice, yet the neural mechanisms
  9635. underlying these 3D behaviors remain unknown. Here, by
  9636. developing a head-mounted inertial sensor for studying 3D head
  9637. rotations and combining it with electrophysiological recordings,
  9638. we show that neurons in the mouse superior colliculus are either
  9639. individually or conjunctively tuned to the three Eulerian
  9640. components of head rotation. The average displacement vectors
  9641. associated with motor-tuned colliculus neurons remain stable
  9642. over time and are unaffected by changes in firing rate or the
  9643. duration of spike trains. Finally, we show that the motor tuning
  9644. of collicular neurons is largely independent from visual or
  9645. landmark cues. By describing the 3D nature of motor tuning in
  9646. the superior colliculus, we contribute to long-standing debate
  9647. on the dimensionality of collicular motor decoding; furthermore,
  9648. by providing an experimental paradigm for the study of the
  9649. metric of motor tuning in mice, this study also paves the way to
  9650. the genetic dissection of the circuits underlying spatially
  9651. targeted motion.",
  9652. journal = "Curr. Biol.",
  9653. publisher = "Elsevier",
  9654. volume = 28,
  9655. number = 11,
  9656. pages = "1744--1755.e12",
  9657. month = jun,
  9658. year = 2018,
  9659. keywords = "3D; motor control; space encoding; superior colliculus",
  9660. language = "en"
  9661. }
  9662. @ARTICLE{Tenenbaum2000-kv,
  9663. title = "A global geometric framework for nonlinear dimensionality
  9664. reduction",
  9665. author = "Tenenbaum, J B and de Silva, V and Langford, J C",
  9666. abstract = "Scientists working with large volumes of high-dimensional data,
  9667. such as global climate patterns, stellar spectra, or human gene
  9668. distributions, regularly confront the problem of dimensionality
  9669. reduction: finding meaningful low-dimensional structures hidden
  9670. in their high-dimensional observations. The human brain confronts
  9671. the same problem in everyday perception, extracting from its
  9672. high-dimensional sensory inputs-30,000 auditory nerve fibers or
  9673. 10(6) optic nerve fibers-a manageably small number of
  9674. perceptually relevant features. Here we describe an approach to
  9675. solving dimensionality reduction problems that uses easily
  9676. measured local metric information to learn the underlying global
  9677. geometry of a data set. Unlike classical techniques such as
  9678. principal component analysis (PCA) and multidimensional scaling
  9679. (MDS), our approach is capable of discovering the nonlinear
  9680. degrees of freedom that underlie complex natural observations,
  9681. such as human handwriting or images of a face under different
  9682. viewing conditions. In contrast to previous algorithms for
  9683. nonlinear dimensionality reduction, ours efficiently computes a
  9684. globally optimal solution, and, for an important class of data
  9685. manifolds, is guaranteed to converge asymptotically to the true
  9686. structure.",
  9687. journal = "Science",
  9688. volume = 290,
  9689. number = 5500,
  9690. pages = "2319--2323",
  9691. month = dec,
  9692. year = 2000,
  9693. language = "en"
  9694. }
  9695. @ARTICLE{Hebert2019-gx,
  9696. title = "Inexperienced preys know when to flee or to freeze in front of a
  9697. threat",
  9698. author = "H{\'e}bert, Marie and Versace, Elisabetta and Vallortigara,
  9699. Giorgio",
  9700. abstract = "Using appropriate antipredatory responses is crucial for
  9701. survival. While slowing down reduces the chances of being
  9702. detected from distant predators, fleeing away is advantageous in
  9703. front of an approaching predator. Whether appropriate responses
  9704. depend on experience with moving objects is still an open
  9705. question. To clarify whether adopting appropriate fleeing or
  9706. freezing responses requires previous experience, we investigated
  9707. responses of chicks naive to movement. When exposed to the moving
  9708. cues mimicking an approaching predator (a rapidly expanding,
  9709. looming stimulus), chicks displayed a fast escape response. In
  9710. contrast, when presented with a distal threat (a small stimulus
  9711. sweeping overhead) they decreased their speed, a maneuver useful
  9712. to avoid detection. The fast expansion of the stimulus toward the
  9713. subject, rather than its size per se or change in luminance,
  9714. triggered the escape response. These results show that young
  9715. animals, in the absence of previous experience, can use motion
  9716. cues to select the appropriate responses to different threats.
  9717. The adaptive needs of young preys are thus matched by spontaneous
  9718. defensive mechanisms that do not require learning.",
  9719. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  9720. month = oct,
  9721. year = 2019,
  9722. keywords = "antipredatory behaviors; defense strategies; motion cues; naive
  9723. animals; threat detection",
  9724. language = "en"
  9725. }
  9726. @ARTICLE{Shin2019-hi,
  9727. title = "Dynamics of Awake {Hippocampal-Prefrontal} Replay for Spatial
  9728. Learning and {Memory-Guided} Decision Making",
  9729. author = "Shin, Justin D and Tang, Wenbo and Jadhav, Shantanu P",
  9730. abstract = "SummarySpatial learning requires remembering and choosing paths
  9731. to goals. Hippocampal place cells replay spatial paths during
  9732. immobility in reverse and forward order, offering a potential
  9733. mechanism. However, how replay supports both goal-directed
  9734. learning and memory-guided decision making is unclear. We
  9735. therefore continuously tracked awake replay in the same
  9736. hippocampal-prefrontal ensembles throughout learning of a
  9737. spatial alternation task. We found that, during pauses between
  9738. behavioral trajectories, reverse and forward hippocampal replay
  9739. supports an internal cognitive search of available past and
  9740. future possibilities and exhibits opposing learning gradients
  9741. for prediction of past and future behavioral paths,
  9742. respectively. Coordinated hippocampal-prefrontal replay
  9743. distinguished correct past and future paths from alternative
  9744. choices, suggesting a role in recall of past paths to guide
  9745. planning of future decisions for spatial working memory. Our
  9746. findings reveal a learning shift from hippocampal
  9747. reverse-replay-based retrospective evaluation to
  9748. forward-replay-based prospective planning, with prefrontal
  9749. readout of memory-guided paths for learning and decision making.",
  9750. journal = "Neuron",
  9751. publisher = "Elsevier",
  9752. volume = 0,
  9753. number = 0,
  9754. month = oct,
  9755. year = 2019,
  9756. keywords = "hippocampus; prefrontal cortex; replay; sharp-wave ripple;
  9757. spatial learning; decision making; working memory; prospection;
  9758. retrospection; planning",
  9759. language = "en"
  9760. }
  9761. @ARTICLE{Harris2019-ht,
  9762. title = "Hierarchical organization of cortical and thalamic connectivity",
  9763. author = "Harris, Julie A and Mihalas, Stefan and Hirokawa, Karla E and
  9764. Whitesell, Jennifer D and Choi, Hannah and Bernard, Amy and Bohn,
  9765. Phillip and Caldejon, Shiella and Casal, Linzy and Cho, Andrew
  9766. and Feiner, Aaron and Feng, David and Gaudreault, Nathalie and
  9767. Gerfen, Charles R and Graddis, Nile and Groblewski, Peter A and
  9768. Henry, Alex M and Ho, Anh and Howard, Robert and Knox, Joseph E
  9769. and Kuan, Leonard and Kuang, Xiuli and Lecoq, Jerome and Lesnar,
  9770. Phil and Li, Yaoyao and Luviano, Jennifer and McConoughey,
  9771. Stephen and Mortrud, Marty T and Naeemi, Maitham and Ng, Lydia
  9772. and Oh, Seung Wook and Ouellette, Benjamin and Shen, Elise and
  9773. Sorensen, Staci A and Wakeman, Wayne and Wang, Quanxin and Wang,
  9774. Yun and Williford, Ali and Phillips, John W and Jones, Allan R
  9775. and Koch, Christof and Zeng, Hongkui",
  9776. abstract = "The mammalian cortex is a laminar structure containing many areas
  9777. and cell types that are densely interconnected in complex ways,
  9778. and for which generalizable principles of organization remain
  9779. mostly unknown. Here we describe a major expansion of the Allen
  9780. Mouse Brain Connectivity Atlas resource1, involving around a
  9781. thousand new tracer experiments in the cortex and its main
  9782. satellite structure, the thalamus. We used Cre driver lines (mice
  9783. expressing Cre recombinase) to comprehensively and selectively
  9784. label brain-wide connections by layer and class of projection
  9785. neuron. Through observations of axon termination patterns, we
  9786. have derived a set of generalized anatomical rules to describe
  9787. corticocortical, thalamocortical and corticothalamic projections.
  9788. We have built a model to assign connection patterns between areas
  9789. as either feedforward or feedback, and generated testable
  9790. predictions of hierarchical positions for individual cortical and
  9791. thalamic areas and for cortical network modules. Our results show
  9792. that cell-class-specific connections are organized in a shallow
  9793. hierarchy within the mouse corticothalamic network.",
  9794. journal = "Nature",
  9795. month = oct,
  9796. year = 2019,
  9797. language = "en"
  9798. }
  9799. @ARTICLE{Buel1935-sn,
  9800. title = "Differential errors in animal mazes",
  9801. author = "Buel, J",
  9802. abstract = "93 different possible factors that may determine the
  9803. differential errors made in maze-running are drawn from a
  9804. literature of 111 titles, and grouped under the headings:
  9805. genetic make-up, physiological determiners, physical
  9806. determiners, route and blind preferences, goal functions,
  9807. emotional factors, extra-maze, pre-maze, and temporal factors,
  9808. maze structure and pattern, general orientation, expectancy,
  9809. organizing factors. A discussion by the author follows.
  9810. (PsycINFO Database Record (c) 2016 APA, all rights reserved)",
  9811. journal = "Psychol. Bull.",
  9812. publisher = "psycnet.apa.org",
  9813. volume = 32,
  9814. number = 1,
  9815. pages = "67--99",
  9816. month = jan,
  9817. year = 1935
  9818. }
  9819. % The entry below contains non-ASCII chars that could not be converted
  9820. % to a LaTeX equivalent.
  9821. @ARTICLE{Tolman1938-aa,
  9822. title = "The determiners of behavior at a choice point",
  9823. author = "Tolman, Edward Chace",
  9824. abstract = "An analysis of the complex of causal determinants on which a
  9825. rat's behavior of turning right or left depends shows them to be
  9826. divided into`` environmental variables,'' such as maintenance
  9827. schedule, appropriate goal-object, types of stimuli provided and
  9828. responses …",
  9829. journal = "Psychol. Rev.",
  9830. publisher = "American Psychological Association",
  9831. volume = 45,
  9832. number = 1,
  9833. pages = "1",
  9834. year = 1938
  9835. }
  9836. % The entry below contains non-ASCII chars that could not be converted
  9837. % to a LaTeX equivalent.
  9838. @ARTICLE{W_R_Boyce_Gibson1900-dh,
  9839. title = "The Principle of Least Action as a Psychological Principle",
  9840. author = "{W. R. Boyce Gibson}",
  9841. abstract = "Action, Leipzig, 1877. 2lMYcanique Analytique, p. 246.'In this
  9842. respect the Principle of Least Action is found wanting;* f.
  9843. Bartholomew Price, Infinitesimal Calculus, vol. iv., p. 150.
  9844. 4Lagrange,(UEluvres, ed. Serret, vol. i., p. 365), in a sequel
  9845. to a paper of his Essai d'unte nouvelle methode pour determiner
  9846. le. s maxima et les minima des formules intJgrales indefinies. 5
  9847. H. v. Helmholtz,`` Iber die physikalische Bedentung des Princips
  9848. der Kleinsten Wirkung,'' Journal ffir die reine und angewandte
  9849. Mathemiatik (usually known as Crelle's …",
  9850. journal = "Mind",
  9851. publisher = "[Oxford University Press, Mind Association]",
  9852. volume = 9,
  9853. number = 36,
  9854. pages = "469--495",
  9855. year = 1900
  9856. }
  9857. @ARTICLE{Gengerelli1930-zo,
  9858. title = "The principle of maxima and minima in animal learning",
  9859. author = "Gengerelli, J A",
  9860. abstract = "A series of experiments were performed with blinded and normal
  9861. white rats to determine the nature of the path which the animals
  9862. would eventually select from an indefinite number of possible
  9863. paths leading to food. The animals entered by one corner of the
  9864. platform 6 ft. by 6 ft. which was enclosed on all sides by a 5
  9865. in. wall and covered with wire mesh. Food was placed at the
  9866. diagonally opposite corner to the entrance corner, in a small
  9867. food box which was outside of the platform and which the animal
  9868. entered by means of a short alley leading to it. Both food and
  9869. observer were invisible to the animal. It was found in
  9870. practically all cases that the path finally chosen by the
  9871. animals (both normal and blinded), when all possibility of
  9872. olfactory orientation was eliminated, was the path of ``least
  9873. effort,'' namely, the path whose distance was a minimum. In the
  9874. experiments where there were no obstructions on the platform,
  9875. the actual path finally chosen was the diagonal from the
  9876. entrance corner to the food corner. The behavior of the rats can
  9877. be most accurately described by stating that the path which the
  9878. animals finally chose served as a limit to which they
  9879. approximated more and more closely with each successive trail.
  9880. (PsycINFO Database Record (c) 2016 APA, all rights reserved)",
  9881. journal = "J. Comp. Psychol.",
  9882. publisher = "psycnet.apa.org",
  9883. volume = 11,
  9884. number = 2,
  9885. pages = "193--236",
  9886. month = dec,
  9887. year = 1930
  9888. }
  9889. @ARTICLE{De_Camp1920-jm,
  9890. title = "Relative distance as a factor in the white rat's selection of a
  9891. path",
  9892. author = "De Camp, Joseph Edgar",
  9893. abstract = "Elimination of errors and decrease in length of path from
  9894. starting point to goal (usually food) are characteristic of
  9895. animal learning. Of two paths leading to food, one being longer,
  9896. the animal soon chooses the shorter. This has been observed with
  9897. white rats in their learning of mazes. This article is a report
  9898. of an attempt to study this selection by the white rat of the
  9899. shorter of two paths.(PsycINFO Database Record (c) 2017 APA, all
  9900. rights reserved)",
  9901. journal = "Psychobiology",
  9902. publisher = "Williams \& Wilkins Company",
  9903. volume = 2,
  9904. number = 3,
  9905. pages = "245",
  9906. year = 1920
  9907. }
  9908. @ARTICLE{Richards2019-sy,
  9909. title = "A deep learning framework for neuroscience",
  9910. author = "Richards, Blake A and Lillicrap, Timothy P and Beaudoin, Philippe
  9911. and Bengio, Yoshua and Bogacz, Rafal and Christensen, Amelia and
  9912. Clopath, Claudia and Costa, Rui Ponte and de Berker, Archy and
  9913. Ganguli, Surya and Gillon, Colleen J and Hafner, Danijar and
  9914. Kepecs, Adam and Kriegeskorte, Nikolaus and Latham, Peter and
  9915. Lindsay, Grace W and Miller, Kenneth D and Naud, Richard and
  9916. Pack, Christopher C and Poirazi, Panayiota and Roelfsema, Pieter
  9917. and Sacramento, Jo{\~a}o and Saxe, Andrew and Scellier, Benjamin
  9918. and Schapiro, Anna C and Senn, Walter and Wayne, Greg and Yamins,
  9919. Daniel and Zenke, Friedemann and Zylberberg, Joel and Therien,
  9920. Denis and Kording, Konrad P",
  9921. abstract = "Systems neuroscience seeks explanations for how the brain
  9922. implements a wide variety of perceptual, cognitive and motor
  9923. tasks. Conversely, artificial intelligence attempts to design
  9924. computational systems based on the tasks they will have to solve.
  9925. In artificial neural networks, the three components specified by
  9926. design are the objective functions, the learning rules and the
  9927. architectures. With the growing success of deep learning, which
  9928. utilizes brain-inspired architectures, these three designed
  9929. components have increasingly become central to how we model,
  9930. engineer and optimize complex artificial learning systems. Here
  9931. we argue that a greater focus on these components would also
  9932. benefit systems neuroscience. We give examples of how this
  9933. optimization-based framework can drive theoretical and
  9934. experimental progress in neuroscience. We contend that this
  9935. principled perspective on systems neuroscience will help to
  9936. generate more rapid progress.",
  9937. journal = "Nat. Neurosci.",
  9938. volume = 22,
  9939. number = 11,
  9940. pages = "1761--1770",
  9941. month = nov,
  9942. year = 2019,
  9943. keywords = "RNN To read;RNN"
  9944. }
  9945. @ARTICLE{Kendler1943-np,
  9946. title = "The influence of a sub-goal on maze behavior",
  9947. author = "Kendler, H H",
  9948. abstract = "The writer reports an experiment designed to determine the
  9949. effect upon learning of a simple distance discrimination of
  9950. varying the ratio of the total distance to the final turn while
  9951. keeping the absolute distance from starting point to goal
  9952. constant. One group of rats was trained on a maze pattern with a
  9953. ratio of 2.33 to the final turn, the other group being trained
  9954. with a ratio of 1.36. The first group (ratio 2.33) took
  9955. significantly less trials and made significantly less errors in
  9956. reaching the criterion of learning. The author feels that the
  9957. results may be interpreted as being in some way attributable to
  9958. the secondary reinforcing characteristics of the final turn or
  9959. ``sub-goal.'' (PsycINFO Database Record (c) 2016 APA, all rights
  9960. reserved)",
  9961. journal = "J. Comp. Psychol.",
  9962. publisher = "psycnet.apa.org",
  9963. volume = 36,
  9964. number = 2,
  9965. pages = "67--73",
  9966. month = oct,
  9967. year = 1943
  9968. }
  9969. @ARTICLE{Hull1951-zh,
  9970. title = "Essentials of behavior",
  9971. author = "Hull, Clark L",
  9972. abstract = "``This volume is designed to present briefly and in an
  9973. intelligible manner the basic laws [postulates and corollaries]
  9974. of mammalian behavior, and to serve as a useful introduction to
  9975. the current aspects of behavior theory.'' The history of the
  9976. present set of postulates is given in the notes. Glossary of
  9977. symbols. 79-item bibliography. (PsycINFO Database Record (c)
  9978. 2016 APA, all rights reserved)",
  9979. publisher = "Yale University Press Essentials of behavior.",
  9980. volume = 145,
  9981. year = 1951,
  9982. address = "New Haven, CT, US",
  9983. keywords = "books;Books"
  9984. }
  9985. @ARTICLE{Waters1937-tc,
  9986. title = "The Principle of Least Effort in Learning",
  9987. author = "Waters, R H",
  9988. journal = "J. Gen. Psychol.",
  9989. publisher = "Routledge",
  9990. volume = 16,
  9991. number = 1,
  9992. pages = "3--20",
  9993. month = jan,
  9994. year = 1937
  9995. }
  9996. @ARTICLE{Hull1932-hs,
  9997. title = "The goal-gradient hypothesis and maze learning",
  9998. author = "Hull, C L",
  9999. abstract = "The hypothesis, which is an extension of Hull's goal reaction
  10000. hypothesis, is that the goal reaction gets conditioned most
  10001. strongly to the stimuli preceding it, and the other reactions in
  10002. the sequence get conditioned to their stimuli, with a strength
  10003. inversely proportional to their temporal or spatial remoteness
  10004. from the goal reaction. Since this assumes a gradient, which is
  10005. related to the goal, he calls it a goal-gradient. The shape of
  10006. this gradient is shown, by reference to Yoshioka's experiment in
  10007. selection of maze pathways by the rat, to be positively
  10008. accelerated, and to conform to the logarithmic law. The author
  10009. deduces ten actual behavior phenomena from his principle, such
  10010. as choice of shorter path, order of elimination of blind alleys,
  10011. relative rates of locomotion in different parts of the maze,
  10012. etc. (PsycINFO Database Record (c) 2016 APA, all rights
  10013. reserved)",
  10014. journal = "Psychol. Rev.",
  10015. publisher = "psycnet.apa.org",
  10016. volume = 39,
  10017. number = 1,
  10018. pages = "25--43",
  10019. month = jan,
  10020. year = 1932
  10021. }
  10022. @ARTICLE{Sams1925-ng,
  10023. title = "Time discrimination in white rats",
  10024. author = "Sams, C F and Tolman, E C",
  10025. abstract = "With two alternative paths in a maze equal in every respect
  10026. except that the animal was detained in a detention chamber 1
  10027. minute before being allowed to enter one alley and 6 minutes
  10028. before allowed to enter the other alley, with the food now
  10029. reachable by path one and now by path two, the investigators
  10030. discovered that almost invariably the animal will learn to seek
  10031. food through that path, the entrance to which has been delayed
  10032. for the shorter period of time. A threshold of difference of
  10033. delay was worked out for one animal; it was a ratio between 1:4
  10034. and 1:5 minutes. (PsycINFO Database Record (c) 2016 APA, all
  10035. rights reserved)",
  10036. journal = "J. Comp. Psychol.",
  10037. publisher = "psycnet.apa.org",
  10038. volume = 5,
  10039. number = 3,
  10040. pages = "255--263",
  10041. month = jun,
  10042. year = 1925
  10043. }
  10044. % The entry below contains non-ASCII chars that could not be converted
  10045. % to a LaTeX equivalent.
  10046. @ARTICLE{Clements1928-ph,
  10047. title = "The effect of time on distance discrimination in the albino rat",
  10048. author = "Clements, Forrest E",
  10049. abstract = "An asymmetrical T-shaped apparatus with paths giving a distance
  10050. ratio of 1:10 was constructed in such a manner that the rat
  10051. could be detained for varying lengths of time immediately after
  10052. making a choice. One group of rats were allowed to make their
  10053. choice of either alley and continue without delay while the
  10054. other three groups were delayed for 30 seconds, 60 seconds, and
  10055. 120 seconds, respectively. The experiment aimed to determine if
  10056. the time spent in running the pathway was as effective as time
  10057. spent in waiting. Without a delay learning begins immediately
  10058. and proceeds rapidly. But when a time interval exists between
  10059. the moment of choice and the continuation of the path to food
  10060. learning does not commence immediately. ``For several days each
  10061. animal apparently learns nothing, his choice of path seeming due
  10062. to chance. Suddenly, however, he begins to learn and from that
  10063. point on his learning proceeds at approximately the same rate as
  10064. that of those animals where there was no detention. The longer
  10065. the detention the longer the initial period of no apparent
  10066. progress. If the detention is long enough… the animals are
  10067. probably unable ever to make the discrimination.'' The group
  10068. which had been delayed 120 seconds ran for 25 days with no signs
  10069. of learning to discriminate the shorter path. On the 26th day
  10070. the detention was removed for three of these rats, whereupon
  10071. they learned it in much shorter time than the animals with no
  10072. detention and no previous acquaintance with the apparatus. ``The
  10073. longer the detention the longer it takes the rat to 'get the
  10074. idea' until, with a long enough detention, the animal's memory
  10075. is perhaps not long enough to span the gap between turning a
  10076. particular way and the 'realization' of the preferential
  10077. character (shortness) of that way.'' (PsycINFO Database Record
  10078. (c) 2017 APA, all rights reserved)",
  10079. journal = "J. Comp. Psychol.",
  10080. publisher = "psycnet.apa.org",
  10081. volume = 8,
  10082. number = 4,
  10083. pages = "317--324",
  10084. month = oct,
  10085. year = 1928
  10086. }
  10087. @MISC{Tolman1967-ym,
  10088. title = "Purposive Behavior in Animal",
  10089. author = "{Tolman}",
  10090. year = 1967
  10091. }
  10092. % The entry below contains non-ASCII chars that could not be converted
  10093. % to a LaTeX equivalent.
  10094. @ARTICLE{Snygg1936-xx,
  10095. title = "Maze Learning as Perception",
  10096. author = "Snygg, Donald",
  10097. abstract = "Because of the growing tendency to think of problems of
  10098. perception and learning as identical, it seems desirable to
  10099. examine empirically the utility of such a viewpoint. The
  10100. systematic advantages of such a simplification of laws and
  10101. entities are tempting; but the ultimate test of the assumption
  10102. of likeness between learning and perception must be its
  10103. usefulness. During a recent investigation (6) of the relative
  10104. difficulty of various patterns of ten- section Warden multiple-U
  10105. mazes it was found that the relative ilumber of entries into the
  10106. …",
  10107. journal = "The Pedagogical Seminary and Journal of Genetic Psychology",
  10108. publisher = "Routledge",
  10109. volume = 49,
  10110. number = 1,
  10111. pages = "231--239",
  10112. month = sep,
  10113. year = 1936
  10114. }
  10115. % The entry below contains non-ASCII chars that could not be converted
  10116. % to a LaTeX equivalent.
  10117. @ARTICLE{Yoshioka1929-bn,
  10118. title = "Weber's law in the discrimination of maze distance by the white
  10119. rat",
  10120. author = "Yoshioka, Joseph Geno",
  10121. abstract = "Using the method of wrong and right cases and two specially
  10122. constructed mazes which allowed the lengths of the alleys to be
  10123. altered, the author attempted to determine whether Weber's Law
  10124. holds for the discrimination of maze distance by the white
  10125. rat.`` Maze I was so constructed that two paths visually similar
  10126. were offered for choice. One path was 211 inches long and kept
  10127. constant, while the other could be shortened by steps of 13
  10128. inches. Maze II was similarly constructed, but magnified by two.
  10129. Five short paths were used in each maze …",
  10130. journal = "Publ. Psychol.",
  10131. publisher = "psycnet.apa.org",
  10132. year = 1929
  10133. }
  10134. @ARTICLE{Yoshioka1928-ob,
  10135. title = "Pattern Versus Frequency and Recency Factors in Maze Learning: A
  10136. Preliminary Study",
  10137. author = "Yoshioka, Joseph G",
  10138. abstract = "*Received for publication by Calvin P. Stone .of the Editorial
  10139. Board, April 2, 1928.",
  10140. journal = "The Pedagogical Seminary and Journal of Genetic Psychology",
  10141. publisher = "Routledge",
  10142. volume = 35,
  10143. number = 2,
  10144. pages = "193--200",
  10145. month = jun,
  10146. year = 1928
  10147. }
  10148. @ARTICLE{Snygg1935-tx,
  10149. title = "Mazes in Which Rats Take the Longer Path to Food",
  10150. author = "Snygg, Donald",
  10151. journal = "J. Psychol.",
  10152. publisher = "Routledge",
  10153. volume = 1,
  10154. number = 1,
  10155. pages = "153--166",
  10156. month = jan,
  10157. year = 1935
  10158. }
  10159. @ARTICLE{Krim2016-el,
  10160. title = "Discovering the Whole by the Coarse: A topological paradigm for
  10161. data analysis",
  10162. author = "Krim, H and Gentimis, T and Chintakunta, H",
  10163. abstract = "The increasing interest in big data applications is ushering in a
  10164. large effort in seeking new, efficient, and adapted data models
  10165. to reduce complexity, while preserving maximal intrinsic
  10166. information. Graph-based models have recently been getting a lot
  10167. of attention on account of their intuitive and direct connection
  10168. to the data [43]. The cost of these models, however, is to some
  10169. extent giving up geometric insight as well as algebraic
  10170. flexibility.",
  10171. journal = "IEEE Signal Process. Mag.",
  10172. volume = 33,
  10173. number = 2,
  10174. pages = "95--104",
  10175. month = mar,
  10176. year = 2016,
  10177. keywords = "Big Data;data analysis;data models;graph theory;Big Data
  10178. application;adapted data model;graph-based model;data
  10179. analysis;Topology;Three-dimensional displays;Data analysis;Big
  10180. data;Time series analysis;Delays"
  10181. }
  10182. @ARTICLE{Curto2008-cc,
  10183. title = "Cell groups reveal structure of stimulus space",
  10184. author = "Curto, Carina and Itskov, Vladimir",
  10185. abstract = "An important task of the brain is to represent the outside
  10186. world. It is unclear how the brain may do this, however, as it
  10187. can only rely on neural responses and has no independent access
  10188. to external stimuli in order to ``decode'' what those responses
  10189. mean. We investigate what can be learned about a space of
  10190. stimuli using only the action potentials (spikes) of cells with
  10191. stereotyped -- but unknown -- receptive fields. Using
  10192. hippocampal place cells as a model system, we show that one can
  10193. (1) extract global features of the environment and (2) construct
  10194. an accurate representation of space, up to an overall scale
  10195. factor, that can be used to track the animal's position. Unlike
  10196. previous approaches to reconstructing position from place cell
  10197. activity, this information is derived without knowing place
  10198. fields or any other functions relating neural responses to
  10199. position. We find that simply knowing which groups of cells fire
  10200. together reveals a surprising amount of structure in the
  10201. underlying stimulus space; this may enable the brain to
  10202. construct its own internal representations.",
  10203. journal = "PLoS Comput. Biol.",
  10204. publisher = "journals.plos.org",
  10205. volume = 4,
  10206. number = 10,
  10207. pages = "e1000205",
  10208. month = oct,
  10209. year = 2008,
  10210. language = "en"
  10211. }
  10212. @ARTICLE{Giusti2015-fa,
  10213. title = "Clique topology reveals intrinsic geometric structure in neural
  10214. correlations",
  10215. author = "Giusti, Chad and Pastalkova, Eva and Curto, Carina and Itskov,
  10216. Vladimir",
  10217. abstract = "Detecting meaningful structure in neural activity and
  10218. connectivity data is challenging in the presence of hidden
  10219. nonlinearities, where traditional eigenvalue-based methods may
  10220. be misleading. We introduce a novel approach to matrix analysis,
  10221. called clique topology, that extracts features of the data
  10222. invariant under nonlinear monotone transformations. These
  10223. features can be used to detect both random and geometric
  10224. structure, and depend only on the relative ordering of matrix
  10225. entries. We then analyzed the activity of pyramidal neurons in
  10226. rat hippocampus, recorded while the animal was exploring a 2D
  10227. environment, and confirmed that our method is able to detect
  10228. geometric organization using only the intrinsic pattern of
  10229. neural correlations. Remarkably, we found similar results during
  10230. nonspatial behaviors such as wheel running and rapid eye
  10231. movement (REM) sleep. This suggests that the geometric structure
  10232. of correlations is shaped by the underlying hippocampal circuits
  10233. and is not merely a consequence of position coding. We propose
  10234. that clique topology is a powerful new tool for matrix analysis
  10235. in biological settings, where the relationship of observed
  10236. quantities to more meaningful variables is often nonlinear and
  10237. unknown.",
  10238. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  10239. publisher = "National Acad Sciences",
  10240. volume = 112,
  10241. number = 44,
  10242. pages = "13455--13460",
  10243. month = nov,
  10244. year = 2015,
  10245. keywords = "Betti curves; clique topology; neural coding; structure of
  10246. neural correlation; topological data analysis",
  10247. language = "en"
  10248. }
  10249. @ARTICLE{Giusti2016-st,
  10250. title = "Two's company, three (or more) is a simplex :
  10251. Algebraic-topological tools for understanding higher-order
  10252. structure in neural data",
  10253. author = "Giusti, Chad and Ghrist, Robert and Bassett, Danielle S",
  10254. abstract = "The language of graph theory, or network science, has proven to
  10255. be an exceptional tool for addressing myriad problems in
  10256. neuroscience. Yet, the use of networks is predicated on a
  10257. critical simplifying assumption: that the quintessential unit of
  10258. interest in a brain is a dyad - two nodes (neurons or brain
  10259. regions) connected by an edge. While rarely mentioned, this
  10260. fundamental assumption inherently limits the types of neural
  10261. structure and function that graphs can be used to model. Here, we
  10262. describe a generalization of graphs that overcomes these
  10263. limitations, thereby offering a broad range of new possibilities
  10264. in terms of modeling and measuring neural phenomena.
  10265. Specifically, we explore the use of simplicial complexes: a
  10266. structure developed in the field of mathematics known as
  10267. algebraic topology, of increasing applicability to real data due
  10268. to a rapidly growing computational toolset. We review the
  10269. underlying mathematical formalism as well as the budding
  10270. literature applying simplicial complexes to neural data, from
  10271. electrophysiological recordings in animal models to hemodynamic
  10272. fluctuations in humans. Based on the exceptional flexibility of
  10273. the tools and recent ground-breaking insights into neural
  10274. function, we posit that this framework has the potential to
  10275. eclipse graph theory in unraveling the fundamental mysteries of
  10276. cognition.",
  10277. journal = "J. Comput. Neurosci.",
  10278. volume = 41,
  10279. number = 1,
  10280. pages = "1--14",
  10281. month = aug,
  10282. year = 2016,
  10283. keywords = "Filtration; Networks; Simplicial complex; Topology",
  10284. language = "en"
  10285. }
  10286. @ARTICLE{Cox2016-ty,
  10287. title = "Clique Topology Reveals Intrinsic Geometric Structure in
  10288. Neural Correlations: An Overview",
  10289. author = "Cox, David",
  10290. abstract = "This publication serves as an overview of clique topology --
  10291. a novel matrix analysis technique used to extract structural
  10292. features from neural activity data that contains hidden
  10293. nonlinearities. We highlight work done by Gusti et al. which
  10294. introduces clique topology and verifies its applicability to
  10295. neural feature extraction by showing that neural
  10296. correlations in the rat hippocampus are determined by
  10297. geometric structure of hippocampal circuits, rather than
  10298. being a consequence of positional coding.",
  10299. month = aug,
  10300. year = 2016,
  10301. archivePrefix = "arXiv",
  10302. primaryClass = "q-bio.NC",
  10303. eprint = "1608.03463"
  10304. }
  10305. @ARTICLE{Ghrist2007-kx,
  10306. title = "Barcodes: The persistent topology of data",
  10307. author = "Ghrist, Robert",
  10308. abstract = "This article surveys recent work of Carlsson and collaborators on
  10309. applications of computational algebraic topology to problems of
  10310. feature detection and shape recognition in high-dimensional data.
  10311. The primary mathematical tool considered is a homology theory for
  10312. point-cloud data sets--persistent homology--and a novel
  10313. representation of this algebraic characterization--barcodes. We
  10314. sketch an application of these techniques to the classification
  10315. of natural images.",
  10316. journal = "Bull. Am. Math. Soc.",
  10317. volume = 45,
  10318. number = 01,
  10319. pages = "61--76",
  10320. month = oct,
  10321. year = 2007
  10322. }
  10323. @ARTICLE{Dabaghian2014-yp,
  10324. title = "Reconceiving the hippocampal map as a topological template",
  10325. author = "Dabaghian, Yuri and Brandt, Vicky L and Frank, Loren M",
  10326. abstract = "The role of the hippocampus in spatial cognition is
  10327. incontrovertible yet controversial. Place cells, initially
  10328. thought to be location-specifiers, turn out to respond
  10329. promiscuously to a wide range of stimuli. Here we test the idea,
  10330. which we have recently demonstrated in a computational model,
  10331. that the hippocampal place cells may ultimately be interested in
  10332. a space's topological qualities (its connectivity) more than its
  10333. geometry (distances and angles); such higher-order functioning
  10334. would be more consistent with other known hippocampal functions.
  10335. We recorded place cell activity in rats exploring morphing linear
  10336. tracks that allowed us to dissociate the geometry of the track
  10337. from its topology. The resulting place fields preserved the
  10338. relative sequence of places visited along the track but did not
  10339. vary with the metrical features of the track or the direction of
  10340. the rat's movement. These results suggest a reinterpretation of
  10341. previous studies and new directions for future experiments.",
  10342. journal = "Elife",
  10343. volume = 3,
  10344. pages = "e03476",
  10345. month = aug,
  10346. year = 2014,
  10347. keywords = "geometry; hippocampus; place cells; spatial learning; topology",
  10348. language = "en"
  10349. }
  10350. @ARTICLE{Cholvin2013-ad,
  10351. title = "The ventral midline thalamus contributes to strategy shifting in
  10352. a memory task requiring both prefrontal cortical and hippocampal
  10353. functions",
  10354. author = "Cholvin, Thibault and Loureiro, Micha{\"e}l and Cassel, Raphaelle
  10355. and Cosquer, Brigitte and Geiger, Karine and De Sa Nogueira,
  10356. David and Raingard, H{\'e}l{\`e}ne and Robelin, Laura and Kelche,
  10357. Christian and Pereira de Vasconcelos, Anne and Cassel,
  10358. Jean-Christophe",
  10359. abstract = "Electrophysiological and neuroanatomical evidence for reciprocal
  10360. connections with the medial prefrontal cortex (mPFC) and the
  10361. hippocampus make the reuniens and rhomboid (ReRh) thalamic nuclei
  10362. a putatively major functional link for regulations of
  10363. cortico-hippocampal interactions. In a first experiment using a
  10364. new water escape device for rodents, the double-H maze, we
  10365. demonstrated in rats that a bilateral muscimol (MSCI)
  10366. inactivation (0.70 vs 0.26 and 0 nmol) of the mPFC or dorsal
  10367. hippocampus (dHip) induces major deficits in a strategy
  10368. shifting/spatial memory retrieval task. By way of comparison,
  10369. only dHip inactivation impaired recall in a classical spatial
  10370. memory task in the Morris water maze. In the second experiment,
  10371. we showed that ReRh inactivation using 0.70 nmol of MSCI, which
  10372. reduced performance without obliterating memory retrieval in the
  10373. water maze, produces an as large strategy shifting/memory
  10374. retrieval deficit as mPFC or dHip inactivation in the double-H
  10375. maze. Thus, behavioral adaptations to task contingency
  10376. modifications requiring a shift toward the use of a memory for
  10377. place might operate in a distributed circuit encompassing the
  10378. mPFC (as the potential set-shifting structure), the hippocampus
  10379. (as the spatial memory substrate), and the ventral midline
  10380. thalamus, and therein the ReRh (as the coordinator of this
  10381. processing). The results of the current experiments provide a
  10382. significant extension of our understanding of the involvement of
  10383. ventral midline thalamic nuclei in cognitive processes: they
  10384. point to a role of the ReRh in strategy shifting in a memory task
  10385. requiring cortical and hippocampal functions and further
  10386. elucidate the functional system underlying behavioral
  10387. flexibility.",
  10388. journal = "J. Neurosci.",
  10389. volume = 33,
  10390. number = 20,
  10391. pages = "8772--8783",
  10392. month = may,
  10393. year = 2013,
  10394. language = "en"
  10395. }
  10396. @ARTICLE{Ito2018-xs,
  10397. title = "Prefrontal-hippocampal interactions for spatial navigation",
  10398. author = "Ito, Hiroshi T",
  10399. abstract = "Animals have the ability to navigate to a desired location by
  10400. making use of information about environmental landmarks and their
  10401. own movements. While decades of neuroscience research have
  10402. identified neurons in the hippocampus and parahippocampal
  10403. structures that represent an animal's position in space, it is
  10404. still largely unclear how an animal can choose the next movement
  10405. direction to reach a desired goal. As the goal destination is
  10406. typically located somewhere outside of the range of sensory
  10407. perception, the animal is required to rely on the internal metric
  10408. of space to estimate the direction and distance of the
  10409. destination to plan a next action. Therefore, the hippocampal
  10410. spatial map should interact with action-planning systems in other
  10411. cortical regions. In accordance with this idea, several recent
  10412. studies have indicated the importance of functional interactions
  10413. between the hippocampus and the prefrontal cortex for
  10414. goal-directed navigation. In this paper, I will review these
  10415. studies and discuss how an animal can estimate its future
  10416. positions correspond to a next movement. Investigation of the
  10417. navigation problem may further provide general insights into
  10418. internal models of the brain for action planning.",
  10419. journal = "Neurosci. Res.",
  10420. volume = 129,
  10421. pages = "2--7",
  10422. month = apr,
  10423. year = 2018,
  10424. keywords = "Hippocampus; Prefrontal cortex; Spatial navigation",
  10425. language = "en"
  10426. }
  10427. @ARTICLE{McKenna2004-kw,
  10428. title = "Afferent projections to nucleus reuniens of the thalamus",
  10429. author = "McKenna, James Timothy and Vertes, Robert P",
  10430. abstract = "The nucleus reuniens (RE) is the largest of the midline nuclei of
  10431. the thalamus and the major source of thalamic afferents to the
  10432. hippocampus and parahippocampal structures. Nucleus reuniens has
  10433. recently been shown to exert powerful excitatory actions on CA1
  10434. of the hippocampus. Few reports on any species have examined
  10435. afferent projections to nucleus reuniens. By using the retrograde
  10436. anatomical tracer Fluorogold, we examined patterns of afferent
  10437. projections to RE in the rat. We showed that RE receives a
  10438. diverse and widely distributed set of afferents projections. The
  10439. main sources of input to nucleus reuniens were from the
  10440. orbitomedial, insular, ectorhinal, perirhinal, and retrosplenial
  10441. cortices; CA1/subiculum of hippocampus; claustrum, tania tecta,
  10442. lateral septum, substantia innominata, and medial and lateral
  10443. preoptic nuclei of the basal forebrain; medial nucleus of
  10444. amygdala; paraventricular and lateral geniculate nuclei of the
  10445. thalamus; zona incerta; anterior, ventromedial, lateral,
  10446. posterior, supramammillary, and dorsal premammillary nuclei of
  10447. the hypothalamus; and ventral tegmental area, periaqueductal
  10448. gray, medial and posterior pretectal nuclei, superior colliculus,
  10449. precommissural/commissural nuclei, nucleus of the posterior
  10450. commissure, parabrachial nucleus, laterodorsal and
  10451. pedunculopontine tegmental nuclei, nucleus incertus, and dorsal
  10452. and median raphe nuclei of the brainstem. The present findings of
  10453. widespread projections to RE, mainly from
  10454. limbic/limbic-associated structures, suggest that nucleus
  10455. reuniens represents a critical relay in the transfer of limbic
  10456. information (emotional/cognitive) from RE to its major targets,
  10457. namely, to the hippocampus and orbitomedial prefrontal cortex. RE
  10458. appears to be a major link in the two-way exchange of information
  10459. between the hippocampus and the medial prefrontal cortex.",
  10460. journal = "J. Comp. Neurol.",
  10461. volume = 480,
  10462. number = 2,
  10463. pages = "115--142",
  10464. month = dec,
  10465. year = 2004,
  10466. language = "en"
  10467. }
  10468. @ARTICLE{Varela2014-yv,
  10469. title = "Anatomical substrates for direct interactions between
  10470. hippocampus, medial prefrontal cortex, and the thalamic nucleus
  10471. reuniens",
  10472. author = "Varela, C and Kumar, S and Yang, J Y and Wilson, M A",
  10473. abstract = "The reuniens nucleus in the midline thalamus projects to the
  10474. medial prefrontal cortex (mPFC) and the hippocampus, and has been
  10475. suggested to modulate interactions between these regions, such as
  10476. spindle-ripple correlations during sleep and theta band coherence
  10477. during exploratory behavior. Feedback from the hippocampus to the
  10478. nucleus reuniens has received less attention but has the
  10479. potential to influence thalamocortical networks as a function of
  10480. hippocampal activation. We used the retrograde tracer cholera
  10481. toxin B conjugated to two fluorophores to study thalamic
  10482. projections to the dorsal and ventral hippocampus and to the
  10483. prelimbic and infralimbic subregions of mPFC. We also examined
  10484. the feedback connections from the hippocampus to reuniens. The
  10485. goal was to evaluate the anatomical basis for direct coordination
  10486. between reuniens, mPFC, and hippocampus by looking for
  10487. double-labeled cells in reuniens and hippocampus. In confirmation
  10488. of previous reports, the nucleus reuniens was the origin of most
  10489. thalamic afferents to the dorsal hippocampus, whereas both
  10490. reuniens and the lateral dorsal nucleus projected to ventral
  10491. hippocampus. Feedback from hippocampus to reuniens originated
  10492. primarily in the dorsal and ventral subiculum. Thalamic cells
  10493. with collaterals to mPFC and hippocampus were found in reuniens,
  10494. across its anteroposterior axis, and represented, on average,
  10495. about 8 \% of the labeled cells in reuniens. Hippocampal cells
  10496. with collaterals to mPFC and reuniens were less common (~1 \% of
  10497. the labeled subicular cells), and located in the molecular layer
  10498. of the subiculum. The results indicate that a subset of reuniens
  10499. cells can directly coordinate activity in mPFC and hippocampus.
  10500. Cells with collaterals in the hippocampus-reuniens-mPFC network
  10501. may be important for the systems consolidation of memory traces
  10502. and for theta synchronization during exploratory behavior.",
  10503. journal = "Brain Struct. Funct.",
  10504. volume = 219,
  10505. number = 3,
  10506. pages = "911--929",
  10507. month = may,
  10508. year = 2014,
  10509. language = "en"
  10510. }
  10511. @ARTICLE{Kim2016-zg,
  10512. title = "Simultaneous fast measurement of circuit dynamics at multiple
  10513. sites across the mammalian brain",
  10514. author = "Kim, Christina K and Yang, Samuel J and Pichamoorthy, Nandini and
  10515. Young, Noah P and Kauvar, Isaac and Jennings, Joshua H and
  10516. Lerner, Talia N and Berndt, Andre and Lee, Soo Yeun and
  10517. Ramakrishnan, Charu and Davidson, Thomas J and Inoue, Masatoshi
  10518. and Bito, Haruhiko and Deisseroth, Karl",
  10519. abstract = "Real-time activity measurements from multiple specific cell
  10520. populations and projections are likely to be important for
  10521. understanding the brain as a dynamical system. Here we developed
  10522. frame-projected independent-fiber photometry (FIP), which we used
  10523. to record fluorescence activity signals from many brain regions
  10524. simultaneously in freely behaving mice. We explored the
  10525. versatility of the FIP microscope by quantifying real-time
  10526. activity relationships among many brain regions during social
  10527. behavior, simultaneously recording activity along multiple axonal
  10528. pathways during sensory experience, performing simultaneous
  10529. two-color activity recording, and applying optical perturbation
  10530. tuned to elicit dynamics that match naturally occurring patterns
  10531. observed during behavior.",
  10532. journal = "Nat. Methods",
  10533. volume = 13,
  10534. number = 4,
  10535. pages = "325--328",
  10536. month = apr,
  10537. year = 2016,
  10538. language = "en"
  10539. }
  10540. @ARTICLE{Guo2015-ev,
  10541. title = "Multi-channel fiber photometry for population neuronal activity
  10542. recording",
  10543. author = "Guo, Qingchun and Zhou, Jingfeng and Feng, Qiru and Lin, Rui and
  10544. Gong, Hui and Luo, Qingming and Zeng, Shaoqun and Luo, Minmin and
  10545. Fu, Ling",
  10546. abstract = "Fiber photometry has become increasingly popular among
  10547. neuroscientists as a convenient tool for the recording of
  10548. genetically defined neuronal population in behaving animals.
  10549. Here, we report the development of the multi-channel fiber
  10550. photometry system to simultaneously monitor neural activities in
  10551. several brain areas of an animal or in different animals. In this
  10552. system, a galvano-mirror modulates and cyclically couples the
  10553. excitation light to individual multimode optical fiber bundles. A
  10554. single photodetector collects excited light and the configuration
  10555. of fiber bundle assembly and the scanner determines the total
  10556. channel number. We demonstrated that the system exhibited
  10557. negligible crosstalk between channels and optical signals could
  10558. be sampled simultaneously with a sample rate of at least 100 Hz
  10559. for each channel, which is sufficient for recording calcium
  10560. signals. Using this system, we successfully recorded GCaMP6
  10561. fluorescent signals from the bilateral barrel cortices of a
  10562. head-restrained mouse in a dual-channel mode, and the
  10563. orbitofrontal cortices of multiple freely moving mice in a
  10564. triple-channel mode. The multi-channel fiber photometry system
  10565. would be a valuable tool for simultaneous recordings of
  10566. population activities in different brain areas of a given animal
  10567. and different interacting individuals.",
  10568. journal = "Biomed. Opt. Express",
  10569. volume = 6,
  10570. number = 10,
  10571. pages = "3919--3931",
  10572. month = oct,
  10573. year = 2015,
  10574. keywords = "(170.0170) Medical optics and biotechnology; (170.2150)
  10575. Endoscopic imaging; (170.2655) Functional monitoring and imaging;
  10576. (180.2520) Fluorescence microscopy",
  10577. language = "en"
  10578. }
  10579. @ARTICLE{Chaudhuri2019-yh,
  10580. title = "The intrinsic attractor manifold and population dynamics of a
  10581. canonical cognitive circuit across waking and sleep",
  10582. author = "Chaudhuri, Rishidev and Ger{\c c}ek, Berk and Pandey, Biraj and
  10583. Peyrache, Adrien and Fiete, Ila",
  10584. abstract = "Neural circuits construct distributed representations of key
  10585. variables-external stimuli or internal constructs of quantities
  10586. relevant for survival, such as an estimate of one's location in
  10587. the world-as vectors of population activity. Although population
  10588. activity vectors may have thousands of entries (dimensions), we
  10589. consider that they trace out a low-dimensional manifold whose
  10590. dimension and topology match the represented variable. This
  10591. manifold perspective enables blind discovery and decoding of the
  10592. represented variable using only neural population activity
  10593. (without knowledge of the input, output, behavior or
  10594. topography). We characterize and directly visualize manifold
  10595. structure in the mammalian head direction circuit, revealing
  10596. that the states form a topologically nontrivial one-dimensional
  10597. ring. The ring exhibits isometry and is invariant across waking
  10598. and rapid eye movement sleep. This result directly demonstrates
  10599. that there are continuous attractor dynamics and enables
  10600. powerful inference about mechanism. Finally, external rather
  10601. than internal noise limits memory fidelity, and the manifold
  10602. approach reveals new dynamical trajectories during sleep.",
  10603. journal = "Nat. Neurosci.",
  10604. publisher = "nature.com",
  10605. volume = 22,
  10606. number = 9,
  10607. pages = "1512--1520",
  10608. month = sep,
  10609. year = 2019,
  10610. language = "en"
  10611. }
  10612. @ARTICLE{Spreemann2015-as,
  10613. title = "Using persistent homology to reveal hidden information in
  10614. neural data",
  10615. author = "Spreemann, Gard and Dunn, Benjamin and Botnan, Magnus Bakke
  10616. and Baas, Nils A",
  10617. abstract = "We propose a method, based on persistent homology, to
  10618. uncover topological properties of a priori unknown
  10619. covariates of neuron activity. Our input data consist of
  10620. spike train measurements of a set of neurons of interest, a
  10621. candidate list of the known stimuli that govern neuron
  10622. activity, and the corresponding state of the animal
  10623. throughout the experiment performed. Using a generalized
  10624. linear model for neuron activity and simple assumptions on
  10625. the effects of the external stimuli, we infer away any
  10626. contribution to the observed spike trains by the candidate
  10627. stimuli. Persistent homology then reveals useful information
  10628. about any further, unknown, covariates.",
  10629. month = oct,
  10630. year = 2015,
  10631. archivePrefix = "arXiv",
  10632. primaryClass = "q-bio.NC",
  10633. eprint = "1510.06629"
  10634. }
  10635. @ARTICLE{Baas2017-zd,
  10636. title = "On the concept of space in neuroscience",
  10637. author = "Baas, Nils A",
  10638. abstract = "In this paper we study recording of neurons creating spatial
  10639. information in the brain. To sets of spike trains we associate a
  10640. topological space which captures the structure of the space in
  10641. which the movement takes place. This space has an even richer
  10642. structure depending on other than spatial stimuli. We describe a
  10643. method to separate the various stimuli and conclude when they
  10644. describe the structure of the space. We discuss what we should
  10645. mean by neural space and its structure, and come up with some
  10646. speculations for the future.",
  10647. journal = "Current Opinion in Systems Biology",
  10648. volume = 1,
  10649. pages = "32--37",
  10650. month = feb,
  10651. year = 2017,
  10652. keywords = "Neurons; spike trains; place cells; grid cells; persistent
  10653. homology; dynamic Ising model; neural space; hyperstructure"
  10654. }
  10655. @ARTICLE{Chen2014-rg,
  10656. title = "Neural representation of spatial topology in the rodent
  10657. hippocampus",
  10658. author = "Chen, Zhe and Gomperts, Stephen N and Yamamoto, Jun and Wilson,
  10659. Matthew A",
  10660. abstract = "Pyramidal cells in the rodent hippocampus often exhibit clear
  10661. spatial tuning in navigation. Although it has been long suggested
  10662. that pyramidal cell activity may underlie a topological code
  10663. rather than a topographic code, it remains unclear whether an
  10664. abstract spatial topology can be encoded in the ensemble spiking
  10665. activity of hippocampal place cells. Using a statistical approach
  10666. developed previously, we investigate this question and related
  10667. issues in greater detail. We recorded ensembles of hippocampal
  10668. neurons as rodents freely foraged in one- and two-dimensional
  10669. spatial environments and used a ``decode-to-uncover'' strategy to
  10670. examine the temporally structured patterns embedded in the
  10671. ensemble spiking activity in the absence of observed spatial
  10672. correlates during periods of rodent navigation or awake
  10673. immobility. Specifically, the spatial environment was represented
  10674. by a finite discrete state space. Trajectories across spatial
  10675. locations (``states'') were associated with consistent
  10676. hippocampal ensemble spiking patterns, which were characterized
  10677. by a state transition matrix. From this state transition matrix,
  10678. we inferred a topology graph that defined the connectivity in the
  10679. state space. In both one- and two-dimensional environments, the
  10680. extracted behavior patterns from the rodent hippocampal
  10681. population codes were compared against randomly shuffled spike
  10682. data. In contrast to a topographic code, our results support the
  10683. efficiency of topological coding in the presence of sparse sample
  10684. size and fuzzy space mapping. This computational approach allows
  10685. us to quantify the variability of ensemble spiking activity,
  10686. examine hippocampal population codes during off-line states, and
  10687. quantify the topological complexity of the environment.",
  10688. journal = "Neural Comput.",
  10689. volume = 26,
  10690. number = 1,
  10691. pages = "1--39",
  10692. month = jan,
  10693. year = 2014,
  10694. language = "en"
  10695. }
  10696. @ARTICLE{Curto2016-sc,
  10697. title = "What can topology tell us about the neural code?",
  10698. author = "Curto, Carina",
  10699. abstract = "Neuroscience is undergoing a period of rapid experimental
  10700. progress and expansion. New mathematical tools, previously
  10701. unknown in the neuroscience community, are now being used to
  10702. tackle fundamental questions and analyze emerging data sets.
  10703. Consistent with this trend, the last decade has seen an uptick in
  10704. the use of topological ideas and methods in neuroscience. In this
  10705. paper I will survey recent applications of topology in
  10706. neuroscience, and explain why topology is an especially natural
  10707. tool for understanding neural codes.",
  10708. journal = "Bull. Am. Math. Soc.",
  10709. volume = 54,
  10710. number = 1,
  10711. pages = "63--78",
  10712. month = sep,
  10713. year = 2016
  10714. }
  10715. @ARTICLE{Alexander2015-vc,
  10716. title = "Retrosplenial cortex maps the conjunction of internal and
  10717. external spaces",
  10718. author = "Alexander, Andrew S and Nitz, Douglas A",
  10719. abstract = "Intelligent behavior demands not only multiple forms of spatial
  10720. representation, but also coordination among the brain regions
  10721. mediating those representations. Retrosplenial cortex is densely
  10722. interconnected with the majority of cortical and subcortical
  10723. brain structures that register an animal's position in multiple
  10724. internal and external spatial frames of reference. This unique
  10725. anatomy suggests that it functions to integrate distinct forms of
  10726. spatial information and provides an interface for transformations
  10727. between them. Evidence for this was found in rats traversing two
  10728. different routes placed at different environmental locations.
  10729. Retrosplenial ensembles robustly encoded conjunctions of progress
  10730. through the current route, position in the larger environment and
  10731. the left versus right turning behavior of the animal. Thus, the
  10732. retrosplenial cortex has the requisite dynamics to serve as an
  10733. intermediary between brain regions generating different forms of
  10734. spatial mapping, a result that is consistent with navigational
  10735. and episodic memory impairments following damage to this region
  10736. in humans.",
  10737. journal = "Nat. Neurosci.",
  10738. volume = 18,
  10739. number = 8,
  10740. pages = "1143--1151",
  10741. month = aug,
  10742. year = 2015,
  10743. language = "en"
  10744. }
  10745. @UNPUBLISHED{McClain2019-rs,
  10746. title = "Position-theta-phase model of hippocampal place cell activity
  10747. applied to quantification of running speed modulation of firing
  10748. rate",
  10749. author = "McClain, Kathryn and Tingley, David and Heeger, David and
  10750. Buzs{\'a}ki, Gy{\"o}rgy",
  10751. abstract = "Abstract Spiking activity of place cells in the hippocampus
  10752. encodes the animal's position as it moves through an environment.
  10753. Within a cell's place field, both the firing rate and the phase
  10754. of spiking in the local theta oscillation contain spatial
  10755. information. We propose a position-theta-phase (PTP) model that
  10756. captures the simultaneous expression of the firing-rate code and
  10757. theta-phase code in place cell spiking. This model parametrically
  10758. characterizes place fields to compare across cells, time and
  10759. condition, generates realistic place cell simulation data, and
  10760. conceptualizes a framework for principled hypothesis testing to
  10761. identify additional features of place cell activity. We use the
  10762. PTP model to assess the effect of running speed in place cell
  10763. data recorded from rats running on linear tracks. For the
  10764. majority of place fields we do not find evidence for speed
  10765. modulation of the firing rate. For a small subset of place
  10766. fields, we find firing rates significantly increase or decrease
  10767. with speed. We use the PTP model to compare candidate mechanisms
  10768. of speed modulation in significantly modulated fields, and
  10769. determine that speed acts as a gain control on the magnitude of
  10770. firing rate. Our model provides a tool that connects rigorous
  10771. analysis with a computational framework for understanding place
  10772. cell activity.Significance The hippocampus is heavily studied in
  10773. the context of spatial navigation, and the format of spatial
  10774. information in hippocampus is multifaceted and complex.
  10775. Furthermore, the hippocampus is also thought to contain
  10776. information about other important aspects of behavior such as
  10777. running speed, though there is not agreement on the nature and
  10778. magnitude of their effect. To understand how all of these
  10779. variables are simultaneously represented and used to guide
  10780. behavior, a theoretical framework is needed that can be directly
  10781. applied to the data we record. We present a model that captures
  10782. well-established spatial-encoding features of hippocampal
  10783. activity and provides the opportunity to identify and incorporate
  10784. novel features for our collective understanding.",
  10785. journal = "bioRxiv",
  10786. pages = "714105",
  10787. month = jul,
  10788. year = 2019,
  10789. language = "en"
  10790. }
  10791. @ARTICLE{Friedmann_undated-xv,
  10792. title = "Mapping Mesoscale Axonal Projections in the Mouse Brain Using A
  10793. {3D} Convolutional Network",
  10794. author = "Friedmann, Drew and Pun, Albert and Adams, Eliza L and Lui, Jan H
  10795. and Kebschull, Justus M and Grutzner, Sophie M and Castagnola,
  10796. Caitlin and Tessier-Lavigne, Marc and Luo, Liqun"
  10797. }
  10798. @ARTICLE{Rubin2019-tw,
  10799. title = "Revealing neural correlates of behavior without behavioral
  10800. measurements",
  10801. author = "Rubin, Alon and Sheintuch, Liron and Brande-Eilat, Noa and
  10802. Pinchasof, Or and Rechavi, Yoav and Geva, Nitzan and Ziv, Yaniv",
  10803. abstract = "Measuring neuronal tuning curves has been instrumental for many
  10804. discoveries in neuroscience but requires a priori assumptions
  10805. regarding the identity of the encoded variables. We applied
  10806. unsupervised learning to large-scale neuronal recordings in
  10807. behaving mice from circuits involved in spatial cognition and
  10808. uncovered a highly-organized internal structure of ensemble
  10809. activity patterns. This emergent structure allowed defining for
  10810. each neuron an 'internal tuning-curve' that characterizes its
  10811. activity relative to the network activity, rather than relative
  10812. to any predefined external variable, revealing place-tuning and
  10813. head-direction tuning without relying on measurements of place or
  10814. head-direction. Similar investigation in prefrontal cortex
  10815. revealed schematic representations of distances and actions, and
  10816. exposed a previously unknown variable, the 'trajectory-phase'.
  10817. The internal structure was conserved across mice, allowing using
  10818. one animal's data to decode another animal's behavior. Thus, the
  10819. internal structure of neuronal activity itself enables
  10820. reconstructing internal representations and discovering new
  10821. behavioral variables hidden within a neural code.",
  10822. journal = "Nat. Commun.",
  10823. volume = 10,
  10824. number = 1,
  10825. pages = "4745",
  10826. month = oct,
  10827. year = 2019,
  10828. language = "en"
  10829. }
  10830. @UNPUBLISHED{Kastner2019-xb,
  10831. title = "Dynamic preferences account for inter-animal variability during
  10832. the continual learning of a cognitive task",
  10833. author = "Kastner, David B and Miller, Eric A and Yang, Zhounan and Roumis,
  10834. Demetris K and Liu, Daniel F and Frank, Loren M and Dayan, Peter",
  10835. abstract = "In novel situations, behavior necessarily reduces to latent
  10836. biases. How these biases interact with new experiences to enable
  10837. subsequent behavior remains poorly understood. We exposed rats to
  10838. a family of spatial alternation contingencies and developed a
  10839. series of reinforcement learning agents to describe the behavior.
  10840. The performance of these agents shows that accurately describing
  10841. the learning of individual animals requires accounting for their
  10842. individual dynamic preferences as well as general, shared,
  10843. cognitive processes. Agents that include only memory of past
  10844. choice do not account for the behavior. Adding an explicit
  10845. representation of biases allows agents to perform the task as
  10846. rapidly as the rats, to accurately predict critical facets of
  10847. their behavior on which it was not fitted, and to capture
  10848. individual differences quantitatively. Our results illustrate the
  10849. value of making explicit models of learning and highlight the
  10850. importance of considering the initial state of each animal in
  10851. understanding behavior.",
  10852. journal = "bioRxiv",
  10853. pages = "808006",
  10854. month = oct,
  10855. year = 2019,
  10856. language = "en"
  10857. }
  10858. @ARTICLE{Faisal2008-as,
  10859. title = "Noise in the nervous system",
  10860. author = "Faisal, A Aldo and Selen, Luc P J and Wolpert, Daniel M",
  10861. abstract = "Noise--random disturbances of signals--poses a fundamental
  10862. problem for information processing and affects all aspects of
  10863. nervous-system function. However, the nature, amount and impact
  10864. of noise in the nervous system have only recently been addressed
  10865. in a quantitative manner. Experimental and computational methods
  10866. have shown that multiple noise sources contribute to cellular
  10867. and behavioural trial-to-trial variability. We review the
  10868. sources of noise in the nervous system, from the molecular to
  10869. the behavioural level, and show how noise contributes to
  10870. trial-to-trial variability. We highlight how noise affects
  10871. neuronal networks and the principles the nervous system applies
  10872. to counter detrimental effects of noise, and briefly discuss
  10873. noise's potential benefits.",
  10874. journal = "Nat. Rev. Neurosci.",
  10875. publisher = "nature.com",
  10876. volume = 9,
  10877. number = 4,
  10878. pages = "292--303",
  10879. month = apr,
  10880. year = 2008,
  10881. language = "en"
  10882. }
  10883. @ARTICLE{Beck2012-xf,
  10884. title = "Not noisy, just wrong: the role of suboptimal inference in
  10885. behavioral variability",
  10886. author = "Beck, Jeffrey M and Ma, Wei Ji and Pitkow, Xaq and Latham, Peter
  10887. E and Pouget, Alexandre",
  10888. abstract = "Behavior varies from trial to trial even when the stimulus is
  10889. maintained as constant as possible. In many models, this
  10890. variability is attributed to noise in the brain. Here, we
  10891. propose that there is another major source of variability:
  10892. suboptimal inference. Importantly, we argue that in most tasks
  10893. of interest, and particularly complex ones, suboptimal inference
  10894. is likely to be the dominant component of behavioral
  10895. variability. This perspective explains a variety of intriguing
  10896. observations, including why variability appears to be larger on
  10897. the sensory than on the motor side, and why our sensors are
  10898. sometimes surprisingly unreliable.",
  10899. journal = "Neuron",
  10900. publisher = "Elsevier",
  10901. volume = 74,
  10902. number = 1,
  10903. pages = "30--39",
  10904. month = apr,
  10905. year = 2012,
  10906. language = "en"
  10907. }
  10908. @ARTICLE{Ellard2009-vi,
  10909. title = "Spatial cognition in the gerbil: computing optimal escape routes
  10910. from visual threats",
  10911. author = "Ellard, Colin G and Eller, Meghan C",
  10912. abstract = "Previous studies in our laboratory have shown that when presented
  10913. with a sudden stimulus simulating an oncoming predator, Mongolian
  10914. gerbils can compute the optimal trajectory to a safe refuge,
  10915. taking into account the position of the threat, the location of a
  10916. clearly visible refuge, and several other contextual variables as
  10917. well. In the present studies, the main goal was to explore the
  10918. abilities of gerbils to use mental representations of spaces that
  10919. were visually occluded by opaque barriers to compute efficient
  10920. escape trajectories. In all studies, gerbils were placed into a
  10921. round open field containing a single refuge. On each trial, an
  10922. overhead visual stimulus was caused to 'fly' overhead, eliciting
  10923. robust escape movements from the gerbils. By manipulating the
  10924. shape and position of a series of opaque barriers that were
  10925. interposed between the gerbils and the refuge, we were able to
  10926. show that gerbils can compute the shortest route to an invisible
  10927. target, even when the available routes to the target are made
  10928. complex by using elaborate barrier shapes. These findings suggest
  10929. that gerbils can maintain representations of their locations with
  10930. respect to salient environmental landmarks and refuges, even when
  10931. such locations are not continuously visible.",
  10932. journal = "Anim. Cogn.",
  10933. volume = 12,
  10934. number = 2,
  10935. pages = "333--345",
  10936. month = mar,
  10937. year = 2009,
  10938. language = "en"
  10939. }
  10940. @ARTICLE{Chen2019-pb,
  10941. title = "{High-Throughput} Mapping of {Long-Range} Neuronal Projection
  10942. Using In Situ Sequencing",
  10943. author = "Chen, Xiaoyin and Sun, Yu-Chi and Zhan, Huiqing and Kebschull,
  10944. Justus M and Fischer, Stephan and Matho, Katherine and Huang, Z
  10945. Josh and Gillis, Jesse and Zador, Anthony M",
  10946. abstract = "Summary Understanding neural circuits requires deciphering
  10947. interactions among myriad cell types defined by spatial
  10948. organization, connectivity, gene expression, and other
  10949. properties. Resolving these cell types requires both
  10950. single-neuron resolution and high throughput, a challenging
  10951. combination with conventional methods. Here, we introduce
  10952. barcoded anatomy resolved by sequencing (BARseq), a multiplexed
  10953. method based on RNA barcoding for mapping projections of
  10954. thousands of spatially resolved neurons in a single brain and
  10955. relating those projections to other properties such as gene or
  10956. Cre expression. Mapping the projections to 11 areas of 3,579
  10957. neurons in mouse auditory cortex using BARseq confirmed the
  10958. laminar organization of the three top classes (intratelencephalic
  10959. [IT], pyramidal tract-like [PT-like], and corticothalamic [CT])
  10960. of projection neurons. In depth analysis uncovered a projection
  10961. type restricted almost exclusively to transcriptionally defined
  10962. subtypes of IT neurons. By bridging anatomical and transcriptomic
  10963. approaches at cellular resolution with high throughput, BARseq
  10964. can potentially uncover the organizing principles underlying the
  10965. structure and formation of neural circuits.",
  10966. journal = "Cell",
  10967. volume = 179,
  10968. number = 3,
  10969. pages = "772--786.e19",
  10970. month = oct,
  10971. year = 2019,
  10972. keywords = "high throughput; projection mapping; cellular barcoding;
  10973. sequencing; auditory cortex"
  10974. }
  10975. @UNPUBLISHED{Van_Wijngaarden2019-md,
  10976. title = "Representation of Distance and Direction of Nearby Boundaries in
  10977. Retrosplenial Cortex",
  10978. author = "van Wijngaarden, Joeri B G and Babl, Susanne S and Ito, Hiroshi T",
  10979. abstract = "Borders and edges are salient and behaviourally relevant features
  10980. for navigating the environment. The brain forms dedicated neural
  10981. representations of environmental boundaries, which are assumed to
  10982. serve as a reference for spatial coding. Here we expand this
  10983. border coding network to include the retrosplenial cortex (RSC)
  10984. in which we identified neurons that increase their firing near
  10985. all boundaries of an arena. RSC border cells specifically encode
  10986. walls, but not objects, and maintain their tuning in the absence
  10987. of direct sensory detection. Unlike border cells in the medial
  10988. entorhinal cortex (MEC), RSC border cells are sensitive to the
  10989. animal9s direction to nearby walls located contralateral to the
  10990. recorded hemisphere. Pharmacogenetic inactivation of MEC led to a
  10991. disruption of RSC border coding, but not vice versa, indicating
  10992. network directionality. Together these data shed light on how
  10993. information about distance and direction of boundaries is
  10994. generated in the brain for guiding navigation behaviour.",
  10995. journal = "bioRxiv",
  10996. pages = "807453",
  10997. month = oct,
  10998. year = 2019,
  10999. language = "en"
  11000. }
  11001. @ARTICLE{Headley2019-if,
  11002. title = "Embracing Complexity in Defensive Networks",
  11003. author = "Headley, Drew B and Kanta, Vasiliki and Kyriazi, Pinelopi and
  11004. Par{\'e}, Denis",
  11005. abstract = "The neural basis of defensive behaviors continues to attract
  11006. much interest, not only because they are important for survival
  11007. but also because their dysregulation may be at the origin of
  11008. anxiety disorders. Recently, a dominant approach in the field
  11009. has been the optogenetic manipulation of specific circuits or
  11010. cell types within these circuits to dissect their role in
  11011. different defensive behaviors. While the usefulness of
  11012. optogenetics is unquestionable, we argue that this method, as
  11013. currently applied, fosters an atomistic conceptualization of
  11014. defensive behaviors, which hinders progress in understanding the
  11015. integrated responses of nervous systems to threats. Instead, we
  11016. advocate for a holistic approach to the problem, including
  11017. observational study of natural behaviors and their neuronal
  11018. correlates at multiple sites, coupled to the use of
  11019. optogenetics, not to globally turn on or off neurons of
  11020. interest, but to manipulate specific activity patterns
  11021. hypothesized to regulate defensive behaviors.",
  11022. journal = "Neuron",
  11023. publisher = "Elsevier",
  11024. volume = 103,
  11025. number = 2,
  11026. pages = "189--201",
  11027. month = jul,
  11028. year = 2019,
  11029. keywords = "amygdala; defensive behaviors; extinction; fear; infralimbic;
  11030. medial prefrontal cortex; prelimbic",
  11031. language = "en"
  11032. }
  11033. @ARTICLE{Honegger2018-xu,
  11034. title = "Stochasticity, individuality and behavior",
  11035. author = "Honegger, Kyle and de Bivort, Benjamin",
  11036. abstract = "No two individuals are exactly alike. More than a simple
  11037. platitude, this observation reflects the fundamentally stochastic
  11038. nature of biological systems. The term 'stochastic' describes
  11039. features that cannot be predicted a priori from readily
  11040. measurable variables. In the dichotomous framework in which
  11041. biological variation arises from genetic or environmental
  11042. effects, stochastic effects are classified as environmental
  11043. because they are not passed on to offspring - any non-heritable
  11044. cause is, by definition, environmental. But non-heritable effects
  11045. can be subdivided into those which can be predicted from
  11046. measurable variables, and those that cannot. These latter effects
  11047. are stochastic.",
  11048. journal = "Curr. Biol.",
  11049. volume = 28,
  11050. number = 1,
  11051. pages = "R8--R12",
  11052. month = jan,
  11053. year = 2018,
  11054. language = "en"
  11055. }
  11056. @ARTICLE{Calhoun2017-ur,
  11057. title = "Quantifying behavior to solve sensorimotor transformations:
  11058. advances from worms and flies",
  11059. author = "Calhoun, Adam J and Murthy, Mala",
  11060. abstract = "The development of new computational tools has recently opened up
  11061. the study of natural behaviors at a precision that was previously
  11062. unachievable. These tools permit a highly quantitative analysis
  11063. of behavioral dynamics at timescales that are well matched to the
  11064. timescales of neural activity. Here we examine how combining
  11065. these methods with established techniques for estimating an
  11066. animal's sensory experience presents exciting new opportunities
  11067. for dissecting the sensorimotor transformations performed by the
  11068. nervous system. We focus this review primarily on examples from
  11069. Caenorhabditis elegans and Drosophila melanogaster-for these
  11070. model systems, computational approaches to characterize behavior,
  11071. in combination with unparalleled genetic tools for neural
  11072. activation, silencing, and recording, have already proven
  11073. instrumental for illuminating underlying neural mechanisms.",
  11074. journal = "Curr. Opin. Neurobiol.",
  11075. volume = 46,
  11076. pages = "90--98",
  11077. month = oct,
  11078. year = 2017,
  11079. language = "en"
  11080. }
  11081. @ARTICLE{Chiel1997-dl,
  11082. title = "The brain has a body: adaptive behavior emerges from interactions
  11083. of nervous system, body and environment",
  11084. author = "Chiel, H J and Beer, R D",
  11085. abstract = "Studies of mechanisms of adaptive behavior generally focus on
  11086. neurons and circuits. But adaptive behavior also depends on
  11087. interactions among the nervous system, body and environment:
  11088. sensory preprocessing and motor post-processing filter inputs to
  11089. and outputs from the nervous system; co-evolution and
  11090. co-development of nervous system and periphery create matching
  11091. and complementarity between them; body structure creates
  11092. constraints and opportunities for neural control; and continuous
  11093. feedback between nervous system, body and environment are
  11094. essential for normal behavior. This broader view of adaptive
  11095. behavior has been a major underpinning of ecological psychology
  11096. and has influenced behavior-based robotics. Computational
  11097. neuroethology, which jointly models neural control and periphery
  11098. of animals, is a promising methodology for understanding adaptive
  11099. behavior.",
  11100. journal = "Trends Neurosci.",
  11101. volume = 20,
  11102. number = 12,
  11103. pages = "553--557",
  11104. month = dec,
  11105. year = 1997,
  11106. language = "en"
  11107. }
  11108. @UNPUBLISHED{Jackson2019-yb,
  11109. title = "Many paths to the same goal: metaheuristic operation of brains
  11110. during natural behavior",
  11111. author = "Jackson, Brian J and Fatima, Gusti Lulu and Oh, Sujean and Gire,
  11112. David H",
  11113. abstract = "During self-guided behaviors animals rapidly identify the
  11114. constraints of the problems they face and adaptively employ
  11115. appropriate cognitive strategies and heuristics to solve these
  11116. problems[1][1],[2][2]. This ability is currently an area of
  11117. active investigation in artificial intelligence[3][3]. Recent
  11118. work in computer science has suggested that this type of flexible
  11119. problem solving could be achievable with metaheuristic approaches
  11120. in which specific algorithms are selected based upon the
  11121. identified demands of the problem to be
  11122. solved[4][4],[5][5],[6][6],[7][7]. Investigating how animals
  11123. employ such metaheuristics while solving self-guided natural
  11124. problems is a fertile area for biologically inspired algorithm
  11125. development. Here we show that animals adaptively shift cognitive
  11126. resources between sensory and memory systems during natural
  11127. behavior to optimize performance under uncertainty. We
  11128. demonstrate this using a new, laboratory-based discovery method
  11129. to define the strategies used to solve a difficult optimization
  11130. scenario, the stochastic ``traveling salesman''
  11131. problem[5][5],[8][8],[9][9]. Using this system we precisely
  11132. manipulated the strength of prior information available to
  11133. animals as well as the complexity of the problem. We find that
  11134. rats are capable of efficiently solving this problem, even under
  11135. conditions in which prior information is unreliable and the space
  11136. of possible solutions is large. We compared animal performance to
  11137. a Bayesian search and found that performance is consistent with a
  11138. metaheuristic approach that adaptively allocates cognitive
  11139. resources between sensory processing and memory, enhancing
  11140. sensory acuity and reducing memory load under conditions in which
  11141. prior information is unreliable. Our findings set the foundation
  11142. for new approaches to understand the neural substrates of natural
  11143. behavior as well as the rational development of biologically
  11144. inspired metaheuristic approaches for complex real-world
  11145. optimization. [1]: \#ref-1 [2]: \#ref-2 [3]: \#ref-3 [4]: \#ref-4
  11146. [5]: \#ref-5 [6]: \#ref-6 [7]: \#ref-7 [8]: \#ref-8 [9]: \#ref-9",
  11147. journal = "bioRxiv",
  11148. pages = "697607",
  11149. month = jul,
  11150. year = 2019,
  11151. language = "en"
  11152. }
  11153. @ARTICLE{Spiro1998-pw,
  11154. title = "Neuroethology: a meeting of brain and behavior",
  11155. author = "Spiro, J E and White, S A",
  11156. journal = "Neuron",
  11157. volume = 21,
  11158. number = 5,
  11159. pages = "981--989",
  11160. month = nov,
  11161. year = 1998,
  11162. language = "en"
  11163. }
  11164. @UNPUBLISHED{Findling2019-yc,
  11165. title = "Imprecise neural computations as source of human adaptive
  11166. behavior in volatile environments",
  11167. author = "Findling, Charles and Chopin, Nicolas and Koechlin, Etienne",
  11168. abstract = "Everyday life features uncertain and ever-changing situations. In
  11169. such environments, optimal adaptive behavior requires
  11170. higher-order inferential capabilities to grasp the volatility of
  11171. external contingencies. These capabilities however involve
  11172. complex and rapidly intractable computations, so that we poorly
  11173. understand how humans develop efficient adaptive behaviors in
  11174. such environments. Here we demonstrate this counterintuitive
  11175. result: simple, low-level inferential processes involving
  11176. imprecise computations conforming to the psychophysical Weber Law
  11177. actually lead to near-optimal adaptive behavior, regardless of
  11178. the environment volatility. Using volatile experimental settings,
  11179. we further show that such imprecise, low-level inferential
  11180. processes accounted for observed human adaptive performances,
  11181. unlike optimal adaptive models involving higher-order inferential
  11182. capabilities, their biologically more plausible, algorithmic
  11183. approximations and non-inferential adaptive models like
  11184. reinforcement learning. Thus, minimal inferential capabilities
  11185. may have evolved along with imprecise neural computations as
  11186. contributing to near-optimal adaptive behavior in real-life
  11187. environments, while leading humans to make suboptimal choices in
  11188. canonical decision-making tasks.",
  11189. journal = "bioRxiv",
  11190. pages = "799239",
  11191. month = oct,
  11192. year = 2019,
  11193. language = "en"
  11194. }
  11195. @ARTICLE{Datta2019-ph,
  11196. title = "Computational Neuroethology: A Call to Action",
  11197. author = "Datta, Sandeep Robert and Anderson, David J and Branson, Kristin
  11198. and Perona, Pietro and Leifer, Andrew",
  11199. abstract = "The brain is worthy of study because it is in charge of
  11200. behavior. A flurry of recent technical advances in measuring and
  11201. quantifying naturalistic behaviors provide an important
  11202. opportunity for advancing brain science. However, the problem of
  11203. understanding unrestrained behavior in the context of neural
  11204. recordings and manipulations remains unsolved, and developing
  11205. approaches to addressing this challenge is critical. Here we
  11206. discuss considerations in computational neuroethology---the
  11207. science of quantifying naturalistic behaviors for understanding
  11208. the brain---and propose strategies to evaluate progress. We
  11209. point to open questions that require resolution and call upon
  11210. the broader systems neuroscience community to further develop
  11211. and leverage measures of naturalistic, unrestrained behavior,
  11212. which will enable us to more effectively probe the richness and
  11213. complexity of the brain.",
  11214. journal = "Neuron",
  11215. publisher = "Elsevier",
  11216. volume = 104,
  11217. number = 1,
  11218. pages = "11--24",
  11219. month = oct,
  11220. year = 2019,
  11221. language = "en"
  11222. }
  11223. @ARTICLE{Gomez-Marin2019-nv,
  11224. title = "The Life of Behavior",
  11225. author = "Gomez-Marin, Alex and Ghazanfar, Asif A",
  11226. abstract = "Neuroscience needs behavior. However, it is daunting to render
  11227. the behavior of organisms intelligible without suppressing most,
  11228. if not all, references to life. When animals are treated as
  11229. passive stimulus-response, disembodied and identical machines,
  11230. the life of behavior perishes. Here, we distill three biological
  11231. principles (materiality, agency, and historicity), spell out
  11232. their consequences for the study of animal behavior, and
  11233. illustrate them with various examples from the literature. We
  11234. propose to put behavior back into context, with the brain in a
  11235. species-typical body and with the animal's body situated in the
  11236. world; stamp Newtonian time with nested ontogenetic and
  11237. phylogenetic processes that give rise to individuals with their
  11238. own histories; and supplement linear cause-and-effect chains and
  11239. information processing with circular loops of purpose and
  11240. meaning. We believe that conceiving behavior in these ways is
  11241. imperative for neuroscience.",
  11242. journal = "Neuron",
  11243. publisher = "Elsevier",
  11244. volume = 104,
  11245. number = 1,
  11246. pages = "25--36",
  11247. month = oct,
  11248. year = 2019,
  11249. language = "en"
  11250. }
  11251. @ARTICLE{Balleine2019-si,
  11252. title = "The Meaning of Behavior: Discriminating Reflex and Volition in
  11253. the Brain",
  11254. author = "Balleine, Bernard W",
  11255. abstract = "The ability to establish behaviorally what psychological
  11256. capacity an animal is deploying---to discern accurately what an
  11257. animal is doing---is key to functional analyses of the brain.
  11258. Our current understanding of these capacities suggests, however,
  11259. that this task is complex; there is evidence that multiple
  11260. capacities are engaged simultaneously and contribute
  11261. independently to the control of behavior. As such, establishing
  11262. the contribution of a cell, circuit, or neural system to any one
  11263. function requires careful dissection of that role from its
  11264. influence on other functions and, therefore, the careful
  11265. selection and design of behavioral tasks fit for that purpose.
  11266. Here I describe recent research that has sought to utilize
  11267. behavioral tools to investigate the neural bases of instrumental
  11268. conditioning, particularly the circuits and systems supporting
  11269. the capacity for goal-directed action, as opposed to conditioned
  11270. reflexes and habits, and how these sources of action control
  11271. interact to generate adaptive behavior.",
  11272. journal = "Neuron",
  11273. publisher = "Elsevier",
  11274. volume = 104,
  11275. number = 1,
  11276. pages = "47--62",
  11277. month = oct,
  11278. year = 2019,
  11279. keywords = "goal-directed action; habitual action; incentive learning;
  11280. reward; reinforcement; experienced value; predicted value;
  11281. corticostriatal circuits; behavioral analysis",
  11282. language = "en"
  11283. }
  11284. @ARTICLE{Keum2019-dz,
  11285. title = "Neural Basis of Observational Fear Learning: A Potential Model
  11286. of Affective Empathy",
  11287. author = "Keum, Sehoon and Shin, Hee-Sup",
  11288. abstract = "Observational fear learning in rodents is a type of
  11289. context-dependent fear conditioning in which an unconditioned
  11290. stimulus (US) is provided vicariously by observing conspecific
  11291. others receiving foot shocks. This suggests the involvement of
  11292. affective empathy, with several recent studies showing many
  11293. similarities between this behavior and human empathy.
  11294. Neurobiologically, it is important to understand the neural
  11295. mechanisms by which the vicarious US activates the fear circuit
  11296. via the affective pain system, obviating the sensory pain
  11297. pathway and eventually leading to fear memory formation. This
  11298. paper reviews current studies on the neural mechanisms
  11299. underlying observational fear learning and provides a
  11300. perspective on future research on this subject.",
  11301. journal = "Neuron",
  11302. publisher = "Elsevier",
  11303. volume = 104,
  11304. number = 1,
  11305. pages = "78--86",
  11306. month = oct,
  11307. year = 2019,
  11308. keywords = "empathy; observational fear learning; social fear; affective
  11309. pain; and anterior cingulate cortex",
  11310. language = "en"
  11311. }
  11312. @ARTICLE{Ma2019-uk,
  11313. title = "Bayesian Decision Models: A Primer",
  11314. author = "Ma, Wei Ji",
  11315. abstract = "To understand decision-making behavior in simple, controlled
  11316. environments, Bayesian models are often useful. First, optimal
  11317. behavior is always Bayesian. Second, even when behavior deviates
  11318. from optimality, the Bayesian approach offers candidate models
  11319. to account for suboptimalities. Third, a realist interpretation
  11320. of Bayesian models opens the door to studying the neural
  11321. representation of uncertainty. In this tutorial, we review the
  11322. principles of Bayesian models of decision making and then focus
  11323. on five case studies with exercises. We conclude with
  11324. reflections and future directions.",
  11325. journal = "Neuron",
  11326. publisher = "Elsevier",
  11327. volume = 104,
  11328. number = 1,
  11329. pages = "164--175",
  11330. month = oct,
  11331. year = 2019,
  11332. language = "en"
  11333. }
  11334. @BOOK{Cooper_Jr2015-qw,
  11335. title = "Escaping From Predators: An Integrative View of Escape Decisions",
  11336. author = "Cooper (Jr., William and Cooper, Jr, William E and Blumstein,
  11337. Daniel T",
  11338. abstract = "When a predator attacks, prey are faced with a series of 'if',
  11339. 'when' and 'how' escape decisions - these critical questions are
  11340. the foci of this book. Cooper and Blumstein bring together a
  11341. balance of theory and empirical research to summarise over fifty
  11342. years of scattered research and benchmark current thinking in
  11343. the rapidly expanding literature on the behavioural ecology of
  11344. escaping. The book consolidates current and new behaviour models
  11345. with taxonomically divided empirical chapters that demonstrate
  11346. the application of escape theory to different groups. The
  11347. chapters integrate behaviour with physiology, genetics and
  11348. evolution to lead the reader through the complex decisions faced
  11349. by prey during a predator attack, examining how these decisions
  11350. interact with life history and individual variation. The chapter
  11351. on best practice field methodology and the ideas for future
  11352. research presented throughout, ensure this volume is practical
  11353. as well as informative.",
  11354. publisher = "Cambridge University Press",
  11355. month = may,
  11356. year = 2015,
  11357. keywords = "books",
  11358. language = "en"
  11359. }
  11360. @ARTICLE{Zingg2017-bx,
  11361. title = "{AAV-Mediated} Anterograde Transsynaptic Tagging: Mapping
  11362. Corticocollicular {Input-Defined} Neural Pathways for Defense
  11363. Behaviors",
  11364. author = "Zingg, Brian and Chou, Xiao-Lin and Zhang, Zheng-Gang and Mesik,
  11365. Lukas and Liang, Feixue and Tao, Huizhong Whit and Zhang, Li I",
  11366. abstract = "To decipher neural circuits underlying brain functions, viral
  11367. tracers are widely applied to map input and output connectivity
  11368. of neuronal populations. Despite the successful application of
  11369. retrograde transsynaptic viruses for identifying presynaptic
  11370. neurons of transduced neurons, analogous anterograde
  11371. transsynaptic tools for tagging postsynaptically targeted neurons
  11372. remain under development. Here, we discovered that
  11373. adeno-associated viruses (AAV1 and AAV9) exhibit anterograde
  11374. transsynaptic spread properties. AAV1-Cre from transduced
  11375. presynaptic neurons effectively and specifically drives
  11376. Cre-dependent transgene expression in selected postsynaptic
  11377. neuronal targets, thus allowing axonal tracing and functional
  11378. manipulations of the latter input-defined neuronal population.
  11379. Its application in superior colliculus (SC) reveals that SC
  11380. neuron subpopulations receiving corticocollicular projections
  11381. from auditory and visual cortex specifically drive flight and
  11382. freezing, two different types of defense behavior, respectively.
  11383. Together with an intersectional approach, AAV-mediated
  11384. anterograde transsynaptic tagging can categorize neurons by their
  11385. inputs and molecular identity, and allow forward screening of
  11386. distinct functional neural pathways embedded in complex brain
  11387. circuits.",
  11388. journal = "Neuron",
  11389. volume = 93,
  11390. number = 1,
  11391. pages = "33--47",
  11392. month = jan,
  11393. year = 2017,
  11394. keywords = "AAV serotypes; Cre and Flp system; corticofugal projection;
  11395. defensive behavior; flight and freezing; intersectional strategy;
  11396. mapping neural circuits; superior colliculus;
  11397. transsynaptic/transneuronal tracer",
  11398. language = "en"
  11399. }
  11400. @ARTICLE{White2019-do,
  11401. title = "The Future Is Open: {Open-Source} Tools for Behavioral
  11402. Neuroscience Research",
  11403. author = "White, Samantha R and Amarante, Linda M and Kravitz, Alexxai V
  11404. and Laubach, Mark",
  11405. journal = "eNeuro",
  11406. volume = 6,
  11407. number = 4,
  11408. month = aug,
  11409. year = 2019,
  11410. keywords = "behavior; designs; methods; open source; protocols; tools",
  11411. language = "en"
  11412. }
  11413. @ARTICLE{Krumin2018-vd,
  11414. title = "Decision and navigation in mouse parietal cortex",
  11415. author = "Krumin, Michael and Lee, Julie J and Harris, Kenneth D and
  11416. Carandini, Matteo",
  11417. abstract = "Posterior parietal cortex (PPC) has been implicated in
  11418. navigation, in the control of movement, and in visually-guided
  11419. decisions. To relate these views, we measured activity in PPC
  11420. while mice performed a virtual navigation task driven by visual
  11421. decisions. PPC neurons were selective for specific combinations
  11422. of the animal's spatial position and heading angle. This
  11423. selectivity closely predicted both the activity of individual
  11424. PPC neurons, and the arrangement of their collective firing
  11425. patterns in choice-selective sequences. These sequences
  11426. reflected PPC encoding of the animal's navigation trajectory.
  11427. Using decision as a predictor instead of heading yielded worse
  11428. fits, and using it in addition to heading only slightly improved
  11429. the fits. Alternative models based on visual or motor variables
  11430. were inferior. We conclude that when mice use vision to choose
  11431. their trajectories, a large fraction of parietal cortex activity
  11432. can be predicted from simple attributes such as spatial position
  11433. and heading.",
  11434. journal = "Elife",
  11435. publisher = "cdn.elifesciences.org",
  11436. volume = 7,
  11437. month = nov,
  11438. year = 2018,
  11439. keywords = "cortex; decision; mouse; navigation; neuroscience; visual
  11440. processing",
  11441. language = "en"
  11442. }
  11443. @ARTICLE{Harvey2012-th,
  11444. title = "Choice-specific sequences in parietal cortex during a
  11445. virtual-navigation decision task",
  11446. author = "Harvey, Christopher D and Coen, Philip and Tank, David W",
  11447. abstract = "The posterior parietal cortex (PPC) has an important role in many
  11448. cognitive behaviours; however, the neural circuit dynamics
  11449. underlying PPC function are not well understood. Here we
  11450. optically imaged the spatial and temporal activity patterns of
  11451. neuronal populations in mice performing a PPC-dependent task that
  11452. combined a perceptual decision and memory-guided navigation in a
  11453. virtual environment. Individual neurons had transient activation
  11454. staggered relative to one another in time, forming a sequence of
  11455. neuronal activation spanning the entire length of a task trial.
  11456. Distinct sequences of neurons were triggered on trials with
  11457. opposite behavioural choices and defined divergent,
  11458. choice-specific trajectories through a state space of neuronal
  11459. population activity. Cells participating in the different
  11460. sequences and at distinct time points in the task were
  11461. anatomically intermixed over microcircuit length scales (<100
  11462. micrometres). During working memory decision tasks, the PPC may
  11463. therefore perform computations through sequence-based circuit
  11464. dynamics, rather than long-lived stable states, implemented using
  11465. anatomically intermingled microcircuits.",
  11466. journal = "Nature",
  11467. volume = 484,
  11468. number = 7392,
  11469. pages = "62--68",
  11470. month = mar,
  11471. year = 2012,
  11472. language = "en"
  11473. }
  11474. @ARTICLE{Whitlock2014-oa,
  11475. title = "Navigating actions through the rodent parietal cortex",
  11476. author = "Whitlock, Jonathan R",
  11477. abstract = "The posterior parietal cortex (PPC) participates in a manifold of
  11478. cognitive functions, including visual attention, working memory,
  11479. spatial processing, and movement planning. Given the vast
  11480. interconnectivity of PPC with sensory and motor areas, it is not
  11481. surprising that neuronal recordings show that PPC often encodes
  11482. mixtures of spatial information as well as the movements required
  11483. to reach a goal. Recent work sought to discern the relative
  11484. strength of spatial vs. motor signaling in PPC by recording
  11485. single unit activity in PPC of freely behaving rats during
  11486. selective changes in either the spatial layout of the local
  11487. environment or in the pattern of locomotor behaviors executed
  11488. during navigational tasks. The results revealed unequivocally a
  11489. predominant sensitivity of PPC neurons to locomotor action
  11490. structure, with subsets of cells even encoding upcoming movements
  11491. more than 1 s in advance. In light of these and other recent
  11492. findings in the field, I propose that one of the key
  11493. contributions of PPC to navigation is the synthesis of
  11494. goal-directed behavioral sequences, and that the rodent PPC may
  11495. serve as an apt system to investigate cellular mechanisms for
  11496. spatial motor planning as traditionally studied in humans and
  11497. monkeys.",
  11498. journal = "Front. Hum. Neurosci.",
  11499. volume = 8,
  11500. pages = "293",
  11501. month = may,
  11502. year = 2014,
  11503. keywords = "cognitive motor function; parietal cortex; parieto-frontal
  11504. network; rodent model; spatial navigation",
  11505. language = "en"
  11506. }
  11507. @ARTICLE{Mitchell2018-wv,
  11508. title = "Retrosplenial cortex and its role in spatial cognition",
  11509. author = "Mitchell, Anna S and Czajkowski, Rafal and Zhang, Ningyu and
  11510. Jeffery, Kate and Nelson, Andrew J D",
  11511. abstract = "Retrosplenial cortex is a region within the posterior neocortical
  11512. system, heavily interconnected with an array of brain networks,
  11513. both cortical and subcortical, that is, engaged by a myriad of
  11514. cognitive tasks. Although there is no consensus as to its precise
  11515. function, evidence from both human and animal studies clearly
  11516. points to a role in spatial cognition. However, the spatial
  11517. processing impairments that follow retrosplenial cortex damage
  11518. are not straightforward to characterise, leading to difficulties
  11519. in defining the exact nature of its role. In this article, we
  11520. review this literature and classify the types of ideas that have
  11521. been put forward into three broad, somewhat overlapping classes:
  11522. (1) learning of landmark location, stability and permanence; (2)
  11523. integration between spatial reference frames; and (3)
  11524. consolidation and retrieval of spatial knowledge (schemas). We
  11525. evaluate these models and suggest ways to test them, before
  11526. briefly discussing whether the spatial function may be a subset
  11527. of a more general function in episodic memory.",
  11528. journal = "Brain Neurosci Adv",
  11529. volume = 2,
  11530. pages = "2398212818757098",
  11531. month = mar,
  11532. year = 2018,
  11533. keywords = "Learning; cingulate cortex; default mode network;
  11534. electrophysiology; hippocampal formation; immediate-early genes;
  11535. memory; neuroimaging; primate; thalamus",
  11536. language = "en"
  11537. }
  11538. @ARTICLE{Kaufman2014-cw,
  11539. title = "Cortical activity in the null space: permitting preparation
  11540. without movement",
  11541. author = "Kaufman, Matthew T and Churchland, Mark M and Ryu, Stephen I and
  11542. Shenoy, Krishna V",
  11543. abstract = "Neural circuits must perform computations and then selectively
  11544. output the results to other circuits. Yet synapses do not change
  11545. radically at millisecond timescales. A key question then is: how
  11546. is communication between neural circuits controlled? In motor
  11547. control, brain areas directly involved in driving movement are
  11548. active well before movement begins. Muscle activity is some
  11549. readout of neural activity, yet it remains largely unchanged
  11550. during preparation. Here we find that during preparation, while
  11551. the monkey holds still, changes in motor cortical activity cancel
  11552. out at the level of these population readouts. Motor cortex can
  11553. thereby prepare the movement without prematurely causing it.
  11554. Further, we found evidence that this mechanism also operates in
  11555. dorsal premotor cortex, largely accounting for how preparatory
  11556. activity is attenuated in primary motor cortex. Selective use of
  11557. 'output-null' vs. 'output-potent' patterns of activity may thus
  11558. help control communication to the muscles and between these brain
  11559. areas.",
  11560. journal = "Nat. Neurosci.",
  11561. volume = 17,
  11562. number = 3,
  11563. pages = "440--448",
  11564. month = mar,
  11565. year = 2014,
  11566. language = "en"
  11567. }
  11568. @ARTICLE{Furth2018-lg,
  11569. title = "An interactive framework for whole-brain maps at cellular
  11570. resolution",
  11571. author = "F{\"u}rth, Daniel and Vaissi{\`e}re, Thomas and Tzortzi, Ourania
  11572. and Xuan, Yang and M{\"a}rtin, Antje and Lazaridis, Iakovos and
  11573. Spigolon, Giada and Fisone, Gilberto and Tomer, Raju and
  11574. Deisseroth, Karl and Carl{\'e}n, Marie and Miller, Courtney A and
  11575. Rumbaugh, Gavin and Meletis, Konstantinos",
  11576. abstract = "To deconstruct the architecture and function of brain circuits,
  11577. it is necessary to generate maps of neuronal connectivity and
  11578. activity on a whole-brain scale. New methods now enable
  11579. large-scale mapping of the mouse brain at cellular and
  11580. subcellular resolution. We developed a framework to automatically
  11581. annotate, analyze, visualize and easily share whole-brain data at
  11582. cellular resolution, based on a scale-invariant, interactive
  11583. mouse brain atlas. This framework enables connectivity and
  11584. mapping projects in individual laboratories and across imaging
  11585. platforms, as well as multiplexed quantitative information on the
  11586. molecular identity of single neurons. As a proof of concept, we
  11587. generated a comparative connectivity map of five major neuron
  11588. types in the corticostriatal circuit, as well as an
  11589. activity-based map to identify hubs mediating the behavioral
  11590. effects of cocaine. Thus, this computational framework provides
  11591. the necessary tools to generate brain maps that integrate data
  11592. from connectivity, neuron identity and function.",
  11593. journal = "Nat. Neurosci.",
  11594. volume = 21,
  11595. number = 1,
  11596. pages = "139--149",
  11597. month = jan,
  11598. year = 2018,
  11599. language = "en"
  11600. }
  11601. @UNPUBLISHED{Huang2018-ny,
  11602. title = "High-throughput mapping of mesoscale connectomes in individual
  11603. mice",
  11604. author = "Huang, Longwen and Kebschull, Justus M and Furth, Daniel and
  11605. Musall, Simon and Kaufman, Matthew T and Churchland, Anne K and
  11606. Zador, Anthony M",
  11607. abstract = "Abstract Brain function is determined by connectivity among brain
  11608. areas, and disruption of this connectivity leads to
  11609. neuropsychiatric disorders. Understanding connectivity is
  11610. essential to modern neuroscience, but mesoscale connectivity
  11611. atlases are currently slow and expensive to generate, exist for
  11612. few model systems, and require pooling across many brains. Here
  11613. we present a method, muMAPseq (multisource Multiplexed Analysis
  11614. of Projections by sequencing), which leverages barcoding and
  11615. high-throughput sequencing to generate atlases from single
  11616. animals rapidly and at low cost. We apply muMAPseq to tracing the
  11617. neocortical connectome of individual mice, and demonstrate high
  11618. reproducibility, and accuracy. Applying muMAPseq to the mutant
  11619. BTBR mouse strain, which lacks a corpus callosum, we recapitulate
  11620. its known connectopathies, and also uncover novel deficits.
  11621. muMAPseq allows individual laboratories to generate atlases
  11622. tailored to individuals, disease models, and new model species,
  11623. and will facilitate quantitative comparative connectomics,
  11624. permitting examination of how age, sex, environment, genetics and
  11625. species affect neuronal wiring.",
  11626. journal = "bioRxiv",
  11627. pages = "422477",
  11628. month = sep,
  11629. year = 2018,
  11630. language = "en"
  11631. }
  11632. @ARTICLE{Wolpert2012-ch,
  11633. title = "Motor control is decision-making",
  11634. author = "Wolpert, Daniel M and Landy, Michael S",
  11635. abstract = "Motor behavior may be viewed as a problem of maximizing the
  11636. utility of movement outcome in the face of sensory, motor and
  11637. task uncertainty. Viewed in this way, and allowing for the
  11638. availability of prior knowledge in the form of a probability
  11639. distribution over possible states of the world, the choice of a
  11640. movement plan and strategy for motor control becomes an
  11641. application of statistical decision theory. This point of view
  11642. has proven successful in recent years in accounting for movement
  11643. under risk, inferring the loss function used in motor tasks, and
  11644. explaining motor behavior in a wide variety of circumstances.",
  11645. journal = "Curr. Opin. Neurobiol.",
  11646. volume = 22,
  11647. number = 6,
  11648. pages = "996--1003",
  11649. month = dec,
  11650. year = 2012,
  11651. language = "en"
  11652. }
  11653. @ARTICLE{Wong2015-pi,
  11654. title = "Motor Planning",
  11655. author = "Wong, Aaron L and Haith, Adrian M and Krakauer, John W",
  11656. abstract = "Motor planning colloquially refers to any process related to the
  11657. preparation of a movement that occurs during the reaction time
  11658. prior to movement onset. However, this broad definition
  11659. encompasses processes that are not strictly motor-related, such
  11660. as decision-making about the identity of task-relevant stimuli in
  11661. the environment. Furthermore, the assumption that all
  11662. motor-planning processes require processing time, and can
  11663. therefore be studied behaviorally by measuring changes in the
  11664. reaction time, needs to be reexamined. In this review, we take a
  11665. critical look at the processes leading from perception to action
  11666. and suggest a definition of motor planning that encompasses only
  11667. those processes necessary for a movement to be executed-that is,
  11668. processes that are strictly movement related. These processes
  11669. resolve the ambiguity inherent in an abstract goal by defining a
  11670. specific movement to achieve it. We propose that the majority of
  11671. processes that meet this definition can be completed nearly
  11672. instantaneously, which means that motor planning itself in fact
  11673. consumes only a small fraction of the reaction time.",
  11674. journal = "Neuroscientist",
  11675. volume = 21,
  11676. number = 4,
  11677. pages = "385--398",
  11678. month = aug,
  11679. year = 2015,
  11680. keywords = "attention; decision making; dynamical systems model; motor
  11681. control; motor goal; optimal control theory; reaction time",
  11682. language = "en"
  11683. }
  11684. @ARTICLE{Mitrofanis2005-lq,
  11685. title = "Some certainty for the ``zone of uncertainty''? Exploring the
  11686. function of the zona incerta",
  11687. author = "Mitrofanis, J",
  11688. abstract = "The zona incerta (ZI), first described over a century ago by
  11689. Auguste Forel as a ``region of which nothing certain can be
  11690. said,'' forms a collection of cells that derives from the
  11691. diencephalon. To this day, we are still not certain of the
  11692. precise function of this ``zone of uncertainty'' although many
  11693. have been proposed, from controlling visceral activity to
  11694. shifting attention and from influencing arousal to maintaining
  11695. posture and locomotion. In this review, I shall outline the
  11696. recent advances in the understanding of the structure,
  11697. connectivity and functions of the ZI. I will then focus on a
  11698. possible and often neglected global role for the ZI, one that
  11699. links its diverse functions together. In particular, I aim to
  11700. highlight the idea that the ZI forms a primal center of the
  11701. diencephalon for generating direct responses (visceral, arousal,
  11702. attention and/or posture-locomotion) to a given sensory (somatic
  11703. and/or visceral) stimulus. With this global role in mind, I will
  11704. then address recent results indicating that abnormal ZI activity
  11705. manifests in clinical symptoms of Parkinson disease.",
  11706. journal = "Neuroscience",
  11707. volume = 130,
  11708. number = 1,
  11709. pages = "1--15",
  11710. year = 2005,
  11711. language = "en"
  11712. }
  11713. @ARTICLE{Chou2018-sq,
  11714. title = "Inhibitory gain modulation of defense behaviors by zona incerta",
  11715. author = "Chou, Xiao-Lin and Wang, Xiyue and Zhang, Zheng-Gang and Shen, Li
  11716. and Zingg, Brian and Huang, Junxiang and Zhong, Wen and Mesik,
  11717. Lukas and Zhang, Li I and Tao, Huizhong Whit",
  11718. abstract = "Zona incerta (ZI) is a functionally mysterious subthalamic
  11719. nucleus containing mostly inhibitory neurons. Here, we discover
  11720. that GABAergic neurons in the rostral sector of ZI (ZIr) directly
  11721. innervate excitatory but not inhibitory neurons in the
  11722. dorsolateral and ventrolateral compartments of periaqueductal
  11723. gray (PAG), which can drive flight and freezing behaviors
  11724. respectively. Optogenetic activation of ZIr neurons or their
  11725. projections to PAG reduces both sound-induced innate flight
  11726. response and conditioned freezing response, while optogenetic
  11727. suppression of these neurons enhances these defensive behaviors,
  11728. likely through a mechanism of gain modulation. ZIr activity
  11729. progressively increases during extinction of conditioned freezing
  11730. response, and suppressing ZIr activity impairs the expression of
  11731. fear extinction. Furthermore, ZIr is innervated by the medial
  11732. prefrontal cortex (mPFC), and silencing mPFC prevents the
  11733. increase of ZIr activity during extinction and the expression of
  11734. fear extinction. Together, our results suggest that ZIr is
  11735. engaged in modulating defense behaviors.",
  11736. journal = "Nat. Commun.",
  11737. volume = 9,
  11738. number = 1,
  11739. pages = "1151",
  11740. month = mar,
  11741. year = 2018,
  11742. language = "en"
  11743. }
  11744. @ARTICLE{Svoboda2018-rr,
  11745. title = "Neural mechanisms of movement planning: motor cortex and beyond",
  11746. author = "Svoboda, Karel and Li, Nuo",
  11747. abstract = "Neurons in motor cortex and connected brain regions fire in
  11748. anticipation of specific movements, long before movement occurs.
  11749. This neural activity reflects internal processes by which the
  11750. brain plans and executes volitional movements. The study of motor
  11751. planning offers an opportunity to understand how the structure
  11752. and dynamics of neural circuits support persistent internal
  11753. states and how these states influence behavior. Recent advances
  11754. in large-scale neural recordings are beginning to decipher the
  11755. relationship of the dynamics of populations of neurons during
  11756. motor planning and movements. New behavioral tasks in rodents,
  11757. together with quantified perturbations, link dynamics in specific
  11758. nodes of neural circuits to behavior. These studies reveal a
  11759. neural network distributed across multiple brain regions that
  11760. collectively supports motor planning. We review recent advances
  11761. and highlight areas where further work is needed to achieve a
  11762. deeper understanding of the mechanisms underlying motor planning
  11763. and related cognitive processes.",
  11764. journal = "Curr. Opin. Neurobiol.",
  11765. volume = 49,
  11766. pages = "33--41",
  11767. month = apr,
  11768. year = 2018,
  11769. language = "en"
  11770. }
  11771. @ARTICLE{Esposito2014-ls,
  11772. title = "Brainstem nucleus {MdV} mediates skilled forelimb motor tasks",
  11773. author = "Esposito, Maria Soledad and Capelli, Paolo and Arber, Silvia",
  11774. abstract = "Translating the behavioural output of the nervous system into
  11775. movement involves interaction between brain and spinal cord. The
  11776. brainstem provides an essential bridge between the two
  11777. structures, but circuit-level organization and function of this
  11778. intermediary system remain poorly understood. Here we use
  11779. intersectional virus tracing and genetic strategies in mice to
  11780. reveal a selective synaptic connectivity matrix between brainstem
  11781. substructures and functionally distinct spinal motor neurons that
  11782. regulate limb movement. The brainstem nucleus medullary reticular
  11783. formation ventral part (MdV) stands out as specifically targeting
  11784. subpopulations of forelimb-innervating motor neurons. Its
  11785. glutamatergic premotor neurons receive synaptic input from key
  11786. upper motor centres and are recruited during motor tasks.
  11787. Selective neuronal ablation or silencing experiments reveal that
  11788. MdV is critically important specifically for skilled motor
  11789. behaviour, including accelerating rotarod and single-food-pellet
  11790. reaching tasks. Our results indicate that distinct premotor
  11791. brainstem nuclei access spinal subcircuits to mediate
  11792. task-specific aspects of motor programs.",
  11793. journal = "Nature",
  11794. volume = 508,
  11795. number = 7496,
  11796. pages = "351--356",
  11797. month = apr,
  11798. year = 2014,
  11799. language = "en"
  11800. }
  11801. @ARTICLE{Sul2011-yj,
  11802. title = "Role of rodent secondary motor cortex in value-based action
  11803. selection",
  11804. author = "Sul, Jung Hoon and Jo, Suhyun and Lee, Daeyeol and Jung, Min Whan",
  11805. abstract = "Despite widespread neural activity related to reward values,
  11806. signals related to upcoming choice have not been clearly
  11807. identified in the rodent brain. Here we examined neuronal
  11808. activity in the lateral (AGl) and medial (AGm) agranular cortex,
  11809. corresponding to the primary and secondary motor cortex,
  11810. respectively, in rats performing a dynamic foraging task. Choice
  11811. signals, before behavioral manifestation of the rat's choice,
  11812. arose in the AGm earlier than in any other areas of the rat brain
  11813. previously studied under free-choice conditions. The AGm also
  11814. conveyed neural signals for decision value and chosen value. By
  11815. contrast, upcoming choice signals arose later, and value signals
  11816. were weaker, in the AGl. We also found that AGm lesions made the
  11817. rats' choices less dependent on dynamically updated values. These
  11818. results suggest that rodent secondary motor cortex might be
  11819. uniquely involved in both representing and reading out value
  11820. signals for flexible action selection.",
  11821. journal = "Nat. Neurosci.",
  11822. volume = 14,
  11823. number = 9,
  11824. pages = "1202--1208",
  11825. month = aug,
  11826. year = 2011,
  11827. keywords = "Locomotion;navigation",
  11828. language = "en"
  11829. }
  11830. @ARTICLE{Wang2015-yh,
  11831. title = "Collateral pathways from the ventromedial hypothalamus mediate
  11832. defensive behaviors",
  11833. author = "Wang, Li and Chen, Irene Z and Lin, Dayu",
  11834. abstract = "The ventromedial hypothalamus (VMH) was thought to be essential
  11835. for coping with threat, although its circuit mechanism remains
  11836. unclear. To investigate this, we optogenetically activated
  11837. steroidogenic factor 1 (SF1)-expressing neurons in the
  11838. dorsomedial and central parts of the VMH (VMHdm/c), and observed
  11839. a range of context-dependent somatomotor and autonomic responses
  11840. resembling animals' natural defensive behaviors. By activating
  11841. independent pathways emanating from the VMHdm/c, we demonstrated
  11842. that VMHdm/c projection to the dorsolateral periaqueductal gray
  11843. (dlPAG) induces inflexible immobility, while the VMHdm/c to
  11844. anterior hypothalamic nucleus (AHN) pathway promotes avoidance.
  11845. Consistent with the behavior changes induced by VMH to AHN
  11846. pathway activation, direct activation of the AHN elicited
  11847. avoidance and escape jumping, but not immobility. Retrograde
  11848. tracing studies revealed that nearly 50\% of PAG-projecting
  11849. VMHdm/c neurons send collateral projection to the AHN and vice
  11850. versa. Thus, VMHdm/c neurons employ a one-to-many wiring
  11851. configuration to orchestrate multiple aspects of defensive
  11852. behaviors.",
  11853. journal = "Neuron",
  11854. volume = 85,
  11855. number = 6,
  11856. pages = "1344--1358",
  11857. month = mar,
  11858. year = 2015,
  11859. language = "en"
  11860. }
  11861. @ARTICLE{Ferreira-Pinto2018-kr,
  11862. title = "Connecting Circuits for Supraspinal Control of Locomotion",
  11863. author = "Ferreira-Pinto, Manuel J and Ruder, Ludwig and Capelli, Paolo
  11864. and Arber, Silvia",
  11865. abstract = "Locomotion is regulated by distributed circuits and achieved by
  11866. the concerted activation of body musculature. While the basic
  11867. properties of executive circuits in the spinal cord are fairly
  11868. well understood, the precise mechanisms by which the brain
  11869. impacts locomotion are much less clear. This Review discusses
  11870. recent work unraveling the cellular identity, connectivity, and
  11871. function of supraspinal circuits. We focus on their involvement
  11872. in the regulation of the different phases of locomotion and
  11873. their interaction with spinal circuits. Dedicated neuronal
  11874. populations in the brainstem carry locomotor instructions,
  11875. including initiation, speed, and termination. To align
  11876. locomotion with behavioral needs, brainstem output structures
  11877. are recruited by midbrain and forebrain circuits that compute
  11878. and infer volitional, innate, and context-dependent locomotor
  11879. properties. We conclude that the emerging logic of supraspinal
  11880. circuit organization helps to understand how locomotor programs
  11881. from exploration to hunting and escape are regulated by the
  11882. brain.",
  11883. journal = "Neuron",
  11884. publisher = "Elsevier",
  11885. volume = 100,
  11886. number = 2,
  11887. pages = "361--374",
  11888. month = oct,
  11889. year = 2018,
  11890. keywords = "Locomotion",
  11891. language = "en"
  11892. }
  11893. @ARTICLE{Noga2003-nn,
  11894. title = "Mechanism for activation of locomotor centers in the spinal cord
  11895. by stimulation of the mesencephalic locomotor region",
  11896. author = "Noga, Brian R and Kriellaars, Dean J and Brownstone, Robert M
  11897. and Jordan, Larry M",
  11898. abstract = "The synaptic pathways of mesencephalic locomotor region
  11899. (MLR)-evoked excitatory and inhibitory postsynaptic potentials
  11900. (EPSPs and IPSPs) recorded from lumbar motoneurons of
  11901. unanesthetized decerebrate cats during fictive locomotion were
  11902. analyzed prior to, during, and after cold block of the medial
  11903. reticular formation (MedRF) or the low thoracic ventral
  11904. funiculus (VF). As others have shown, electrical stimulation of
  11905. the MLR typically evoked short-latency excitatory or mixed
  11906. excitatory/inhibitory PSPs in flexor and extensor motoneurons.
  11907. The bulbospinal conduction velocities averaged approximately 88
  11908. m/s (range: 62-145 m/s) and segmental latencies for EPSPs ranged
  11909. from 1.2 to 10.9 ms. The histogram of segmental latencies showed
  11910. three peaks, suggesting di-, tri-, and polysynaptic linkages.
  11911. Segmental latencies for IPSPs suggested trisynaptic or
  11912. polysynaptic transmission. Most EPSPs (69/77) were significantly
  11913. larger during the depolarized phase of the intracellular
  11914. locomotor drive potential (LDP), and most IPSPs (35/46) were
  11915. larger during the corresponding hyperpolarized phase. Bilateral
  11916. cooling of the MedRF reversibly abolished locomotion of both
  11917. hindlimbs as measured from the electroneurogram (ENG) activity
  11918. of muscle nerves and simultaneously abolished or diminished the
  11919. motoneuron PSPs and LDPs. Unilateral cooling of the VF blocked
  11920. locomotion ipsilaterally and diminished it contralaterally with
  11921. concomitant loss or decrease the motoneuron PSPs and LDPs.
  11922. Relative to the side of motoneuron recording, cooling of the
  11923. ipsilateral VF sometimes uncovered longer-latency EPSPs, whereas
  11924. cooling of the contralateral VF abolished longer-latency EPSPs.
  11925. It is concluded that MLR stimulation activates a pathway that
  11926. relays in the MedRF and descends bilaterally in the VF to
  11927. contact spinal interneurons that project to motoneurons. Local
  11928. segmental pathways that activate or inhibit motoneurons during
  11929. MLR-evoked fictive locomotion appear to be both ipsilateral and
  11930. contralateral.",
  11931. journal = "J. Neurophysiol.",
  11932. publisher = "physiology.org",
  11933. volume = 90,
  11934. number = 3,
  11935. pages = "1464--1478",
  11936. month = sep,
  11937. year = 2003,
  11938. language = "en"
  11939. }
  11940. @ARTICLE{Thompson2013-zi,
  11941. title = "Activity in mouse pedunculopontine tegmental nucleus reflects
  11942. action and outcome in a decision-making task",
  11943. author = "Thompson, John A and Felsen, Gidon",
  11944. abstract = "Recent studies across several mammalian species have revealed a
  11945. distributed network of cortical and subcortical brain regions
  11946. responsible for sensorimotor decision making. Many of these
  11947. regions have been shown to be interconnected with the
  11948. pedunculopontine tegmental nucleus (PPTg), a brain stem structure
  11949. characterized by neuronal heterogeneity and thought to be
  11950. involved in several cognitive and behavioral functions. However,
  11951. whether this structure plays a general functional role in
  11952. sensorimotor decision making is unclear. We hypothesized that, in
  11953. the context of a sensorimotor task, activity in the PPTg would
  11954. reflect task-related variables in a similar manner as do the
  11955. cortical and subcortical regions with which it is anatomically
  11956. associated. To examine this hypothesis, we recorded PPTg activity
  11957. in mice performing an odor-cued spatial choice task requiring a
  11958. stereotyped leftward or rightward orienting movement to obtain a
  11959. reward. We studied single-neuron activity during epochs of the
  11960. task related to movement preparation, execution, and outcome
  11961. (i.e., whether or not the movement was rewarded). We found that a
  11962. substantial proportion of neurons in the PPTg exhibited
  11963. direction-selective activity during one or more of these epochs.
  11964. In addition, an overlapping population of neurons reflected
  11965. movement direction and reward outcome. These results suggest that
  11966. the PPTg should be considered within the network of brain areas
  11967. responsible for sensorimotor decision making and lay the
  11968. foundation for future experiments to examine how the PPTg
  11969. interacts with other regions to control sensory-guided motor
  11970. output.",
  11971. journal = "J. Neurophysiol.",
  11972. volume = 110,
  11973. number = 12,
  11974. pages = "2817--2829",
  11975. month = dec,
  11976. year = 2013,
  11977. keywords = "basal ganglia; decision making; pedunculopontine tegmental
  11978. nucleus; sensorimotor",
  11979. language = "en"
  11980. }
  11981. @ARTICLE{Li2018-xo,
  11982. title = "Hypothalamic Circuits for Predation and Evasion",
  11983. author = "Li, Yi and Zeng, Jiawei and Zhang, Juen and Yue, Chenyu and
  11984. Zhong, Weixin and Liu, Zhixiang and Feng, Qiru and Luo, Minmin",
  11985. abstract = "The interactions between predator and prey represent some of the
  11986. most dramatic events in nature and constitute a matter of life
  11987. and death for both sides. The hypothalamus has been implicated in
  11988. driving predation and evasion; however, the exact hypothalamic
  11989. neural circuits underlying these behaviors remain poorly defined.
  11990. Here, we demonstrate that inhibitory and excitatory projections
  11991. from the mouse lateral hypothalamus (LH) to the periaqueductal
  11992. gray (PAG) in the midbrain drive, respectively, predation and
  11993. evasion. LH GABA neurons were activated during predation.
  11994. Optogenetically stimulating PAG-projecting LH GABA neurons drove
  11995. strong predatory attack, and inhibiting these cells reversibly
  11996. blocked predation. In contrast, LH glutamate neurons were
  11997. activated during evasion. Stimulating PAG-projecting LH glutamate
  11998. neurons drove evasion and inhibiting them impeded predictive
  11999. evasion. Therefore, the seemingly opposite behaviors of predation
  12000. and evasion are tightly regulated by two dissociable modular
  12001. command systems within a single neural projection from the LH to
  12002. the PAG. VIDEO ABSTRACT.",
  12003. journal = "Neuron",
  12004. volume = 97,
  12005. number = 4,
  12006. pages = "911--924.e5",
  12007. month = feb,
  12008. year = 2018,
  12009. keywords = "GABA; chemogenetics; escape behavior; fiber photometry;
  12010. glutamate; hunting behavior; lateral hypothalamus; optogenetics;
  12011. periaqueductal gray",
  12012. language = "en"
  12013. }
  12014. @ARTICLE{Sharma2019-bp,
  12015. title = "Towards a connectome of descending commands controlling
  12016. locomotion",
  12017. author = "Sharma, Sandeep and Kim, Linda H and Whelan, Patrick J",
  12018. abstract = "Understanding the neural basis for locomotion is of critical
  12019. importance since it subserves many behaviours necessary for
  12020. survival. The spinal cord contains all the elements required to
  12021. produce the basic locomotor pattern. These elements which compose
  12022. the central pattern generator for locomotion are activated and
  12023. sculpted by descending inputs from the brainstem, subcortical and
  12024. cortical structures. In this review, we examine the aspects of
  12025. descending control of spinal cord circuits, focusing on the
  12026. spinal cord, brainstem, and the diencephalon--hypothalamus. In
  12027. this short review, we discuss recent data and consider
  12028. opportunities for incorporating connectomics and optogenetic
  12029. advances to continue the progress in deciphering the descending
  12030. locomotor connectome.",
  12031. journal = "Current Opinion in Physiology",
  12032. volume = 8,
  12033. pages = "70--75",
  12034. month = apr,
  12035. year = 2019,
  12036. keywords = "Locomotion"
  12037. }
  12038. @ARTICLE{Roseberry2016-xm,
  12039. title = "{Cell-Type-Specific} Control of Brainstem Locomotor Circuits by
  12040. Basal Ganglia",
  12041. author = "Roseberry, Thomas K and Lee, A Moses and Lalive, Arnaud L and
  12042. Wilbrecht, Linda and Bonci, Antonello and Kreitzer, Anatol C",
  12043. abstract = "The basal ganglia (BG) are critical for adaptive motor control,
  12044. but the circuit principles underlying their pathway-specific
  12045. modulation of target regions are not well understood. Here, we
  12046. dissect the mechanisms underlying BG direct and indirect
  12047. pathway-mediated control of the mesencephalic locomotor region
  12048. (MLR), a brainstem target of BG that is critical for locomotion.
  12049. We optogenetically dissect the locomotor function of the three
  12050. neurochemically distinct cell types within the MLR:
  12051. glutamatergic, GABAergic, and cholinergic neurons. We find that
  12052. the glutamatergic subpopulation encodes locomotor state and
  12053. speed, is necessary and sufficient for locomotion, and is
  12054. selectively innervated by BG. We further show activation and
  12055. suppression, respectively, of MLR glutamatergic neurons by direct
  12056. and indirect pathways, which is required for bidirectional
  12057. control of locomotion by BG circuits. These findings provide a
  12058. fundamental understanding of how BG can initiate or suppress a
  12059. motor program through cell-type-specific regulation of neurons
  12060. linked to specific actions.",
  12061. journal = "Cell",
  12062. volume = 164,
  12063. number = 3,
  12064. pages = "526--537",
  12065. month = jan,
  12066. year = 2016,
  12067. language = "en"
  12068. }
  12069. @ARTICLE{Josset2018-oj,
  12070. title = "Distinct Contributions of Mesencephalic Locomotor Region Nuclei
  12071. to Locomotor Control in the Freely Behaving Mouse",
  12072. author = "Josset, Nicolas and Roussel, Marie and Lemieux, Maxime and
  12073. Lafrance-Zoubga, David and Rastqar, Ali and Bretzner, Frederic",
  12074. abstract = "The mesencephalic locomotor region (MLR) has been initially
  12075. identified as a supraspinal center capable of initiating and
  12076. modulating locomotion. Whereas its functional contribution to
  12077. locomotion has been widely documented throughout the phylogeny
  12078. from the lamprey to humans, there is still debate about its exact
  12079. organization. Combining kinematic and electrophysiological
  12080. recordings in mouse genetics, our study reveals that
  12081. glutamatergic neurons of the cuneiform nucleus initiate
  12082. locomotion and induce running gaits, whereas glutamatergic and
  12083. cholinergic neurons of the pedunculopontine nucleus modulate
  12084. locomotor pattern and rhythm, contributing to slow-walking gaits.
  12085. By initiating, modulating, and accelerating locomotion, our study
  12086. identifies and characterizes distinct neuronal populations of
  12087. this functional region important to locomotor command.",
  12088. journal = "Curr. Biol.",
  12089. volume = 28,
  12090. number = 6,
  12091. pages = "884--901.e3",
  12092. month = mar,
  12093. year = 2018,
  12094. keywords = "cuneiform nucleus; electrophysiology; glutamatergic and
  12095. cholinergic neurons; kinematic analysis; locomotor command;
  12096. locomotor pattern rhythm and gait; mesencephalic locomotor
  12097. region; optogenetic tools; pedunculopontine nucleus;Locomotion",
  12098. language = "en"
  12099. }
  12100. @ARTICLE{Kiehn2016-nj,
  12101. title = "Decoding the organization of spinal circuits that control
  12102. locomotion",
  12103. author = "Kiehn, Ole",
  12104. abstract = "Unravelling the functional operation of neuronal networks and
  12105. linking cellular activity to specific behavioural outcomes are
  12106. among the biggest challenges in neuroscience. In this broad field
  12107. of research, substantial progress has been made in studies of the
  12108. spinal networks that control locomotion. Through united efforts
  12109. using electrophysiological and molecular genetic network
  12110. approaches and behavioural studies in phylogenetically diverse
  12111. experimental models, the organization of locomotor networks has
  12112. begun to be decoded. The emergent themes from this research are
  12113. that the locomotor networks have a modular organization with
  12114. distinct transmitter and molecular codes and that their
  12115. organization is reconfigured with changes to the speed of
  12116. locomotion or changes in gait.",
  12117. journal = "Nat. Rev. Neurosci.",
  12118. volume = 17,
  12119. number = 4,
  12120. pages = "224--238",
  12121. month = apr,
  12122. year = 2016,
  12123. keywords = "Locomotion",
  12124. language = "en"
  12125. }
  12126. @ARTICLE{Azim2014-xn,
  12127. title = "Skilled reaching relies on a V2a propriospinal internal copy
  12128. circuit",
  12129. author = "Azim, Eiman and Jiang, Juan and Alstermark, Bror and Jessell,
  12130. Thomas M",
  12131. abstract = "The precision of skilled forelimb movement has long been
  12132. presumed to rely on rapid feedback corrections triggered by
  12133. internally directed copies of outgoing motor commands, but the
  12134. functional relevance of inferred internal copy circuits has
  12135. remained unclear. One class of spinal interneurons implicated in
  12136. the control of mammalian forelimb movement, cervical
  12137. propriospinal neurons (PNs), has the potential to convey an
  12138. internal copy of premotor signals through dual innervation of
  12139. forelimb-innervating motor neurons and precerebellar neurons of
  12140. the lateral reticular nucleus. Here we examine whether the PN
  12141. internal copy pathway functions in the control of goal-directed
  12142. reaching. In mice, PNs include a genetically accessible
  12143. subpopulation of cervical V2a interneurons, and their targeted
  12144. ablation perturbs reaching while leaving intact other elements
  12145. of forelimb movement. Moreover, optogenetic activation of the PN
  12146. internal copy branch recruits a rapid cerebellar feedback loop
  12147. that modulates forelimb motor neuron activity and severely
  12148. disrupts reaching kinematics. Our findings implicate V2a PNs as
  12149. the focus of an internal copy pathway assigned to the rapid
  12150. updating of motor output during reaching behaviour.",
  12151. journal = "Nature",
  12152. publisher = "nature.com",
  12153. volume = 508,
  12154. number = 7496,
  12155. pages = "357--363",
  12156. month = apr,
  12157. year = 2014,
  12158. language = "en"
  12159. }
  12160. @ARTICLE{Reinhold2019-pr,
  12161. title = "Behavioral and neural correlates of hide-and-seek in rats",
  12162. author = "Reinhold, Annika Stefanie and Sanguinetti-Scheck, Juan Ignacio
  12163. and Hartmann, Konstantin and Brecht, Michael",
  12164. abstract = "There is controversy regarding how widespread animal play
  12165. behavior is and what its evolutionary function might be.
  12166. Reinhold et al. demonstrated that rats can play hide-and-seek
  12167. with a human. In the ``seek'' condition, rats learned to look
  12168. for the hidden humans and kept seeking until they found them. In
  12169. the ``hide'' condition, they learned to hide in one of several
  12170. locations and waited there until being found. In both cases, the
  12171. rats were rewarded by social interaction with the human. Rats
  12172. vocalized when seeking and finding and were silent when hiding.
  12173. Recordings in the medial prefrontal cortex detected neurons that
  12174. were sensitive to the game structure. Science , this issue p.
  12175. [1180][1] Evolutionary, cognitive, and neural underpinnings of
  12176. mammalian play are not yet fully elucidated. We played
  12177. hide-and-seek, an elaborate role-play game, with rats. We did
  12178. not offer food rewards but engaged in playful interactions after
  12179. finding or being found. Rats quickly learned the game and
  12180. learned to alternate between hiding versus seeking roles. They
  12181. guided seeking by vision and memories of past hiding locations
  12182. and emitted game event--specific vocalizations. When hiding,
  12183. rats vocalized infrequently and they preferred opaque over
  12184. transparent hiding enclosures, a preference not observed during
  12185. seeking. Neuronal recordings revealed intense prefrontal cortex
  12186. activity that varied with game events and trial types (``hide''
  12187. versus ``seek'') and might instruct role play. The elaborate
  12188. cognitive capacities for hide-and-seek in rats suggest that this
  12189. game might be evolutionarily old. [1]:
  12190. /lookup/doi/10.1126/science.aax4705",
  12191. journal = "Science",
  12192. publisher = "American Association for the Advancement of Science",
  12193. volume = 365,
  12194. number = 6458,
  12195. pages = "1180--1183",
  12196. month = sep,
  12197. year = 2019,
  12198. language = "en"
  12199. }
  12200. @INCOLLECTION{May2006-kw,
  12201. title = "The mammalian superior colliculus: laminar structure and
  12202. connections",
  12203. booktitle = "Progress in Brain Research",
  12204. author = "May, Paul J",
  12205. editor = "B{\"u}ttner-Ennever, J A",
  12206. abstract = "The superior colliculus is a laminated midbrain structure that
  12207. acts as one of the centers organizing gaze movements. This
  12208. review will concentrate on sensory and motor inputs to the
  12209. superior colliculus, on its internal circuitry, and on its
  12210. connections with other brainstem gaze centers, as well as its
  12211. extensive outputs to those structures with which it is
  12212. reciprocally connected. This will be done in the context of its
  12213. laminar arrangement. Specifically, the superficial layers
  12214. receive direct retinal input, and are primarily visual sensory
  12215. in nature. They project upon the visual thalamus and pretectum
  12216. to influence visual perception. These visual layers also project
  12217. upon the deeper layers, which are both multimodal, and premotor
  12218. in nature. Thus, the deep layers receive input from both
  12219. somatosensory and auditory sources, as well as from the basal
  12220. ganglia and cerebellum. Sensory, association, and motor areas of
  12221. cerebral cortex provide another major source of collicular
  12222. input, particularly in more encephalized species. For example,
  12223. visual sensory cortex terminates superficially, while the eye
  12224. fields target the deeper layers. The deeper layers are
  12225. themselves the source of a major projection by way of the
  12226. predorsal bundle which contributes collicular target information
  12227. to the brainstem structures containing gaze-related burst
  12228. neurons, and the spinal cord and medullary reticular formation
  12229. regions that produce head turning.",
  12230. publisher = "Elsevier",
  12231. volume = 151,
  12232. pages = "321--378",
  12233. month = jan,
  12234. year = 2006
  12235. }
  12236. @ARTICLE{Tovote2016-fr,
  12237. title = "Midbrain circuits for defensive behaviour",
  12238. author = "Tovote, Philip and Esposito, Maria Soledad and Botta, Paolo and
  12239. Chaudun, Fabrice and Fadok, Jonathan P and Markovic, Milica and
  12240. Wolff, Steffen B E and Ramakrishnan, Charu and Fenno, Lief and
  12241. Deisseroth, Karl and Herry, Cyril and Arber, Silvia and
  12242. L{\"u}thi, Andreas",
  12243. abstract = "Survival in threatening situations depends on the selection and
  12244. rapid execution of an appropriate active or passive defensive
  12245. response, yet the underlying brain circuitry is not understood.
  12246. Here we use circuit-based optogenetic, in vivo and in vitro
  12247. electrophysiological, and neuroanatomical tracing methods to
  12248. define midbrain periaqueductal grey circuits for specific
  12249. defensive behaviours. We identify an inhibitory pathway from the
  12250. central nucleus of the amygdala to the ventrolateral
  12251. periaqueductal grey that produces freezing by disinhibition of
  12252. ventrolateral periaqueductal grey excitatory outputs to pre-motor
  12253. targets in the magnocellular nucleus of the medulla. In addition,
  12254. we provide evidence for anatomical and functional interaction of
  12255. this freezing pathway with long-range and local circuits
  12256. mediating flight. Our data define the neuronal circuitry
  12257. underlying the execution of freezing, an evolutionarily conserved
  12258. defensive behaviour, which is expressed by many species including
  12259. fish, rodents and primates. In humans, dysregulation of this
  12260. 'survival circuit' has been implicated in anxiety-related
  12261. disorders.",
  12262. journal = "Nature",
  12263. volume = 534,
  12264. number = 7606,
  12265. pages = "206--212",
  12266. month = jun,
  12267. year = 2016,
  12268. language = "en"
  12269. }
  12270. @ARTICLE{Kim2017-mo,
  12271. title = "Integration of Descending Command Systems for the Generation of
  12272. {Context-Specific} Locomotor Behaviors",
  12273. author = "Kim, Linda H and Sharma, Sandeep and Sharples, Simon A and Mayr,
  12274. Kyle A and Kwok, Charlie H T and Whelan, Patrick J",
  12275. abstract = "Over the past decade there has been a renaissance in our
  12276. understanding of spinal cord circuits; new technologies are
  12277. beginning to provide key insights into descending circuits which
  12278. project onto spinal cord central pattern generators. By
  12279. integrating work from both the locomotor and animal behavioral
  12280. fields, we can now examine context-specific control of
  12281. locomotion, with an emphasis on descending modulation arising
  12282. from various regions of the brainstem. Here we examine approach
  12283. and avoidance behaviors and the circuits that lead to the
  12284. production and arrest of locomotion.",
  12285. journal = "Front. Neurosci.",
  12286. volume = 11,
  12287. pages = "581",
  12288. month = oct,
  12289. year = 2017,
  12290. keywords = "approach; aversion; descending; goal-directed; locomotor
  12291. behavior; supraspinal",
  12292. language = "en"
  12293. }
  12294. @ARTICLE{Capelli2017-hq,
  12295. title = "Locomotor speed control circuits in the caudal brainstem",
  12296. author = "Capelli, Paolo and Pivetta, Chiara and Soledad Esposito, Maria
  12297. and Arber, Silvia",
  12298. abstract = "Locomotion is a universal behaviour that provides animals with
  12299. the ability to move between places. Classical experiments have
  12300. used electrical microstimulation to identify brain regions that
  12301. promote locomotion, but the identity of neurons that act as key
  12302. intermediaries between higher motor planning centres and
  12303. executive circuits in the spinal cord has remained controversial.
  12304. Here we show that the mouse caudal brainstem encompasses
  12305. functionally heterogeneous neuronal subpopulations that have
  12306. differential effects on locomotion. These subpopulations are
  12307. distinguishable by location, neurotransmitter identity and
  12308. connectivity. Notably, glutamatergic neurons within the lateral
  12309. paragigantocellular nucleus (LPGi), a small subregion in the
  12310. caudal brainstem, are essential to support high-speed locomotion,
  12311. and can positively tune locomotor speed through inputs from
  12312. glutamatergic neurons of the upstream midbrain locomotor region.
  12313. By contrast, glycinergic inhibitory neurons can induce different
  12314. forms of behavioural arrest mapping onto distinct caudal
  12315. brainstem regions. Anatomically, descending pathways of
  12316. glutamatergic and glycinergic LPGi subpopulations communicate
  12317. with distinct effector circuits in the spinal cord. Our results
  12318. reveal that behaviourally opposing locomotor functions in the
  12319. caudal brainstem were historically masked by the unexposed
  12320. diversity of intermingled neuronal subpopulations. We demonstrate
  12321. how specific brainstem neuron populations represent essential
  12322. substrates to implement key parameters in the execution of motor
  12323. programs.",
  12324. journal = "Nature",
  12325. volume = 551,
  12326. number = 7680,
  12327. pages = "373--377",
  12328. month = nov,
  12329. year = 2017,
  12330. keywords = "Locomotion",
  12331. language = "en"
  12332. }
  12333. @INCOLLECTION{Knudsen2017-ji,
  12334. title = "1.21 - The Optic Tectum: A Structure Evolved for Stimulus
  12335. Selection",
  12336. booktitle = "Evolution of Nervous Systems (Second Edition)",
  12337. author = "Knudsen, E I and Schwarz, J S",
  12338. editor = "Kaas, Jon H",
  12339. abstract = "The core function of the optic tectum (OT) in all vertebrates is
  12340. to collect information about the location and immediate
  12341. relevance of stimuli in the environment and, based on this
  12342. information, to compute the ``highest priority'' stimulus at
  12343. each moment in time. The OT transmits the location of this
  12344. stimulus to the forebrain to help direct spatial attention and,
  12345. when appropriate, to the brain stem and spinal cord to guide
  12346. immediate orienting or defensive behaviors. This chapter
  12347. describes how a stimulus selection network in the midbrain
  12348. generates the OT signal that identifies the location of the
  12349. ``highest priority'' stimulus.",
  12350. publisher = "Academic Press",
  12351. pages = "387--408",
  12352. month = jan,
  12353. year = 2017,
  12354. address = "Oxford",
  12355. keywords = "Amphibians; Attention; Birds; Fish; Gamma oscillations; Gaze
  12356. control; Isthmic nuclei; Mammals; Orienting response; Primates;
  12357. Reptiles; Retinotectal; Selective attention; Superior
  12358. colliculus; Visual pathways"
  12359. }
  12360. @UNPUBLISHED{Cregg2019-ah,
  12361. title = "Brainstem Neurons that Command {Left/Right} Locomotor Asymmetries",
  12362. author = "Cregg, Jared M and Leiras, Roberto and Montalant, Alexia and
  12363. Wickersham, Ian R and Kiehn, Ole",
  12364. abstract = "Descending command neurons instruct spinal networks to execute
  12365. basic locomotor functions, such as which gait and what speed. The
  12366. command functions for gait and speed are symmetric, implying that
  12367. a separate unknown system directs asymmetric movements---the
  12368. ability to move left or right. Here we report the discovery that
  12369. Chx10-lineage reticulospinal neurons act to control the direction
  12370. of locomotor movements in mammals. Chx10 neurons exhibit
  12371. ipsilateral projection, and can decrease spinal limb-based
  12372. locomotor activity ipsilaterally. This circuit mechanism acts as
  12373. the basis for left or right locomotor movements in freely moving
  12374. animals: selective unilateral activation of Chx10 neurons causes
  12375. ipsilateral movements whereas inhibition causes contralateral
  12376. movements. Spontaneous forward locomotion is thus transformed
  12377. into an ipsilateral movement by braking locomotion on the
  12378. ipsilateral side. We identify sensorimotor brain regions that
  12379. project onto Chx10 reticulospinal neurons, and demonstrate that
  12380. their unilateral activation can impart left/right directional
  12381. commands. Together these data identify the descending motor
  12382. system which commands left/right locomotor asymmetries in
  12383. mammals.",
  12384. journal = "bioRxiv",
  12385. pages = "754812",
  12386. month = sep,
  12387. year = 2019,
  12388. keywords = "Locomotion",
  12389. language = "en"
  12390. }
  12391. @ARTICLE{Comer2009-mh,
  12392. title = "Behavioral biology: inside the mind of proteus?",
  12393. author = "Comer, Christopher",
  12394. abstract = "A new study of the escape behavior of the cockroach has found
  12395. that its spatial variability is based on some underlying
  12396. regularity. This constrained variability may maximise the
  12397. effectiveness of the escape strategy.",
  12398. journal = "Curr. Biol.",
  12399. volume = 19,
  12400. number = 1,
  12401. pages = "R27--8",
  12402. month = jan,
  12403. year = 2009,
  12404. language = "en"
  12405. }
  12406. @ARTICLE{Arnott1999-vi,
  12407. title = "Escape trajectories of the brown shrimp crangon crangon, and a
  12408. theoretical consideration of initial escape angles from predators",
  12409. author = "Arnott, S A and Neil, D M and Ansell, A D",
  12410. abstract = "Tail-flip escape trajectories of the brown shrimp Crangon crangon
  12411. have been investigated in response to a natural predator, the cod
  12412. Gadus morhua, and an artificial stimulus. Shrimps escaped by
  12413. rolling to their left or right during the initial tail-flip of a
  12414. response, and thereafter swam on their side. As a result of the
  12415. laterally directed first tail-flip, initial escape angles always
  12416. lay between 75 degrees and 156 degrees with respect to the
  12417. (pre-escape) longitudinal axis (anterior=0 degrees) of the
  12418. shrimp. Symmetrical attacks from either head-on or tail-on
  12419. produced escapes to the shrimp's left or right in equal
  12420. proportions, although a contralateral bias did occur if the
  12421. shrimp experienced a looming object from one side before a
  12422. symmetrical attack was applied. Lateral attacks produced a
  12423. significantly greater proportion of contralateral responses than
  12424. ipsilateral ones. Empirical and theoretical analyses indicate
  12425. that the initial escape direction is influenced by an interaction
  12426. between the range of first tail-flip escape angles that the
  12427. shrimp is capable of performing and the risk of being intercepted
  12428. by a predator during the initial stage of an escape. Thus, the
  12429. unpredictability ('protean behaviour') of the response may be
  12430. affected by the conditions of the interaction. Subsequent
  12431. tail-flips of an escape usually directed the response away from
  12432. the stimulus, but sometimes escapes were instead steered to the
  12433. side of the stimulus and then behind it. The probability of each
  12434. type of escape occurring changed with attack direction. The
  12435. elements of protean behaviour that have been identified in both
  12436. the initial and subsequent stages of the escape may prevent
  12437. predators from learning a fixed pattern of response, but a
  12438. trade-off occurs when escape trajectories infringe upon zones of
  12439. high capture risk.",
  12440. journal = "J. Exp. Biol.",
  12441. volume = "202 (Pt 2)",
  12442. pages = "193--209",
  12443. month = jan,
  12444. year = 1999,
  12445. language = "en"
  12446. }
  12447. @ARTICLE{Poucet2004-wh,
  12448. title = "Spatial navigation and hippocampal place cell firing: the problem
  12449. of goal encoding",
  12450. author = "Poucet, B and Lenck-Santini, P P and Hok, V and Save, E and
  12451. Banquet, J P and Gaussier, P and Muller, R U",
  12452. abstract = "Place cells are hippocampal neurons whose discharge is strongly
  12453. related to a rat's location in the environment. The existence of
  12454. such cells, combined with the reliable impairments seen in
  12455. spatial tasks after hippocampal damage, has led to the proposal
  12456. that place cells form part of an integrated neural system
  12457. dedicated to spatial navigation. This hypothesis is supported by
  12458. the strong relationships between place cell activity and spatial
  12459. problem solving, which indicate that the place cell
  12460. representation must be both functional and in register with the
  12461. surroundings for the animal to perform correctly in spatial
  12462. tasks. The place cell system nevertheless requires other
  12463. essential elements to be competent, such as a component that
  12464. specifies the overall goal of the animal and computes the path
  12465. required to take the rat from its current location to the goal.
  12466. Here, we propose a model of the neural network responsible for
  12467. spatial navigation that includes goal coding and path selection.
  12468. In this model, the hippocampal formation allows for place
  12469. recognition, and stores the set of places that can be accessed
  12470. from each position in the environment. The prefrontal cortex is
  12471. responsible for encoding goal location and for route planning.
  12472. The nucleus accumbens translates paths in neural space into
  12473. appropriate locomotor activity that moves the animal towards the
  12474. goal in real space. The complete model assumes that the
  12475. hippocampal output to nucleus accumbens and prefrontal cortex
  12476. provides information for generating solutions to spatial
  12477. problems. In support of this model, we finally present
  12478. preliminary evidence that the goal representation necessary for
  12479. path planning might be encoded in the prelimbic/infralimbic
  12480. region of the medial prefrontal cortex.",
  12481. journal = "Rev. Neurosci.",
  12482. volume = 15,
  12483. number = 2,
  12484. pages = "89--107",
  12485. year = 2004,
  12486. language = "en"
  12487. }
  12488. @ARTICLE{Alyan1994-ev,
  12489. title = "Short-range homing in the house mouse, Mus musculus: stages in
  12490. the learning of directions",
  12491. author = "Alyan, Sofyan and Jander, Rudolf",
  12492. abstract = "Abstract. Female house mice readily learn to retrieve their pups
  12493. 50 cm from the centre of an open arena and take them to their
  12494. nest outside the arena's periphery. Experimental manipulation to
  12495. reveal the spatial-orientation constituents of this behaviour
  12496. disclosed thus submechanisms. Guided orientation, the direct
  12497. response to objects. Path integration, the continuous monitoring
  12498. of spatial displacements combined with computation of the
  12499. locomotor vector to the starting point of the path. Landmark
  12500. navigation, the movement by means of distal visual cues toward a
  12501. goal not directly perceived. Learning to home passes through
  12502. three stages. First, the exploring mouse is directly guided to
  12503. objects of interest. Second, the homing mouse adds path
  12504. integration; that is, it keeps a running, integrated spatial
  12505. record derived from locomotion. Finally (circumstances
  12506. permitting) the homing mouse links path integration with spatial
  12507. references to distal visual landmarks. Sparse comparative
  12508. evidence from other species of rodents suggests that such a
  12509. system of short-range topographical orientation is universal
  12510. among rodents.",
  12511. journal = "Anim. Behav.",
  12512. volume = 48,
  12513. number = 2,
  12514. pages = "285--298",
  12515. month = aug,
  12516. year = 1994
  12517. }
  12518. @ARTICLE{Ito2015-ld,
  12519. title = "A prefrontal-thalamo-hippocampal circuit for goal-directed
  12520. spatial navigation",
  12521. author = "Ito, Hiroshi T and Zhang, Sheng-Jia and Witter, Menno P and
  12522. Moser, Edvard I and Moser, May-Britt",
  12523. abstract = "Spatial navigation requires information about the relationship
  12524. between current and future positions. The activity of hippocampal
  12525. neurons appears to reflect such a relationship, representing not
  12526. only instantaneous position but also the path towards a goal
  12527. location. However, how the hippocampus obtains information about
  12528. goal direction is poorly understood. Here we report a
  12529. prefrontal-thalamic neural circuit that is required for
  12530. hippocampal representation of routes or trajectories through the
  12531. environment. Trajectory-dependent firing was observed in medial
  12532. prefrontal cortex, the nucleus reuniens of the thalamus, and the
  12533. CA1 region of the hippocampus in rats. Lesioning or optogenetic
  12534. silencing of the nucleus reuniens substantially reduced
  12535. trajectory-dependent CA1 firing. Trajectory-dependent activity
  12536. was almost absent in CA3, which does not receive nucleus reuniens
  12537. input. The data suggest that projections from medial prefrontal
  12538. cortex, via the nucleus reuniens, are crucial for representation
  12539. of the future path during goal-directed behaviour and point to
  12540. the thalamus as a key node in networks for long-range
  12541. communication between cortical regions involved in navigation.",
  12542. journal = "Nature",
  12543. volume = 522,
  12544. number = 7554,
  12545. pages = "50--55",
  12546. month = jun,
  12547. year = 2015,
  12548. language = "en"
  12549. }
  12550. @ARTICLE{Vedder2017-qz,
  12551. title = "Retrosplenial Cortical Neurons Encode Navigational Cues,
  12552. Trajectories and Reward Locations During Goal Directed Navigation",
  12553. author = "Vedder, Lindsey C and Miller, Adam M P and Harrison, Marc B and
  12554. Smith, David M",
  12555. abstract = "The retrosplenial cortex (RSC) plays an important role in memory
  12556. and spatial navigation. It shares functional similarities with
  12557. the hippocampus, including the presence of place fields and
  12558. lesion-induced impairments in spatial navigation, and the RSC is
  12559. an important source of visual-spatial input to the hippocampus.
  12560. Recently, the RSC has been the target of intense scrutiny among
  12561. investigators of human memory and navigation. fMRI and lesion
  12562. data suggest an RSC role in the ability to use landmarks to
  12563. navigate to goal locations. However, no direct neurophysiological
  12564. evidence of encoding navigational cues has been reported so the
  12565. specific RSC contribution to spatial cognition has been
  12566. uncertain. To examine this, we trained rats on a T-maze task in
  12567. which the reward location was explicitly cued by a flashing light
  12568. and we recorded RSC neurons as the rats learned. We found that
  12569. RSC neurons rapidly encoded the light cue. Additionally, RSC
  12570. neurons encoded the reward and its location, and they showed
  12571. distinct firing patterns along the left and right trajectories to
  12572. the goal. These responses may provide key information for
  12573. goal-directed navigation, and the loss of these signals may
  12574. underlie navigational impairments in subjects with RSC damage.",
  12575. journal = "Cereb. Cortex",
  12576. volume = 27,
  12577. number = 7,
  12578. pages = "3713--3723",
  12579. month = jul,
  12580. year = 2017,
  12581. keywords = "cingulate cortex; landmark; learning and memory; visual cue",
  12582. language = "en"
  12583. }
  12584. @ARTICLE{Feierstein2006-hs,
  12585. title = "Representation of spatial goals in rat orbitofrontal cortex",
  12586. author = "Feierstein, Claudia E and Quirk, Michael C and Uchida, Naoshige
  12587. and Sosulski, Dara L and Mainen, Zachary F",
  12588. abstract = "The orbitofrontal cortex (OFC) is thought to participate in
  12589. making and evaluating goal-directed decisions. In rodents,
  12590. spatial navigation is a major mode of goal-directed behavior, and
  12591. anatomical and lesion studies implicate the OFC in spatial
  12592. processing, but there is little direct evidence for coding of
  12593. spatial or motor variables. Here, we recorded from ventrolateral
  12594. and lateral OFC in an odor-cued two-alternative choice task
  12595. requiring orientation and approach to spatial goal ports. In this
  12596. context, over half of OFC neurons encoded choice direction or
  12597. goal port location. A subset of neurons was jointly selective for
  12598. the trial outcome and port location, information useful for the
  12599. selection or evaluation of spatial goals. These observations show
  12600. that the rodent OFC not only encodes information relating to
  12601. general motivational significance, as shown previously, but also
  12602. encodes spatiomotor variables needed to define specific
  12603. behavioral goals and the locomotor actions required to attain
  12604. them.",
  12605. journal = "Neuron",
  12606. volume = 51,
  12607. number = 4,
  12608. pages = "495--507",
  12609. month = aug,
  12610. year = 2006,
  12611. language = "en"
  12612. }
  12613. @ARTICLE{Domenici2008-so,
  12614. title = "Cockroaches keep predators guessing by using preferred escape
  12615. trajectories",
  12616. author = "Domenici, Paolo and Booth, David and Blagburn, Jonathan M and
  12617. Bacon, Jonathan P",
  12618. abstract = "Antipredator behavior is vital for most animals and calls for
  12619. accurate timing and swift motion. Whereas fast reaction times [1]
  12620. and predictable, context-dependent escape-initiation distances
  12621. [2] are common features of most escape systems, previous work has
  12622. highlighted the need for unpredictability in escape directions,
  12623. in order to prevent predators from learning a repeated, fixed
  12624. pattern [3-5]. Ultimate unpredictability would result from random
  12625. escape trajectories. Although this strategy would deny any
  12626. predictive power to the predator, it would also result in some
  12627. escape trajectories toward the threat. Previous work has shown
  12628. that escape trajectories are in fact generally directed away from
  12629. the threat, although with a high variability [5-8]. However, the
  12630. rules governing this variability are largely unknown. Here, we
  12631. demonstrate that individual cockroaches (Periplaneta americana, a
  12632. much-studied model prey species [9-14]) keep each escape
  12633. unpredictable by running along one of a set of preferred
  12634. trajectories at fixed angles from the direction of the
  12635. threatening stimulus. These results provide a new paradigm for
  12636. understanding the behavioral strategies for escape responses,
  12637. underscoring the need to revisit the neural mechanisms
  12638. controlling escape directions in the cockroach and similar animal
  12639. models, and the evolutionary forces driving unpredictable, or
  12640. ``protean''[3], antipredator behavior.",
  12641. journal = "Curr. Biol.",
  12642. volume = 18,
  12643. number = 22,
  12644. pages = "1792--1796",
  12645. month = nov,
  12646. year = 2008,
  12647. language = "en"
  12648. }
  12649. @ARTICLE{Moore2017-ge,
  12650. title = "Unpredictability of escape trajectory explains predator evasion
  12651. ability and microhabitat preference of desert rodents",
  12652. author = "Moore, Talia Y and Cooper, Kimberly L and Biewener, Andrew A and
  12653. Vasudevan, Ramanarayan",
  12654. abstract = "Mechanistically linking movement behaviors and ecology is key to
  12655. understanding the adaptive evolution of locomotion. Predator
  12656. evasion, a behavior that enhances fitness, may depend upon short
  12657. bursts or complex patterns of locomotion. However, such movements
  12658. are poorly characterized by existing biomechanical metrics. We
  12659. present methods based on the entropy measure of randomness from
  12660. Information Theory to quantitatively characterize the
  12661. unpredictability of non-steady-state locomotion. We then apply
  12662. the method by examining sympatric rodent species whose escape
  12663. trajectories differ in dimensionality. Unlike the speed-regulated
  12664. gait use of cursorial animals to enhance locomotor economy,
  12665. bipedal jerboa (family Dipodidae) gait transitions likely enhance
  12666. maneuverability. In field-based observations, jerboa trajectories
  12667. are significantly less predictable than those of quadrupedal
  12668. rodents, likely increasing predator evasion ability. Consistent
  12669. with this hypothesis, jerboas exhibit lower anxiety in open
  12670. fields than quadrupedal rodents, a behavior that varies inversely
  12671. with predator evasion ability. Our unpredictability metric
  12672. expands the scope of quantitative biomechanical studies to
  12673. include non-steady-state locomotion in a variety of evolutionary
  12674. and ecologically significant contexts.Biomechanical understanding
  12675. of animal gait and maneuverability has primarily been limited to
  12676. species with more predictable, steady-state movement patterns.
  12677. Here, the authors develop a method to quantify movement
  12678. predictability, and apply the method to study escape-related
  12679. movement in several species of desert rodents.",
  12680. journal = "Nat. Commun.",
  12681. volume = 8,
  12682. number = 1,
  12683. pages = "440",
  12684. month = sep,
  12685. year = 2017,
  12686. language = "en"
  12687. }
  12688. @ARTICLE{Card2012-fz,
  12689. title = "Escape behaviors in insects",
  12690. author = "Card, Gwyneth M",
  12691. abstract = "Escape behaviors are, by necessity, fast and robust, making them
  12692. excellent systems with which to study the neural basis of
  12693. behavior. This is especially true in insects, which have
  12694. comparatively tractable nervous systems and members who are
  12695. amenable to manipulation with genetic tools. Recent technical
  12696. developments in high-speed video reveal that, despite their short
  12697. duration, insect escape behaviors are more complex than
  12698. previously appreciated. For example, before initiating an escape
  12699. jump, a fly performs sophisticated posture and stimulus-dependent
  12700. preparatory leg movements that enable it to jump away from a
  12701. looming threat. This newfound flexibility raises the question of
  12702. how the nervous system generates a behavior that is both rapid
  12703. and flexible. Recordings from the cricket nervous system suggest
  12704. that synchrony between the activity of specific interneuron pairs
  12705. may provide a rapid cue for the cricket to detect the direction
  12706. of an approaching predator and thus which direction it should
  12707. run. Technical advances make possible wireless recording from
  12708. neurons while locusts escape from a looming threat, enabling, for
  12709. the first time, a direct correlation between the activity of
  12710. multiple neurons and the time-course of an insect escape
  12711. behavior.",
  12712. journal = "Curr. Opin. Neurobiol.",
  12713. volume = 22,
  12714. number = 2,
  12715. pages = "180--186",
  12716. month = apr,
  12717. year = 2012,
  12718. language = "en"
  12719. }
  12720. @INCOLLECTION{Balakrishnan2014-zj,
  12721. title = "Escape Trajectory",
  12722. booktitle = "Wiley {StatsRef}: Statistics Reference Online",
  12723. editor = "Balakrishnan, N and Colton, Theodore and Everitt, Brian and
  12724. Piegorsch, Walter and Ruggeri, Fabrizio and Teugels, Jozef L",
  12725. abstract = "Abstract Escape responses consist of sudden accelerations as
  12726. reactions to threatening stimuli and are present in most animals
  12727. as a means of avoiding predation. Escape responses are currently
  12728. receiving increased attention as valuable models for animal
  12729. behavior, ecology, and neurobiology. The path along which an
  12730. animal moves during an escape response (its escape trajectory)
  12731. has attracted the interest of ecologists and behaviorists in
  12732. relation to its functional role in predator?prey interactions.
  12733. Escape trajectories (ETs) as considered in this article refer to
  12734. the initial prey response to a predator's attack. Such a
  12735. trajectory should be measured at the end of the main rotational
  12736. motion present during the escape response, which usually
  12737. corresponds to a specific kinematic stage of the animal's
  12738. locomotion. Beyond this stage, prey may continue escaping along
  12739. a zigzag path, especially if predators follow up their attack
  12740. with a chase.",
  12741. publisher = "John Wiley \& Sons, Ltd",
  12742. volume = 200,
  12743. pages = "1",
  12744. month = apr,
  12745. year = 2014,
  12746. address = "Chichester, UK"
  12747. }
  12748. @ARTICLE{Domenici2011-op,
  12749. title = "Animal escapology {II}: escape trajectory case studies",
  12750. author = "Domenici, Paolo and Blagburn, Jonathan M and Bacon, Jonathan P",
  12751. abstract = "Escape trajectories (ETs; measured as the angle relative to the
  12752. direction of the threat) have been studied in many taxa using a
  12753. variety of methodologies and definitions. Here, we provide a
  12754. review of methodological issues followed by a survey of ET
  12755. studies across animal taxa, including insects, crustaceans,
  12756. molluscs, lizards, fish, amphibians, birds and mammals.
  12757. Variability in ETs is examined in terms of ecological
  12758. significance and morpho-physiological constraints. The survey
  12759. shows that certain escape strategies (single ETs and highly
  12760. variable ETs within a limited angular sector) are found in most
  12761. taxa reviewed here, suggesting that at least some of these ET
  12762. distributions are the result of convergent evolution. High
  12763. variability in ETs is found to be associated with multiple
  12764. preferred trajectories in species from all taxa, and is suggested
  12765. to provide unpredictability in the escape response. Random ETs
  12766. are relatively rare and may be related to constraints in the
  12767. manoeuvrability of the prey. Similarly, reports of the effect of
  12768. refuges in the immediate environment are relatively uncommon, and
  12769. mainly confined to lizards and mammals. This may be related to
  12770. the fact that work on ETs carried out in laboratory settings has
  12771. rarely provided shelters. Although there are a relatively large
  12772. number of examples in the literature that suggest trends in the
  12773. distribution of ETs, our understanding of animal escape
  12774. strategies would benefit from a standardization of the analytical
  12775. approach in the study of ETs, using circular statistics and
  12776. related tests, in addition to the generation of large data sets.",
  12777. journal = "J. Exp. Biol.",
  12778. volume = 214,
  12779. number = "Pt 15",
  12780. pages = "2474--2494",
  12781. month = aug,
  12782. year = 2011,
  12783. language = "en"
  12784. }
  12785. @ARTICLE{Humphries1970-rt,
  12786. title = "Protean defence by prey animals",
  12787. author = "Humphries, D A and Driver, P M",
  12788. abstract = "Attention is drawn to the widespread occurrence ofprotean
  12789. phenomena, in which the appearance and behaviour of prey animals
  12790. are rendered variable and irregular, as a weapon in the
  12791. biological arms race between predators and their prey. Protean
  12792. behaviour is defined as that behaviour which is sufficiently
  12793. unsystematic to prevent a reactor predicting in detail the
  12794. position or actions of the actor.Single prey animals frequently
  12795. flee from a predator in an irregular manner, zigzagging,
  12796. spinning, looping, or bouncing. Thissingle erratic display occurs
  12797. widely in the Animal Kingdom, and may also be utilised in
  12798. everyday movements of potential prey as insurance against
  12799. possible attack. Examples are given.In a group of prey animals
  12800. the protean aspect of escape is enhanced by the effect of
  12801. numbers. In scatter reactions the effect is of multiple choice
  12802. and of the simultaneous operation of several single erratics. In
  12803. mobbing displays there are also successive changes in the actors'
  12804. behavioural role. In protean deterrence the shuffling of
  12805. individuals within a tightly packed group prevents a predator
  12806. from singling one out for attack.In many species the confusing
  12807. effect of changes in movement and behavioural role is enhanced by
  12808. rapid changes in appearance, particularly colour.It is suggested
  12809. that those prey individuals which employ escape patterns
  12810. unfamiliar to the predator will tend to be at a selective
  12811. advantage. During phylogeny this is likely to lead to
  12812. intra-specific and inter-specific increase in the number and
  12813. diversity of escape behaviours. Apostatic polymorphism is seen as
  12814. a special case of protean variation within populations.There is
  12815. evidence that protean displays operate by arousing neurological
  12816. conflict, thereby delaying the predator's reactions and reducing
  12817. the effectiveness of predatory mechanisms. Also they insure
  12818. against learned countermeasures by incorporating irregularities
  12819. as a basic principle. It is stressed that the irregular
  12820. variability of protean displays is not accidental but has been
  12821. selected for in phylogeny. A number of poorly understood
  12822. behavioural aspects of the ecology of predator-prey relationships
  12823. are thus united in a single theory.",
  12824. journal = "Oecologia",
  12825. volume = 5,
  12826. number = 4,
  12827. pages = "285--302",
  12828. month = dec,
  12829. year = 1970,
  12830. language = "en"
  12831. }
  12832. @ARTICLE{Card2008-kq,
  12833. title = "Visually mediated motor planning in the escape response of
  12834. Drosophila",
  12835. author = "Card, Gwyneth and Dickinson, Michael H",
  12836. abstract = "A key feature of reactive behaviors is the ability to spatially
  12837. localize a salient stimulus and act accordingly. Such
  12838. sensory-motor transformations must be particularly fast and well
  12839. tuned in escape behaviors, in which both the speed and accuracy
  12840. of the evasive response determine whether an animal successfully
  12841. avoids predation [1]. We studied the escape behavior of the fruit
  12842. fly, Drosophila, and found that flies can use visual information
  12843. to plan a jump directly away from a looming threat. This is
  12844. surprising, given the architecture of the pathway thought to
  12845. mediate escape [2, 3]. Using high-speed videography, we found
  12846. that approximately 200 ms before takeoff, flies begin a series of
  12847. postural adjustments that determine the direction of their
  12848. escape. These movements position their center of mass so that leg
  12849. extension will push them away from the expanding visual stimulus.
  12850. These preflight movements are not the result of a simple
  12851. feed-forward motor program because their magnitude and direction
  12852. depend on the flies' initial postural state. Furthermore, flies
  12853. plan a takeoff direction even in instances when they choose not
  12854. to jump. This sophisticated motor program is evidence for a form
  12855. of rapid, visually mediated motor planning in a genetically
  12856. accessible model organism.",
  12857. journal = "Curr. Biol.",
  12858. volume = 18,
  12859. number = 17,
  12860. pages = "1300--1307",
  12861. month = sep,
  12862. year = 2008,
  12863. language = "en"
  12864. }
  12865. @ARTICLE{Santer2005-km,
  12866. title = "Motor activity and trajectory control during escape jumping in
  12867. the locust Locusta migratoria",
  12868. author = "Santer, Roger D and Yamawaki, Yoshifumi and Rind, F Claire and
  12869. Simmons, Peter J",
  12870. abstract = "We investigated the escape jumps that locusts produce in response
  12871. to approaching objects. Hindleg muscular activity during an
  12872. escape jump is similar to that during a defensive kick. Locusts
  12873. can direct their escape jumps up to 50 degrees either side of the
  12874. direction of their long axis at the time of hindleg flexion,
  12875. allowing them to consistently jump away from the side towards
  12876. which an object is approaching. Variation in jump trajectory is
  12877. achieved by rolling and yawing movements of the body that are
  12878. controlled by the fore- and mesothoracic legs. During hindleg
  12879. flexion, a locust flexes the foreleg ipsilateral to its eventual
  12880. jump trajectory and then extends the contralateral foreleg. These
  12881. foreleg movements continue throughout co-contraction of the
  12882. hindleg tibial muscles, pivoting the locust's long axis towards
  12883. its eventual jump trajectory. However, there are no bilateral
  12884. differences in the motor programs of the left and right hindlegs
  12885. that correlate with jump trajectory. Foreleg movements enable a
  12886. locust to control its jump trajectory independent of the hindleg
  12887. motor program, allowing a decision on jump trajectory to be made
  12888. after the hindlegs have been cocked in preparation for a jump.",
  12889. journal = "J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol.",
  12890. volume = 191,
  12891. number = 10,
  12892. pages = "965--975",
  12893. month = oct,
  12894. year = 2005,
  12895. language = "en"
  12896. }
  12897. @ARTICLE{Juechems2019-kv,
  12898. title = "Where Does Value Come From?",
  12899. author = "Juechems, Keno and Summerfield, Christopher",
  12900. abstract = "The computational framework of reinforcement learning (RL) has
  12901. allowed us to both understand biological brains and build
  12902. successful artificial agents. However, in this opinion, we
  12903. highlight open challenges for RL as a model of animal behaviour
  12904. in natural environments. We ask how the external reward function
  12905. is designed for biological systems, and how we can account for
  12906. the context sensitivity of valuation. We summarise both old and
  12907. new theories proposing that animals track current and desired
  12908. internal states and seek to minimise the distance to a goal
  12909. across multiple value dimensions. We suggest that this framework
  12910. readily accounts for canonical phenomena observed in the fields
  12911. of psychology, behavioural ecology, and economics, and recent
  12912. findings from brain-imaging studies of value-guided
  12913. decision-making.",
  12914. journal = "Trends Cogn. Sci.",
  12915. month = sep,
  12916. year = 2019,
  12917. keywords = "goal-directed decision-making; homeostasis; medial prefrontal
  12918. cortex; reinforcement learning; reward; value",
  12919. language = "en"
  12920. }
  12921. @ARTICLE{Fanselow2018-bx,
  12922. title = "The Role of Learning in Threat Imminence and Defensive Behaviors",
  12923. author = "Fanselow, Michael S",
  12924. abstract = "Life threatening situations as urgent as defending against a
  12925. predator precludes the use of slow trial and error strategies.
  12926. Natural selection has led to the evolution of a behavioral system
  12927. that has 3 critical elements. 1) When it is activated it limits
  12928. the behaviors available to the organism to a set of prewired
  12929. responses that have proven over phylogeny to be effective at
  12930. defense. 2) A rapid learning system, called Pavlovian fear
  12931. conditioning, that has the ability to immediately identify
  12932. threats and promote prewired defensive behaviors. 3) That
  12933. learning system has the ability to integrate several
  12934. informational dimensions to determine threat imminence and this
  12935. allows the organism to match the most effective defensive
  12936. behavior to the current situation. The adaptive significance of
  12937. conscious experiential states is also considered.",
  12938. journal = "Curr Opin Behav Sci",
  12939. volume = 24,
  12940. pages = "44--49",
  12941. month = dec,
  12942. year = 2018,
  12943. keywords = "Anxiety; Defensive Behavior; Fear; Fear Conditioning; Innate
  12944. Fear; Panic; Predatory Imminence; Threat Imminence; amygdala;
  12945. avoidance; consciousness; freezing; selective association",
  12946. language = "en"
  12947. }
  12948. @ARTICLE{Dunn2016-xl,
  12949. title = "Brain-wide mapping of neural activity controlling zebrafish
  12950. exploratory locomotion",
  12951. author = "Dunn, Timothy W and Mu, Yu and Narayan, Sujatha and Randlett,
  12952. Owen and Naumann, Eva A and Yang, Chao-Tsung and Schier,
  12953. Alexander F and Freeman, Jeremy and Engert, Florian and Ahrens,
  12954. Misha B",
  12955. abstract = "In the absence of salient sensory cues to guide behavior,
  12956. animals must still execute sequences of motor actions in order
  12957. to forage and explore. How such successive motor actions are
  12958. coordinated to form global locomotion trajectories is unknown.
  12959. We mapped the structure of larval zebrafish swim trajectories in
  12960. homogeneous environments and found that trajectories were
  12961. characterized by alternating sequences of repeated turns to the
  12962. left and to the right. Using whole-brain light-sheet imaging, we
  12963. identified activity relating to the behavior in specific neural
  12964. populations that we termed the anterior rhombencephalic turning
  12965. region (ARTR). ARTR perturbations biased swim direction and
  12966. reduced the dependence of turn direction on turn history,
  12967. indicating that the ARTR is part of a network generating the
  12968. temporal correlations in turn direction. We also find suggestive
  12969. evidence for ARTR mutual inhibition and ARTR projections to
  12970. premotor neurons. Finally, simulations suggest the observed turn
  12971. sequences may underlie efficient exploration of local
  12972. environments.",
  12973. journal = "Elife",
  12974. publisher = "cdn.elifesciences.org",
  12975. volume = 5,
  12976. pages = "e12741",
  12977. month = mar,
  12978. year = 2016,
  12979. keywords = "exploration strategies; higher-order motor control; larval
  12980. zebrafish; neural basis of behavior; neuroscience; spontaneous
  12981. brain activity; whole-brain functional imaging; zebrafish",
  12982. language = "en"
  12983. }
  12984. @UNPUBLISHED{Bolton2019-kt,
  12985. title = "Elements of a stochastic {3D} prediction engine in larval
  12986. zebrafish prey capture",
  12987. author = "Bolton, Andrew D and Haesemeyer, Martin and Jordi, Josua and
  12988. Schaechtle, Ulrich and Saad, Feras and Mansinghka, Vikash K and
  12989. Tenenbaum, Joshua B and Engert, Florian",
  12990. abstract = "Many predatory animals rely on accurate sensory perception,
  12991. predictive models, and precise pursuits to catch moving prey.
  12992. Larval zebrafish intercept paramecia during their hunting
  12993. behavior, but the precise trajectories of their prey have never
  12994. been recorded in relation to fish movements in three dimensions.
  12995. As a means of uncovering what a simple organism understands about
  12996. its physical world, we have constructed a 3D-imaging setup to
  12997. simultaneously record the behavior of larval zebrafish, as well
  12998. as their moving prey, during hunting. We show that zebrafish
  12999. robustly transform their 3D displacement and rotation according
  13000. to the position of their prey while modulating both of these
  13001. variables depending on prey velocity. This is true for both
  13002. azimuth and altitude, but particulars of the hunting algorithm in
  13003. the two planes are slightly different to accommodate an
  13004. asymmetric strike zone. We show that the combination of position
  13005. and velocity perception provides the fish with a preferred future
  13006. positional estimate, indicating an ability to project
  13007. trajectories forward in time. Using computational models, we show
  13008. that this projection ability is critical for prey capture
  13009. efficiency and success. Further, we demonstrate that fish use a
  13010. graded stochasticity algorithm where the variance around the mean
  13011. result of each swim scales with distance from the target.
  13012. Notably, this strategy provides the animal with a considerable
  13013. improvement over equivalent noise-free strategies. In sum, our
  13014. quantitative and probabilistic modeling shows that zebrafish are
  13015. equipped with a stochastic recursive algorithm that embodies an
  13016. implicit predictive model of the world. This algorithm, built by
  13017. a simple set of behavioral rules, allows the fish to optimize
  13018. their hunting strategy in a naturalistic three-dimensional
  13019. environment.",
  13020. journal = "bioRxiv",
  13021. pages = "755777",
  13022. month = sep,
  13023. year = 2019,
  13024. language = "en"
  13025. }
  13026. @ARTICLE{Neftci2019-pw,
  13027. title = "Reinforcement learning in artificial and biological systems",
  13028. author = "Neftci, Emre O and Averbeck, Bruno B",
  13029. abstract = "There is and has been a fruitful flow of concepts and ideas
  13030. between studies of learning in biological and artificial systems.
  13031. Much early work that led to the development of reinforcement
  13032. learning (RL) algorithms for artificial systems was inspired by
  13033. learning rules first developed in biology by Bush and Mosteller,
  13034. and Rescorla and Wagner. More recently, temporal-difference RL,
  13035. developed for learning in artificial agents, has provided a
  13036. foundational framework for interpreting the activity of dopamine
  13037. neurons. In this Review, we describe state-of-the-art work on RL
  13038. in biological and artificial agents. We focus on points of
  13039. contact between these disciplines and identify areas where future
  13040. research can benefit from information flow between these fields.
  13041. Most work in biological systems has focused on simple learning
  13042. problems, often embedded in dynamic environments where
  13043. flexibility and ongoing learning are important, similar to
  13044. real-world learning problems faced by biological systems. In
  13045. contrast, most work in artificial agents has focused on learning
  13046. a single complex problem in a static environment. Moving forward,
  13047. work in each field will benefit from a flow of ideas that
  13048. represent the strengths within each discipline.",
  13049. journal = "Nature Machine Intelligence",
  13050. volume = 1,
  13051. number = 3,
  13052. pages = "133--143",
  13053. month = mar,
  13054. year = 2019
  13055. }
  13056. @BOOK{Burkhardt2005-ej,
  13057. title = "Patterns of Behavior: Konrad Lorenz, Niko Tinbergen, and the
  13058. Founding of Ethology",
  13059. author = "Burkhardt, Richard W",
  13060. abstract = "It is hard to imagine, by their very name, the life sciences not
  13061. involving the study of living things, but until the twentieth
  13062. century much of what was known in the field was based primarily
  13063. on specimens that had long before taken their last breaths. Only
  13064. in the last century has ethology---the study of animal
  13065. behavior---emerged as a major field of the life sciences. In
  13066. Patterns of Behavior, Richard W. Burkhardt Jr. traces the
  13067. scientific theories, practices, subjects, and settings integral
  13068. to the construction of a discipline pivotal to our understanding
  13069. of the diversity of life. Central to this tale are Konrad Lorenz
  13070. and Niko Tinbergen, 1973 Nobel laureates whose research helped
  13071. legitimize the field of ethology and bring international
  13072. attention to the culture of behavioral research. Demonstrating
  13073. how matters of practice, politics, and place all shaped
  13074. ``ethology's ecologies,'' Burkhardt's book offers a sensitive
  13075. reading of the complex interplay of the field's celebrated
  13076. pioneers and a richly textured reconstruction of ethology's
  13077. transformation from a quiet backwater of natural history to the
  13078. forefront of the biological sciences. Winner of the 2006 Pfizer
  13079. Awad from the History of Science Society",
  13080. publisher = "University of Chicago Press",
  13081. month = mar,
  13082. year = 2005,
  13083. language = "en"
  13084. }
  13085. @ARTICLE{Swartz2006-xa,
  13086. title = "Inverse Decision Theory",
  13087. author = "Swartz, Richard J and Cox, Dennis D and Cantor, Scott B and
  13088. Davies, Kalatu and Follen, Michele",
  13089. abstract = "Identifying an optimal decision rule using Bayesian decision
  13090. theory requires priors, likelihoods, and losses. In many medical
  13091. settings, we can develop priors and likelihoods, but specifying
  13092. losses can be difficult, especially when considering both
  13093. patient outcomes and economic costs. If there is a widely
  13094. accepted treatment strategy, then we can consider the inverse
  13095. problem and find a region in the space of losses where the
  13096. procedure is optimal. We call this approach inverse decision
  13097. theory (IDT). We apply IDT to the standard of care for diagnosis
  13098. and treatment of precancerous lesions of the cervix, and
  13099. consider an alternative procedure that has been proposed. We use
  13100. a Bayesian approach to estimate the probabilities associated
  13101. with the diagnostic tests and make inferences about the region
  13102. in loss space where these medical procedures are optimal. In
  13103. particular, we find evidence supporting the current standard of
  13104. care.",
  13105. journal = "J. Am. Stat. Assoc.",
  13106. publisher = "Taylor \& Francis",
  13107. volume = 101,
  13108. number = 473,
  13109. pages = "1--8",
  13110. month = mar,
  13111. year = 2006
  13112. }
  13113. @ARTICLE{Richards1974-wh,
  13114. title = "The Innate and the Learned: The Evolution of Konrad Lorenz's
  13115. Theory of Instinct",
  13116. author = "Richards, Robert J",
  13117. journal = "Philos. Soc. Sci.",
  13118. publisher = "SAGE Publications Inc",
  13119. volume = 4,
  13120. number = "2-3",
  13121. pages = "111--133",
  13122. month = jun,
  13123. year = 1974
  13124. }
  13125. % The entry below contains non-ASCII chars that could not be converted
  13126. % to a LaTeX equivalent.
  13127. @BOOK{Belew2018-rz,
  13128. title = "Adaptive individuals in evolving populations: models and
  13129. algorithms",
  13130. author = "Belew, Richard K",
  13131. abstract = "A Model of Individual Adaptive Behavior in a Fluctuating
  13132. Environment Individual behavioral strategies that use
  13133. conditional probabilities for future environments and
  13134. information about past environments are studied. The
  13135. environments are random and Markovian. The individual uses the
  13136. available information to prepare for the next environmental
  13137. state in order to increase its fitness. The fitness depends on
  13138. the discrepancy between the realized environment and that for
  13139. which the individual is prepared. Additive and multiplicative …",
  13140. publisher = "Routledge",
  13141. year = 2018
  13142. }
  13143. @ARTICLE{Turney1996-uv,
  13144. title = "Evolution, Learning, and Instinct: 100 Years of the Baldwin
  13145. Effect",
  13146. author = "Turney, Peter and Whitley, Darrell and Anderson, Russell W",
  13147. journal = "Evol. Comput.",
  13148. publisher = "MIT Press",
  13149. volume = 4,
  13150. number = 3,
  13151. pages = "iv--viii",
  13152. month = sep,
  13153. year = 1996
  13154. }
  13155. @ARTICLE{Zador2019-wh,
  13156. title = "A critique of pure learning and what artificial neural networks
  13157. can learn from animal brains",
  13158. author = "Zador, Anthony M",
  13159. abstract = "Artificial neural networks (ANNs) have undergone a revolution,
  13160. catalyzed by better supervised learning algorithms. However, in
  13161. stark contrast to young animals (including humans), training
  13162. such networks requires enormous numbers of labeled examples,
  13163. leading to the belief that animals must rely instead mainly on
  13164. unsupervised learning. Here we argue that most animal behavior
  13165. is not the result of clever learning algorithms-supervised or
  13166. unsupervised-but is encoded in the genome. Specifically, animals
  13167. are born with highly structured brain connectivity, which
  13168. enables them to learn very rapidly. Because the wiring diagram
  13169. is far too complex to be specified explicitly in the genome, it
  13170. must be compressed through a ``genomic bottleneck''. The genomic
  13171. bottleneck suggests a path toward ANNs capable of rapid
  13172. learning.",
  13173. journal = "Nat. Commun.",
  13174. publisher = "nature.com",
  13175. volume = 10,
  13176. number = 1,
  13177. pages = "3770",
  13178. month = aug,
  13179. year = 2019,
  13180. language = "en"
  13181. }
  13182. @BOOK{Hebb2005-wq,
  13183. title = "The Organization of Behavior: A Neuropsychological Theory",
  13184. author = "Hebb, D O",
  13185. abstract = "Since its publication in 1949, D.O. Hebb's, The Organization of
  13186. Behavior has been one of the most influential books in the
  13187. fields of psychology and neuroscience. However, the original
  13188. edition has been unavailable since 1966, ensuring that Hebb's
  13189. comment that a classic normally means ``cited but not read'' is
  13190. true in his case. This new edition rectifies a long-standing
  13191. problem for behavioral neuroscientists--the inability to obtain
  13192. one of the most cited publications in the field. The
  13193. Organization of Behavior played a significant part in
  13194. stimulating the investigation of the neural foundations of
  13195. behavior and continues to be inspiring because it provides a
  13196. general framework for relating behavior to synaptic organization
  13197. through the dynamics of neural networks. D.O. Hebb was also the
  13198. first to examine the mechanisms by which environment and
  13199. experience can influence brain structure and function, and his
  13200. ideas formed the basis for work on enriched environments as
  13201. stimulants for behavioral development. References to Hebb, the
  13202. Hebbian cell assembly, the Hebb synapse, and the Hebb rule
  13203. increase each year. These forceful ideas of 1949 are now applied
  13204. in engineering, robotics, and computer science, as well as
  13205. neurophysiology, neuroscience, and psychology--a tribute to
  13206. Hebb's foresight in developing a foundational neuropsychological
  13207. theory of the organization of behavior.",
  13208. publisher = "Psychology Press",
  13209. month = apr,
  13210. year = 2005,
  13211. language = "en"
  13212. }
  13213. @BOOK{Kirk2012-ji,
  13214. title = "Optimal Control Theory: An Introduction",
  13215. author = "Kirk, Donald E",
  13216. abstract = "Optimal control theory is the science of maximizing the returns
  13217. from and minimizing the costs of the operation of physical,
  13218. social, and economic processes. Geared toward upper-level
  13219. undergraduates, this text introduces three aspects of optimal
  13220. control theory: dynamic programming, Pontryagin's minimum
  13221. principle, and numerical techniques for trajectory
  13222. optimization.Chapters 1 and 2 focus on describing systems and
  13223. evaluating their performances. Chapter 3 deals with dynamic
  13224. programming. The calculus of variations and Pontryagin's minimum
  13225. principle are the subjects of chapters 4 and 5, and chapter 6
  13226. examines iterative numerical techniques for finding optimal
  13227. controls and trajectories. Numerous problems, intended to
  13228. introduce additional topics as well as to illustrate basic
  13229. concepts, appear throughout the text.",
  13230. publisher = "Courier Corporation",
  13231. month = apr,
  13232. year = 2012,
  13233. language = "en"
  13234. }
  13235. @ARTICLE{Vicente2019-kq,
  13236. title = "The many faces of deep learning",
  13237. author = "Vicente, Raul",
  13238. abstract = "Deep learning has sparked a network of mutual interactions
  13239. between different disciplines and AI. Naturally, each
  13240. discipline focuses and interprets the workings of deep
  13241. learning in different ways. This diversity of perspectives
  13242. on deep learning, from neuroscience to statistical physics,
  13243. is a rich source of inspiration that fuels novel
  13244. developments in the theory and applications of machine
  13245. learning. In this perspective, we collect and synthesize
  13246. different intuitions scattered across several communities as
  13247. for how deep learning works. In particular, we will briefly
  13248. discuss the different perspectives that disciplines across
  13249. mathematics, physics, computation, and neuroscience take on
  13250. how deep learning does its tricks. Our discussion on each
  13251. perspective is necessarily shallow due to the multiple views
  13252. that had to be covered. The deepness in this case should
  13253. come from putting all these faces of deep learning together
  13254. in the reader's mind, so that one can look at the same
  13255. problem from different angles.",
  13256. month = aug,
  13257. year = 2019,
  13258. archivePrefix = "arXiv",
  13259. primaryClass = "cs.LG",
  13260. eprint = "1908.10206"
  13261. }
  13262. @ARTICLE{Stephens2008-um,
  13263. title = "Decision ecology: foraging and the ecology of animal decision
  13264. making",
  13265. author = "Stephens, David W",
  13266. abstract = "In this article, I review the approach taken by behavioral
  13267. ecologists to the study of animal foraging behavior and explore
  13268. connections with general analyses of decision making. I use the
  13269. example of patch exploitation decisions in this article in order
  13270. to develop several key points about the properties of naturally
  13271. occurring foraging decisions. First, I argue that experimental
  13272. preparations based on binary, mutually exclusive choice are not
  13273. good models of foraging decisions. Instead, foraging choices have
  13274. a sequential foreground-background structure, in which one option
  13275. is in the background of all other options. Second, behavioral
  13276. ecologists view foraging as a hierarchy of decisions that range
  13277. from habitat selection to food choice. Finally, data suggest that
  13278. foraging animals are sensitive to several important trade-offs.
  13279. These trade-offs include the effects of competitors and group
  13280. mates, as well as the problem of predator avoidance.",
  13281. journal = "Cogn. Affect. Behav. Neurosci.",
  13282. volume = 8,
  13283. number = 4,
  13284. pages = "475--484",
  13285. month = dec,
  13286. year = 2008,
  13287. language = "en"
  13288. }
  13289. % The entry below contains non-ASCII chars that could not be converted
  13290. % to a LaTeX equivalent.
  13291. @ARTICLE{Pezzulo2014-up,
  13292. title = "The principles of goal-directed decision-making: from neural
  13293. mechanisms to computation and robotics",
  13294. author = "Pezzulo, Giovanni and Verschure, Paul F M J and Balkenius,
  13295. Christian and Pennartz, Cyriel M A",
  13296. abstract = "… Verschure et al … Footnotes. One contribution of 18 to a Theme
  13297. Issue 'The principles of goal-directed decision - making : from
  13298. neural mechanisms to computation and robotics'. \copyright{}
  13299. 2014 The Author(s) Published by the Royal Society. All rights
  13300. reserved. References …",
  13301. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  13302. publisher = "royalsocietypublishing.org",
  13303. volume = 369,
  13304. number = 1655,
  13305. month = nov,
  13306. year = 2014,
  13307. keywords = "computational model; decision-making; goal-directed; neural
  13308. mechanism; prediction; robotics",
  13309. language = "en"
  13310. }
  13311. @ARTICLE{Bogacz2006-lr,
  13312. title = "The physics of optimal decision making: a formal analysis of
  13313. models of performance in two-alternative forced-choice tasks",
  13314. author = "Bogacz, Rafal and Brown, Eric and Moehlis, Jeff and Holmes,
  13315. Philip and Cohen, Jonathan D",
  13316. abstract = "In this article, the authors consider optimal decision making in
  13317. two-alternative forced-choice (TAFC) tasks. They begin by
  13318. analyzing 6 models of TAFC decision making and show that all but
  13319. one can be reduced to the drift diffusion model, implementing
  13320. the statistically optimal algorithm (most accurate for a given
  13321. speed or fastest for a given accuracy). They prove further that
  13322. there is always an optimal trade-off between speed and accuracy
  13323. that maximizes various reward functions, including reward rate
  13324. (percentage of correct responses per unit time), as well as
  13325. several other objective functions, including ones weighted for
  13326. accuracy. They use these findings to address empirical data and
  13327. make novel predictions about performance under optimality.",
  13328. journal = "Psychol. Rev.",
  13329. publisher = "psycnet.apa.org",
  13330. volume = 113,
  13331. number = 4,
  13332. pages = "700--765",
  13333. month = oct,
  13334. year = 2006,
  13335. language = "en"
  13336. }
  13337. @ARTICLE{Cande2018-le,
  13338. title = "Optogenetic dissection of descending behavioral control in
  13339. Drosophila",
  13340. author = "Cande, Jessica and Namiki, Shigehiro and Qiu, Jirui and Korff,
  13341. Wyatt and Card, Gwyneth M and Shaevitz, Joshua W and Stern, David
  13342. L and Berman, Gordon J",
  13343. abstract = "In most animals, the brain makes behavioral decisions that are
  13344. transmitted by descending neurons to the nerve cord circuitry
  13345. that produces behaviors. In insects, only a few descending
  13346. neurons have been associated with specific behaviors. To explore
  13347. how descending neurons control an insect's movements, we
  13348. developed a novel method to systematically assay the behavioral
  13349. effects of activating individual neurons on freely behaving
  13350. terrestrial D. melanogaster. We calculated a two-dimensional
  13351. representation of the entire behavior space explored by these
  13352. flies, and we associated descending neurons with specific
  13353. behaviors by identifying regions of this space that were visited
  13354. with increased frequency during optogenetic activation. Applying
  13355. this approach across a large collection of descending neurons, we
  13356. found that (1) activation of most of the descending neurons drove
  13357. stereotyped behaviors, (2) in many cases multiple descending
  13358. neurons activated similar behaviors, and (3) optogenetically
  13359. activated behaviors were often dependent on the behavioral state
  13360. prior to activation.",
  13361. journal = "Elife",
  13362. volume = 7,
  13363. month = jun,
  13364. year = 2018,
  13365. keywords = "D. melanogaster; behavior; descending interneurons; neuroscience;
  13366. optogenetics",
  13367. language = "en"
  13368. }
  13369. @INCOLLECTION{Gomez-Marin2017-xa,
  13370. title = "Causal Circuit Explanations of Behavior: Are Necessity and
  13371. Sufficiency Necessary and Sufficient?",
  13372. booktitle = "Decoding Neural Circuit Structure and Function: Cellular
  13373. Dissection Using Genetic Model Organisms",
  13374. author = "Gomez-Marin, Alex",
  13375. editor = "{\c C}elik, Arzu and Wernet, Mathias F",
  13376. abstract = "In the current advent of technological innovation allowing for
  13377. precise neural manipulations and copious data collection, it is
  13378. hardly questioned that the explanation of behavioral processes
  13379. is to be chiefly found in neural circuits. Such belief, rooted
  13380. in the exhausted dualism of cause and effect, is enacted by a
  13381. methodology that promotes ``necessity and sufficiency'' claims
  13382. as the goal-standard in neuroscience, thus instructing young
  13383. students on what shall reckon as explanation. Here I wish to
  13384. deconstruct and explicate the difference between what is done,
  13385. what is said, and what is meant by such causal circuit
  13386. explanations of behavior. Well-known to most philosophers, yet
  13387. ignored or at least hardly ever made explicit by
  13388. neuroscientists, the original grand claim of ``understanding the
  13389. brain'' is imperceptibly substituted by the methodologically
  13390. sophisticated task of empirically establishing counterfactual
  13391. dependencies. But for the twenty-first century neuroscientist,
  13392. after so much pride, this is really an excess of humility. I
  13393. argue that to upgrade intervention to explanation is prone to
  13394. logical fallacies, interpretational leaps and carries a weak
  13395. explanatory force, thus settling and maintaining low standards
  13396. for intelligibility in neuroscience. To claim that behavior is
  13397. explained by a ``necessary and sufficient'' neural circuit is,
  13398. at best, misleading. In that, my critique (rather than
  13399. criticism) is indeed mainly negative. Positively, I briefly
  13400. suggest some available alternatives for conceptual progress,
  13401. such as adopting circular causality (rather than lineal
  13402. causality in the flavor of top-down reductionism), searching for
  13403. principles of behavior (rather than taking an arbitrary
  13404. definition of behavior and rushing to dissect its ``underlying''
  13405. neural mechanisms), and embracing process philosophy (rather
  13406. than substance-mechanistic ontologies). Overall, if the goal of
  13407. neuroscience is to understand the relation between brain and
  13408. behavior then, in addition to excruciating neural studies (one
  13409. pillar), we will need a strong theory of behavior (the other
  13410. pillar) and a solid foundation to establish their relation (the
  13411. bridge).",
  13412. publisher = "Springer International Publishing",
  13413. pages = "283--306",
  13414. year = 2017,
  13415. address = "Cham"
  13416. }
  13417. @UNPUBLISHED{Fakhar2019-ew,
  13418. title = "Neuronal Causes and Behavioural Effects: a Review on Logical,
  13419. Methodological, and Technical Issues With Respect to Causal
  13420. Explanations of Behaviour in Neuroscience",
  13421. author = "Fakhar, Kayson and Gonschorek, Dominic and Schmors, Lisa and
  13422. Bielczyk, Natalia Z",
  13423. abstract = "Elucidating causal, neurobiological underpinnings of behaviour is
  13424. an ultimate goal of every neuroscientific study. However, due to
  13425. the complexity of the brain as well as the complexity of the
  13426. human environment, finding a~causal architecture that underlies
  13427. behaviour remains a~formidable challenge. In this manuscript, we
  13428. review the logical and conceptual issues with respect to causal
  13429. research in neuroscience. First, we review the state of the art
  13430. interventional and computational approaches to infer causal
  13431. brain-behaviour relationships. We provide an~overview of
  13432. potential issues, flaws, and confounds in these studies. We
  13433. conclude that studies on the causal structure underlying
  13434. behaviour should be performed by accumulating evidence coming
  13435. from several lines of experimental and modelling studies. Lastly,
  13436. we also propose computational models including artificial
  13437. neuronal networks and simulated animats as a~potential
  13438. breakthrough to causal brain-behaviour investigations.",
  13439. month = aug,
  13440. year = 2019,
  13441. keywords = "brain-behaviour relations; brain interventions; causal inference;
  13442. causality; cause and effect; computational models; Necessity and
  13443. Sufficiency"
  13444. }
  13445. @ARTICLE{Carey2019-gk,
  13446. title = "Reward revaluation biases hippocampal replay content away from
  13447. the preferred outcome",
  13448. author = "Carey, Alyssa A and Tanaka, Youki and van der Meer, Matthijs A A",
  13449. abstract = "The rodent hippocampus spontaneously generates bursts of neural
  13450. activity (replay) that can depict spatial trajectories to reward
  13451. locations, suggesting a role in model-based behavioral control. A
  13452. largely separate literature emphasizes reward revaluation as the
  13453. litmus test for such control, yet the content of hippocampal
  13454. replay under revaluation conditions is unknown. We examined the
  13455. content of awake replay events following motivational shifts
  13456. between hunger and thirst. On a T-maze offering free choice
  13457. between food and water outcomes, rats shifted their behavior
  13458. toward the restricted outcome, but replay content was shifted
  13459. away from the restricted outcome. This effect preceded experience
  13460. on the task each day and did not reverse with experience. These
  13461. results demonstrate that replay content is not limited to
  13462. reflecting recent experience or trajectories toward the preferred
  13463. goal and suggest a role for motivational states in determining
  13464. replay content.",
  13465. journal = "Nat. Neurosci.",
  13466. month = aug,
  13467. year = 2019
  13468. }
  13469. @ARTICLE{Pouget2003-qe,
  13470. title = "Inference and computation with population codes",
  13471. author = "Pouget, Alexandre and Dayan, Peter and Zemel, Richard S",
  13472. abstract = "In the vertebrate nervous system, sensory stimuli are typically
  13473. encoded through the concerted activity of large populations of
  13474. neurons. Classically, these patterns of activity have been
  13475. treated as encoding the value of the stimulus (e.g., the
  13476. orientation of a contour), and computation has been formalized in
  13477. terms of function approximation. More recently, there have been
  13478. several suggestions that neural computation is akin to a Bayesian
  13479. inference process, with population activity patterns representing
  13480. uncertainty about stimuli in the form of probability
  13481. distributions (e.g., the probability density function over the
  13482. orientation of a contour). This paper reviews both approaches,
  13483. with a particular emphasis on the latter, which we see as a very
  13484. promising framework for future modeling and experimental work.",
  13485. journal = "Annu. Rev. Neurosci.",
  13486. volume = 26,
  13487. pages = "381--410",
  13488. month = apr,
  13489. year = 2003,
  13490. language = "en"
  13491. }
  13492. @ARTICLE{Garrett_undated-lg,
  13493. title = "Biased belief updating and suboptimal choice in foraging decisions",
  13494. author = "Garrett, Neil and Daw, Nathaniel D"
  13495. }
  13496. @ARTICLE{Edwards1954-hk,
  13497. title = "The theory of decision making",
  13498. author = "Edwards, W",
  13499. journal = "Psychol. Bull.",
  13500. volume = 51,
  13501. number = 4,
  13502. pages = "380--417",
  13503. month = jul,
  13504. year = 1954,
  13505. keywords = "THINKING",
  13506. language = "en"
  13507. }
  13508. % The entry below contains non-ASCII chars that could not be converted
  13509. % to a LaTeX equivalent.
  13510. @ARTICLE{McNamara2014-xz,
  13511. title = "Natural selection can favour `irrational'behaviour",
  13512. author = "McNamara, John M and Trimmer, Pete C and Houston, A I",
  13513. abstract = "Understanding decisions is the fundamental aim of the
  13514. behavioural sciences. The theory of rational choice is based on
  13515. axiomatic principles such as transitivity and independence of
  13516. irrelevant alternatives (IIA). Empirical studies have
  13517. demonstrated that the behaviour of humans and other animals
  13518. often seems irrational; there can be a lack of transitivity in
  13519. choice and seemingly irrelevant alternatives can alter
  13520. decisions. These violations of transitivity and IIA undermine
  13521. rational choice theory. However, we show that an individual that
  13522. is …",
  13523. journal = "Biol. Lett.",
  13524. publisher = "The Royal Society",
  13525. volume = 10,
  13526. number = 1,
  13527. pages = "20130935",
  13528. year = 2014
  13529. }
  13530. @UNPUBLISHED{Gupta2019-hk,
  13531. title = "A context-free grammar for Caenorhabditis elegans behavior",
  13532. author = "Gupta, Saurabh and Gomez-Marin, Alex",
  13533. abstract = "Hierarchy is a candidate organizing principle of ethology, where
  13534. actions grouped into higher order chunks combine in specific ways
  13535. to generate adaptive behavior. However, demonstrations of
  13536. hierarchical organization in behavior have been scarce. Moreover,
  13537. it remains unclear how such underlying organization allows for
  13538. behavioral flexibility. Here we uncover the hierarchical and
  13539. flexible nature of Caenorhabditis elegans behavior. By describing
  13540. worm locomotion as a sequence of discrete postural templates, we
  13541. identified chunks containing mutually substitutable postures
  13542. along the dynamics. We then elucidated the rules governing their
  13543. interactions. We found that stereotypical roaming can be
  13544. described by a specific sequence of postural chunks, which
  13545. exhibit flexibility at the lowest postural level. The same chunks
  13546. get combined differently to produce dwelling, capturing
  13547. non-stereotypical actions across timescales. We show that worm
  13548. foraging is organized hierarchically ---a feature not explainable
  13549. via Markovian dynamics---, and derive a context-free grammar
  13550. governing its behavior ---which is different than a regular
  13551. grammar, or a hidden Markov chain. In sum, in making the analogy
  13552. with human language concrete (but not literal) our work
  13553. demonstrates, in line with the foundational insights of classical
  13554. ethologists, that spontaneous behavior is orderly flexible. Once
  13555. more, investigating the humble nematode suggests that everything
  13556. human has its roots in lower animal behavior. ![Figure][1]
  13557. Graphical abstract [1]: pending:yes",
  13558. journal = "bioRxiv",
  13559. pages = "708891",
  13560. month = jul,
  13561. year = 2019,
  13562. language = "en"
  13563. }
  13564. @ARTICLE{Gershman_undated-ml,
  13565. title = "The generative adversarial brain",
  13566. author = "Gershman, Samuel J"
  13567. }
  13568. @ARTICLE{Iwanir2019-dw,
  13569. title = "Irrational behavior in C. elegans arises from asymmetric
  13570. modulatory effects within single sensory neurons",
  13571. author = "Iwanir, Shachar and Ruach, Rotem and Itskovits, Eyal and Pritz,
  13572. Christian O and Bokman, Eduard and Zaslaver, Alon",
  13573. abstract = "C. elegans worms exhibit a natural chemotaxis towards food cues.
  13574. This provides a potential platform to study the interactions
  13575. between stimulus valence and innate behavioral preferences. Here
  13576. we perform a comprehensive set of choice assays to measure worms'
  13577. relative preference towards various attractants. Surprisingly, we
  13578. find that when facing a combination of choices, worms'
  13579. preferences do not always follow value-based hierarchy. In fact,
  13580. the innate chemotaxis behavior in worms robustly violates key
  13581. rationality paradigms of transitivity, independence of irrelevant
  13582. alternatives and regularity. These violations arise due to
  13583. asymmetric modulatory effects between the presented options.
  13584. Functional analysis of the entire chemosensory system at a
  13585. single-neuron resolution, coupled with analyses of mutants,
  13586. defective in individual neurons, reveals that these asymmetric
  13587. effects originate in specific sensory neurons.",
  13588. journal = "Nat. Commun.",
  13589. volume = 10,
  13590. number = 1,
  13591. pages = "3202",
  13592. month = jul,
  13593. year = 2019,
  13594. language = "en"
  13595. }
  13596. @UNPUBLISHED{Trueblood2019-vi,
  13597. title = "Urgency, Leakage, and the Relative Nature of Information
  13598. Processing in Decision-making",
  13599. author = "Trueblood, Jennifer S and Heathcote, Andrew and Evans, Nathan J
  13600. and Holmes, William R",
  13601. abstract = "Over the last decade, there has been a robust debate in decision
  13602. neuroscience and psychology about what mechanism governs the time
  13603. course of decision making. Historically, the most prominent
  13604. hypothesis is that neural architectures accumulate information
  13605. over time until some threshold is met, the so-called Evidence
  13606. Accumulation hypothesis. However, most applications of this
  13607. theory rely on simplifying assumptions, belying a number of
  13608. potential complexities. Is changing stimulus information
  13609. perceived and processed in an independent manner or is there a
  13610. relative component? Does urgency play a role? What about evidence
  13611. leakage? Although the latter questions have been the subject of
  13612. recent investigations, most studies to date have been piecemeal
  13613. in nature, addressing one aspect of the decision process or
  13614. another. Here we develop a modeling framework, an extension of
  13615. the Urgency Gating Model, in conjunction with a changing
  13616. information experimental paradigm to simultaneously probe these
  13617. aspects of the decision process. Using state-of-the-art Bayesian
  13618. methods to perform parameter-based inference, we find that 1)
  13619. information processing is relative with early information
  13620. influencing the perception of late information, 2) time varying
  13621. urgency and evidence accumulation are of roughly equal importance
  13622. in the decision process, and 3) leakage is present with a time
  13623. scale of ~200-250ms. To our knowledge, this is the first
  13624. comprehensive study to utilize a changing information paradigm to
  13625. jointly and quantitatively estimate the temporal dynamics of
  13626. human decision-making.",
  13627. journal = "bioRxiv",
  13628. pages = "706291",
  13629. month = jul,
  13630. year = 2019,
  13631. language = "en"
  13632. }
  13633. @ARTICLE{Sohn2019-kc,
  13634. title = "Bayesian Computation through Cortical Latent Dynamics",
  13635. author = "Sohn, Hansem and Narain, Devika and Meirhaeghe, Nicolas and
  13636. Jazayeri, Mehrdad",
  13637. abstract = "Statistical regularities in the environment create prior beliefs
  13638. that we rely on to optimize our behavior when sensory information
  13639. is uncertain. Bayesian theory formalizes how prior beliefs can be
  13640. leveraged and has had a major impact on models of perception,
  13641. sensorimotor function, and cognition. However, it is not known
  13642. how recurrent interactions among neurons mediate Bayesian
  13643. integration. By using a time-interval reproduction task in
  13644. monkeys, we found that prior statistics warp neural
  13645. representations in the frontal cortex, allowing the mapping of
  13646. sensory inputs to motor outputs to incorporate prior statistics
  13647. in accordance with Bayesian inference. Analysis of recurrent
  13648. neural network models performing the task revealed that this
  13649. warping was enabled by a low-dimensional curved manifold and
  13650. allowed us to further probe the potential causal underpinnings of
  13651. this computational strategy. These results uncover a simple and
  13652. general principle whereby prior beliefs exert their influence on
  13653. behavior by sculpting cortical latent dynamics.",
  13654. journal = "Neuron",
  13655. month = jul,
  13656. year = 2019,
  13657. keywords = "Bayesian inference; Bayesian integration; frontal cortex; neural
  13658. manifold; neural trajectories; recurrent neural networks",
  13659. language = "en"
  13660. }
  13661. @ARTICLE{Draft2018-dk,
  13662. title = "Carpenter ants use diverse antennae sampling strategies to track
  13663. odor trails",
  13664. author = "Draft, Ryan W and McGill, Matthew R and Kapoor, Vikrant and
  13665. Murthy, Venkatesh N",
  13666. abstract = "Directed and meaningful animal behavior depends on the ability to
  13667. sense key features in the environment. Among the different
  13668. environmental signals, olfactory cues are critically important
  13669. for foraging, navigation and social communication in many
  13670. species, including ants. Ants use their two antennae to explore
  13671. the olfactory world, but how they do so remains largely unknown.
  13672. In this study, we used high-resolution videography to
  13673. characterize the antennae dynamics of carpenter ants (Camponotus
  13674. pennsylvanicus). Antennae are highly active during both odor
  13675. tracking and exploratory behavior. When tracking, ants used
  13676. several distinct behavioral strategies with stereotyped antennae
  13677. sampling patterns (which we call 'sinusoidal', 'probing' and
  13678. 'trail following'). In all behaviors, left and right antennae
  13679. movements were anti-correlated, and tracking ants exhibited
  13680. biases in the use of left versus right antenna to sample the odor
  13681. trail. These results suggest non-redundant roles for the two
  13682. antennae. In one of the behavioral modules (trail following),
  13683. ants used both antennae to detect trail edges and direct
  13684. subsequent turns, suggesting a specialized form of tropotaxis.
  13685. Lastly, removal of an antenna resulted not only in less accurate
  13686. tracking but also in changes in the sampling pattern of the
  13687. remaining antenna. Our quantitative characterization of odor
  13688. trail tracking lays a foundation to build better models of
  13689. olfactory sensory processing and sensorimotor behavior in
  13690. terrestrial insects.",
  13691. journal = "J. Exp. Biol.",
  13692. volume = 221,
  13693. number = "Pt 22",
  13694. month = nov,
  13695. year = 2018,
  13696. keywords = "Behavior; Camponotus; Navigation; Olfaction; Pheromone; Trail
  13697. tracking",
  13698. language = "en"
  13699. }
  13700. @MISC{Hyvonen2019-se,
  13701. title = "Bayesian Inference 2019",
  13702. author = "Hyv{\"o}nen, Ville and Tolonen, Topias",
  13703. abstract = "Lecture notes for Bayesian Inference course lectured at
  13704. University of Helsinki Spring 2019",
  13705. month = mar,
  13706. year = 2019,
  13707. howpublished = "\url{https://vioshyvo.github.io/Bayesian_inference/index.html}",
  13708. note = "Accessed: 2019-7-15",
  13709. keywords = "books;Books"
  13710. }
  13711. @UNPUBLISHED{Mohl2019-dl,
  13712. title = "Sensitivity and specificity of a Bayesian single trial analysis
  13713. for time varying neural signals",
  13714. author = "Mohl, Jeff T and Caruso, Valeria C and Tokdar, Surya T and Groh,
  13715. J M",
  13716. abstract = "We recently reported the existence of fluctuations in neural
  13717. signals that may permit neurons to code multiple simultaneous
  13718. stimuli sequentially across time[1][1]. This required deploying a
  13719. novel statistical approach to permit investigation of neural
  13720. activity at the scale of individual trials. Here we present tests
  13721. using synthetic data to assess the sensitivity and specificity of
  13722. this analysis. Data sets were fabricated to match each of several
  13723. potential response patterns derived from single-stimulus response
  13724. distributions. In particular, we simulated dual stimulus trial
  13725. spike counts that reflected fluctuating mixtures of the single
  13726. stimulus spike counts, stable intermediate averages, single
  13727. stimulus winner-take-all, or response distributions that were
  13728. outside the range defined by the single stimulus responses (such
  13729. as summation or suppression). We then assessed how well the
  13730. analysis recovered the correct response pattern as a function of
  13731. the number of simulated trials and the difference between the
  13732. simulated responses to each ``stimulus'' alone. We found
  13733. excellent recovery of the mixture, intermediate, and outside
  13734. categories (>97\% correct), and good recovery of the
  13735. single/winner-take-all category (>90\% correct) when the number
  13736. of trials was >20 and the single-stimulus response rates were
  13737. 50Hz and 20Hz respectively. Both larger numbers of trials and
  13738. greater separation between the single stimulus firing rates
  13739. improved categorization accuracy. The results provide a valid
  13740. benchmark, and guidelines for data collection, for use of this
  13741. method to investigate coding of multiple items at the
  13742. individual-trial time scale. [1]: \#ref-1",
  13743. journal = "bioRxiv",
  13744. pages = "690958",
  13745. month = jul,
  13746. year = 2019,
  13747. language = "en"
  13748. }
  13749. @ARTICLE{Bari2019-iw,
  13750. title = "Stable Representations of Decision Variables for Flexible
  13751. Behavior",
  13752. author = "Bari, Bilal A and Grossman, Cooper D and Lubin, Emily E and
  13753. Rajagopalan, Adithya E and Cressy, Jianna I and Cohen, Jeremiah Y",
  13754. abstract = "Decisions occur in dynamic environments. In the framework of
  13755. reinforcement learning, the probability of performing an action
  13756. is influenced by decision variables. Discrepancies between
  13757. predicted and obtained rewards (reward prediction errors) update
  13758. these variables, but they are otherwise stable between decisions.
  13759. Although reward prediction errors have been mapped to midbrain
  13760. dopamine neurons, it is unclear how the brain represents decision
  13761. variables themselves. We trained mice on a dynamic foraging task
  13762. in which they chose between alternatives that delivered reward
  13763. with changing probabilities. Neurons in the medial prefrontal
  13764. cortex, including projections to the dorsomedial striatum,
  13765. maintained persistent firing rate changes over long timescales.
  13766. These changes stably represented relative action values (to bias
  13767. choices) and total action values (to bias response times) with
  13768. slow decay. In contrast, decision variables were weakly
  13769. represented in the anterolateral motor cortex, a region necessary
  13770. for generating choices. Thus, we define a stable neural mechanism
  13771. to drive flexible behavior.",
  13772. journal = "Neuron",
  13773. month = jun,
  13774. year = 2019,
  13775. language = "en"
  13776. }
  13777. @ARTICLE{noauthor_undated-eq,
  13778. title = "Decision Field Theory"
  13779. }
  13780. @ARTICLE{Japyassu2017-rx,
  13781. title = "Extended spider cognition",
  13782. author = "Japyass{\'u}, Hilton F and Laland, Kevin N",
  13783. abstract = "There is a tension between the conception of cognition as a
  13784. central nervous system (CNS) process and a view of cognition as
  13785. extending towards the body or the contiguous environment. The
  13786. centralised conception requires large or complex nervous systems
  13787. to cope with complex environments. Conversely, the extended
  13788. conception involves the outsourcing of information processing to
  13789. the body or environment, thus making fewer demands on the
  13790. processing power of the CNS. The evolution of extended cognition
  13791. should be particularly favoured among small, generalist predators
  13792. such as spiders, and here, we review the literature to evaluate
  13793. the fit of empirical data with these contrasting models of
  13794. cognition. Spiders do not seem to be cognitively limited,
  13795. displaying a large diversity of learning processes, from
  13796. habituation to contextual learning, including a sense of
  13797. numerosity. To tease apart the central from the extended
  13798. cognition, we apply the mutual manipulability criterion, testing
  13799. the existence of reciprocal causal links between the putative
  13800. elements of the system. We conclude that the web threads and
  13801. configurations are integral parts of the cognitive systems. The
  13802. extension of cognition to the web helps to explain some puzzling
  13803. features of spider behaviour and seems to promote evolvability
  13804. within the group, enhancing innovation through cognitive
  13805. connectivity to variable habitat features. Graded changes in
  13806. relative brain size could also be explained by outsourcing
  13807. information processing to environmental features. More generally,
  13808. niche-constructed structures emerge as prime candidates for
  13809. extending animal cognition, generating the selective pressures
  13810. that help to shape the evolving cognitive system.",
  13811. journal = "Anim. Cogn.",
  13812. volume = 20,
  13813. number = 3,
  13814. pages = "375--395",
  13815. month = may,
  13816. year = 2017,
  13817. keywords = "Evolvability; Extended cognition; Modular cognition; Niche
  13818. construction; Web building",
  13819. language = "en"
  13820. }
  13821. @ARTICLE{Globus1992-on,
  13822. title = "Toward a noncomputational cognitive neuroscience",
  13823. author = "Globus, G G",
  13824. abstract = "The near universally accepted theory that the brain processes
  13825. information persists in current neural network theory where
  13826. there is ``subsymbolic'' computation (Smolensky, 1988) on
  13827. distributed representations. This theory of brain information
  13828. processing may suffice for simplifying models simulated in
  13829. silicon but not for living neural nets where there is ongoing
  13830. chemical tuning of the input/output transfer function at the
  13831. nodes, connection weights, network parameters, and connectivity.
  13832. Here the brain continually changes itself as it intersects with
  13833. information from the outside. An alternative theory to
  13834. information processing is developed in which the brain permits
  13835. and supports ``participation'' of self and other as constraints
  13836. on the dynamically evolving, self-organizing whole. The
  13837. noncomputational process of ``differing and deferring'' in
  13838. nonlinear dynamic neural systems is contrasted with Black's
  13839. (1991) account of molecular information processing. State
  13840. hyperspace for the noncomputational process of nonlinear
  13841. dynamical systems, unlike classical systems, has a fractal
  13842. dimension. The noncomputational model is supported by suggestive
  13843. evidence for fractal properties of the brain.",
  13844. journal = "J. Cogn. Neurosci.",
  13845. publisher = "MIT Press",
  13846. volume = 4,
  13847. number = 4,
  13848. pages = "299--300",
  13849. year = 1992,
  13850. language = "en"
  13851. }
  13852. @ARTICLE{Turchin1992-ad,
  13853. title = "Complex Dynamics in Ecological Time Series",
  13854. author = "Turchin, Peter and Taylor, Andrew D",
  13855. abstract = "Although the possibility of complex dynamical behaviors?limit
  13856. cycles, quasiperodic oscillations, and aperiodic chaos?has been
  13857. recognized theoretically, most ecologists are skeptical of their
  13858. importance in nature. In this paper we develop a methodology for
  13859. reconstructing endogenous (or deterministic) dynamics from
  13860. ecological time series. Our method consists of fitting a
  13861. response surface to the yearly population change as a function
  13862. of lagged population densities. Using the version of the model
  13863. that includes two lags, we fitted time?series data for 14 insect
  13864. and 22 vertebrate populations. The 14 insect populations were
  13865. classified as: unregulated (1 case), exponentially stable (three
  13866. cases), damped oscillations (six cases), limit cycles (one
  13867. case), quasiperiodic oscillations (two cases), and chaos (one
  13868. case). The vertebrate examples exhibited a similar spectrum of
  13869. dynamics, although there were no cases of chaos. We tested the
  13870. results of the response?surface methodology by calculating
  13871. autocorrelation functions for each time series. Autocorrelation
  13872. patterns were in agreement with our findings of periodic
  13873. behaviors (damped oscillations, limit cycles, and
  13874. quasiperiodicity). On the basis of these results, we conclude
  13875. that the complete spectrum of dynamical behaviors, ranging from
  13876. exponential stability to chaos, is likely to be found among
  13877. natural populations.",
  13878. journal = "Ecology",
  13879. publisher = "Wiley Online Library",
  13880. volume = 73,
  13881. number = 1,
  13882. pages = "289--305",
  13883. month = feb,
  13884. year = 1992
  13885. }
  13886. @INPROCEEDINGS{Koza1991-pg,
  13887. title = "Evolution and co-evolution of computer programs to control
  13888. independently-acting agents",
  13889. booktitle = "Proceedings of the First International Conference on Simulation
  13890. of Adaptive Behavior: From Animals to Animats. {MIT} Press,
  13891. Cambridge, {MA}",
  13892. author = "Koza, John R",
  13893. abstract = "This paper describes the recently developed`` genetic
  13894. programming'' paradigm which genetically breeds populations of
  13895. computer programs to solve problems. In genetic programming, the
  13896. individuals in the population are hierarchical computer programs
  13897. of various sizes and shapes. This paper also extends the genetic
  13898. programming paradigm to a`` co-evolution'' algorithm which
  13899. operates simultaneously on two populations of independently-
  13900. acting hierarchical computer programs of various sizes and
  13901. shapes.",
  13902. publisher = "sci.brooklyn.cuny.edu",
  13903. pages = "366--375",
  13904. year = 1991
  13905. }
  13906. @ARTICLE{Pereira2019-ym,
  13907. title = "Fast animal pose estimation using deep neural networks",
  13908. author = "Pereira, Talmo D and Aldarondo, Diego E and Willmore, Lindsay and
  13909. Kislin, Mikhail and Wang, Samuel S-H and Murthy, Mala and
  13910. Shaevitz, Joshua W",
  13911. abstract = "The need for automated and efficient systems for tracking full
  13912. animal pose has increased with the complexity of behavioral data
  13913. and analyses. Here we introduce LEAP (LEAP estimates animal
  13914. pose), a deep-learning-based method for predicting the positions
  13915. of animal body parts. This framework consists of a graphical
  13916. interface for labeling of body parts and training the network.
  13917. LEAP offers fast prediction on new data, and training with as few
  13918. as 100 frames results in 95\% of peak performance. We validated
  13919. LEAP using videos of freely behaving fruit flies and tracked 32
  13920. distinct points to describe the pose of the head, body, wings and
  13921. legs, with an error rate of <3\% of body length. We recapitulated
  13922. reported findings on insect gait dynamics and demonstrated LEAP's
  13923. applicability for unsupervised behavioral classification.
  13924. Finally, we extended the method to more challenging imaging
  13925. situations and videos of freely moving mice.",
  13926. journal = "Nat. Methods",
  13927. volume = 16,
  13928. number = 1,
  13929. pages = "117--125",
  13930. month = jan,
  13931. year = 2019,
  13932. language = "en"
  13933. }
  13934. @ARTICLE{Garamszegi2011-vk,
  13935. title = "Information-theoretic approaches to statistical analysis in
  13936. behavioural ecology: an introduction",
  13937. author = "Garamszegi, L{\'a}szl{\'o} Zsolt",
  13938. abstract = "Scientific thinking may require the consideration of multiple
  13939. hypotheses, which often call for complex statistical models at
  13940. the level of data analysis. The aim of this introduction is to
  13941. provide a brief overview on how competing hypotheses are
  13942. evaluated statistically in behavioural ecological studies and to
  13943. offer potentially fruitful avenues for future methodological
  13944. developments. Complex models have traditionally been treated by
  13945. model selection approaches using threshold-based removal of
  13946. terms, i.e. stepwise selection. A recently introduced method for
  13947. model selection applies an information-theoretic (IT) approach,
  13948. which simultaneously evaluates hypotheses by balancing between
  13949. model complexity and goodness of fit. The IT method has been
  13950. increasingly propagated in the field of ecology, while a
  13951. literature survey shows that its spread in behavioural ecology
  13952. has been much slower, and model simplification using stepwise
  13953. selection is still more widespread than IT-based model selection.
  13954. Why has the use of IT methods in behavioural ecology lagged
  13955. behind other disciplines? This special issue examines the
  13956. suitability of the IT method for analysing data with multiple
  13957. predictors, which researchers encounter in our field. The volume
  13958. brings together different viewpoints to aid behavioural
  13959. ecologists in understanding the method, with the hope of
  13960. enhancing the statistical integration of our discipline.",
  13961. journal = "Behav. Ecol. Sociobiol.",
  13962. volume = 65,
  13963. number = 1,
  13964. pages = "1--11",
  13965. month = jan,
  13966. year = 2011
  13967. }
  13968. @ARTICLE{Panzeri2007-vl,
  13969. title = "Correcting for the sampling bias problem in spike train
  13970. information measures",
  13971. author = "Panzeri, Stefano and Senatore, Riccardo and Montemurro, Marcelo
  13972. A and Petersen, Rasmus S",
  13973. abstract = "Information Theory enables the quantification of how much
  13974. information a neuronal response carries about external stimuli
  13975. and is hence a natural analytic framework for studying neural
  13976. coding. The main difficulty in its practical application to
  13977. spike train analysis is that estimates of neuronal information
  13978. from experimental data are prone to a systematic error (called
  13979. ``bias''). This bias is an inevitable consequence of the limited
  13980. number of stimulus-response samples that it is possible to
  13981. record in a real experiment. In this paper, we first explain the
  13982. origin and the implications of the bias problem in spike train
  13983. analysis. We then review and evaluate some recent
  13984. general-purpose methods to correct for sampling bias: the
  13985. Panzeri-Treves, Quadratic Extrapolation, Best Universal Bound,
  13986. Nemenman-Shafee-Bialek procedures, and a recently proposed
  13987. shuffling bias reduction procedure. Finally, we make practical
  13988. recommendations for the accurate computation of information from
  13989. spike trains. Our main recommendation is to estimate information
  13990. using the shuffling bias reduction procedure in combination with
  13991. one of the other four general purpose bias reduction procedures
  13992. mentioned in the preceding text. This provides information
  13993. estimates with acceptable variance and which are unbiased even
  13994. when the number of trials per stimulus is as small as the number
  13995. of possible discrete neuronal responses.",
  13996. journal = "J. Neurophysiol.",
  13997. publisher = "physiology.org",
  13998. volume = 98,
  13999. number = 3,
  14000. pages = "1064--1072",
  14001. month = sep,
  14002. year = 2007,
  14003. language = "en"
  14004. }
  14005. @ARTICLE{Quian_Quiroga2009-gj,
  14006. title = "Extracting information from neuronal populations: information
  14007. theory and decoding approaches",
  14008. author = "Quian Quiroga, Rodrigo and Panzeri, Stefano",
  14009. abstract = "To a large extent, progress in neuroscience has been driven by
  14010. the study of single-cell responses averaged over several
  14011. repetitions of stimuli or behaviours. However,the brain
  14012. typically makes decisions based on single events by evaluating
  14013. the activity of large neuronal populations. Therefore, to
  14014. further understand how the brain processes information, it is
  14015. important to shift from a single-neuron, multiple-trial
  14016. framework to multiple-neuron, single-trial methodologies. Two
  14017. related approaches--decoding and information theory--can be used
  14018. to extract single-trial information from the activity of
  14019. neuronal populations. Such population analysis can give us more
  14020. information about how neurons encode stimulus features than
  14021. traditional single-cell studies.",
  14022. journal = "Nat. Rev. Neurosci.",
  14023. publisher = "nature.com",
  14024. volume = 10,
  14025. number = 3,
  14026. pages = "173--185",
  14027. month = mar,
  14028. year = 2009,
  14029. language = "en"
  14030. }
  14031. @ARTICLE{Treves1995-zx,
  14032. title = "The Upward Bias in Measures of Information Derived from Limited
  14033. Data Samples",
  14034. author = "Treves, Alessandro and Panzeri, Stefano",
  14035. abstract = "Extracting information measures from limited experimental
  14036. samples, such as those normally available when using data
  14037. recorded in vivo from mammalian cortical neurons, is known to be
  14038. plagued by a systematic error, which tends to bias the estimate
  14039. upward. We calculate here the average of the bias, under certain
  14040. conditions, as an asymptotic expansion in the inverse of the
  14041. size of the data sample. The result agrees with numerical
  14042. simulations, and is applicable, as an additive correction term,
  14043. to measurements obtained under such conditions. Moreover, we
  14044. discuss the implications for measurements obtained through other
  14045. usual procedures.",
  14046. journal = "Neural Comput.",
  14047. publisher = "MIT Press",
  14048. volume = 7,
  14049. number = 2,
  14050. pages = "399--407",
  14051. month = mar,
  14052. year = 1995
  14053. }
  14054. @ARTICLE{Dimitrov2011-fm,
  14055. title = "Information theory in neuroscience",
  14056. author = "Dimitrov, Alexander G and Lazar, Aurel A and Victor, Jonathan D",
  14057. journal = "J. Comput. Neurosci.",
  14058. volume = 30,
  14059. number = 1,
  14060. pages = "1--5",
  14061. month = feb,
  14062. year = 2011,
  14063. language = "en"
  14064. }
  14065. @INCOLLECTION{Nadel2006-ld,
  14066. title = "Information Theory",
  14067. booktitle = "Encyclopedia of Cognitive Science",
  14068. editor = "Nadel, Lynn",
  14069. abstract = "Abstract Information theory is a mathematical theory defining
  14070. the limits and possibilities of communication. It provides a
  14071. quantitative measure of the information content of a message,
  14072. which is independent of the meaning of the message, in terms of
  14073. the reduction of uncertainty resulting from receiving the
  14074. message.",
  14075. publisher = "John Wiley \& Sons, Ltd",
  14076. volume = 61,
  14077. pages = "183",
  14078. month = jan,
  14079. year = 2006,
  14080. address = "Chichester"
  14081. }
  14082. @ARTICLE{Borst1999-st,
  14083. title = "Information theory and neural coding",
  14084. author = "Borst, A and Theunissen, F E",
  14085. abstract = "Information theory quantifies how much information a neural
  14086. response carries about the stimulus. This can be compared to the
  14087. information transferred in particular models of the
  14088. stimulus-response function and to maximum possible information
  14089. transfer. Such comparisons are crucial because they validate
  14090. assumptions present in any neurophysiological analysis. Here we
  14091. review information-theory basics before demonstrating its use in
  14092. neural coding. We show how to use information theory to validate
  14093. simple stimulus-response models of neural coding of dynamic
  14094. stimuli. Because these models require specification of spike
  14095. timing precision, they can reveal which time scales contain
  14096. information in neural coding. This approach shows that dynamic
  14097. stimuli can be encoded efficiently by single neurons and that
  14098. each spike contributes to information transmission. We argue,
  14099. however, that the data obtained so far do not suggest a temporal
  14100. code, in which the placement of spikes relative to each other
  14101. yields additional information.",
  14102. journal = "Nat. Neurosci.",
  14103. volume = 2,
  14104. number = 11,
  14105. pages = "947--957",
  14106. month = nov,
  14107. year = 1999,
  14108. language = "en"
  14109. }
  14110. @ARTICLE{noauthor_undated-vy,
  14111. title = "mataric.pdf"
  14112. }
  14113. @ARTICLE{Peek2016-wi,
  14114. title = "Comparative approaches to escape",
  14115. author = "Peek, Martin Y and Card, Gwyneth M",
  14116. abstract = "Neural circuits mediating visually evoked escape behaviors are
  14117. promising systems in which to dissect the neural basis of
  14118. behavior. Behavioral responses to predator-like looming stimuli,
  14119. and their underlying neural computations, are remarkably similar
  14120. across species. Recently, genetic tools have been applied in this
  14121. classical paradigm, revealing novel non-cortical pathways that
  14122. connect loom processing to defensive behaviors in mammals and
  14123. demonstrating that loom encoding models from locusts also fit
  14124. vertebrate neural responses. In both invertebrates and
  14125. vertebrates, relative spike-timing in descending pathways is a
  14126. mechanism for escape behavior choice. Current findings suggest
  14127. that experimentally tractable systems, such as Drosophila, may be
  14128. applicable models for sensorimotor processing and persistent
  14129. states in higher organisms.",
  14130. journal = "Curr. Opin. Neurobiol.",
  14131. volume = 41,
  14132. pages = "167--173",
  14133. month = dec,
  14134. year = 2016,
  14135. language = "en"
  14136. }
  14137. @ARTICLE{Ache2019-jc,
  14138. title = "Neural Basis for Looming Size and Velocity Encoding in the
  14139. Drosophila Giant Fiber Escape Pathway",
  14140. author = "Ache, Jan M and Polsky, Jason and Alghailani, Shada and Parekh,
  14141. Ruchi and Breads, Patrick and Peek, Martin Y and Bock, Davi D and
  14142. von Reyn, Catherine R and Card, Gwyneth M",
  14143. abstract = "Identified neuron classes in vertebrate cortical [1-4] and
  14144. subcortical [5-8] areas and invertebrate peripheral [9-11] and
  14145. central [12-14] brain neuropils encode specific visual features
  14146. of a panorama. How downstream neurons integrate these features to
  14147. control vital behaviors, like escape, is unclear [15]. In
  14148. Drosophila, the timing of a single spike in the giant fiber (GF)
  14149. descending neuron [16-18] determines whether a fly uses a short
  14150. or long takeoff when escaping a looming predator [13]. We
  14151. previously proposed that GF spike timing results from summation
  14152. of two visual features whose detection is highly conserved across
  14153. animals [19]: an object's subtended angular size and its angular
  14154. velocity [5-8, 11, 20, 21]. We attributed velocity encoding to
  14155. input from lobula columnar type 4 (LC4) visual projection
  14156. neurons, but the size-encoding source remained unknown. Here, we
  14157. show that lobula plate/lobula columnar, type 2 (LPLC2) visual
  14158. projection neurons anatomically specialized to detect looming
  14159. [22] provide the entire GF size component. We find LPLC2 neurons
  14160. to be necessary for GF-mediated escape and show that LPLC2 and
  14161. LC4 synapse directly onto the GF via reconstruction in a fly
  14162. brain electron microscopy (EM) volume [23]. LPLC2 silencing
  14163. eliminates the size component of the GF looming response in
  14164. patch-clamp recordings, leaving only the velocity component. A
  14165. model summing a linear function of angular velocity (provided by
  14166. LC4) and a Gaussian function of angular size (provided by LPLC2)
  14167. replicates GF looming response dynamics and predicts the peak
  14168. response time. We thus present an identified circuit in which
  14169. information from looming feature-detecting neurons is combined by
  14170. a common post-synaptic target to determine behavioral output.",
  14171. journal = "Curr. Biol.",
  14172. volume = 29,
  14173. number = 6,
  14174. pages = "1073--1081.e4",
  14175. month = mar,
  14176. year = 2019,
  14177. keywords = "Drosophila; descending neuron; electrophysiology; escape; in vivo
  14178. patch clamp; neural circuit reconstruction; sensorimotor
  14179. integration; visual feature detection; visual looming; visual
  14180. projection neuron",
  14181. language = "en"
  14182. }
  14183. @ARTICLE{Domenici2011-ex,
  14184. title = "Animal escapology I: theoretical issues and emerging trends in
  14185. escape trajectories",
  14186. author = "Domenici, Paolo and Blagburn, Jonathan M and Bacon, Jonathan P",
  14187. abstract = "Escape responses are used by many animal species as their main
  14188. defence against predator attacks. Escape success is determined by
  14189. a number of variables; important are the directionality (the
  14190. percentage of responses directed away from the threat) and the
  14191. escape trajectories (ETs) measured relative to the threat.
  14192. Although logic would suggest that animals should always turn away
  14193. from a predator, work on various species shows that these away
  14194. responses occur only approximately 50-90\% of the time. A small
  14195. proportion of towards responses may introduce some
  14196. unpredictability and may be an adaptive feature of the escape
  14197. system. Similar issues apply to ETs. Theoretically, an optimal ET
  14198. can be modelled on the geometry of predator-prey encounters.
  14199. However, unpredictability (and hence high variability) in
  14200. trajectories may be necessary for preventing predators from
  14201. learning a simple escape pattern. This review discusses the
  14202. emerging trends in escape trajectories, as well as the modulating
  14203. key factors, such as the surroundings and body design. The main
  14204. ET patterns identified are: (1) high ET variability within a
  14205. limited angular sector (mainly 90-180 deg away from the threat;
  14206. this variability is in some cases based on multiple peaks of
  14207. ETs), (2) ETs that allow sensory tracking of the threat and (3)
  14208. ETs towards a shelter. These characteristic features are observed
  14209. across various taxa and, therefore, their expression may be
  14210. mainly related to taxon-independent animal design features and to
  14211. the environmental context in which prey live - for example
  14212. whether the immediate surroundings of the prey provide potential
  14213. refuges.",
  14214. journal = "J. Exp. Biol.",
  14215. volume = 214,
  14216. number = "Pt 15",
  14217. pages = "2463--2473",
  14218. month = aug,
  14219. year = 2011,
  14220. language = "en"
  14221. }
  14222. @ARTICLE{Schaefer2001-fc,
  14223. title = "Descending influences on escape behavior and motor pattern in the
  14224. cockroach",
  14225. author = "Schaefer, P L and Ritzmann, R E",
  14226. abstract = "The escape behavior of the cockroach is a ballistic behavior with
  14227. well characterized kinematics. The circuitry known to control the
  14228. behavior lies in the thoracic ganglia, abdominal ganglia, and
  14229. abdominal nerve cord. Some evidence suggests inputs may occur
  14230. from the brain or suboesophageal ganglion. We tested this notion
  14231. by decapitating cockroaches, removing all descending inputs, and
  14232. evoking escape responses. The decapitated cockroaches exhibited
  14233. directionally appropriate escape turns. However, there was a
  14234. front-to-back gradient of change: the front legs moved little if
  14235. at all, the middle legs moved in the proper direction but with
  14236. reduced excursion, and the rear legs moved normally. The same
  14237. pattern was seen when only inputs from the brain were removed,
  14238. the suboesophageal ganglion remaining intact and connected to the
  14239. thoracic ganglia. Electromyogram (EMG) analysis showed that the
  14240. loss of or reduction in excursion was accompanied by a loss of or
  14241. reduction in fast motor neuron activity. The loss of fast motor
  14242. neuron activity was also observed in a reduced preparation in
  14243. which descending neural signals were reversibly blocked via an
  14244. isotonic sucrose solution superfusing the neck connectives,
  14245. indicating that the changes seen were not due to trauma. Our data
  14246. demonstrate that while the thoracic circuitry is sufficient to
  14247. produce directional escape, lesion or blockage of the connective
  14248. affects the excitability of components of the escape circuitry.
  14249. Because of the rapidity of the escape response, such effects are
  14250. likely due to the elimination of tonic descending inputs.",
  14251. journal = "J. Neurobiol.",
  14252. volume = 49,
  14253. number = 1,
  14254. pages = "9--28",
  14255. month = oct,
  14256. year = 2001,
  14257. language = "en"
  14258. }
  14259. % The entry below contains non-ASCII chars that could not be converted
  14260. % to a LaTeX equivalent.
  14261. @INPROCEEDINGS{Attias2003-wk,
  14262. title = "Planning by probabilistic inference",
  14263. booktitle = "{AISTATS}",
  14264. author = "Attias, Hagai",
  14265. abstract = "This paper presents and demonstrates a new approach to the
  14266. problem of planning under uncertainty. Actions are treated as
  14267. hidden variables, with their own prior distributions, in a
  14268. probabilistic generative model involving actions and states.
  14269. Planning is done by computing …",
  14270. publisher = "pdfs.semanticscholar.org",
  14271. year = 2003
  14272. }
  14273. @ARTICLE{Boutilier2011-ui,
  14274. title = "{Decision-Theoretic} Planning: Structural Assumptions and
  14275. Computational Leverage",
  14276. author = "Boutilier, C and Dean, T and Hanks, S",
  14277. abstract = "Planning under uncertainty is a central problem in the study of",
  14278. journal = "arXiv e-prints",
  14279. publisher = "adsabs.harvard.edu",
  14280. month = may,
  14281. year = 2011,
  14282. keywords = "Computer Science - Artificial Intelligence"
  14283. }
  14284. @ARTICLE{Grush2004-db,
  14285. title = "The emulation theory of representation: motor control, imagery,
  14286. and perception",
  14287. author = "Grush, Rick",
  14288. abstract = "The emulation theory of representation is developed and explored
  14289. as a framework that can revealingly synthesize a wide variety of
  14290. representational functions of the brain. The framework is based
  14291. on constructs from control theory (forward models) and signal
  14292. processing (Kalman filters). The idea is that in addition to
  14293. simply engaging with the body and environment, the brain
  14294. constructs neural circuits that act as models of the body and
  14295. environment. During overt sensorimotor engagement, these models
  14296. are driven by efference copies in parallel with the body and
  14297. environment, in order to provide expectations of the sensory
  14298. feedback, and to enhance and process sensory information. These
  14299. models can also be run off-line in order to produce imagery,
  14300. estimate outcomes of different actions, and evaluate and develop
  14301. motor plans. The framework is initially developed within the
  14302. context of motor control, where it has been shown that inner
  14303. models running in parallel with the body can reduce the effects
  14304. of feedback delay problems. The same mechanisms can account for
  14305. motor imagery as the off-line driving of the emulator via
  14306. efference copies. The framework is extended to account for
  14307. visual imagery as the off-line driving of an emulator of the
  14308. motor-visual loop. I also show how such systems can provide for
  14309. amodal spatial imagery. Perception, including visual perception,
  14310. results from such models being used to form expectations of, and
  14311. to interpret, sensory input. I close by briefly outlining other
  14312. cognitive functions that might also be synthesized within this
  14313. framework, including reasoning, theory of mind phenomena, and
  14314. language.",
  14315. journal = "Behav. Brain Sci.",
  14316. publisher = "cambridge.org",
  14317. volume = 27,
  14318. number = 3,
  14319. pages = "377--96; discussion 396--442",
  14320. month = jun,
  14321. year = 2004,
  14322. language = "en"
  14323. }
  14324. % The entry below contains non-ASCII chars that could not be converted
  14325. % to a LaTeX equivalent.
  14326. @ARTICLE{Murphy2002-fs,
  14327. title = "Dynamic bayesian networks: representation, inference and
  14328. learning",
  14329. author = "Murphy, Kevin Patrick and Russell, Stuart",
  14330. abstract = "Sequential data arises in many areas of science and engineering.
  14331. The data may either be a time series, generated by a dynamical
  14332. system, or a sequence generated by a 1-dimensional spatial
  14333. process, eg, biosequences. One may be interested either in
  14334. online analysis, where the data arrives in real-time, or in
  14335. offline analysis, where all the data has already been collected.
  14336. In online analysis, one common task is to predict future
  14337. observations, given all the observations up to the present time,
  14338. which we will denote by y1: t=(y1,..., yt).(In this thesis …",
  14339. publisher = "University of California, Berkeley Dissertation",
  14340. year = 2002
  14341. }
  14342. % The entry below contains non-ASCII chars that could not be converted
  14343. % to a LaTeX equivalent.
  14344. @ARTICLE{Toussaint2009-ac,
  14345. title = "Probabilistic inference as a model of planned behavior",
  14346. author = "Toussaint, Marc",
  14347. abstract = "The problem of planning and goal-directed behavior has been
  14348. addressed in computer science for many years, typically based on
  14349. classical concepts like Bellman's optimality principle, dynamic
  14350. programming, or Reinforcement Learning methods--but is this the
  14351. only way to address the problem? Recently there is growing
  14352. interest in using probabilistic inference methods for decision
  14353. making and planning. Promising about such approaches is that
  14354. they naturally extend to distributed state representations and
  14355. efficiently cope with …",
  14356. journal = "KI",
  14357. publisher = "researchgate.net",
  14358. volume = 23,
  14359. number = 3,
  14360. pages = "23--29",
  14361. year = 2009
  14362. }
  14363. @ARTICLE{Muller2017-ag,
  14364. title = "What Is Morphological Computation? On How the Body Contributes to
  14365. Cognition and Control",
  14366. author = "M{\"u}ller, Vincent C and Hoffmann, Matej",
  14367. abstract = "The contribution of the body to cognition and control in natural
  14368. and artificial agents is increasingly described as ``offloading
  14369. computation from the brain to the body,'' where the body is said
  14370. to perform ``morphological computation.'' Our investigation of
  14371. four characteristic cases of morphological computation in animals
  14372. and robots shows that the ``offloading'' perspective is
  14373. misleading. Actually, the contribution of body morphology to
  14374. cognition and control is rarely computational, in any useful
  14375. sense of the word. We thus distinguish (1) morphology that
  14376. facilitates control, (2) morphology that facilitates perception,
  14377. and the rare cases of (3) morphological computation proper, such
  14378. as reservoir computing, where the body is actually used for
  14379. computation. This result contributes to the understanding of the
  14380. relation between embodiment and computation: The question for
  14381. robot design and cognitive science is not whether computation is
  14382. offloaded to the body, but to what extent the body facilitates
  14383. cognition and control-how it contributes to the overall
  14384. orchestration of intelligent behavior.",
  14385. journal = "Artif. Life",
  14386. volume = 23,
  14387. number = 1,
  14388. pages = "1--24",
  14389. month = jan,
  14390. year = 2017,
  14391. keywords = "Body; cognition; computation; control; embodiment; soft robotics",
  14392. language = "en"
  14393. }
  14394. @ARTICLE{Pezzulo2017-xp,
  14395. title = "{Model-Based} Approaches to Active Perception and Control",
  14396. author = "Pezzulo, Giovanni and Donnarumma, Francesco and Iodice,
  14397. Pierpaolo and Maisto, Domenico and Stoianov, Ivilin",
  14398. abstract = "There is an on-going debate in cognitive (neuro) science and
  14399. philosophy between classical cognitive theory and embodied,
  14400. embedded, extended, and enactive (``4-Es'') views of
  14401. cognition---a family of theories that emphasize the role of the
  14402. body in cognition and the importance of brain-body-environment
  14403. interaction over and above internal representation. This debate
  14404. touches foundational issues, such as whether the brain
  14405. internally represents the external environment, and ``infers''
  14406. or ``computes'' something. Here we focus on two (4-Es-based)
  14407. criticisms to traditional cognitive theories---to the notions of
  14408. passive perception and of serial information processing---and
  14409. discuss alternative ways to address them, by appealing to
  14410. frameworks that use, or do not use, notions of internal
  14411. modelling and inference. Our analysis illustrates that: an
  14412. explicitly inferential framework can capture some key aspects of
  14413. embodied and enactive theories of cognition; some claims of
  14414. computational and dynamical theories can be reconciled rather
  14415. than seen as alternative explanations of cognitive phenomena;
  14416. and some aspects of cognitive processing (e.g., detached
  14417. cognitive operations, such as planning and imagination) that are
  14418. sometimes puzzling to explain from enactive and
  14419. non-representational perspectives can, instead, be captured
  14420. nicely from the perspective that internal generative models and
  14421. predictive processing mediate adaptive control loops.",
  14422. journal = "Entropy",
  14423. publisher = "Multidisciplinary Digital Publishing Institute",
  14424. volume = 19,
  14425. number = 6,
  14426. pages = "266",
  14427. month = jun,
  14428. year = 2017,
  14429. language = "en"
  14430. }
  14431. @ARTICLE{Milkowski2018-bi,
  14432. title = "Morphological Computation: Nothing but Physical Computation",
  14433. author = "Mi{\l}kowski, Marcin",
  14434. abstract = "The purpose of this paper is to argue against the claim that
  14435. morphological computation is substantially different from other
  14436. kinds of physical computation. I show that some (but not all)
  14437. purported cases of morphological computation do not count as
  14438. specifically computational, and that those that do are solely
  14439. physical computational systems. These latter cases are not,
  14440. however, specific enough: all computational systems, not only
  14441. morphological ones, may (and sometimes should) be studied in
  14442. various ways, including their energy efficiency, cost,
  14443. reliability, and durability. Second, I critically analyze the
  14444. notion of ``offloading'' computation to the morphology of an
  14445. agent or robot, by showing that, literally, computation is
  14446. sometimes not offloaded but simply avoided. Third, I point out
  14447. that while the morphology of any agent is indicative of the
  14448. environment that it is adapted to, or informative about that
  14449. environment, it does not follow that every agent has access to
  14450. its morphology as the model of its environment.",
  14451. journal = "Entropy",
  14452. publisher = "Multidisciplinary Digital Publishing Institute",
  14453. volume = 20,
  14454. number = 12,
  14455. pages = "942",
  14456. month = dec,
  14457. year = 2018,
  14458. language = "en"
  14459. }
  14460. @UNPUBLISHED{Calhoun2019-qk,
  14461. title = "Unsupervised identification of the internal states that shape
  14462. natural behavior",
  14463. author = "Calhoun, Adam J and Pillow, Jonathan and Murthy, Mala",
  14464. abstract = "Internal states can shape stimulus responses and decision-making,
  14465. but we lack methods to identify internal states and how they
  14466. evolve over time. To address this gap, we have developed an
  14467. unsupervised method to identify internal states from behavioral
  14468. data, and have applied it to the study of a dynamic social
  14469. interaction. During courtship, Drosophila melanogaster males
  14470. pattern their songs using feedback cues from their partner. Our
  14471. model uncovers three latent states underlying this behavior, and
  14472. is able to predict the moment-to-moment variation in natural song
  14473. patterning decisions. These distinct behavioral states correspond
  14474. to different sensorimotor strategies, each of which is
  14475. characterized by different mappings from feedback cues to song
  14476. modes. Using the model, we show that a pair of neurons previously
  14477. thought to be command neurons for song production are sufficient
  14478. to drive switching between states. Our results reveal how animals
  14479. compose behavior from previously unidentified internal states, a
  14480. necessary step for quantitative descriptions of animal behavior
  14481. that link environmental cues, internal needs, neuronal activity,
  14482. and motor outputs.",
  14483. journal = "bioRxiv",
  14484. pages = "691196",
  14485. month = jul,
  14486. year = 2019,
  14487. language = "en"
  14488. }
  14489. @ARTICLE{Gallego2017-bn,
  14490. title = "Neural Manifolds for the Control of Movement",
  14491. author = "Gallego, Juan A and Perich, Matthew G and Miller, Lee E and
  14492. Solla, Sara A",
  14493. abstract = "The analysis of neural dynamics in several brain cortices has
  14494. consistently uncovered low-dimensional manifolds that capture a
  14495. significant fraction of neural variability. These neural
  14496. manifolds are spanned by specific patterns of correlated neural
  14497. activity, the ``neural modes.'' We discuss a model for neural
  14498. control of movement in which the time-dependent activation of
  14499. these neural modes is the generator of motor behavior. This
  14500. manifold-based view of motor cortex may lead to a better
  14501. understanding of how the brain controls movement.",
  14502. journal = "Neuron",
  14503. volume = 94,
  14504. number = 5,
  14505. pages = "978--984",
  14506. month = jun,
  14507. year = 2017,
  14508. language = "en"
  14509. }
  14510. @ARTICLE{Hinman2019-nx,
  14511. title = "Neuronal representation of environmental boundaries in egocentric
  14512. coordinates",
  14513. author = "Hinman, James R and Chapman, G William and Hasselmo, Michael E",
  14514. abstract = "Movement through space is a fundamental behavior for all animals.
  14515. Cognitive maps of environments are encoded in the hippocampal
  14516. formation in an allocentric reference frame, but motor movements
  14517. that comprise physical navigation are represented within an
  14518. egocentric reference frame. Allocentric navigational plans must
  14519. be converted to an egocentric reference frame prior to
  14520. implementation as overt behavior. Here we describe an egocentric
  14521. spatial representation of environmental boundaries in the
  14522. dorsomedial striatum.",
  14523. journal = "Nat. Commun.",
  14524. volume = 10,
  14525. number = 1,
  14526. pages = "2772",
  14527. month = jun,
  14528. year = 2019,
  14529. language = "en"
  14530. }
  14531. @ARTICLE{Groman2019-xc,
  14532. title = "Orbitofrontal Circuits Control Multiple {Reinforcement-Learning}
  14533. Processes",
  14534. author = "Groman, Stephanie M and Keistler, Colby and Keip, Alex J and
  14535. Hammarlund, Emma and DiLeone, Ralph J and Pittenger, Christopher
  14536. and Lee, Daeyeol and Taylor, Jane R",
  14537. abstract = "Adaptive decision making in dynamic environments requires
  14538. multiple reinforcement-learning steps that may be implemented by
  14539. dissociable neural circuits. Here, we used a novel directionally
  14540. specific viral ablation approach to investigate the function of
  14541. several anatomically defined orbitofrontal cortex (OFC) circuits
  14542. during adaptive, flexible decision making in rats trained on a
  14543. probabilistic reversal learning task. Ablation of OFC neurons
  14544. projecting to the nucleus accumbens selectively disrupted
  14545. performance following a reversal, by disrupting the use of
  14546. negative outcomes to guide subsequent choices. Ablation of
  14547. amygdala neurons projecting to the OFC also impaired reversal
  14548. performance, but due to disruptions in the use of positive
  14549. outcomes to guide subsequent choices. Ablation of OFC neurons
  14550. projecting to the amygdala, by contrast, enhanced reversal
  14551. performance by destabilizing action values. Our data are
  14552. inconsistent with a unitary function of the OFC in decision
  14553. making. Rather, distinct OFC-amygdala-striatal circuits mediate
  14554. distinct components of the action-value updating and maintenance
  14555. necessary for decision making.",
  14556. journal = "Neuron",
  14557. month = jun,
  14558. year = 2019,
  14559. keywords = "amygdala; decision making; nucleus accumbens; orbitofrontal
  14560. cortex; reinforcement learning",
  14561. language = "en"
  14562. }
  14563. @UNPUBLISHED{Johnson2019-wt,
  14564. title = "Probabilistic Models of Larval Zebrafish Behavior: Structure on
  14565. Many Scales",
  14566. author = "Johnson, Robert Evan and Linderman, Scott and Panier, Thomas and
  14567. Wee, Caroline Lei and Song, Erin and Herrera, Kristian Joseph and
  14568. Miller, Andrew and Engert, Florian",
  14569. abstract = "Nervous systems have evolved to combine environmental information
  14570. with internal state to select and generate adaptive behavioral
  14571. sequences. To better understand these computations and their
  14572. implementation in neural circuits, natural behavior must be
  14573. carefully measured and quantified. Here, we collect high spatial
  14574. resolution video of single zebrafish larvae swimming in a
  14575. naturalistic environment and develop models of their action
  14576. selection across exploration and hunting. Zebrafish larvae swim
  14577. in punctuated bouts separated by longer periods of rest called
  14578. interbout intervals. We take advantage of this structure by
  14579. categorizing bouts into discrete types and representing their
  14580. behavior as labeled sequences of bout-types emitted over time. We
  14581. then construct probabilistic models - specifically, marked
  14582. renewal processes - to evaluate how bout-types and interbout
  14583. intervals are selected by the fish as a function of its internal
  14584. hunger state, behavioral history, and the locations and
  14585. properties of nearby prey. Finally, we evaluate the models by
  14586. their predictive likelihood and their ability to generate
  14587. realistic trajectories of virtual fish swimming through simulated
  14588. environments. Our simulations capture multiple timescales of
  14589. structure in larval zebrafish behavior and expose many ways in
  14590. which hunger state influences their action selection to promote
  14591. food seeking during hunger and safety during satiety.",
  14592. journal = "bioRxiv",
  14593. pages = "672246",
  14594. month = jun,
  14595. year = 2019,
  14596. language = "en"
  14597. }
  14598. @UNPUBLISHED{De_Groot2019-zn,
  14599. title = "{NINscope}: a versatile miniscope for multi-region circuit
  14600. investigations",
  14601. author = "de Groot, Andres and van den Boom, Bastijn J G and van Genderen,
  14602. Romano M and Coppens, Joris and van Veldhuijzen, John and Bos,
  14603. Joop and Hoedemaker, Hugo and Negrello, Mario and Wiluhn, Ingo
  14604. and De Zeeuw, Chris I and Hoogland, Tycho M",
  14605. abstract = "Miniaturized fluorescence microscopes (miniscopes) have been
  14606. instrumental to monitor neural activity during unrestrained
  14607. behavior and their open-source versions have helped to distribute
  14608. them at an affordable cost. Generally, the footprint and weight
  14609. of open-source miniscopes is sacrificed for added functionality.
  14610. Here, we present NINscope: a light-weight, small footprint
  14611. open-source miniscope that incorporates a high-sensitivity image
  14612. sensor, an inertial measurement unit (IMU), and an LED driver for
  14613. an external optogenetic probe. We highlight the advantages of
  14614. NINscope by performing the first simultaneous cellular resolution
  14615. (dual scope) recordings from cerebellum and cerebral cortex in
  14616. unrestrained mice, revealing that the activity of both regions
  14617. generally precede the onset of behavioral acceleration. At the
  14618. same time, we demonstrate the optogenetic stimulation
  14619. capabilities of NINscope and show that cerebral cortical activity
  14620. can be driven strongly by cerebellar stimulation. Finally, we
  14621. combine optogenetic stimulation of cortex with imaging in the
  14622. dorsal striatum and replicate previous studies that show action
  14623. space is encoded by neurons in this subcortical region. In
  14624. combination with cross-platform control software NINscope is a
  14625. versatile addition to the expanding toolbox of open-source
  14626. miniscopes and will aid multi-region circuit investigations
  14627. during unrestrained behavior.",
  14628. journal = "bioRxiv",
  14629. pages = "685909",
  14630. month = jun,
  14631. year = 2019,
  14632. language = "en"
  14633. }
  14634. @UNPUBLISHED{Vertechi2019-bo,
  14635. title = "Inference based decisions in a hidden state foraging task:
  14636. differential contributions of prefrontal cortical areas",
  14637. author = "Vertechi, Pietro and Lottem, Eran and Sarra, Dario and Godinho,
  14638. Beatriz and Treves, Isaac and Quendera, Tiago and Lohuis,
  14639. Matthijs Nicolai Oude and Mainen, Zachary F",
  14640. abstract = "Essential features of the world are often hidden and must be
  14641. inferred by constructing internal models based on indirect
  14642. evidence. Here, to study the mechanisms of inference we
  14643. established a foraging task that is naturalistic and easily
  14644. learned, yet can distinguish inference from simpler strategies
  14645. such as the direct integration of sensory data. We show that both
  14646. mice and humans learn a strategy consistent with optimal
  14647. inference of a hidden state. However, humans acquire this
  14648. strategy more than an order of magnitude faster than mice. Using
  14649. optogenetics in mice we show that orbitofrontal and anterior
  14650. cingulate cortex inactivation impact task performance, but only
  14651. orbitofrontal inactivation reverts mice from an inference-based
  14652. to a stimulus-bound decision strategy. These results establish a
  14653. cross-species paradigm for studying the problem of
  14654. inference-based decision-making and begin to dissect the network
  14655. of brain regions crucial for its performance.",
  14656. journal = "bioRxiv",
  14657. pages = "679142",
  14658. month = jun,
  14659. year = 2019,
  14660. language = "en"
  14661. }
  14662. @MISC{noauthor_undated-pm,
  14663. title = "Refactoring {UI} v1.0.1.pdf",
  14664. keywords = "books"
  14665. }
  14666. @ARTICLE{Etienne2004-dr,
  14667. title = "Path integration in mammals",
  14668. author = "Etienne, Ariane S and Jeffery, Kathryn J",
  14669. abstract = "It is often assumed that navigation implies the use, by animals,
  14670. of landmarks indicating the location of the goal. However, many
  14671. animals (including humans) are able to return to the starting
  14672. point of a journey, or to other goal sites, by relying on
  14673. self-motion cues only. This process is known as path integration,
  14674. and it allows an agent to calculate a route without making use of
  14675. landmarks. We review the current literature on path integration
  14676. and its interaction with external, location-based cues. Special
  14677. importance is given to the correlation between observable
  14678. behavior and the activity pattern of particular neural cell
  14679. populations that implement the internal representation of space.
  14680. In mammals, the latter may well be the first high-level cognitive
  14681. representation to be understood at the neural level.",
  14682. journal = "Hippocampus",
  14683. volume = 14,
  14684. number = 2,
  14685. pages = "180--192",
  14686. year = 2004,
  14687. language = "en"
  14688. }
  14689. % The entry below contains non-ASCII chars that could not be converted
  14690. % to a LaTeX equivalent.
  14691. @ARTICLE{Tarsitano1997-jo,
  14692. title = "Araneophagic jumping spiders discriminate between detour routes
  14693. that do and do not lead to prey",
  14694. author = "Tarsitano, Michael S and Jackson, Robert R",
  14695. abstract = "In a laboratory study, 12 different experimental set-ups were
  14696. used to examine the ability ofPortia fimbriataa web-invading
  14697. araneophagic jumping spider from Queensland, Australia, to choose
  14698. between two detour paths, only one of which led to a lure (a
  14699. dead, dried spider). Regardless of set-up, the spider could see
  14700. the lure when on the starting platform of the apparatus, but not
  14701. after leaving the starting platform. The spider consistently
  14702. chose the `correct route' (the route that led to the lure) more
  14703. often than the `wrong route' (the route that did not lead to the
  14704. lure). In these tests, the spider was able to make detours that
  14705. required walking about 180° away from the lure and walking past
  14706. where the incorrect route began. There was also a pronounced
  14707. relationship between time of day when tests were carried out and
  14708. the spider's tendency to choose a route. Furthermore, those
  14709. spiders that chose the wrong route abandoned the detour more
  14710. frequently than those that chose the correct route, despite both
  14711. groups being unable to see the lure when the decision was made to
  14712. abandon the detour.",
  14713. journal = "Anim. Behav.",
  14714. volume = 53,
  14715. number = 2,
  14716. pages = "257--266",
  14717. month = feb,
  14718. year = 1997
  14719. }
  14720. @ARTICLE{Todorov2004-re,
  14721. title = "Optimality principles in sensorimotor control",
  14722. author = "Todorov, Emanuel",
  14723. abstract = "The sensorimotor system is a product of evolution, development,
  14724. learning and adaptation-which work on different time scales to
  14725. improve behavioral performance. Consequently, many theories of
  14726. motor function are based on 'optimal performance': they quantify
  14727. task goals as cost functions, and apply the sophisticated tools
  14728. of optimal control theory to obtain detailed behavioral
  14729. predictions. The resulting models, although not without
  14730. limitations, have explained more empirical phenomena than any
  14731. other class. Traditional emphasis has been on optimizing desired
  14732. movement trajectories while ignoring sensory feedback. Recent
  14733. work has redefined optimality in terms of feedback control laws,
  14734. and focused on the mechanisms that generate behavior online. This
  14735. approach has allowed researchers to fit previously unrelated
  14736. concepts and observations into what may become a unified
  14737. theoretical framework for interpreting motor function. At the
  14738. heart of the framework is the relationship between high-level
  14739. goals, and the real-time sensorimotor control strategies most
  14740. suitable for accomplishing those goals.",
  14741. journal = "Nat. Neurosci.",
  14742. volume = 7,
  14743. number = 9,
  14744. pages = "907--915",
  14745. month = sep,
  14746. year = 2004,
  14747. language = "en"
  14748. }
  14749. @ARTICLE{Todorov2009-gp,
  14750. title = "Efficient computation of optimal actions",
  14751. author = "Todorov, Emanuel",
  14752. abstract = "Optimal choice of actions is a fundamental problem relevant to
  14753. fields as diverse as neuroscience, psychology, economics,
  14754. computer science, and control engineering. Despite this broad
  14755. relevance the abstract setting is similar: we have an agent
  14756. choosing actions over time, an uncertain dynamical system whose
  14757. state is affected by those actions, and a performance criterion
  14758. that the agent seeks to optimize. Solving problems of this kind
  14759. remains hard, in part, because of overly generic formulations.
  14760. Here, we propose a more structured formulation that greatly
  14761. simplifies the construction of optimal control laws in both
  14762. discrete and continuous domains. An exhaustive search over
  14763. actions is avoided and the problem becomes linear. This yields
  14764. algorithms that outperform Dynamic Programming and Reinforcement
  14765. Learning, and thereby solve traditional problems more
  14766. efficiently. Our framework also enables computations that were
  14767. not possible before: composing optimal control laws by mixing
  14768. primitives, applying deterministic methods to stochastic systems,
  14769. quantifying the benefits of error tolerance, and inferring goals
  14770. from behavioral data via convex optimization. Development of a
  14771. general class of easily solvable problems tends to accelerate
  14772. progress--as linear systems theory has done, for example. Our
  14773. framework may have similar impact in fields where optimal choice
  14774. of actions is relevant.",
  14775. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  14776. volume = 106,
  14777. number = 28,
  14778. pages = "11478--11483",
  14779. month = jul,
  14780. year = 2009,
  14781. language = "en"
  14782. }
  14783. @ARTICLE{Sanfey2006-em,
  14784. title = "Neuroeconomics: cross-currents in research on decision-making",
  14785. author = "Sanfey, Alan G and Loewenstein, George and McClure, Samuel M and
  14786. Cohen, Jonathan D",
  14787. abstract = "Despite substantial advances, the question of how we make
  14788. decisions and judgments continues to pose important challenges
  14789. for scientific research. Historically, different disciplines have
  14790. approached this problem using different techniques and
  14791. assumptions, with few unifying efforts made. However, the field
  14792. of neuroeconomics has recently emerged as an inter-disciplinary
  14793. effort to bridge this gap. Research in neuroscience and
  14794. psychology has begun to investigate neural bases of decision
  14795. predictability and value, central parameters in the economic
  14796. theory of expected utility. Economics, in turn, is being
  14797. increasingly influenced by a multiple-systems approach to
  14798. decision-making, a perspective strongly rooted in psychology and
  14799. neuroscience. The integration of these disparate theoretical
  14800. approaches and methodologies offers exciting potential for the
  14801. construction of more accurate models of decision-making.",
  14802. journal = "Trends Cogn. Sci.",
  14803. volume = 10,
  14804. number = 3,
  14805. pages = "108--116",
  14806. month = mar,
  14807. year = 2006,
  14808. language = "en"
  14809. }
  14810. @ARTICLE{Padoa-Schioppa2006-ps,
  14811. title = "Neurons in the orbitofrontal cortex encode economic value",
  14812. author = "Padoa-Schioppa, Camillo and Assad, John A",
  14813. abstract = "Economic choice is the behaviour observed when individuals
  14814. select one among many available options. There is no
  14815. intrinsically 'correct' answer: economic choice depends on
  14816. subjective preferences. This behaviour is traditionally the
  14817. object of economic analysis and is also of primary interest in
  14818. psychology. However, the underlying mental processes and
  14819. neuronal mechanisms are not well understood. Theories of human
  14820. and animal choice have a cornerstone in the concept of 'value'.
  14821. Consider, for example, a monkey offered one raisin versus one
  14822. piece of apple: behavioural evidence suggests that the animal
  14823. chooses by assigning values to the two options. But where and
  14824. how values are represented in the brain is unclear. Here we show
  14825. that, during economic choice, neurons in the orbitofrontal
  14826. cortex (OFC) encode the value of offered and chosen goods.
  14827. Notably, OFC neurons encode value independently of visuospatial
  14828. factors and motor responses. If a monkey chooses between A and
  14829. B, neurons in the OFC encode the value of the two goods
  14830. independently of whether A is presented on the right and B on
  14831. the left, or vice versa. This trait distinguishes the OFC from
  14832. other brain areas in which value modulates activity related to
  14833. sensory or motor processes. Our results have broad implications
  14834. for possible psychological models, suggesting that economic
  14835. choice is essentially choice between goods rather than choice
  14836. between actions. In this framework, neurons in the OFC seem to
  14837. be a good candidate network for value assignment underlying
  14838. economic choice.",
  14839. journal = "Nature",
  14840. publisher = "nature.com",
  14841. volume = 441,
  14842. number = 7090,
  14843. pages = "223--226",
  14844. month = may,
  14845. year = 2006,
  14846. language = "en"
  14847. }
  14848. @ARTICLE{Clark_undated-kw,
  14849. title = "Putting Brain, Body, and World Together Again",
  14850. author = "Clark, Andy",
  14851. keywords = "Books"
  14852. }
  14853. @UNPUBLISHED{Obaid2019-jn,
  14854. title = "Massively Parallel Microwire Arrays Integrated with {CMOS} chips
  14855. for Neural Recording",
  14856. author = "Obaid, Abdulmalik and Hanna, Mina-Elraheb and Wu, Yu-Wei and
  14857. Kollo, Mihaly and Racz, Romeo and Angle, Matthew R and
  14858. M{\"u}ller, Jan and Brackbill, Nora and Wray, William and Franke,
  14859. Felix and Chichilinsky, E J and Hierlemann, Andreas and Ding, Jun
  14860. B and Schaefer, Andreas T and Melosh, Nicholas A",
  14861. abstract = "Abstract Multiple-channel count neural recordings of brain
  14862. activity are a powerful technique that is increasingly uncovering
  14863. new aspects of neural communication, computation, and prosthetic
  14864. interfaces. However, while silicon CMOS devices continue to scale
  14865. rapidly in number and power in planar geometries, this scaling
  14866. has not been followed for large-scale mapping along three
  14867. dimensions. Here, we present a new strategy to interface
  14868. CMOS-based devices with a three-dimensional microwire array,
  14869. providing the link between rapidly-developing electronics, and
  14870. high density neural interfaces. The system consists of a bundle
  14871. of insulated and spaced microwires perpendicularly mated to a
  14872. commercial large-scale CMOS microelectrode array, such as a
  14873. camera chip. The modular nature of the design enables a variety
  14874. of microwire types and sizes to be integrated with different
  14875. types of silicon-based arrays, allowing channel counts to be
  14876. scaled from a few dozen to thousands of electrodes using the same
  14877. fundamental platform. This system has excellent recording
  14878. performance, demonstrated via single unit and local-field
  14879. potential recordings in isolated retina, and in the motor cortex
  14880. and striatum of awake moving mice. This concept links the rapid
  14881. progress and power of commercial multiplexing, digitisation and
  14882. data acquisition hardware together with a three-dimensional
  14883. neural interface.",
  14884. journal = "bioRxiv",
  14885. pages = "573295",
  14886. month = mar,
  14887. year = 2019,
  14888. language = "en"
  14889. }
  14890. @ARTICLE{Gold2007-tn,
  14891. title = "The neural basis of decision making",
  14892. author = "Gold, Joshua I and Shadlen, Michael N",
  14893. abstract = "The study of decision making spans such varied fields as
  14894. neuroscience, psychology, economics, statistics, political
  14895. science, and computer science. Despite this diversity of
  14896. applications, most decisions share common elements including
  14897. deliberation and commitment. Here we evaluate recent progress in
  14898. understanding how these basic elements of decision formation are
  14899. implemented in the brain. We focus on simple decisions that can
  14900. be studied in the laboratory but emphasize general principles
  14901. likely to extend to other settings.",
  14902. journal = "Annu. Rev. Neurosci.",
  14903. publisher = "annualreviews.org",
  14904. volume = 30,
  14905. pages = "535--574",
  14906. year = 2007,
  14907. language = "en"
  14908. }
  14909. @ARTICLE{Smith2004-ti,
  14910. title = "Psychology and neurobiology of simple decisions",
  14911. author = "Smith, Philip L and Ratcliff, Roger",
  14912. abstract = "Patterns of neural firing linked to eye movement decisions show
  14913. that behavioral decisions are predicted by the differential
  14914. firing rates of cells coding selected and nonselected stimulus
  14915. alternatives. These results can be interpreted using models
  14916. developed in mathematical psychology to model behavioral
  14917. decisions. Current models assume that decisions are made by
  14918. accumulating noisy stimulus information until sufficient
  14919. information for a response is obtained. Here, the models, and the
  14920. techniques used to test them against response-time distribution
  14921. and accuracy data, are described. Such models provide a
  14922. quantitative link between the time-course of behavioral decisions
  14923. and the growth of stimulus information in neural firing data.",
  14924. journal = "Trends Neurosci.",
  14925. volume = 27,
  14926. number = 3,
  14927. pages = "161--168",
  14928. month = mar,
  14929. year = 2004,
  14930. language = "en"
  14931. }
  14932. @ARTICLE{Gold2002-rg,
  14933. title = "Banburismus and the brain: decoding the relationship between
  14934. sensory stimuli, decisions, and reward",
  14935. author = "Gold, Joshua I and Shadlen, Michael N",
  14936. abstract = "This article relates a theoretical framework developed by British
  14937. codebreakers in World War II to the neural computations thought
  14938. to be responsible for forming categorical decisions about sensory
  14939. stimuli. In both, a weight of evidence is computed and
  14940. accumulated to support or oppose the alternative interpretations.
  14941. A decision is reached when the evidence reaches a threshold
  14942. value. In the codebreaking scheme, the threshold determined the
  14943. speed and accuracy of the decision process. Here we propose that
  14944. in the brain, the threshold may be controlled by neural circuits
  14945. that calculate the rate of reward.",
  14946. journal = "Neuron",
  14947. volume = 36,
  14948. number = 2,
  14949. pages = "299--308",
  14950. month = oct,
  14951. year = 2002,
  14952. language = "en"
  14953. }
  14954. @ARTICLE{Constantinople2019-hd,
  14955. title = "An Analysis of Decision under Risk in Rats",
  14956. author = "Constantinople, Christine M and Piet, Alex T and Brody, Carlos D",
  14957. abstract = "In 1979, Daniel Kahneman and Amos Tversky published a
  14958. ground-breaking paper titled ``Prospect Theory: An Analysis of
  14959. Decision under Risk,'' which presented a behavioral economic
  14960. theory that accounted for the ways in which humans deviate from
  14961. economists' normative workhorse model, Expected Utility Theory
  14962. [1, 2]. For example, people exhibit probability distortion (they
  14963. overweight low probabilities), loss aversion (losses loom larger
  14964. than gains), and reference dependence (outcomes are evaluated as
  14965. gains or losses relative to an internal reference point). We
  14966. found that rats exhibited many of these same biases, using a task
  14967. in which rats chose between guaranteed and probabilistic rewards.
  14968. However, prospect theory assumes stable preferences in the
  14969. absence of learning, an assumption at odds with alternative
  14970. frameworks such as animal learning theory and reinforcement
  14971. learning [3-7]. Rats also exhibited trial history effects,
  14972. consistent with ongoing learning. A reinforcement learning model
  14973. in which state-action values were updated by the subjective value
  14974. of outcomes according to prospect theory reproduced rats'
  14975. nonlinear utility and probability weighting functions and also
  14976. captured trial-by-trial learning dynamics.",
  14977. journal = "Curr. Biol.",
  14978. volume = 29,
  14979. number = 12,
  14980. pages = "2066--2074.e5",
  14981. month = jun,
  14982. year = 2019,
  14983. keywords = "computational model; decision-making; prospect theory; rat
  14984. behavior; reinforcement learning; reward; subjective value",
  14985. language = "en"
  14986. }
  14987. @ARTICLE{Trommershauser2003-gp,
  14988. title = "Statistical decision theory and the selection of rapid,
  14989. goal-directed movements",
  14990. author = "Trommersh{\"a}user, Julia and Maloney, Laurence T and Landy,
  14991. Michael S",
  14992. abstract = "We present two experiments that test the range of applicability
  14993. of a movement planning model (MEGaMove) based on statistical
  14994. decision theory. Subjects attempted to earn money by rapidly
  14995. touching a green target region on a computer screen while
  14996. avoiding nearby red penalty regions. In two experiments we
  14997. varied the magnitudes of penalties, the degree of overlap of
  14998. target and penalty regions, and the number of penalty regions.
  14999. Overall, subjects acted so as to maximize gain in a wide variety
  15000. of stimulus configurations, in good agreement with predictions
  15001. of the model.",
  15002. journal = "J. Opt. Soc. Am. A Opt. Image Sci. Vis.",
  15003. publisher = "osapublishing.org",
  15004. volume = 20,
  15005. number = 7,
  15006. pages = "1419--1433",
  15007. month = jul,
  15008. year = 2003,
  15009. language = "en"
  15010. }
  15011. @ARTICLE{Vilares2011-lp,
  15012. title = "Bayesian models: the structure of the world, uncertainty,
  15013. behavior, and the brain",
  15014. author = "Vilares, Iris and Kording, Konrad",
  15015. abstract = "Experiments on humans and other animals have shown that
  15016. uncertainty due to unreliable or incomplete information affects
  15017. behavior. Recent studies have formalized uncertainty and asked
  15018. which behaviors would minimize its effect. This formalization
  15019. results in a wide range of Bayesian models that derive from
  15020. assumptions about the world, and it often seems unclear how these
  15021. models relate to one another. In this review, we use the concept
  15022. of graphical models to analyze differences and commonalities
  15023. across Bayesian approaches to the modeling of behavioral and
  15024. neural data. We review behavioral and neural data associated with
  15025. each type of Bayesian model and explain how these models can be
  15026. related. We finish with an overview of different theories that
  15027. propose possible ways in which the brain can represent
  15028. uncertainty.",
  15029. journal = "Ann. N. Y. Acad. Sci.",
  15030. volume = 1224,
  15031. pages = "22--39",
  15032. month = apr,
  15033. year = 2011,
  15034. language = "en"
  15035. }
  15036. @ARTICLE{Kording2004-ii,
  15037. title = "The loss function of sensorimotor learning",
  15038. author = "K{\"o}rding, Konrad Paul and Wolpert, Daniel M",
  15039. abstract = "Motor learning can be defined as changing performance so as to
  15040. optimize some function of the task, such as accuracy. The measure
  15041. of accuracy that is optimized is called a loss function and
  15042. specifies how the CNS rates the relative success or cost of a
  15043. particular movement outcome. Models of pointing in sensorimotor
  15044. control and learning usually assume a quadratic loss function in
  15045. which the mean squared error is minimized. Here we develop a
  15046. technique for measuring the loss associated with errors. Subjects
  15047. were required to perform a task while we experimentally
  15048. controlled the skewness of the distribution of errors they
  15049. experienced. Based on the change in the subjects' average
  15050. performance, we infer the loss function. We show that people use
  15051. a loss function in which the cost increases approximately
  15052. quadratically with error for small errors and significantly less
  15053. than quadratically for large errors. The system is thus robust to
  15054. outliers. This suggests that models of sensorimotor control and
  15055. learning that have assumed minimizing squared error are a good
  15056. approximation but tend to penalize large errors excessively.",
  15057. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  15058. volume = 101,
  15059. number = 26,
  15060. pages = "9839--9842",
  15061. month = jun,
  15062. year = 2004,
  15063. language = "en"
  15064. }
  15065. @ARTICLE{Ortega_Pedro_A2013-ub,
  15066. title = "Thermodynamics as a theory of decision-making with
  15067. information-processing costs",
  15068. author = "{Ortega Pedro A.} and {Braun Daniel A.}",
  15069. journal = "Proceedings of the Royal Society A: Mathematical, Physical and
  15070. Engineering Sciences",
  15071. publisher = "Royal Society",
  15072. volume = 469,
  15073. number = 2153,
  15074. pages = "20120683",
  15075. month = may,
  15076. year = 2013
  15077. }
  15078. @ARTICLE{Kording2004-ui,
  15079. title = "A neuroeconomics approach to inferring utility functions in
  15080. sensorimotor control",
  15081. author = "K{\"o}rding, Konrad P and Fukunaga, Izumi and Howard, Ian S and
  15082. Ingram, James N and Wolpert, Daniel M",
  15083. abstract = "Making choices is a fundamental aspect of human life. For over a
  15084. century experimental economists have characterized the decisions
  15085. people make based on the concept of a utility function. This
  15086. function increases with increasing desirability of the outcome,
  15087. and people are assumed to make decisions so as to maximize
  15088. utility. When utility depends on several variables, indifference
  15089. curves arise that represent outcomes with identical utility that
  15090. are therefore equally desirable. Whereas in economics utility is
  15091. studied in terms of goods and services, the sensorimotor system
  15092. may also have utility functions defining the desirability of
  15093. various outcomes. Here, we investigate the indifference curves
  15094. when subjects experience forces of varying magnitude and
  15095. duration. Using a two-alternative forced-choice paradigm, in
  15096. which subjects chose between different magnitude-duration
  15097. profiles, we inferred the indifference curves and the utility
  15098. function. Such a utility function defines, for example, whether
  15099. subjects prefer to lift a 4-kg weight for 30 s or a 1-kg weight
  15100. for a minute. The measured utility function depends nonlinearly
  15101. on the force magnitude and duration and was remarkably conserved
  15102. across subjects. This suggests that the utility function, a
  15103. central concept in economics, may be applicable to the study of
  15104. sensorimotor control.",
  15105. journal = "PLoS Biol.",
  15106. volume = 2,
  15107. number = 10,
  15108. pages = "e330",
  15109. month = oct,
  15110. year = 2004,
  15111. language = "en"
  15112. }
  15113. @ARTICLE{Kording2006-ty,
  15114. title = "Bayesian decision theory in sensorimotor control",
  15115. author = "K{\"o}rding, Konrad P and Wolpert, Daniel M",
  15116. abstract = "Action selection is a fundamental decision process for us, and
  15117. depends on the state of both our body and the environment.
  15118. Because signals in our sensory and motor systems are corrupted by
  15119. variability or noise, the nervous system needs to estimate these
  15120. states. To select an optimal action these state estimates need to
  15121. be combined with knowledge of the potential costs or rewards of
  15122. different action outcomes. We review recent studies that have
  15123. investigated the mechanisms used by the nervous system to solve
  15124. such estimation and decision problems, which show that human
  15125. behaviour is close to that predicted by Bayesian Decision Theory.
  15126. This theory defines optimal behaviour in a world characterized by
  15127. uncertainty, and provides a coherent way of describing
  15128. sensorimotor processes.",
  15129. journal = "Trends Cogn. Sci.",
  15130. volume = 10,
  15131. number = 7,
  15132. pages = "319--326",
  15133. month = jul,
  15134. year = 2006,
  15135. language = "en"
  15136. }
  15137. @ARTICLE{McNamara2006-sr,
  15138. title = "Bayes' theorem and its applications in animal behaviour",
  15139. author = "McNamara, John M and Green, Richard F and Olsson, Ola",
  15140. abstract = "Bayesian decision theory can be used to model animal behaviour.
  15141. In this paper we give an overview of the theoretical concepts in
  15142. such models. We also review the biological contexts in which
  15143. Bayesian models have been applied, and outline some directions
  15144. where future studies would be useful. Bayesian decision theory,
  15145. when applied to animal behaviour, is based on the assumption that
  15146. the individual has some sort of ?prior opinion? of the possible
  15147. states of the world. This may, for example, be a previously
  15148. experienced distribution of qualities of food patches, or
  15149. qualities of potential mates. The animal is then assumed to be
  15150. able use sampling information to arrive at a ?posterior opinion?,
  15151. concerning e.g. the quality of a given food patch, or the average
  15152. qualities of mates in a year. A correctly formulated Bayesian
  15153. model predicts how animals may combine previous experience with
  15154. sampling information to make optimal decisions. We argue that the
  15155. assumption that animals may have ?prior opinions? is reasonable.
  15156. Their priors may come from one or both of two sources: either
  15157. from their own individual experience, gained while sampling the
  15158. environment, or from an adaptation to the environment experienced
  15159. by previous generations. This means that we should often expect
  15160. to see ?Bayesian-like? decision-making in nature.",
  15161. journal = "Oikos",
  15162. volume = 112,
  15163. number = 2,
  15164. pages = "243--251",
  15165. month = feb,
  15166. year = 2006
  15167. }
  15168. @ARTICLE{Bogacz2007-hx,
  15169. title = "Optimal decision-making theories: linking neurobiology with
  15170. behaviour",
  15171. author = "Bogacz, Rafal",
  15172. abstract = "This article reviews recently proposed theories postulating that,
  15173. during simple choices, the brain performs statistically optimal
  15174. decision making. These theories are ecologically motivated by
  15175. evolutionary pressures to optimize the speed and accuracy of
  15176. decisions and to maximize the rate of receiving rewards for
  15177. correct choices. This article suggests that the models of
  15178. decision making that are proposed on different levels of
  15179. abstraction can be linked by virtue of the same optimal
  15180. computation. Also reviewed here are recent observations that many
  15181. aspects of the circuit that involves the cortex and basal ganglia
  15182. are the same as those that are required to perform statistically
  15183. optimal choice. This review illustrates how optimal-decision
  15184. theories elucidate current data and provide experimental
  15185. predictions that concern both neurobiology and behaviour.",
  15186. journal = "Trends Cogn. Sci.",
  15187. volume = 11,
  15188. number = 3,
  15189. pages = "118--125",
  15190. month = mar,
  15191. year = 2007,
  15192. language = "en"
  15193. }
  15194. @ARTICLE{Clarke2015-kk,
  15195. title = "Human and machine learning in non-Markovian decision making",
  15196. author = "Clarke, Aaron Michael and Friedrich, Johannes and Tartaglia,
  15197. Elisa M and Marchesotti, Silvia and Senn, Walter and Herzog,
  15198. Michael H",
  15199. abstract = "Humans can learn under a wide variety of feedback conditions.
  15200. Reinforcement learning (RL), where a series of rewarded decisions
  15201. must be made, is a particularly important type of learning.
  15202. Computational and behavioral studies of RL have focused mainly on
  15203. Markovian decision processes, where the next state depends on
  15204. only the current state and action. Little is known about
  15205. non-Markovian decision making, where the next state depends on
  15206. more than the current state and action. Learning is
  15207. non-Markovian, for example, when there is no unique mapping
  15208. between actions and feedback. We have produced a model based on
  15209. spiking neurons that can handle these non-Markovian conditions by
  15210. performing policy gradient descent [1]. Here, we examine the
  15211. model's performance and compare it with human learning and a
  15212. Bayes optimal reference, which provides an upper-bound on
  15213. performance. We find that in all cases, our population of spiking
  15214. neurons model well-describes human performance.",
  15215. journal = "PLoS One",
  15216. volume = 10,
  15217. number = 4,
  15218. pages = "e0123105",
  15219. month = apr,
  15220. year = 2015,
  15221. language = "en"
  15222. }
  15223. % The entry below contains non-ASCII chars that could not be converted
  15224. % to a LaTeX equivalent.
  15225. @ARTICLE{McNamara1980-xn,
  15226. title = "The application of statistical decision theory to animal
  15227. behaviour",
  15228. author = "McNamara, J and Houston, A",
  15229. abstract = "Statistical decision theory is discussed as a general framework
  15230. for analysing how animals should learn. Attention is focused on
  15231. optimal foraging behaviour in stochastic environments. We
  15232. emphasise the distinction between the mathematical procedure
  15233. that can be used to find optimal solutions and the mechanism an
  15234. animal might use to implement such solutions. The mechanisms
  15235. might be specific to a restricted class of problems and produce
  15236. suboptimal behaviour when faced with problems outside this
  15237. class. We illustrate this point by an …",
  15238. journal = "J. Theor. Biol.",
  15239. publisher = "Elsevier",
  15240. volume = 85,
  15241. number = 4,
  15242. pages = "673--690",
  15243. month = aug,
  15244. year = 1980,
  15245. language = "en"
  15246. }
  15247. @INPROCEEDINGS{Rothkopf2011-xk,
  15248. title = "Preference Elicitation and Inverse Reinforcement Learning",
  15249. booktitle = "Machine Learning and Knowledge Discovery in Databases",
  15250. author = "Rothkopf, Constantin A and Dimitrakakis, Christos",
  15251. abstract = "We state the problem of inverse reinforcement learning in terms
  15252. of preference elicitation, resulting in a principled (Bayesian)
  15253. statistical formulation. This generalises previous work on
  15254. Bayesian inverse reinforcement learning and allows us to obtain
  15255. a posterior distribution on the agent's preferences, policy and
  15256. optionally, the obtained reward sequence, from observations. We
  15257. examine the relation of the resulting approach to other
  15258. statistical methods for inverse reinforcement learning via
  15259. analysis and experimental results. We show that preferences can
  15260. be determined accurately, even if the observed agent's policy is
  15261. sub-optimal with respect to its own preferences. In that case,
  15262. significantly improved policies with respect to the agent's
  15263. preferences are obtained, compared to both other methods and to
  15264. the performance of the demonstrated policy.",
  15265. publisher = "Springer Berlin Heidelberg",
  15266. pages = "34--48",
  15267. year = 2011
  15268. }
  15269. @ARTICLE{McFarland1977-gm,
  15270. title = "Decision making in animals",
  15271. author = "McFarland, D J",
  15272. abstract = "Animals must make decisions about when to feed, when to court,
  15273. when to sleep, and so on, in such a way as to maximise as far as
  15274. possible their chances of survival and reproductive success. It
  15275. is possible to formulate in mathematical terms the optimal
  15276. strategy for an animal to pursue. The theoretical optimum
  15277. behaviour can be compared with the actual behaviour of the
  15278. animal, and perhaps shed some light on the evolution of
  15279. behaviour.",
  15280. journal = "Nature",
  15281. volume = 269,
  15282. number = 5623,
  15283. pages = "15--21",
  15284. month = sep,
  15285. year = 1977
  15286. }
  15287. @ARTICLE{Daunizeau2010-kj,
  15288. title = "Observing the observer (I): meta-bayesian models of learning and
  15289. decision-making",
  15290. author = "Daunizeau, Jean and den Ouden, Hanneke E M and Pessiglione,
  15291. Matthias and Kiebel, Stefan J and Stephan, Klaas E and Friston,
  15292. Karl J",
  15293. abstract = "In this paper, we present a generic approach that can be used to
  15294. infer how subjects make optimal decisions under uncertainty. This
  15295. approach induces a distinction between a subject's perceptual
  15296. model, which underlies the representation of a hidden ``state of
  15297. affairs'' and a response model, which predicts the ensuing
  15298. behavioural (or neurophysiological) responses to those inputs. We
  15299. start with the premise that subjects continuously update a
  15300. probabilistic representation of the causes of their sensory
  15301. inputs to optimise their behaviour. In addition, subjects have
  15302. preferences or goals that guide decisions about actions given the
  15303. above uncertain representation of these hidden causes or state of
  15304. affairs. From a Bayesian decision theoretic perspective,
  15305. uncertain representations are so-called ``posterior'' beliefs,
  15306. which are influenced by subjective ``prior'' beliefs. Preferences
  15307. and goals are encoded through a ``loss'' (or ``utility'')
  15308. function, which measures the cost incurred by making any
  15309. admissible decision for any given (hidden) state of affair. By
  15310. assuming that subjects make optimal decisions on the basis of
  15311. updated (posterior) beliefs and utility (loss) functions, one can
  15312. evaluate the likelihood of observed behaviour. Critically, this
  15313. enables one to ``observe the observer'', i.e. identify (context-
  15314. or subject-dependent) prior beliefs and utility-functions using
  15315. psychophysical or neurophysiological measures. In this paper, we
  15316. describe the main theoretical components of this meta-Bayesian
  15317. approach (i.e. a Bayesian treatment of Bayesian decision
  15318. theoretic predictions). In a companion paper ('Observing the
  15319. observer (II): deciding when to decide'), we describe a concrete
  15320. implementation of it and demonstrate its utility by applying it
  15321. to simulated and real reaction time data from an associative
  15322. learning task.",
  15323. journal = "PLoS One",
  15324. volume = 5,
  15325. number = 12,
  15326. pages = "e15554",
  15327. month = dec,
  15328. year = 2010,
  15329. language = "en"
  15330. }
  15331. @ARTICLE{noauthor_undated-nr,
  15332. title = "Goal Inference as Inverse Planning"
  15333. }
  15334. @ARTICLE{J_Valone2006-fh,
  15335. title = "Are animals capable of Bayesian updating? An empirical review",
  15336. author = "J. Valone, Thomas",
  15337. abstract = "Numerous behavioral models assume individuals combine knowledge
  15338. in the form of a prior distribution with current sample
  15339. information using Bayesian updating to estimate the quality of
  15340. environmental parameters. I examine this assumption by reviewing
  15341. 11 empirical studies. Six studies compared observed behavior to
  15342. predictions of Bayesian and non-Bayesian models, while five
  15343. studies manipulated prior distributions directly and observed how
  15344. such manipulations altered behavior. Eight species of birds,
  15345. three mammals, one fish and one insect exhibited behavior
  15346. consistent with Bayesian updating models; one studied bird
  15347. species failed to show evidence of Bayesian updating. Most
  15348. studies examined how individuals estimated food patch quality but
  15349. two investigated mating decisions. These studies suggest a
  15350. variety of animals in different ecological contexts behave in
  15351. manners consistent with predictions of Bayesian updating models.
  15352. Future work on decision-making should focus on understanding how
  15353. animals learn prior distributions and on decision-making in
  15354. additional ecological contexts.",
  15355. journal = "Oikos",
  15356. volume = 112,
  15357. number = 2,
  15358. pages = "252--259",
  15359. month = feb,
  15360. year = 2006
  15361. }
  15362. @ARTICLE{Franks2006-ul,
  15363. title = "Not everything that counts can be counted: ants use multiple
  15364. metrics for a single nest trait",
  15365. author = "Franks, Nigel R and Dornhaus, Anna and Metherell, Bonnie G and
  15366. Nelson, Toby R and Lanfear, Sophie A J and Symes, William S",
  15367. abstract = "There are claims in the literature that certain insects can
  15368. count. We question the generality of these claims and suggest
  15369. that summation rather than counting (sensu stricto) is a more
  15370. likely explanation. We show that Temnothorax albipennis ant
  15371. colonies can discriminate between potential nest sites with
  15372. different numbers of entrances. However, our experiments suggest
  15373. that the ants use ambient light levels within the nest cavity to
  15374. assess the abundance of nest entrances rather than counting per
  15375. se. Intriguingly, Weber's Law cannot explain the ants'
  15376. inaccuracy. The ants also use a second metric, independent of
  15377. light, to assess and discriminate against wide entrances. Thus,
  15378. these ants use at least two metrics to evaluate one nest trait:
  15379. the configuration of the portals to their potential homes.",
  15380. journal = "Proc. Biol. Sci.",
  15381. publisher = "royalsocietypublishing.org",
  15382. volume = 273,
  15383. number = 1583,
  15384. pages = "165--169",
  15385. month = jan,
  15386. year = 2006,
  15387. language = "en"
  15388. }
  15389. @ARTICLE{Trimmer_Pete_C2008-ht,
  15390. title = "Mammalian choices: combining fast-but-inaccurate and
  15391. slow-but-accurate decision-making systems",
  15392. author = "{Trimmer Pete C} and {Houston Alasdair I} and {Marshall James
  15393. A.R} and {Bogacz Rafal} and {Paul Elizabeth S} and {Mendl Mike
  15394. T} and {McNamara John M}",
  15395. journal = "Proceedings of the Royal Society B: Biological Sciences",
  15396. publisher = "Royal Society",
  15397. volume = 275,
  15398. number = 1649,
  15399. pages = "2353--2361",
  15400. month = oct,
  15401. year = 2008
  15402. }
  15403. @ARTICLE{Trimmer2011-pv,
  15404. title = "Decision-making under uncertainty: biases and Bayesians",
  15405. author = "Trimmer, Pete C and Houston, Alasdair I and Marshall, James A R
  15406. and Mendl, Mike T and Paul, Elizabeth S and McNamara, John M",
  15407. abstract = "Animals (including humans) often face circumstances in which the
  15408. best choice of action is not certain. Environmental cues may be
  15409. ambiguous, and choices may be risky. This paper reviews the
  15410. theoretical side of decision-making under uncertainty,
  15411. particularly with regard to unknown risk (ambiguity). We use
  15412. simple models to show that, irrespective of pay-offs, whether it
  15413. is optimal to bias probability estimates depends upon how those
  15414. estimates have been generated. In particular, if estimates have
  15415. been calculated in a Bayesian framework with a sensible prior,
  15416. it is best to use unbiased estimates. We review the extent of
  15417. evidence for and against viewing animals (including humans) as
  15418. Bayesian decision-makers. We pay particular attention to the
  15419. Ellsberg Paradox, a classic result from experimental economics,
  15420. in which human subjects appear to deviate from optimal
  15421. decision-making by demonstrating an apparent aversion to
  15422. ambiguity in a choice between two options with equal expected
  15423. rewards. The paradox initially seems to be an example where
  15424. decision-making estimates are biased relative to the Bayesian
  15425. optimum. We discuss the extent to which the Bayesian paradigm
  15426. might be applied to the evolution of decision-makers and how the
  15427. Ellsberg Paradox may, with a deeper understanding, be resolved.",
  15428. journal = "Anim. Cogn.",
  15429. publisher = "Springer",
  15430. volume = 14,
  15431. number = 4,
  15432. pages = "465--476",
  15433. month = jul,
  15434. year = 2011,
  15435. language = "en"
  15436. }
  15437. @ARTICLE{Kording2007-va,
  15438. title = "Decision theory: what ``should'' the nervous system do?",
  15439. author = "K{\"o}rding, Konrad",
  15440. abstract = "The purpose of our nervous system is to allow us to successfully
  15441. interact with our environment. This normative idea is formalized
  15442. by decision theory that defines which choices would be most
  15443. beneficial. We live in an uncertain world, and each decision may
  15444. have many possible outcomes; choosing the best decision is thus
  15445. complicated. Bayesian decision theory formalizes these problems
  15446. in the presence of uncertainty and often provides compact models
  15447. that predict observed behavior. With its elegant formalization of
  15448. the problems faced by the nervous system, it promises to become a
  15449. major inspiration for studies in neuroscience.",
  15450. journal = "Science",
  15451. volume = 318,
  15452. number = 5850,
  15453. pages = "606--610",
  15454. month = oct,
  15455. year = 2007,
  15456. language = "en"
  15457. }
  15458. @ARTICLE{Seth_Anil_K2007-vj,
  15459. title = "The ecology of action selection: insights from artificial life",
  15460. author = "{Seth Anil K}",
  15461. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  15462. publisher = "Royal Society",
  15463. volume = 362,
  15464. number = 1485,
  15465. pages = "1545--1558",
  15466. month = sep,
  15467. year = 2007
  15468. }
  15469. @ARTICLE{Bogacz_Rafal2007-kq,
  15470. title = "Extending a biologically inspired model of choice:
  15471. multi-alternatives, nonlinearity and value-based
  15472. multidimensional choice",
  15473. author = "{Bogacz Rafal} and {Usher Marius} and {Zhang Jiaxiang} and
  15474. {McClelland James L}",
  15475. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  15476. publisher = "Royal Society",
  15477. volume = 362,
  15478. number = 1485,
  15479. pages = "1655--1670",
  15480. month = sep,
  15481. year = 2007
  15482. }
  15483. @INPROCEEDINGS{Gordon2004-gu,
  15484. title = "Evolving sparse direction maps for maze pathfinding",
  15485. booktitle = "Proceedings of the 2004 Congress on Evolutionary Computation
  15486. ({IEEE} Cat. {No.04TH8753})",
  15487. author = "Gordon, V S and Matley, Z",
  15488. abstract = "A genetic algorithm is used to solve a class of maze pathfinding
  15489. problems. In particular, we find a complete set of paths
  15490. directing an agent from any position in the maze towards a
  15491. single goal. To this end, we define a sparse direction map,
  15492. wherein the maze is divided into sectors, each of which contains
  15493. a direction indicator. Maps are evolved using a simple genetic
  15494. algorithm. The fitness function samples the efficacy of the map
  15495. from random starting points, this estimating the likelihood that
  15496. agents find the goal. The framework was effective in evolving
  15497. successful maps for three different mazes of varying size and
  15498. complexity, resulting in interesting and lifelike agent behavior
  15499. suitable for games, but not always the shortest paths.",
  15500. volume = 1,
  15501. pages = "835--838 Vol.1",
  15502. month = jun,
  15503. year = 2004,
  15504. keywords = "genetic algorithms;path planning;game theory;graph theory;sparse
  15505. direction maps;maze pathfinding;genetic algorithm;fitness
  15506. function;shortest paths;direction indicator;agent
  15507. behavior;Genetic algorithms;Filling;Counting circuits;Computer
  15508. science;Delay"
  15509. }
  15510. @ARTICLE{Jaafar2008-us,
  15511. title = "A Fuzzy Action Selection Method for Virtual Agent Navigation in
  15512. Unknown Virtual Environments",
  15513. author = "Jaafar, Jafreezal and Mc Kenzie, Eric",
  15514. journal = "Plan. Perspect.",
  15515. volume = 144,
  15516. pages = "154",
  15517. year = 2008
  15518. }
  15519. @ARTICLE{Caillerie_undated-bp,
  15520. title = "Geodesic trail formation in a two-dimensional model of foraging
  15521. ants with directed pheromones",
  15522. author = "Caillerie, Nils"
  15523. }
  15524. @ARTICLE{Deneubourg1989-ff,
  15525. title = "Collective patterns and decision-making",
  15526. author = "Deneubourg, J L and Goss, S",
  15527. abstract = "Autocatalytic interactions between the members of an animal
  15528. group or society, and particularly chemically or visually
  15529. mediated allelomimesis, can be an important factor in the
  15530. organisation of their collective activity. Furthermore, the
  15531. interactions between the individuals and the environment allow
  15532. different collective patterns and decisions to appear under
  15533. different conditions, with the same individual behaviour. While
  15534. most clearly demonstrable in social insects, these principles
  15535. are fundamental to schools of fishes, flocks of birds, groups of
  15536. mammals, and many other social aggregates. The analysis of
  15537. collective behaviour in these terms implies detailed observation
  15538. of both individual and collective behaviour, combined with
  15539. mathematical modelling to link the two.",
  15540. journal = "Ethol. Ecol. Evol.",
  15541. publisher = "Taylor \& Francis",
  15542. volume = 1,
  15543. number = 4,
  15544. pages = "295--311",
  15545. month = dec,
  15546. year = 1989
  15547. }
  15548. @MISC{Anil_K_Seth_Tony_J_Prescott_Joanna_J_Bryson_undated-eg,
  15549. title = "Modelling Natural Action Selection",
  15550. author = "{Anil K. Seth, Tony J. Prescott, Joanna J. Bryson}",
  15551. keywords = "books;Books"
  15552. }
  15553. @ARTICLE{Maisto_Domenico2015-mt,
  15554. title = "Divide et impera: subgoaling reduces the complexity of
  15555. probabilistic inference and problem solving",
  15556. author = "{Maisto Domenico} and {Donnarumma Francesco} and {Pezzulo
  15557. Giovanni}",
  15558. journal = "J. R. Soc. Interface",
  15559. publisher = "Royal Society",
  15560. volume = 12,
  15561. number = 104,
  15562. pages = "20141335",
  15563. month = mar,
  15564. year = 2015
  15565. }
  15566. @ARTICLE{Russek2017-vt,
  15567. title = "Predictive representations can link model-based reinforcement
  15568. learning to model-free mechanisms",
  15569. author = "Russek, Evan M and Momennejad, Ida and Botvinick, Matthew M and
  15570. Gershman, Samuel J and Daw, Nathaniel D",
  15571. abstract = "Humans and animals are capable of evaluating actions by
  15572. considering their long-run future rewards through a process
  15573. described using model-based reinforcement learning (RL)
  15574. algorithms. The mechanisms by which neural circuits perform the
  15575. computations prescribed by model-based RL remain largely unknown;
  15576. however, multiple lines of evidence suggest that neural circuits
  15577. supporting model-based behavior are structurally homologous to
  15578. and overlapping with those thought to carry out model-free
  15579. temporal difference (TD) learning. Here, we lay out a family of
  15580. approaches by which model-based computation may be built upon a
  15581. core of TD learning. The foundation of this framework is the
  15582. successor representation, a predictive state representation that,
  15583. when combined with TD learning of value predictions, can produce
  15584. a subset of the behaviors associated with model-based learning,
  15585. while requiring less decision-time computation than dynamic
  15586. programming. Using simulations, we delineate the precise
  15587. behavioral capabilities enabled by evaluating actions using this
  15588. approach, and compare them to those demonstrated by biological
  15589. organisms. We then introduce two new algorithms that build upon
  15590. the successor representation while progressively mitigating its
  15591. limitations. Because this framework can account for the full
  15592. range of observed putatively model-based behaviors while still
  15593. utilizing a core TD framework, we suggest that it represents a
  15594. neurally plausible family of mechanisms for model-based
  15595. evaluation.",
  15596. journal = "PLoS Comput. Biol.",
  15597. volume = 13,
  15598. number = 9,
  15599. pages = "e1005768",
  15600. month = sep,
  15601. year = 2017,
  15602. language = "en"
  15603. }
  15604. @ARTICLE{Daw2014-uh,
  15605. title = "The algorithmic anatomy of model-based evaluation",
  15606. author = "Daw, Nathaniel D and Dayan, Peter",
  15607. abstract = "Despite many debates in the first half of the twentieth century,
  15608. it is now largely a truism that humans and other animals build
  15609. models of their environments and use them for prediction and
  15610. control. However, model-based (MB) reasoning presents severe
  15611. computational challenges. Alternative, computationally simpler,
  15612. model-free (MF) schemes have been suggested in the reinforcement
  15613. learning literature, and have afforded influential accounts of
  15614. behavioural and neural data. Here, we study the realization of MB
  15615. calculations, and the ways that this might be woven together with
  15616. MF values and evaluation methods. There are as yet mostly only
  15617. hints in the literature as to the resulting tapestry, so we offer
  15618. more preview than review.",
  15619. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  15620. volume = 369,
  15621. number = 1655,
  15622. month = nov,
  15623. year = 2014,
  15624. keywords = "Monte Carlo tree search; model-based reasoning; model-free
  15625. reasoning; orbitofrontal cortex; reinforcement learning; striatum",
  15626. language = "en"
  15627. }
  15628. @ARTICLE{Friston2017-ib,
  15629. title = "Active Inference: A Process Theory",
  15630. author = "Friston, Karl and FitzGerald, Thomas and Rigoli, Francesco and
  15631. Schwartenbeck, Philipp and Pezzulo, Giovanni",
  15632. abstract = "This article describes a process theory based on active inference
  15633. and belief propagation. Starting from the premise that all
  15634. neuronal processing (and action selection) can be explained by
  15635. maximizing Bayesian model evidence-or minimizing variational free
  15636. energy-we ask whether neuronal responses can be described as a
  15637. gradient descent on variational free energy. Using a standard
  15638. (Markov decision process) generative model, we derive the
  15639. neuronal dynamics implicit in this description and reproduce a
  15640. remarkable range of well-characterized neuronal phenomena. These
  15641. include repetition suppression, mismatch negativity, violation
  15642. responses, place-cell activity, phase precession, theta
  15643. sequences, theta-gamma coupling, evidence accumulation,
  15644. race-to-bound dynamics, and transfer of dopamine responses.
  15645. Furthermore, the (approximately Bayes' optimal) behavior
  15646. prescribed by these dynamics has a degree of face validity,
  15647. providing a formal explanation for reward seeking, context
  15648. learning, and epistemic foraging. Technically, the fact that a
  15649. gradient descent appears to be a valid description of neuronal
  15650. activity means that variational free energy is a Lyapunov
  15651. function for neuronal dynamics, which therefore conform to
  15652. Hamilton's principle of least action.",
  15653. journal = "Neural Comput.",
  15654. volume = 29,
  15655. number = 1,
  15656. pages = "1--49",
  15657. month = jan,
  15658. year = 2017,
  15659. language = "en"
  15660. }
  15661. @UNPUBLISHED{Miller2018-ag,
  15662. title = "Habits without Values",
  15663. author = "Miller, Kevin and Shenhav, Amitai and Ludvig, Elliot",
  15664. abstract = "Habits form a crucial component of behavior. In recent years, key
  15665. computational models have conceptualized habits as arising from
  15666. model-free reinforcement learning (RL) mechanisms, which
  15667. typically select between available actions based on the future
  15668. value expected to result from each. Traditionally, however,
  15669. habits have been understood as behaviors that can be triggered
  15670. directly by a stimulus, without requiring the animal to evaluate
  15671. expected outcomes. Here, we develop a computational model
  15672. instantiating this traditional view, in which habits develop
  15673. through the direct strengthening of recently taken actions rather
  15674. than through the encoding of outcomes. We demonstrate that this
  15675. model accounts for key behavioral manifestations of habits,
  15676. including insensitivity to outcome devaluation and contingency
  15677. degradation, as well as the effects of reinforcement schedule on
  15678. the rate of habit formation. The model also explains the
  15679. prevalent observation of perseveration in repeated-choice tasks
  15680. as an additional behavioral manifestation of the habit system. We
  15681. suggest that mapping habitual behaviors onto value-free
  15682. mechanisms provides a parsimonious account of existing behavioral
  15683. and neural data. This mapping may provide a new foundation for
  15684. building robust and comprehensive models of the interaction of
  15685. habits with other, more goal-directed types of behaviors and help
  15686. to better guide research into the neural mechanisms underlying
  15687. control of instrumental behavior more generally.",
  15688. journal = "bioRxiv",
  15689. pages = "067603",
  15690. month = mar,
  15691. year = 2018,
  15692. language = "en"
  15693. }
  15694. @ARTICLE{Doll2012-qb,
  15695. title = "The ubiquity of model-based reinforcement learning",
  15696. author = "Doll, Bradley B and Simon, Dylan A and Daw, Nathaniel D",
  15697. abstract = "The reward prediction error (RPE) theory of dopamine (DA)
  15698. function has enjoyed great success in the neuroscience of
  15699. learning and decision-making. This theory is derived from
  15700. model-free reinforcement learning (RL), in which choices are made
  15701. simply on the basis of previously realized rewards. Recently,
  15702. attention has turned to correlates of more flexible, albeit
  15703. computationally complex, model-based methods in the brain. These
  15704. methods are distinguished from model-free learning by their
  15705. evaluation of candidate actions using expected future outcomes
  15706. according to a world model. Puzzlingly, signatures from these
  15707. computations seem to be pervasive in the very same regions
  15708. previously thought to support model-free learning. Here, we
  15709. review recent behavioral and neural evidence about these two
  15710. systems, in attempt to reconcile their enigmatic cohabitation in
  15711. the brain.",
  15712. journal = "Curr. Opin. Neurobiol.",
  15713. volume = 22,
  15714. number = 6,
  15715. pages = "1075--1081",
  15716. month = dec,
  15717. year = 2012,
  15718. language = "en"
  15719. }
  15720. @ARTICLE{Miller_undated-ho,
  15721. title = "Re-aligning models of habitual and goal-directed decision-making",
  15722. author = "Miller, Kevin and Ludvig, Elliot A and Pezzulo, Giovanni and
  15723. Shenhav, Amitai"
  15724. }
  15725. @ARTICLE{Forstmeier2011-tz,
  15726. title = "Cryptic multiple hypotheses testing in linear models:
  15727. overestimated effect sizes and the winner's curse",
  15728. author = "Forstmeier, Wolfgang and Schielzeth, Holger",
  15729. abstract = "Fitting generalised linear models (GLMs) with more than one
  15730. predictor has become the standard method of analysis in
  15731. evolutionary and behavioural research. Often, GLMs are used for
  15732. exploratory data analysis, where one starts with a complex full
  15733. model including interaction terms and then simplifies by
  15734. removing non-significant terms. While this approach can be
  15735. useful, it is problematic if significant effects are interpreted
  15736. as if they arose from a single a priori hypothesis test. This is
  15737. because model selection involves cryptic multiple hypothesis
  15738. testing, a fact that has only rarely been acknowledged or
  15739. quantified. We show that the probability of finding at least one
  15740. 'significant' effect is high, even if all null hypotheses are
  15741. true (e.g. 40\% when starting with four predictors and their
  15742. two-way interactions). This probability is close to theoretical
  15743. expectations when the sample size (N) is large relative to the
  15744. number of predictors including interactions (k). In contrast,
  15745. type I error rates strongly exceed even those expectations when
  15746. model simplification is applied to models that are over-fitted
  15747. before simplification (low N/k ratio). The increase in
  15748. false-positive results arises primarily from an overestimation
  15749. of effect sizes among significant predictors, leading to
  15750. upward-biased effect sizes that often cannot be reproduced in
  15751. follow-up studies ('the winner's curse'). Despite having their
  15752. own problems, full model tests and P value adjustments can be
  15753. used as a guide to how frequently type I errors arise by
  15754. sampling variation alone. We favour the presentation of full
  15755. models, since they best reflect the range of predictors
  15756. investigated and ensure a balanced representation also of
  15757. non-significant results.",
  15758. journal = "Behav. Ecol. Sociobiol.",
  15759. publisher = "Springer",
  15760. volume = 65,
  15761. number = 1,
  15762. pages = "47--55",
  15763. month = jan,
  15764. year = 2011,
  15765. language = "en"
  15766. }
  15767. @ARTICLE{Bolker2009-az,
  15768. title = "Generalized linear mixed models: a practical guide for ecology
  15769. and evolution",
  15770. author = "Bolker, Benjamin M and Brooks, Mollie E and Clark, Connie J and
  15771. Geange, Shane W and Poulsen, John R and Stevens, M Henry H and
  15772. White, Jada-Simone S",
  15773. abstract = "How should ecologists and evolutionary biologists analyze
  15774. nonnormal data that involve random effects? Nonnormal data such
  15775. as counts or proportions often defy classical statistical
  15776. procedures. Generalized linear mixed models (GLMMs) provide a
  15777. more flexible approach for analyzing nonnormal data when random
  15778. effects are present. The explosion of research on GLMMs in the
  15779. last decade has generated considerable uncertainty for
  15780. practitioners in ecology and evolution. Despite the availability
  15781. of accurate techniques for estimating GLMM parameters in simple
  15782. cases, complex GLMMs are challenging to fit and statistical
  15783. inference such as hypothesis testing remains difficult. We review
  15784. the use (and misuse) of GLMMs in ecology and evolution, discuss
  15785. estimation and inference and summarize 'best-practice' data
  15786. analysis procedures for scientists facing this challenge.",
  15787. journal = "Trends Ecol. Evol.",
  15788. volume = 24,
  15789. number = 3,
  15790. pages = "127--135",
  15791. month = mar,
  15792. year = 2009,
  15793. language = "en"
  15794. }
  15795. @ARTICLE{Czaczkes2015-tx,
  15796. title = "Trail pheromones: an integrative view of their role in social
  15797. insect colony organization",
  15798. author = "Czaczkes, Tomer J and Gr{\"u}ter, Christoph and Ratnieks, Francis
  15799. L W",
  15800. abstract = "Trail pheromones do more than simply guide social insect workers
  15801. from point A to point B. Recent research has revealed additional
  15802. ways in which they help to regulate colony foraging, often via
  15803. positive and negative feedback processes that influence the
  15804. exploitation of the different resources that a colony has
  15805. knowledge of. Trail pheromones are often complementary or
  15806. synergistic with other information sources, such as individual
  15807. memory. Pheromone trails can be composed of two or more
  15808. pheromones with different functions, and information may be
  15809. embedded in the trail network geometry. These findings indicate
  15810. remarkable sophistication in how trail pheromones are used to
  15811. regulate colony-level behavior, and how trail pheromones are used
  15812. and deployed at the individual level.",
  15813. journal = "Annu. Rev. Entomol.",
  15814. volume = 60,
  15815. pages = "581--599",
  15816. month = jan,
  15817. year = 2015,
  15818. keywords = "ants; complex adaptive systems; complexity; organization;
  15819. recruitment; review",
  15820. language = "en"
  15821. }
  15822. @ARTICLE{Ramsch2012-so,
  15823. title = "A mathematical model of foraging in a dynamic environment by
  15824. trail-laying Argentine ants",
  15825. author = "Ramsch, Kai and Reid, Chris R and Beekman, Madeleine and
  15826. Middendorf, Martin",
  15827. abstract = "Ants live in dynamically changing environments, where food
  15828. sources become depleted and alternative sources appear. Yet most
  15829. mathematical models of ant foraging assume that the ants'
  15830. foraging environment is static. Here we describe a mathematical
  15831. model of ant foraging in a dynamic environment. Our model
  15832. attempts to explain recent empirical data on dynamic foraging in
  15833. the Argentine ant Linepithema humile (Mayr). The ants are able to
  15834. find the shortest path in a Towers of Hanoi maze, a complex
  15835. network containing 32,768 alternative paths, even when the maze
  15836. is altered dynamically. We modify existing models developed to
  15837. explain ant foraging in static environments, to elucidate what
  15838. possible mechanisms allow the ants to quickly adapt to changes in
  15839. their foraging environment. Our results suggest that navigation
  15840. of individual ants based on a combination of one pheromone
  15841. deposited during foraging and directional information enables the
  15842. ants to adapt their foraging trails and recreates the
  15843. experimental results.",
  15844. journal = "J. Theor. Biol.",
  15845. volume = 306,
  15846. pages = "32--45",
  15847. month = aug,
  15848. year = 2012,
  15849. language = "en"
  15850. }
  15851. @ARTICLE{Holcombe2012-qk,
  15852. title = "Modelling complex biological systems using an agent-based
  15853. approach",
  15854. author = "Holcombe, Mike and Adra, Salem and Bicak, Mesude and Chin, Shawn
  15855. and Coakley, Simon and Graham, Alison I and Green, Jeffrey and
  15856. Greenough, Chris and Jackson, Duncan and Kiran, Mariam and
  15857. MacNeil, Sheila and Maleki-Dizaji, Afsaneh and McMinn, Phil and
  15858. Pogson, Mark and Poole, Robert and Qwarnstrom, Eva and Ratnieks,
  15859. Francis and Rolfe, Matthew D and Smallwood, Rod and Sun, Tao and
  15860. Worth, David",
  15861. abstract = "Many of the complex systems found in biology are comprised of
  15862. numerous components, where interactions between individual agents
  15863. result in the emergence of structures and function, typically in
  15864. a highly dynamic manner. Often these entities have limited
  15865. lifetimes but their interactions both with each other and their
  15866. environment can have profound biological consequences. We will
  15867. demonstrate how modelling these entities, and their interactions,
  15868. can lead to a new approach to experimental biology bringing new
  15869. insights and a deeper understanding of biological systems.",
  15870. journal = "Integr. Biol.",
  15871. volume = 4,
  15872. number = 1,
  15873. pages = "53--64",
  15874. month = jan,
  15875. year = 2012,
  15876. language = "en"
  15877. }
  15878. @ARTICLE{Czaczkes2013-tr,
  15879. title = "Ant foraging on complex trails: route learning and the role of
  15880. trail pheromones in Lasius niger",
  15881. author = "Czaczkes, Tomer J and Gr{\"u}ter, Christoph and Ellis, Laura and
  15882. Wood, Elizabeth and Ratnieks, Francis L W",
  15883. abstract = "Ants are central place foragers and use multiple information
  15884. sources to navigate between the nest and feeding sites.
  15885. Individual ants rapidly learn a route, and often prioritize
  15886. memory over pheromone trails when tested on a simple trail with a
  15887. single bifurcation. However, in nature, ants often forage at
  15888. locations that are reached via more complex routes with multiple
  15889. trail bifurcations. Such routes may be more difficult to learn,
  15890. and thus ants would benefit from additional information. We
  15891. hypothesized that trail pheromones play a more significant role
  15892. in ant foraging on complex routes, either by assisting in
  15893. navigation or route learning or both. We studied Lasius niger
  15894. workers foraging on a doubly bifurcating trail with four end
  15895. points. Route learning was slower and errors greater on
  15896. alternating (e.g. left-right) versus repeating routes (e.g.
  15897. left-left), with error rates of 32 and 3\%, respectively.
  15898. However, errors on alternating routes decreased by 30\% when
  15899. trail pheromone was present. Trail pheromones also aid route
  15900. learning, leading to reduced errors in subsequent journeys
  15901. without pheromone. If an experienced forager makes an error when
  15902. returning to a food source, it reacts by increasing pheromone
  15903. deposition on the return journey. In addition, high levels of
  15904. trail pheromone suppress further pheromone deposition. This
  15905. negative feedback mechanism may act to conserve pheromone or to
  15906. regulate recruitment. Taken together, these results demonstrate
  15907. further complexity and sophistication in the foraging system of
  15908. ant colonies, especially in the role of trail pheromones and
  15909. their relationship with learning and the use of private
  15910. information (memory) in a complex environment.",
  15911. journal = "J. Exp. Biol.",
  15912. volume = 216,
  15913. number = "Pt 2",
  15914. pages = "188--197",
  15915. month = jan,
  15916. year = 2013,
  15917. language = "en"
  15918. }
  15919. @ARTICLE{Vittori2006-qd,
  15920. title = "Path efficiency of ant foraging trails in an artificial network",
  15921. author = "Vittori, Karla and Talbot, Gr{\'e}goire and Gautrais, Jacques and
  15922. Fourcassi{\'e}, Vincent and Ara{\'u}jo, Aluizio F R and
  15923. Theraulaz, Guy",
  15924. abstract = "In this paper we present an individual-based model describing the
  15925. foraging behavior of ants moving in an artificial network of
  15926. tunnels in which several interconnected paths can be used to
  15927. reach a single food source. Ants lay a trail pheromone while
  15928. moving in the network and this pheromone acts as a system of mass
  15929. recruitment that attracts other ants in the network. The rules
  15930. implemented in the model are based on measures of the decisions
  15931. taken by ants at tunnel bifurcations during real experiments. The
  15932. collective choice of the ants is estimated by measuring their
  15933. probability to take a given path in the network. Overall, we
  15934. found a good agreement between the results of the simulations and
  15935. those of the experiments, showing that simple behavioral rules
  15936. can lead ants to find the shortest paths in the network. The
  15937. match between the experiments and the model, however, was better
  15938. for nestbound than for outbound ants. A sensitivity study of the
  15939. model suggests that the bias observed in the choice of the ants
  15940. at asymmetrical bifurcations is a key behavior to reproduce the
  15941. collective choice observed in the experiments.",
  15942. journal = "J. Theor. Biol.",
  15943. volume = 239,
  15944. number = 4,
  15945. pages = "507--515",
  15946. month = apr,
  15947. year = 2006,
  15948. language = "en"
  15949. }
  15950. % The entry below contains non-ASCII chars that could not be converted
  15951. % to a LaTeX equivalent.
  15952. @ARTICLE{Forster2014-ms,
  15953. title = "Effect of Trail Bifurcation Asymmetry and Pheromone Presence or
  15954. Absence on Trail Choice by Lasius niger Ants",
  15955. author = "Forster, Antonia and Czaczkes, Tomer J and Warner, Emma and
  15956. Woodall, Tom and Martin, Emily and Ratnieks, Francis L W and
  15957. Herberstein, M",
  15958. abstract = "During foraging, ant workers are known to make use of multiple
  15959. information sources, such as private information (personal
  15960. memory) and social information (trail pheromones). Environmental
  15961. effects on foraging, and how these interact with other
  15962. information sources, have, however, been little studied. One
  15963. environmental effect is trail bifurcation asymmetry. Ants forage
  15964. on branching trail networks and must often decide which branch to
  15965. take at a junction (bifurcation). This is an important decision,
  15966. as finding food sources relies on making the correct choices at
  15967. bifurcations. Bifurcation angle may provide important information
  15968. when making this choice. We used a Y-maze with a pivoting 90°
  15969. bifurcation to study trail choice of Lasius niger foragers at
  15970. varying branch asymmetries (0°, [both branches 45° from straight
  15971. ahead], 30° [branches at 30° and 60° from straight ahead], 45°,
  15972. 60° and 90° [one branch straight ahead, the other at 90°]). The
  15973. experiment was carried out either with equal amounts of trail
  15974. pheromone on both branches of the bifurcation or with pheromone
  15975. present on only one branch. Our results show that with equal
  15976. pheromone, trail asymmetry has a significant effect on trail
  15977. choice. Ants preferentially follow the branch deviating least
  15978. from straight, and this effect increases as asymmetry increases
  15979. (47\% at 0°, 54\% at 30°, 57\% at 45°, 66\% at 60° and 73\% at
  15980. 90°). However, when pheromone is only present on one branch, the
  15981. graded effect of asymmetry disappears. Overall, however, there is
  15982. an effect of asymmetry as the preference of ants for the
  15983. pheromone-marked branch over the unmarked branch is reduced from
  15984. 65\%, when it is the less deviating branch, to 53\%, when it is
  15985. the more deviating branch. These results demonstrate that trail
  15986. asymmetry influences ant decision-making at bifurcations and that
  15987. this information interacts with trail pheromone presence in a
  15988. non-hierarchical manner.",
  15989. journal = "Ethology",
  15990. volume = 120,
  15991. number = 8,
  15992. pages = "768--775",
  15993. month = aug,
  15994. year = 2014,
  15995. keywords = "Lasius niger; asymmetry; environmental effects; foraging;
  15996. pheromone; trail choice",
  15997. language = "en"
  15998. }
  15999. @ARTICLE{Garnier2013-cv,
  16000. title = "Do ants need to estimate the geometrical properties of trail
  16001. bifurcations to find an efficient route? A swarm robotics test
  16002. bed",
  16003. author = "Garnier, Simon and Combe, Maud and Jost, Christian and Theraulaz,
  16004. Guy",
  16005. abstract = "Interactions between individuals and the structure of their
  16006. environment play a crucial role in shaping self-organized
  16007. collective behaviors. Recent studies have shown that ants
  16008. crossing asymmetrical bifurcations in a network of galleries tend
  16009. to follow the branch that deviates the least from their incoming
  16010. direction. At the collective level, the combination of this
  16011. tendency and the pheromone-based recruitment results in a greater
  16012. likelihood of selecting the shortest path between the colony's
  16013. nest and a food source in a network containing asymmetrical
  16014. bifurcations. It was not clear however what the origin of this
  16015. behavioral bias is. Here we propose that it results from a simple
  16016. interaction between the behavior of the ants and the geometry of
  16017. the network, and that it does not require the ability to measure
  16018. the angle of the bifurcation. We tested this hypothesis using
  16019. groups of ant-like robots whose perceptual and cognitive
  16020. abilities can be fully specified. We programmed them only to lay
  16021. down and follow light trails, avoid obstacles and move according
  16022. to a correlated random walk, but not to use more sophisticated
  16023. orientation methods. We recorded the behavior of the robots in
  16024. networks of galleries presenting either only symmetrical
  16025. bifurcations or a combination of symmetrical and asymmetrical
  16026. bifurcations. Individual robots displayed the same pattern of
  16027. branch choice as individual ants when crossing a bifurcation,
  16028. suggesting that ants do not actually measure the geometry of the
  16029. bifurcations when travelling along a pheromone trail. Finally at
  16030. the collective level, the group of robots was more likely to
  16031. select one of the possible shorter paths between two designated
  16032. areas when moving in an asymmetrical network, as observed in
  16033. ants. This study reveals the importance of the shape of trail
  16034. networks for foraging in ants and emphasizes the underestimated
  16035. role of the geometrical properties of transportation networks in
  16036. general.",
  16037. journal = "PLoS Comput. Biol.",
  16038. volume = 9,
  16039. number = 3,
  16040. pages = "e1002903",
  16041. month = mar,
  16042. year = 2013,
  16043. language = "en"
  16044. }
  16045. @ARTICLE{Sych2019-ds,
  16046. title = "High-density multi-fiber photometry for studying large-scale
  16047. brain circuit dynamics",
  16048. author = "Sych, Yaroslav and Chernysheva, Maria and Sumanovski, Lazar T and
  16049. Helmchen, Fritjof",
  16050. abstract = "Animal behavior originates from neuronal activity distributed
  16051. across brain-wide networks. However, techniques available to
  16052. assess large-scale neural dynamics in behaving animals remain
  16053. limited. Here we present compact, chronically implantable,
  16054. high-density arrays of optical fibers that enable multi-fiber
  16055. photometry and optogenetic perturbations across many regions in
  16056. the mammalian brain. In mice engaged in a texture discrimination
  16057. task, we achieved simultaneous photometric calcium recordings
  16058. from networks of 12-48 brain regions, including striatal,
  16059. thalamic, hippocampal and cortical areas. Furthermore, we
  16060. optically perturbed subsets of regions in VGAT-ChR2 mice by
  16061. targeting specific fiber channels with a spatial light modulator.
  16062. Perturbation of ventral thalamic nuclei caused distributed
  16063. network modulation and behavioral deficits. Finally, we
  16064. demonstrate multi-fiber photometry in freely moving animals,
  16065. including simultaneous recordings from two mice during social
  16066. interaction. High-density multi-fiber arrays are versatile tools
  16067. for the investigation of large-scale brain dynamics during
  16068. behavior.",
  16069. journal = "Nat. Methods",
  16070. month = may,
  16071. year = 2019,
  16072. language = "en"
  16073. }
  16074. @ARTICLE{Pezzulo2019-dg,
  16075. title = "Planning at decision time and in the background during spatial
  16076. navigation",
  16077. author = "Pezzulo, Giovanni and Donnarumma, Francesco and Maisto, Domenico
  16078. and Stoianov, Ivilin",
  16079. abstract = "Planning is the model-based approach to solving control problems.
  16080. The hallmark of planning is the endogenous generation of
  16081. dynamical representations of future states, like goal locations,
  16082. or state sequences, like trajectories to the goal location, using
  16083. an internal model of the task. We review recent evidence of
  16084. model-based planning processes and the representation of future
  16085. goal states in the brain of rodents and humans engaged in spatial
  16086. navigation tasks. We highlight two distinct but complementary
  16087. usages of planning as identified in artificial intelligence: `at
  16088. decision time', to support goal-directed choices and sequential
  16089. memory encoding, and `in the background', to learn behavioral
  16090. policies and to optimize internal models. We discuss how two
  16091. kinds of internally generated sequences in the hippocampus --
  16092. theta and SWR sequences -- might participate in the neuronal
  16093. implementation of these two planning modes, thus supporting a
  16094. flexible model-based system for adaptive cognition and action.",
  16095. journal = "Current Opinion in Behavioral Sciences",
  16096. volume = 29,
  16097. pages = "69--76",
  16098. month = oct,
  16099. year = 2019
  16100. }
  16101. @ARTICLE{Blum2005-ut,
  16102. title = "Ant colony optimization: Introduction and recent trends",
  16103. author = "Blum, Christian",
  16104. abstract = "Ant colony optimization is a technique for optimization that was
  16105. introduced in the early 1990's. The inspiring source of ant
  16106. colony optimization is the foraging behavior of real ant
  16107. colonies. This behavior is exploited in artificial ant colonies
  16108. for the search of approximate solutions to discrete optimization
  16109. problems, to continuous optimization problems, and to important
  16110. problems in telecommunications, such as routing and load
  16111. balancing. First, we deal with the biological inspiration of ant
  16112. colony optimization algorithms. We show how this biological
  16113. inspiration can be transfered into an algorithm for discrete
  16114. optimization. Then, we outline ant colony optimization in more
  16115. general terms in the context of discrete optimization, and
  16116. present some of the nowadays best-performing ant colony
  16117. optimization variants. After summarizing some important
  16118. theoretical results, we demonstrate how ant colony optimization
  16119. can be applied to continuous optimization problems. Finally, we
  16120. provide examples of an interesting recent research direction: The
  16121. hybridization with more classical techniques from artificial
  16122. intelligence and operations research.",
  16123. journal = "Phys. Life Rev.",
  16124. volume = 2,
  16125. number = 4,
  16126. pages = "353--373",
  16127. month = dec,
  16128. year = 2005,
  16129. keywords = "Ant colony optimization; Discrete optimization; Hybridization"
  16130. }
  16131. @ARTICLE{Gronenberg2008-rj,
  16132. title = "Structure and function of ant (Hymenoptera: Formicidae) brains:
  16133. strength in numbers",
  16134. author = "Gronenberg, Wulfila",
  16135. journal = "Myrmecol. News",
  16136. volume = 11,
  16137. pages = "25--36",
  16138. year = 2008
  16139. }
  16140. @ARTICLE{Daw2011-jx,
  16141. title = "Model-based influences on humans' choices and striatal
  16142. prediction errors",
  16143. author = "Daw, Nathaniel D and Gershman, Samuel J and Seymour, Ben and
  16144. Dayan, Peter and Dolan, Raymond J",
  16145. abstract = "The mesostriatal dopamine system is prominently implicated in
  16146. model-free reinforcement learning, with fMRI BOLD signals in
  16147. ventral striatum notably covarying with model-free prediction
  16148. errors. However, latent learning and devaluation studies show
  16149. that behavior also shows hallmarks of model-based planning, and
  16150. the interaction between model-based and model-free values,
  16151. prediction errors, and preferences is underexplored. We designed
  16152. a multistep decision task in which model-based and model-free
  16153. influences on human choice behavior could be distinguished. By
  16154. showing that choices reflected both influences we could then
  16155. test the purity of the ventral striatal BOLD signal as a
  16156. model-free report. Contrary to expectations, the signal
  16157. reflected both model-free and model-based predictions in
  16158. proportions matching those that best explained choice behavior.
  16159. These results challenge the notion of a separate model-free
  16160. learner and suggest a more integrated computational architecture
  16161. for high-level human decision-making.",
  16162. journal = "Neuron",
  16163. publisher = "Elsevier",
  16164. volume = 69,
  16165. number = 6,
  16166. pages = "1204--1215",
  16167. month = mar,
  16168. year = 2011,
  16169. language = "en"
  16170. }
  16171. @ARTICLE{Gao2015-ik,
  16172. title = "On simplicity and complexity in the brave new world of
  16173. large-scale neuroscience",
  16174. author = "Gao, Peiran and Ganguli, Surya",
  16175. abstract = "Technological advances have dramatically expanded our ability to
  16176. probe multi-neuronal dynamics and connectivity in the brain.
  16177. However, our ability to extract a simple conceptual understanding
  16178. from complex data is increasingly hampered by the lack of
  16179. theoretically principled data analytic procedures, as well as
  16180. theoretical frameworks for how circuit connectivity and dynamics
  16181. can conspire to generate emergent behavioral and cognitive
  16182. functions. We review and outline potential avenues for progress,
  16183. including new theories of high dimensional data analysis, the
  16184. need to analyze complex artificial networks, and methods for
  16185. analyzing entire spaces of circuit models, rather than one model
  16186. at a time. Such interplay between experiments, data analysis and
  16187. theory will be indispensable in catalyzing conceptual advances in
  16188. the age of large-scale neuroscience.",
  16189. journal = "Curr. Opin. Neurobiol.",
  16190. volume = 32,
  16191. pages = "148--155",
  16192. month = jun,
  16193. year = 2015,
  16194. language = "en"
  16195. }
  16196. @INPROCEEDINGS{Lobo1997-qb,
  16197. title = "Decision making in a hybrid genetic algorithm",
  16198. booktitle = "Proceedings of 1997 {IEEE} International Conference on
  16199. Evolutionary Computation ({ICEC} '97)",
  16200. author = "Lobo, F G and Goldberg, D E",
  16201. abstract = "There are several issues that need to be taken into
  16202. consideration when designing a hybrid problem solver. The paper
  16203. focuses on one of them-decision making. More specifically, we
  16204. address the following questions: given two different methods,
  16205. how to get the most out of both of them? When should we use one
  16206. and when should we use the other in order to get maximum
  16207. efficiency? We present a model for hybridizing genetic
  16208. algorithms (GAs) based on a concept that decision theorists call
  16209. probability matching and we use it to combine an elitist
  16210. selecto-recombinative GA with a simple hill climber (HC). Tests
  16211. on an easy problem with a small population size match our
  16212. intuition that both GA and HC are needed to solve the problem
  16213. efficiently.",
  16214. pages = "121--125",
  16215. month = apr,
  16216. year = 1997,
  16217. keywords = "decision theory;genetic algorithms;probability;decision
  16218. making;hybrid genetic algorithm;hybrid problem solver;maximum
  16219. efficiency;decision theorists;probability matching;elitist
  16220. selecto-recombinative GA;simple hill climber;population
  16221. size;Decision making;Genetic algorithms;Algorithm design and
  16222. analysis;Testing;Expert systems;Turbines;Jet engines;Diversity
  16223. reception;Maintenance engineering;Mathematical analysis"
  16224. }
  16225. @ARTICLE{Balaguer2016-ho,
  16226. title = "Neural Mechanisms of Hierarchical Planning in a Virtual Subway
  16227. Network",
  16228. author = "Balaguer, Jan and Spiers, Hugo and Hassabis, Demis and
  16229. Summerfield, Christopher",
  16230. abstract = "Planning allows actions to be structured in pursuit of a future
  16231. goal. However, in natural environments, planning over multiple
  16232. possible future states incurs prohibitive computational costs. To
  16233. represent plans efficiently, states can be clustered
  16234. hierarchically into ``contexts''. For example, representing a
  16235. journey through a subway network as a succession of individual
  16236. states (stations) is more costly than encoding a sequence of
  16237. contexts (lines) and context switches (line changes). Here, using
  16238. functional brain imaging, we asked humans to perform a planning
  16239. task in a virtual subway network. Behavioral analyses revealed
  16240. that humans executed a hierarchically organized plan. Brain
  16241. activity in the dorsomedial prefrontal cortex and premotor cortex
  16242. scaled with the cost of hierarchical plan representation and
  16243. unique neural signals in these regions signaled contexts and
  16244. context switches. These results suggest that humans represent
  16245. hierarchical plans using a network of caudal prefrontal
  16246. structures. VIDEO ABSTRACT.",
  16247. journal = "Neuron",
  16248. volume = 90,
  16249. number = 4,
  16250. pages = "893--903",
  16251. month = may,
  16252. year = 2016,
  16253. language = "en"
  16254. }
  16255. @ARTICLE{Spiers2015-fq,
  16256. title = "Neural systems supporting navigation",
  16257. author = "Spiers, Hugo J and Barry, Caswell",
  16258. abstract = "Much is known about how neural systems determine current spatial
  16259. position and orientation in the environment. By contrast little
  16260. is understood about how the brain represents future goal
  16261. locations or computes the distance and direction to such goals.
  16262. Recent electrophysiology, computational modelling and
  16263. neuroimaging research have shed new light on how the spatial
  16264. relationship to a goal may be determined and represented during
  16265. navigation. This research suggests that the hippocampus may code
  16266. the path to the goal while the entorhinal cortex represents the
  16267. vector to the goal. It also reveals that the engagement of the
  16268. hippocampus and entorhinal cortex varies across the different
  16269. operational stages of navigation, such as during travel, route
  16270. planning, and decision-making at waypoints.",
  16271. journal = "Current Opinion in Behavioral Sciences",
  16272. volume = 1,
  16273. pages = "47--55",
  16274. month = feb,
  16275. year = 2015
  16276. }
  16277. @ARTICLE{Spiers2008-aw,
  16278. title = "Keeping the goal in mind: prefrontal contributions to spatial
  16279. navigation",
  16280. author = "Spiers, Hugo J",
  16281. journal = "Neuropsychologia",
  16282. volume = 46,
  16283. number = 7,
  16284. pages = "2106--2108",
  16285. month = feb,
  16286. year = 2008,
  16287. language = "en"
  16288. }
  16289. @ARTICLE{Beekman2001-xe,
  16290. title = "Phase transition between disordered and ordered foraging in
  16291. Pharaoh's ants",
  16292. author = "Beekman, M and Sumpter, D J and Ratnieks, F L",
  16293. abstract = "The complex collective behavior seen in many insect societies
  16294. strongly suggests that a minimum number of workers are required
  16295. for these societies to function effectively. Here we investigated
  16296. the transition between disordered and ordered foraging in the
  16297. Pharaoh's ant. We show that small colonies forage in a
  16298. disorganized manner, with a transition to organized
  16299. pheromone-based foraging in larger colonies. We also show that
  16300. when food sources are difficult to locate through independent
  16301. searching, this transition is first-order and exhibits
  16302. hysteresis, comparable to a first-order phase transition found in
  16303. many physical systems. To our knowledge, this is the first
  16304. experimental evidence of a behavioral phase transition between a
  16305. maladaptive (disorganized) and an adaptive (organized) state.",
  16306. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  16307. volume = 98,
  16308. number = 17,
  16309. pages = "9703--9706",
  16310. month = aug,
  16311. year = 2001,
  16312. language = "en"
  16313. }
  16314. @ARTICLE{Trimmer_Pete_C2008-ub,
  16315. title = "Mammalian choices: combining fast-but-inaccurate and
  16316. slow-but-accurate decision-making systems",
  16317. author = "{Trimmer Pete C} and {Houston Alasdair I} and {Marshall James
  16318. A.R} and {Bogacz Rafal} and {Paul Elizabeth S} and {Mendl Mike
  16319. T} and {McNamara John M}",
  16320. journal = "Proceedings of the Royal Society B: Biological Sciences",
  16321. publisher = "Royal Society",
  16322. volume = 275,
  16323. number = 1649,
  16324. pages = "2353--2361",
  16325. month = oct,
  16326. year = 2008
  16327. }
  16328. @ARTICLE{Gremel2013-yb,
  16329. title = "Orbitofrontal and striatal circuits dynamically encode the shift
  16330. between goal-directed and habitual actions",
  16331. author = "Gremel, Christina M and Costa, Rui M",
  16332. abstract = "Shifting between goal-directed and habitual actions allows for
  16333. efficient and flexible decision making. Here we demonstrate a
  16334. novel, within-subject instrumental lever-pressing paradigm, in
  16335. which mice shift between goal-directed and habitual actions. We
  16336. identify a role for orbitofrontal cortex (OFC) in actions
  16337. following outcome revaluation, and confirm that dorsal medial
  16338. (DMS) and lateral striatum (DLS) mediate different action
  16339. strategies. Simultaneous in vivo recordings of OFC, DMS and DLS
  16340. neuronal ensembles during shifting reveal that the same neurons
  16341. display different activities depending on whether presses are
  16342. goal-directed or habitual, with DMS and OFC becoming more and
  16343. DLS less engaged during goal-directed actions. Importantly, the
  16344. magnitude of neural activity changes in OFC following changes in
  16345. outcome value positively correlates with the level of
  16346. goal-directed behavior. Chemogenetic inhibition of OFC disrupts
  16347. goal-directed actions, whereas optogenetic activation of OFC
  16348. specifically increases goal-directed pressing. These results
  16349. also reveal a role for OFC in action revaluation, which has
  16350. implications for understanding compulsive behavior.",
  16351. journal = "Nat. Commun.",
  16352. publisher = "nature.com",
  16353. volume = 4,
  16354. pages = "2264",
  16355. year = 2013,
  16356. language = "en"
  16357. }
  16358. @ARTICLE{McDannald2011-df,
  16359. title = "Ventral striatum and orbitofrontal cortex are both required for
  16360. model-based, but not model-free, reinforcement learning",
  16361. author = "McDannald, Michael A and Lucantonio, Federica and Burke, Kathryn
  16362. A and Niv, Yael and Schoenbaum, Geoffrey",
  16363. abstract = "In many cases, learning is thought to be driven by differences
  16364. between the value of rewards we expect and rewards we actually
  16365. receive. Yet learning can also occur when the identity of the
  16366. reward we receive is not as expected, even if its value remains
  16367. unchanged. Learning from changes in reward identity implies
  16368. access to an internal model of the environment, from which
  16369. information about the identity of the expected reward can be
  16370. derived. As a result, such learning is not easily accounted for
  16371. by model-free reinforcement learning theories such as temporal
  16372. difference reinforcement learning (TDRL), which predicate
  16373. learning on changes in reward value, but not identity. Here, we
  16374. used unblocking procedures to assess learning driven by value-
  16375. versus identity-based prediction errors. Rats were trained to
  16376. associate distinct visual cues with different food quantities
  16377. and identities. These cues were subsequently presented in
  16378. compound with novel auditory cues and the reward quantity or
  16379. identity was selectively changed. Unblocking was assessed by
  16380. presenting the auditory cues alone in a probe test. Consistent
  16381. with neural implementations of TDRL models, we found that the
  16382. ventral striatum was necessary for learning in response to
  16383. changes in reward value. However, this area, along with
  16384. orbitofrontal cortex, was also required for learning driven by
  16385. changes in reward identity. This observation requires that
  16386. existing models of TDRL in the ventral striatum be modified to
  16387. include information about the specific features of expected
  16388. outcomes derived from model-based representations, and that the
  16389. role of orbitofrontal cortex in these models be clearly
  16390. delineated.",
  16391. journal = "J. Neurosci.",
  16392. publisher = "Soc Neuroscience",
  16393. volume = 31,
  16394. number = 7,
  16395. pages = "2700--2705",
  16396. month = feb,
  16397. year = 2011,
  16398. language = "en"
  16399. }
  16400. @UNPUBLISHED{Kay2019-yl,
  16401. title = "Regular cycling between representations of alternatives in the
  16402. hippocampus",
  16403. author = "Kay, Kenneth and Chung, Jason E and Sosa, Marielena and Schor,
  16404. Jonathan S and Karlsson, Mattias P and Larkin, Margaret C and
  16405. Liu, Daniel F and Frank, Loren M",
  16406. abstract = "Cognitive faculties such as imagination, planning, and
  16407. decision-making require the ability to represent alternative
  16408. scenarios. In animals, split-second decision-making implies that
  16409. the brain can represent alternatives at a commensurate speed. Yet
  16410. despite this insight, it has remained unknown whether there
  16411. exists neural activity that can consistently represent
  16412. alternatives in",
  16413. journal = "bioRxiv",
  16414. pages = "528976",
  16415. month = jan,
  16416. year = 2019,
  16417. language = "en"
  16418. }
  16419. @ARTICLE{Karlsson2009-wx,
  16420. title = "Awake replay of remote experiences in the hippocampus",
  16421. author = "Karlsson, Mattias P and Frank, Loren M",
  16422. abstract = "Hippocampal replay is thought to be essential for the
  16423. consolidation of event memories in hippocampal-neocortical
  16424. networks. Replay is present during both sleep and waking
  16425. behavior, but although sleep replay involves the reactivation of
  16426. stored representations in the absence of specific sensory inputs,
  16427. awake replay is thought to depend on sensory input from the
  16428. current environment. Here, we show that stored representations
  16429. are reactivated during both waking and sleep replay. We found
  16430. frequent awake replay of sequences of rat hippocampal place cells
  16431. from a previous experience. This spatially remote replay was as
  16432. common as local replay of the current environment and was more
  16433. robust when the rat had recently been in motion than during
  16434. extended periods of quiescence. Our results indicate that the
  16435. hippocampus consistently replays past experiences during brief
  16436. pauses in waking behavior, suggesting a role for waking replay in
  16437. memory consolidation and retrieval.",
  16438. journal = "Nat. Neurosci.",
  16439. volume = 12,
  16440. number = 7,
  16441. pages = "913--918",
  16442. month = jul,
  16443. year = 2009,
  16444. language = "en"
  16445. }
  16446. @ARTICLE{Miller2017-em,
  16447. title = "Dorsal hippocampus contributes to model-based planning",
  16448. author = "Miller, Kevin J and Botvinick, Matthew M and Brody, Carlos D",
  16449. abstract = "Planning can be defined as action selection that leverages an
  16450. internal model of the outcomes likely to follow each possible
  16451. action. Its neural mechanisms remain poorly understood. Here we
  16452. adapt recent advances from human research for rats, presenting
  16453. for the first time an animal task that produces many trials of
  16454. planned behavior per session, making multitrial rodent
  16455. experimental tools available to study planning. We use part of
  16456. this toolkit to address a perennially controversial issue in
  16457. planning: the role of the dorsal hippocampus. Although
  16458. prospective hippocampal representations have been proposed to
  16459. support planning, intact planning in animals with damaged
  16460. hippocampi has been repeatedly observed. Combining formal
  16461. algorithmic behavioral analysis with muscimol inactivation, we
  16462. provide causal evidence directly linking dorsal hippocampus with
  16463. planning behavior. Our results and methods open the door to new
  16464. and more detailed investigations of the neural mechanisms of
  16465. planning in the hippocampus and throughout the brain.",
  16466. journal = "Nat. Neurosci.",
  16467. volume = 20,
  16468. number = 9,
  16469. pages = "1269--1276",
  16470. month = sep,
  16471. year = 2017,
  16472. language = "en"
  16473. }
  16474. @ARTICLE{Botvinick2019-ac,
  16475. title = "Reinforcement Learning, Fast and Slow",
  16476. author = "Botvinick, Matthew and Ritter, Sam and Wang, Jane X and
  16477. Kurth-Nelson, Zeb and Blundell, Charles and Hassabis, Demis",
  16478. abstract = "Deep reinforcement learning (RL) methods have driven impressive
  16479. advances in artificial intelligence in recent years, exceeding
  16480. human performance in domains ranging from Atari to Go to no-limit
  16481. poker. This progress has drawn the attention of cognitive
  16482. scientists interested in understanding human learning. However,
  16483. the concern has been raised that deep RL may be too
  16484. sample-inefficient - that is, it may simply be too slow - to
  16485. provide a plausible model of how humans learn. In the present
  16486. review, we counter this critique by describing recently developed
  16487. techniques that allow deep RL to operate more nimbly, solving
  16488. problems much more quickly than previous methods. Although these
  16489. techniques were developed in an AI context, we propose that they
  16490. may have rich implications for psychology and neuroscience. A key
  16491. insight, arising from these AI methods, concerns the fundamental
  16492. connection between fast RL and slower, more incremental forms of
  16493. learning.",
  16494. journal = "Trends Cogn. Sci.",
  16495. month = apr,
  16496. year = 2019,
  16497. language = "en"
  16498. }
  16499. @UNPUBLISHED{Huda2018-bn,
  16500. title = "Bidirectional control of goal-oriented action selection by
  16501. distinct prefrontal cortex circuits",
  16502. author = "Huda, Rafiq and Sipe, Grayson O and Adam, Elie and
  16503. Breton-Provencher, Vincent and Pho, Gerald N and Gunter, Liadan M
  16504. and Wickersham, Ian R and Sur, Mriganka",
  16505. abstract = "Summary The immense behavioral repertoire of animals necessitates
  16506. mechanisms that select and suppress specific actions depending on
  16507. current goals. The prefrontal cortex (PFC) has been suggested to
  16508. orchestrate these processes by biasing activity in its target
  16509. structures, but how its vastly converging inputs and diverging
  16510. outputs are coordinated to control goal-oriented actions remains
  16511. unclear. Here we use a bilateral task in which mice select
  16512. between symmetric but opposing actions to show that distinct
  16513. outputs from a subdivision of the PFC, the anterior cingulate
  16514. cortex (ACC), promote correct and suppress incorrect actions.
  16515. Surprisingly, ACC outputs to the superior colliculus principally
  16516. inhibit incorrect actions. Optogenetic analyses and a
  16517. projection-based activity model make the unexpected prediction
  16518. that feedback from the ACC to the visual cortex promotes correct
  16519. actions, which we confirm. Our results show that anatomically
  16520. non-overlapping but functionally complementary PFC outputs
  16521. bidirectionally control actions, and suggest a candidate
  16522. organizing principle for PFC circuits.",
  16523. journal = "bioRxiv",
  16524. pages = "307009",
  16525. month = jul,
  16526. year = 2018,
  16527. language = "en"
  16528. }
  16529. @ARTICLE{Coutureau2003-sj,
  16530. title = "Inactivation of the infralimbic prefrontal cortex reinstates
  16531. goal-directed responding in overtrained rats",
  16532. author = "Coutureau, Etienne and Killcross, Simon",
  16533. abstract = "Over the course of extended training, instrumental responding in
  16534. rats shows a transition from goal-dependent performance to
  16535. goal-independent performance, as assessed by sensitivity to
  16536. reward-devaluation induced by taste aversions or specific
  16537. satiety. It has been suggested that this reflects the gradual
  16538. dominance of reflexive, habit-based responding over voluntary,
  16539. goal-directed actions. Previous research suggests that lesions
  16540. of the medial prefrontal cortex disrupt this interaction between
  16541. goal-directed and habitual responding. More specifically,
  16542. whereas lesions of the prelimbic prefrontal cortex appear to
  16543. disrupt normal goal-directed responding, lesions of the
  16544. infralimbic prefrontal cortex cause animals to remain
  16545. goal-directed even after substantial overtraining. The current
  16546. experiment explored further the nature of this interaction
  16547. between actions and habits. Rats were given extended training of
  16548. an instrumental lever press response before bilateral
  16549. intracerebral cannulae giving access to the infralimbic cortex
  16550. were implanted. Following further reminder training all animals
  16551. were given a test of goal sensitivity by specific-satiety
  16552. devaluation of the instrumental outcome, or a matched reward,
  16553. prior to extinction tests. Before these tests, half of the
  16554. animals received bilateral infusions of muscimol into the
  16555. infralimbic cortex, and the remainder, control vehicle
  16556. infusions. As expected after extended instrumental training,
  16557. control-infused animals showed habitual performance that was not
  16558. selectively influenced by devaluation of the instrumental
  16559. outcome. In contrast, animals receiving temporary inactivation
  16560. of the infralimbic cortex by muscimol showed selective
  16561. sensitivity to devaluation of the instrumental outcome,
  16562. indicating a reinstatement of goal-directed responding in these
  16563. animals. This suggests that the development of habitual
  16564. responding reflects the active inhibition of goal-directed
  16565. responses that are mediated by action-outcome associations.",
  16566. journal = "Behav. Brain Res.",
  16567. publisher = "Elsevier",
  16568. volume = 146,
  16569. number = "1-2",
  16570. pages = "167--174",
  16571. month = nov,
  16572. year = 2003,
  16573. language = "en"
  16574. }
  16575. @ARTICLE{Rao2010-cq,
  16576. title = "Decision making under uncertainty: a neural model based on
  16577. partially observable markov decision processes",
  16578. author = "Rao, Rajesh P N",
  16579. abstract = "A fundamental problem faced by animals is learning to select
  16580. actions based on noisy sensory information and incomplete
  16581. knowledge of the world. It has been suggested that the brain
  16582. engages in Bayesian inference during perception but how such
  16583. probabilistic representations are used to select actions has
  16584. remained unclear. Here we propose a neural model of action
  16585. selection and decision making based on the theory of partially
  16586. observable Markov decision processes (POMDPs). Actions are
  16587. selected based not on a single ``optimal'' estimate of state but
  16588. on the posterior distribution over states (the ``belief'' state).
  16589. We show how such a model provides a unified framework for
  16590. explaining experimental results in decision making that involve
  16591. both information gathering and overt actions. The model utilizes
  16592. temporal difference (TD) learning for maximizing expected reward.
  16593. The resulting neural architecture posits an active role for the
  16594. neocortex in belief computation while ascribing a role to the
  16595. basal ganglia in belief representation, value computation, and
  16596. action selection. When applied to the random dots motion
  16597. discrimination task, model neurons representing belief exhibit
  16598. responses similar to those of LIP neurons in primate neocortex.
  16599. The appropriate threshold for switching from information
  16600. gathering to overt actions emerges naturally during reward
  16601. maximization. Additionally, the time course of reward prediction
  16602. error in the model shares similarities with dopaminergic
  16603. responses in the basal ganglia during the random dots task. For
  16604. tasks with a deadline, the model learns a decision making
  16605. strategy that changes with elapsed time, predicting a collapsing
  16606. decision threshold consistent with some experimental studies. The
  16607. model provides a new framework for understanding neural decision
  16608. making and suggests an important role for interactions between
  16609. the neocortex and the basal ganglia in learning the mapping
  16610. between probabilistic sensory representations and actions that
  16611. maximize rewards.",
  16612. journal = "Front. Comput. Neurosci.",
  16613. volume = 4,
  16614. pages = "146",
  16615. month = nov,
  16616. year = 2010,
  16617. keywords = "Bayesian inference; basal ganglia; decision theory; dopamine;
  16618. parietal cortex; probabilistic models; reinforcement learning;
  16619. temporal difference learning",
  16620. language = "en"
  16621. }
  16622. @ARTICLE{Van_der_Meer2010-ka,
  16623. title = "Triple dissociation of information processing in dorsal striatum,
  16624. ventral striatum, and hippocampus on a learned spatial decision
  16625. task",
  16626. author = "van der Meer, Matthijs A A and Johnson, Adam and
  16627. Schmitzer-Torbert, Neil C and Redish, A David",
  16628. abstract = "Decision-making studies across different domains suggest that
  16629. decisions can arise from multiple, parallel systems in the brain:
  16630. a flexible system utilizing action-outcome expectancies and a
  16631. more rigid system based on situation-action associations. The
  16632. hippocampus, ventral striatum, and dorsal striatum make unique
  16633. contributions to each system, but how information processing in
  16634. each of these structures supports these systems is unknown.
  16635. Recent work has shown covert representations of future paths in
  16636. hippocampus and of future rewards in ventral striatum. We
  16637. developed analyses in order to use a comparative methodology and
  16638. apply the same analyses to all three structures. Covert
  16639. representations of future paths and reward were both absent from
  16640. the dorsal striatum. In contrast, dorsal striatum slowly
  16641. developed situation representations that selectively represented
  16642. action-rich parts of the task. This triple dissociation suggests
  16643. that the different roles these structures play are due to
  16644. differences in information-processing mechanisms.",
  16645. journal = "Neuron",
  16646. volume = 67,
  16647. number = 1,
  16648. pages = "25--32",
  16649. month = jul,
  16650. year = 2010,
  16651. language = "en"
  16652. }
  16653. @ARTICLE{Yin2006-jc,
  16654. title = "The role of the basal ganglia in habit formation",
  16655. author = "Yin, Henry H and Knowlton, Barbara J",
  16656. abstract = "Many organisms, especially humans, are characterized by their
  16657. capacity for intentional, goal-directed actions. However,
  16658. similar behaviours often proceed automatically, as habitual
  16659. responses to antecedent stimuli. How are goal-directed actions
  16660. transformed into habitual responses? Recent work combining
  16661. modern behavioural assays and neurobiological analysis of the
  16662. basal ganglia has begun to yield insights into the neural basis
  16663. of habit formation.",
  16664. journal = "Nat. Rev. Neurosci.",
  16665. publisher = "nature.com",
  16666. volume = 7,
  16667. number = 6,
  16668. pages = "464--476",
  16669. month = jun,
  16670. year = 2006,
  16671. language = "en"
  16672. }
  16673. @ARTICLE{Fiore2015-om,
  16674. title = "Evolutionarily conserved mechanisms for the selection and
  16675. maintenance of behavioural activity",
  16676. author = "Fiore, Vincenzo G and Dolan, Raymond J and Strausfeld, Nicholas J
  16677. and Hirth, Frank",
  16678. abstract = "Survival and reproduction entail the selection of adaptive
  16679. behavioural repertoires. This selection manifests as
  16680. phylogenetically acquired activities that depend on evolved
  16681. nervous system circuitries. Lorenz and Tinbergen already
  16682. postulated that heritable behaviours and their reliable
  16683. performance are specified by genetically determined programs.
  16684. Here we compare the functional anatomy of the insect central
  16685. complex and vertebrate basal ganglia to illustrate their role in
  16686. mediating selection and maintenance of adaptive behaviours.
  16687. Comparative analyses reveal that central complex and basal
  16688. ganglia circuitries share comparable lineage relationships within
  16689. clusters of functionally integrated neurons. These clusters are
  16690. specified by genetic mechanisms that link birth time and order to
  16691. their neuronal identities and functions. Their subsequent
  16692. connections and associated functions are characterized by similar
  16693. mechanisms that implement dimensionality reduction and transition
  16694. through attractor states, whereby spatially organized
  16695. parallel-projecting loops integrate and convey sensorimotor
  16696. representations that select and maintain behavioural activity. In
  16697. both taxa, these neural systems are modulated by dopamine
  16698. signalling that also mediates memory-like processes. The
  16699. multiplicity of similarities between central complex and basal
  16700. ganglia suggests evolutionarily conserved computational
  16701. mechanisms for action selection. We speculate that these may have
  16702. originated from ancestral ground pattern circuitries present in
  16703. the brain of the last common ancestor of insects and vertebrates.",
  16704. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  16705. volume = 370,
  16706. number = 1684,
  16707. month = dec,
  16708. year = 2015,
  16709. keywords = "action selection; attractor state; basal ganglia; brain
  16710. evolution; central complex; sensorimotor representation",
  16711. language = "en"
  16712. }
  16713. @ARTICLE{Patrick_undated-xh,
  16714. title = "Computational model of habit learning and reversal",
  16715. author = "Patrick, Sean and Bullock, Daniel"
  16716. }
  16717. @ARTICLE{Harvey2009-ql,
  16718. title = "Intracellular dynamics of hippocampal place cells during virtual
  16719. navigation",
  16720. author = "Harvey, Christopher D and Collman, Forrest and Dombeck, Daniel A
  16721. and Tank, David W",
  16722. abstract = "Hippocampal place cells encode spatial information in rate and
  16723. temporal codes. To examine the mechanisms underlying hippocampal
  16724. coding, here we measured the intracellular dynamics of place
  16725. cells by combining in vivo whole-cell recordings with a
  16726. virtual-reality system. Head-restrained mice, running on a
  16727. spherical treadmill, interacted with a computer-generated visual
  16728. environment to perform spatial behaviours. Robust place-cell
  16729. activity was present during movement along a virtual linear
  16730. track. From whole-cell recordings, we identified three
  16731. subthreshold signatures of place fields: an asymmetric ramp-like
  16732. depolarization of the baseline membrane potential, an increase
  16733. in the amplitude of intracellular theta oscillations, and a
  16734. phase precession of the intracellular theta oscillation relative
  16735. to the extracellularly recorded theta rhythm. These
  16736. intracellular dynamics underlie the primary features of
  16737. place-cell rate and temporal codes. The virtual-reality system
  16738. developed here will enable new experimental approaches to study
  16739. the neural circuits underlying navigation.",
  16740. journal = "Nature",
  16741. publisher = "nature.com",
  16742. volume = 461,
  16743. number = 7266,
  16744. pages = "941--946",
  16745. month = oct,
  16746. year = 2009,
  16747. language = "en"
  16748. }
  16749. @ARTICLE{Buschman2014-yb,
  16750. title = "Goal-direction and top-down control",
  16751. author = "Buschman, Timothy J and Miller, Earl K",
  16752. abstract = "We review the neural mechanisms that support top-down control of
  16753. behaviour and suggest that goal-directed behaviour uses two
  16754. systems that work in concert. A basal ganglia-centred system
  16755. quickly learns simple, fixed goal-directed behaviours while a
  16756. prefrontal cortex-centred system gradually learns more complex
  16757. (abstract or long-term) goal-directed behaviours. Interactions
  16758. between these two systems allow top-down control mechanisms to
  16759. learn how to direct behaviour towards a goal but also how to
  16760. guide behaviour when faced with a novel situation.",
  16761. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  16762. publisher = "royalsocietypublishing.org",
  16763. volume = 369,
  16764. number = 1655,
  16765. month = nov,
  16766. year = 2014,
  16767. keywords = "basal ganglia; cognition; frontal lobe; goal direction; learning",
  16768. language = "en"
  16769. }
  16770. @ARTICLE{Kawato2007-vk,
  16771. title = "Efficient reinforcement learning: computational theories,
  16772. neuroscience and robotics",
  16773. author = "Kawato, Mitsuo and Samejima, Kazuyuki",
  16774. abstract = "Reinforcement learning algorithms have provided some of the most
  16775. influential computational theories for behavioral learning that
  16776. depends on reward and penalty. After briefly reviewing
  16777. supporting experimental data, this paper tackles three difficult
  16778. theoretical issues that remain to be explored. First, plain
  16779. reinforcement learning is much too slow to be considered a
  16780. plausible brain model. Second, although the temporal-difference
  16781. error has an important role both in theory and in experiments,
  16782. how to compute it remains an enigma. Third, function of all
  16783. brain areas, including the cerebral cortex, cerebellum,
  16784. brainstem and basal ganglia, seems to necessitate a new
  16785. computational framework. Computational studies that emphasize
  16786. meta-parameters, hierarchy, modularity and supervised learning
  16787. to resolve these issues are reviewed here, together with the
  16788. related experimental data.",
  16789. journal = "Curr. Opin. Neurobiol.",
  16790. publisher = "Elsevier",
  16791. volume = 17,
  16792. number = 2,
  16793. pages = "205--212",
  16794. month = apr,
  16795. year = 2007,
  16796. language = "en"
  16797. }
  16798. @ARTICLE{Ghahramani2015-gx,
  16799. title = "Probabilistic machine learning and artificial intelligence",
  16800. author = "Ghahramani, Zoubin",
  16801. abstract = "How can a machine learn from experience? Probabilistic modelling
  16802. provides a framework for understanding what learning is, and has
  16803. therefore emerged as one of the principal theoretical and
  16804. practical approaches for designing machines that learn from data
  16805. acquired through experience. The probabilistic framework, which
  16806. describes how to represent and manipulate uncertainty about
  16807. models and predictions, has a central role in scientific data
  16808. analysis, machine learning, robotics, cognitive science and
  16809. artificial intelligence. This Review provides an introduction to
  16810. this framework, and discusses some of the state-of-the-art
  16811. advances in the field, namely, probabilistic programming,
  16812. Bayesian optimization, data compression and automatic model
  16813. discovery.",
  16814. journal = "Nature",
  16815. publisher = "nature.com",
  16816. volume = 521,
  16817. number = 7553,
  16818. pages = "452--459",
  16819. month = may,
  16820. year = 2015,
  16821. language = "en"
  16822. }
  16823. % The entry below contains non-ASCII chars that could not be converted
  16824. % to a LaTeX equivalent.
  16825. @ARTICLE{Becker1964-sy,
  16826. title = "Measuring utility by a single-response sequential method",
  16827. author = "Becker, Gordon M and DeGroot, Morris H and Marschak, Jacob",
  16828. abstract = "A person deciding on a career, a wife, or a place to live bases
  16829. his choice on two factors:(1) How much do I like each of the
  16830. available alternatives? and (2) What are the chances for a
  16831. successful outcome of each alternative? These two factors
  16832. comprise the utility of each outcome for the person making the
  16833. choice. This notion of utility is fundamental to most current
  16834. theories of decision behavior. According to the expected utility
  16835. hypothesis, if we could know the utility function of a person,
  16836. we could predict his choice from among any set of …",
  16837. journal = "Behav. Sci.",
  16838. publisher = "Wiley Online Library",
  16839. volume = 9,
  16840. number = 3,
  16841. pages = "226--232",
  16842. year = 1964
  16843. }
  16844. @ARTICLE{Rangel2010-ch,
  16845. title = "Neural computations associated with goal-directed choice",
  16846. author = "Rangel, Antonio and Hare, Todd",
  16847. abstract = "In goal-directed decision-making, animals choose between actions
  16848. that are associated with different reward outcomes (e.g., foods)
  16849. and with different costs (e.g., effort). Rapid advances have
  16850. been made over the past few years in our understanding of the
  16851. computations associated with goal-directed choices, and of how
  16852. those computations are implemented in the brain. We review some
  16853. important findings, with an emphasis on computational models,
  16854. human fMRI, and monkey neurophysiology studies.",
  16855. journal = "Curr. Opin. Neurobiol.",
  16856. publisher = "rnl.caltech.edu",
  16857. volume = 20,
  16858. number = 2,
  16859. pages = "262--270",
  16860. month = apr,
  16861. year = 2010,
  16862. language = "en"
  16863. }
  16864. @MISC{noauthor_undated-xv,
  16865. title = "elegantscipy.pdf",
  16866. keywords = "Python books;Books/Python"
  16867. }
  16868. @ARTICLE{Ramalho_undated-wv,
  16869. title = "{CLEAR}, {CONCISE}, {AND} {EFFECTIVE} {PROGRAMMING}",
  16870. author = "Ramalho, Luciano",
  16871. keywords = "Python books;Books/Python"
  16872. }
  16873. @MISC{noauthor_undated-qc,
  16874. title = "introducingpython.pdf",
  16875. keywords = "Python books;Books/Python"
  16876. }
  16877. @MISC{noauthor_undated-jh,
  16878. title = "naturallanguageprocessingwithpython.pdf",
  16879. keywords = "Python books;Books/Python"
  16880. }
  16881. @ARTICLE{By_undated-ov,
  16882. title = "{ESSENTIAL} {TOOLS} {FOR} {WORKING} {WITH} {DATA}",
  16883. author = "By, Powered",
  16884. keywords = "Python books;Books/Python"
  16885. }
  16886. @MISC{noauthor_undated-rx,
  16887. title = "thinkbayes.pdf",
  16888. keywords = "Python books;Books/Python"
  16889. }
  16890. @ARTICLE{Kirk_undated-qt,
  16891. title = "A {TEST-DRIVEN} {APPROACHPython}",
  16892. author = "Kirk, Matthew",
  16893. keywords = "Python books;Books/Python"
  16894. }
  16895. @MISC{noauthor_undated-pt,
  16896. title = "twistednetworkprogrammingessentials.pdf",
  16897. keywords = "Python books;Books/Python"
  16898. }
  16899. @MISC{noauthor_undated-wa,
  16900. title = "webscrapingwithpython.pdf",
  16901. keywords = "Python books;Books/Python"
  16902. }
  16903. @ARTICLE{Altshuler2005-ci,
  16904. title = "Symmetry breaking in escaping ants",
  16905. author = "Altshuler, E and Ramos, O and N{\'u}{\~n}ez, Y and
  16906. Fern{\'a}ndez, J and Batista-Leyva, A J and Noda, C",
  16907. abstract = "The phenomenon of herding is a very general feature of the
  16908. collective behavior of many species in panic conditions,
  16909. including humans. It has been predicted theoretically that
  16910. panic-induced herding in individuals confined to a room can
  16911. produce a nonsymmetrical use of two identical exit doors. Here
  16912. we demonstrate the existence of that phenomenon in experiments,
  16913. using ants as a model of pedestrians. We show that ants confined
  16914. to a cell with two symmetrically located exits use both exits in
  16915. approximately equal proportions to abandon it in normal
  16916. conditions but prefer one of the exits if panic is created by
  16917. adding a repellent fluid. In addition, we are able to reproduce
  16918. the observed escape dynamics in detail using a modification of a
  16919. previous theoretical model that includes herding associated with
  16920. a panic parameter as a central ingredient. Our experimental
  16921. results, combined with theoretical models, suggest that some
  16922. features of the collective behavior of humans and ants can be
  16923. quite similar when escaping under panic.",
  16924. journal = "Am. Nat.",
  16925. publisher = "journals.uchicago.edu",
  16926. volume = 166,
  16927. number = 6,
  16928. pages = "643--649",
  16929. month = dec,
  16930. year = 2005,
  16931. language = "en"
  16932. }
  16933. @ARTICLE{Nicolis1999-xi,
  16934. title = "Emerging patterns and food recruitment in ants: an analytical
  16935. study",
  16936. author = "Nicolis, S C and Deneubourg, J L",
  16937. abstract = "A model of food recruitment by social insects accounting for the
  16938. competition between trails in the presence of an arbitrary
  16939. number of sources is developed and analysed in detail. Both the
  16940. case of identical environmental characteristics and the case
  16941. where one source and the corresponding trail are different from
  16942. the others are considered. Different collective responses
  16943. depending on the environmental conditions, and without change of
  16944. individual behaviour, are shown to exist, associated with the
  16945. possibility that the colony may be led to exploit one source or
  16946. a group of sources preferentially. The full bifurcation diagram
  16947. of steady-state solutions is constructed from which the dominant
  16948. exploitation patterns are identified. The biological relevance
  16949. of the results is discussed and suggestions are made for their
  16950. experimental testing in connection with the recruitment behavior
  16951. of species using trail recruitment. The same phenomenological
  16952. model can be used for different trail-laying species since the
  16953. predictions are generic and not restricted to a given species,
  16954. except for the parameter values used. Copyright 1999 Academic
  16955. Press.",
  16956. journal = "J. Theor. Biol.",
  16957. publisher = "Elsevier",
  16958. volume = 198,
  16959. number = 4,
  16960. pages = "575--592",
  16961. month = jun,
  16962. year = 1999,
  16963. language = "en"
  16964. }
  16965. @TECHREPORT{Millonas1993-ss,
  16966. title = "Swarms, Phase Transitions, and Collective Intelligence (Paper
  16967. 1); and A Nonequilibrium Statistical Field Theory of Swarms and
  16968. Other Spatially Extended Complex Systems (Paper 2)",
  16969. author = "Millonas, Mark M",
  16970. abstract = "(Paper 1) A spacially extended model of the collective behavior
  16971. of a large number of locally acting organisms is proposed in
  16972. which organisms move probabilistically between local cells in
  16973. space, but with weights dependent on local morphogenetic
  16974. substances, or morphogens. The morphogens are in turn effected
  16975. by the passage of an organism. The evolution of the morphogens,
  16976. and the corresponding flow of the organisms constitutes the
  16977. collective behavior of the group. Such models have various types
  16978. of phase transitions and self-organizing properties controlled
  16979. both by the level of the noise, and other parameters. The model
  16980. is then applied to the specific case of ants moving on a
  16981. lattice. The local behavior of the ants is inspired by the
  16982. actual behavior observed in the laboratory, and analytic results
  16983. for the collective behavior are compared to the corresponding
  16984. laboratory results. It is hoped that the present model might
  16985. serve as a paradigmatic example of a complex cooperative system
  16986. in nature. In particular swarm models can be used to explore the
  16987. relation of nonequilibrium phase transitions to at least three
  16988. important issues encountered in artificial life. Firstly, that
  16989. of emergence as complex adaptive behavior. Secondly, as an
  16990. exporation of continuous phase transitions in biological
  16991. systems. Lastly, to derive behavioral criteria for the evolution
  16992. of collective behavior in social organisms. (Paper 2) A class of
  16993. models with applications to swarm behavior as well as many other
  16994. types of spatially extended complex biological and physical
  16995. systems is studied. Internal fluctuations can play an active
  16996. role in the organization of the phase structure of such systems.
  16997. Consequently, it is not possible to fully understand the
  16998. behavior of these systems without explicitly incorporating the
  16999. fluctuations. In particular, for the class of models studied
  17000. here the effect of internal fluctuations due to finite size is a
  17001. renormalized \{\textbackslashit decrease\} in the temperature
  17002. near the point of spontaneous symmetry breaking. We briefly
  17003. outline how these models can be applied to the behavior of an
  17004. ant swarm.",
  17005. publisher = "Santa Fe Institute",
  17006. number = "93-06-039",
  17007. month = jun,
  17008. year = 1993
  17009. }
  17010. @ARTICLE{Marshall2009-pa,
  17011. title = "On optimal decision-making in brains and social insect colonies",
  17012. author = "Marshall, James A R and Bogacz, Rafal and Dornhaus, Anna and
  17013. Planqu{\'e}, Robert and Kovacs, Tim and Franks, Nigel R",
  17014. abstract = "The problem of how to compromise between speed and accuracy in
  17015. decision-making faces organisms at many levels of biological
  17016. complexity. Striking parallels are evident between
  17017. decision-making in primate brains and collective decision-making
  17018. in social insect colonies: in both systems, separate populations
  17019. accumulate evidence for alternative choices; when one population
  17020. reaches a threshold, a decision is made for the corresponding
  17021. alternative, and this threshold may be varied to compromise
  17022. between the speed and the accuracy of decision-making. In
  17023. primate decision-making, simple models of these processes have
  17024. been shown, under certain parametrizations, to implement the
  17025. statistically optimal procedure that minimizes decision time for
  17026. any given error rate. In this paper, we adapt these same
  17027. analysis techniques and apply them to new models of collective
  17028. decision-making in social insect colonies. We show that social
  17029. insect colonies may also be able to achieve statistically
  17030. optimal collective decision-making in a very similar way to
  17031. primate brains, via direct competition between
  17032. evidence-accumulating populations. This optimality result makes
  17033. testable predictions for how collective decision-making in
  17034. social insects should be organized. Our approach also represents
  17035. the first attempt to identify a common theoretical framework for
  17036. the study of decision-making in diverse biological systems.",
  17037. journal = "J. R. Soc. Interface",
  17038. publisher = "royalsocietypublishing.org",
  17039. volume = 6,
  17040. number = 40,
  17041. pages = "1065--1074",
  17042. month = nov,
  17043. year = 2009,
  17044. language = "en"
  17045. }
  17046. @ARTICLE{Detrain2006-vd,
  17047. title = "Self-organized structures in a superorganism: do ants ``behave''
  17048. like molecules?",
  17049. author = "Detrain, Claire and Deneubourg, Jean-Louis",
  17050. abstract = "While the striking structures (e.g. nest architecture, trail
  17051. networks) of insect societies may seem familiar to many of us,
  17052. the understanding of pattern formation still constitutes a
  17053. challenging problem. Over the last two decades,
  17054. self-organization has dramatically changed our view on how
  17055. collective decision-making and structures may emerge out of a
  17056. population of ant workers having each their own individuality as
  17057. well as a limited access to information. A variety of collective
  17058. behaviour spontaneously outcome from multiple interactions
  17059. between nestmates, even when there is no directing influence
  17060. imposed by an external template, a pacemaker or a leader. By
  17061. focussing this review on foraging structures, we show that ant
  17062. societies display some properties which are usually considered
  17063. in physico-chemical systems, as typical signatures of
  17064. self-organization. We detail the key role played by feed-back
  17065. loops, fluctuations, number of interacting units and sensitivity
  17066. to environmental factors in the emergence of a structured
  17067. collective behaviour. Nonetheless, going beyond simple analogies
  17068. with non-living self-organized patterns, we stress on the
  17069. specificities of social structures made of complex living units
  17070. of which the biological features have been selected throughout
  17071. the evolution depending on their adaptive value. In particular,
  17072. we consider the ability of each ant individual to process
  17073. information about environmental and social parameters, to
  17074. accordingly tune its interactions with nestmates and ultimately
  17075. to determine the final pattern emerging at the collective level.
  17076. We emphasize on the parsimony and simplicity of behavioural
  17077. rules at the individual level which allow an efficient
  17078. processing of information, energy and matter within the whole
  17079. colony.",
  17080. journal = "Phys. Life Rev.",
  17081. publisher = "Elsevier",
  17082. volume = 3,
  17083. number = 3,
  17084. pages = "162--187",
  17085. month = sep,
  17086. year = 2006,
  17087. keywords = "Self-organization; Decision-making; Pattern formation; Social
  17088. insects; Foraging; Trail"
  17089. }
  17090. @ARTICLE{Mora2011-th,
  17091. title = "Are Biological Systems Poised at Criticality?",
  17092. author = "Mora, Thierry and Bialek, William",
  17093. abstract = "Many of life's most fascinating phenomena emerge from
  17094. interactions among many elements---many amino acids determine
  17095. the structure of a single protein, many genes determine the fate
  17096. of a cell, many neurons are involved in shaping our thoughts and
  17097. memories. Physicists have long hoped that these collective
  17098. behaviors could be described using the ideas and methods of
  17099. statistical mechanics. In the past few years, new, larger scale
  17100. experiments have made it possible to construct statistical
  17101. mechanics models of biological systems directly from real data.
  17102. We review the surprising successes of this ``inverse'' approach,
  17103. using examples from families of proteins, networks of neurons,
  17104. and flocks of birds. Remarkably, in all these cases the models
  17105. that emerge from the data are poised near a very special point
  17106. in their parameter space---a critical point. This suggests there
  17107. may be some deeper theoretical principle behind the behavior of
  17108. these diverse systems.",
  17109. journal = "J. Stat. Phys.",
  17110. publisher = "Springer",
  17111. volume = 144,
  17112. number = 2,
  17113. pages = "268--302",
  17114. month = jul,
  17115. year = 2011
  17116. }
  17117. @ARTICLE{Sole1995-op,
  17118. title = "Information at the edge of chaos in fluid neural networks",
  17119. author = "Sol{\'e}, Ricard V and Miramontes, Octavio",
  17120. abstract = "Fluid neural networks, defined as neural nets of mobile elements
  17121. with random activation, are studied by means of several
  17122. approaches. They are proposed as a theoretical framework for a
  17123. wide class of systems as insect societies, collectives of robots
  17124. or the immune system. The critical properties of this model are
  17125. also analysed, showing the existence of a critical boundary in
  17126. parameter space where maximum information transfer occurs. In
  17127. this sense, this boundary is in fact an example of the ``edge of
  17128. chaos'' in systems like those described in our approach. Recent
  17129. experiments with ant colonies seem to confirm our result.",
  17130. journal = "Physica D",
  17131. publisher = "Elsevier",
  17132. volume = 80,
  17133. number = 1,
  17134. pages = "171--180",
  17135. month = jan,
  17136. year = 1995
  17137. }
  17138. @ARTICLE{Baluska2016-po,
  17139. title = "On Having No Head: Cognition throughout Biological Systems",
  17140. author = "Balu{\v s}ka, Franti{\v s}ek and Levin, Michael",
  17141. abstract = "The central nervous system (CNS) underlies memory, perception,
  17142. decision-making, and behavior in numerous organisms. However,
  17143. neural networks have no monopoly on the signaling functions that
  17144. implement these remarkable algorithms. It is often forgotten
  17145. that neurons optimized cellular signaling modes that existed
  17146. long before the CNS appeared during evolution, and were used by
  17147. somatic cellular networks to orchestrate physiology, embryonic
  17148. development, and behavior. Many of the key dynamics that enable
  17149. information processing can, in fact, be implemented by different
  17150. biological hardware. This is widely exploited by organisms
  17151. throughout the tree of life. Here, we review data on memory,
  17152. learning, and other aspects of cognition in a range of models,
  17153. including single celled organisms, plants, and tissues in animal
  17154. bodies. We discuss current knowledge of the molecular mechanisms
  17155. at work in these systems, and suggest several hypotheses for
  17156. future investigation. The study of cognitive processes
  17157. implemented in aneural contexts is a fascinating, highly
  17158. interdisciplinary topic that has many implications for
  17159. evolution, cell biology, regenerative medicine, computer
  17160. science, and synthetic bioengineering.",
  17161. journal = "Front. Psychol.",
  17162. publisher = "frontiersin.org",
  17163. volume = 7,
  17164. pages = "902",
  17165. month = jun,
  17166. year = 2016,
  17167. keywords = "aneural; bioelectric signaling; cognition; computation;
  17168. information; learning; memory; plants",
  17169. language = "en"
  17170. }
  17171. @ARTICLE{Pinero_Jordi2019-zs,
  17172. title = "Statistical physics of liquid brains",
  17173. author = "{Pi{\~n}ero Jordi} and {Sol{\'e} Ricard}",
  17174. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  17175. publisher = "Royal Society",
  17176. volume = 374,
  17177. number = 1774,
  17178. pages = "20180376",
  17179. month = jun,
  17180. year = 2019
  17181. }
  17182. @ARTICLE{Niv2009-gn,
  17183. title = "Reinforcement learning in the brain",
  17184. author = "Niv, Yael",
  17185. abstract = "A wealth of research focuses on the decision-making processes
  17186. that animals and humans employ when selecting actions in the
  17187. face of reward and punishment. Initially such work stemmed from
  17188. psychological investigations of conditioned behavior, and
  17189. explanations of these in terms of computational models.
  17190. Increasingly, analysis at the computational level has drawn on
  17191. ideas from reinforcement learning, which provide a normative
  17192. framework within which decision-making can be analyzed. More
  17193. recently, the fruits of these extensive lines of research have
  17194. made contact with investigations into the neural basis of
  17195. decision making. Converging evidence now links reinforcement
  17196. learning to specific neural substrates, assigning them precise
  17197. computational roles. Specifically, electrophysiological
  17198. recordings in behaving animals and functional imaging of human
  17199. decision-making have revealed in the brain the existence of a
  17200. key reinforcement learning signal, the temporal difference
  17201. reward prediction error. Here, we first introduce the formal
  17202. reinforcement learning framework. We then review the multiple
  17203. lines of evidence linking reinforcement learning to the function
  17204. of dopaminergic neurons in the mammalian midbrain and to more
  17205. recent data from human imaging experiments. We further extend
  17206. the discussion to aspects of learning not associated with phasic
  17207. dopamine signals, such as learning of goal-directed responding
  17208. that may not be dopamine-dependent, and learning about the vigor
  17209. (or rate) with which actions should be performed that has been
  17210. linked to tonic aspects of dopaminergic signaling. We end with a
  17211. brief discussion of some of the limitations of the reinforcement
  17212. learning framework, highlighting questions for future research.",
  17213. journal = "J. Math. Psychol.",
  17214. publisher = "Elsevier",
  17215. volume = 53,
  17216. number = 3,
  17217. pages = "139--154",
  17218. month = jun,
  17219. year = 2009
  17220. }
  17221. @ARTICLE{Friston2010-yo,
  17222. title = "The free-energy principle: a unified brain theory?",
  17223. author = "Friston, Karl",
  17224. abstract = "A free-energy principle has been proposed recently that accounts
  17225. for action, perception and learning. This Review looks at some
  17226. key brain theories in the biological (for example, neural
  17227. Darwinism) and physical (for example, information theory and
  17228. optimal control theory) sciences from the free-energy
  17229. perspective. Crucially, one key theme runs through each of these
  17230. theories - optimization. Furthermore, if we look closely at what
  17231. is optimized, the same quantity keeps emerging, namely value
  17232. (expected reward, expected utility) or its complement, surprise
  17233. (prediction error, expected cost). This is the quantity that is
  17234. optimized under the free-energy principle, which suggests that
  17235. several global brain theories might be unified within a
  17236. free-energy framework.",
  17237. journal = "Nat. Rev. Neurosci.",
  17238. volume = 11,
  17239. number = 2,
  17240. pages = "127--138",
  17241. month = feb,
  17242. year = 2010,
  17243. language = "en"
  17244. }
  17245. @MISC{Murphy_undated-zd,
  17246. title = "Machine Learning - A Probabilistic Perspective",
  17247. author = "{Murphy}",
  17248. keywords = "books;Books"
  17249. }
  17250. @ARTICLE{Botvinick2012-ho,
  17251. title = "Planning as inference",
  17252. author = "Botvinick, Matthew and Toussaint, Marc",
  17253. abstract = "Recent developments in decision-making research are bringing the
  17254. topic of planning back to center stage in cognitive science. This
  17255. renewed interest reopens an old, but still unanswered question:
  17256. how exactly does planning happen? What are the underlying
  17257. information processing operations and how are they implemented in
  17258. the brain? Although a range of interesting possibilities exists,
  17259. recent work has introduced a potentially transformative new idea,
  17260. according to which planning is accomplished through probabilistic
  17261. inference.",
  17262. journal = "Trends Cogn. Sci.",
  17263. volume = 16,
  17264. number = 10,
  17265. pages = "485--488",
  17266. month = oct,
  17267. year = 2012,
  17268. language = "en"
  17269. }
  17270. @ARTICLE{Kraus2013-yt,
  17271. title = "Hippocampal ``time cells'': time versus path integration",
  17272. author = "Kraus, Benjamin J and Robinson, 2nd, Robert J and White, John A
  17273. and Eichenbaum, Howard and Hasselmo, Michael E",
  17274. abstract = "Recent studies have reported the existence of hippocampal ``time
  17275. cells,'' neurons that fire at particular moments during periods
  17276. when behavior and location are relatively constant. However, an
  17277. alternative explanation of apparent time coding is that
  17278. hippocampal neurons ``path integrate'' to encode the distance an
  17279. animal has traveled. Here, we examined hippocampal neuronal
  17280. firing patterns as rats ran in place on a treadmill, thus
  17281. ``clamping'' behavior and location, while we varied the treadmill
  17282. speed to distinguish time elapsed from distance traveled.
  17283. Hippocampal neurons were strongly influenced by time and
  17284. distance, and less so by minor variations in location.
  17285. Furthermore, the activity of different neurons reflected
  17286. integration over time and distance to varying extents, with most
  17287. neurons strongly influenced by both factors and some
  17288. significantly influenced by only time or distance. Thus,
  17289. hippocampal neuronal networks captured both the organization of
  17290. time and distance in a situation where these dimensions dominated
  17291. an ongoing experience.",
  17292. journal = "Neuron",
  17293. volume = 78,
  17294. number = 6,
  17295. pages = "1090--1101",
  17296. month = jun,
  17297. year = 2013,
  17298. language = "en"
  17299. }
  17300. @ARTICLE{Smith2012-nb,
  17301. title = "Reversible online control of habitual behavior by optogenetic
  17302. perturbation of medial prefrontal cortex",
  17303. author = "Smith, Kyle S and Virkud, Arti and Deisseroth, Karl and Graybiel,
  17304. Ann M",
  17305. abstract = "Habits tend to form slowly but, once formed, can have great
  17306. stability. We probed these temporal characteristics of habitual
  17307. behaviors by intervening optogenetically in forebrain habit
  17308. circuits as rats performed well-ingrained habitual runs in a
  17309. T-maze. We trained rats to perform a maze habit, confirmed the
  17310. habitual behavior by devaluation tests, and then, during the maze
  17311. runs (ca. 3 s), we disrupted population activity in a small
  17312. region in the medial prefrontal cortex, the infralimbic cortex.
  17313. In accordance with evidence that this region is necessary for the
  17314. expression of habits, we found that this cortical disruption
  17315. blocked habitual behavior. Notably, however, this blockade of
  17316. habitual performance occurred on line, within an average of three
  17317. trials (ca. 9 s of inhibition), and as soon as during the first
  17318. trial (<3 s). During subsequent weeks of training, the rats
  17319. acquired a new behavioral pattern. When we again imposed the same
  17320. cortical perturbation, the rats regained the suppressed
  17321. maze-running that typified the original habit, and,
  17322. simultaneously, the more recently acquired habit was blocked.
  17323. These online changes occurred within an average of two trials
  17324. (ca. 6 s of infralimbic inhibition). Measured changes in
  17325. generalized performance ability and motivation to consume reward
  17326. were unaffected. This immediate toggling between breaking old
  17327. habits and returning to them demonstrates that even semiautomatic
  17328. behaviors are under cortical control and that this control occurs
  17329. online, second by second. These temporal characteristics define a
  17330. framework for uncovering cellular transitions between fixed and
  17331. flexible behaviors, and corresponding disturbances in
  17332. pathologies.",
  17333. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  17334. volume = 109,
  17335. number = 46,
  17336. pages = "18932--18937",
  17337. month = nov,
  17338. year = 2012,
  17339. language = "en"
  17340. }
  17341. @ARTICLE{Gruber2012-nm,
  17342. title = "Context, emotion, and the strategic pursuit of goals:
  17343. interactions among multiple brain systems controlling motivated
  17344. behavior",
  17345. author = "Gruber, Aaron J and McDonald, Robert J",
  17346. abstract = "Motivated behavior exhibits properties that change with
  17347. experience and partially dissociate among a number of brain
  17348. structures. Here, we review evidence from rodent experiments
  17349. demonstrating that multiple brain systems acquire information in
  17350. parallel and either cooperate or compete for behavioral control.
  17351. We propose a conceptual model of systems interaction wherein a
  17352. ventral emotional memory network involving ventral striatum (VS),
  17353. amygdala, ventral hippocampus, and ventromedial prefrontal cortex
  17354. triages behavioral responding to stimuli according to their
  17355. associated affective outcomes. This system engages autonomic and
  17356. postural responding (avoiding, ignoring, approaching) in
  17357. accordance with associated stimulus valence (negative, neutral,
  17358. positive), but does not engage particular operant responses.
  17359. Rather, this emotional system suppresses or invigorates actions
  17360. that are selected through competition between goal-directed
  17361. control involving dorsomedial striatum (DMS) and habitual control
  17362. involving dorsolateral striatum (DLS). The hippocampus provides
  17363. contextual specificity to the emotional system, and provides an
  17364. information rich input to the goal-directed system for navigation
  17365. and discriminations involving ambiguous contexts, complex sensory
  17366. configurations, or temporal ordering. The rapid acquisition and
  17367. high capacity for episodic associations in the emotional system
  17368. may unburden the more complex goal-directed system and reduce
  17369. interference in the habit system from processing contingencies of
  17370. neutral stimuli. Interactions among these systems likely involve
  17371. inhibitory mechanisms and neuromodulation in the striatum to form
  17372. a dominant response strategy. Innate traits, training methods,
  17373. and task demands contribute to the nature of these interactions,
  17374. which can include incidental learning in non-dominant systems.
  17375. Addition of these features to reinforcement learning models of
  17376. decision-making may better align theoretical predictions with
  17377. behavioral and neural correlates in animals.",
  17378. journal = "Front. Behav. Neurosci.",
  17379. volume = 6,
  17380. pages = "50",
  17381. month = aug,
  17382. year = 2012,
  17383. keywords = "Pavlovian-instrumental transfer; amygdala; dopamine; emotion;
  17384. hippocampus; inhibition; reinforcement learning; striatum",
  17385. language = "en"
  17386. }
  17387. @ARTICLE{Bornstein2011-xi,
  17388. title = "Multiplicity of control in the basal ganglia: computational roles
  17389. of striatal subregions",
  17390. author = "Bornstein, Aaron M and Daw, Nathaniel D",
  17391. abstract = "The basal ganglia, in particular the striatum, are central to
  17392. theories of behavioral control, and often identified as a seat of
  17393. action selection. Reinforcement learning (RL) models--which have
  17394. driven much recent experimental work on this region--cast
  17395. striatum as a dynamic controller, integrating sensory and
  17396. motivational information to construct efficient and enriching
  17397. behavioral policies. Befitting this informationally central role,
  17398. the BG sit at the nexus of multiple anatomical 'loops' of
  17399. synaptic projections, connecting a wide range of cortical and
  17400. subcortical structures. Numerous pioneering anatomical studies
  17401. conducted over the past several decades have meticulously
  17402. catalogued these loops, and labeled them according to the
  17403. inferred functions of the connected regions. The specific
  17404. cotermina of the projections are highly localized to several
  17405. different subregions of the striatum, leading to the suggestion
  17406. that these subregions perform complementary but distinct
  17407. functions. However, until recently, the dominant computational
  17408. framework outlined only a bipartite, dorsal/ventral, division of
  17409. striatum. We review recent computational and experimental
  17410. advances that argue for a more finely fractionated delineation.
  17411. In particular, experimental data provide extensive insight into
  17412. unique functions subserved by the dorsomedial striatum (DMS).
  17413. These functions appear to correspond well with theories of a
  17414. 'model-based' RL subunit, and may also shed light on the
  17415. suborganization of ventral striatum. Finally, we discuss the
  17416. limitations of these ideas and how they point the way toward
  17417. future refinements of neurocomputational theories of striatal
  17418. function, bringing them into contact with other areas of
  17419. computational theory and other regions of the brain.",
  17420. journal = "Curr. Opin. Neurobiol.",
  17421. volume = 21,
  17422. number = 3,
  17423. pages = "374--380",
  17424. month = jun,
  17425. year = 2011,
  17426. language = "en"
  17427. }
  17428. @ARTICLE{Cooper2006-gu,
  17429. title = "Hierarchical schemas and goals in the control of sequential
  17430. behavior",
  17431. author = "Cooper, Richard P and Shallice, Tim",
  17432. abstract = "Traditional accounts of sequential behavior assume that schemas
  17433. and goals play a causal role in the control of behavior. In
  17434. contrast, M. Botvinick and D. C. Plaut argued that, at least in
  17435. routine behavior, schemas and goals are epiphenomenal. The
  17436. authors evaluate the Botvinick and Plaut account by contrasting
  17437. the simple recurrent network model of Botvinick and Plaut with
  17438. their own more traditional hierarchically structured interactive
  17439. activation model (R. P. Cooper \& T. Shallice, 2000). The
  17440. authors present a range of arguments and additional simulations
  17441. that demonstrate theoretical and empirical difficulties for both
  17442. Botvinick and Plaut's model and their theoretical position. The
  17443. authors conclude that explicit hierarchically organized and
  17444. causally efficacious schema and goal representations are
  17445. required to provide an adequate account of the flexibility of
  17446. sequential behavior.",
  17447. journal = "Psychol. Rev.",
  17448. publisher = "psycnet.apa.org",
  17449. volume = 113,
  17450. number = 4,
  17451. pages = "887--916; discussion 917--31",
  17452. month = oct,
  17453. year = 2006,
  17454. language = "en"
  17455. }
  17456. @ARTICLE{Botvinick2006-ux,
  17457. title = "Such stuff as habits are made on: A reply to Cooper and Shallice
  17458. (2006)",
  17459. author = "Botvinick, Matthew M and Plaut, David C",
  17460. abstract = "The representations and mechanisms guiding everyday routine
  17461. sequential action remain incompletely understood. In recent
  17462. work, the authors proposed a computational model of routine
  17463. sequential behavior that took the form of a recurrent neural
  17464. network (M. Botvinick \& D. C. Plaut, 2004; see record
  17465. 2004-12248-005). Subsequently, R. P. Cooper and T. Shallice
  17466. (2006; see record 2006-12689-008) put forth a detailed critique
  17467. of that work, contrasting it with their own account, which
  17468. assumes a strict hierarchical processing system (R. P. Cooper \&
  17469. T. Shallice, 2000; see record 2000-03986-001). The authors
  17470. respond here to the main points of R. P. Cooper and T.
  17471. Shallice's (2006) critique. Although careful and constructive,
  17472. the arguments offered by R. P. Cooper and T. Shallice (2006)
  17473. mistook several superficial implementational issues for
  17474. fundamental theoretical ones, underestimated the computational
  17475. power of recurrent networks as a class, and in some ways
  17476. mischaracterized the relationship between the accounts they
  17477. compare. In responding to these points, the authors articulate
  17478. several key theoretical choices facing models of routine
  17479. sequential behavior. (PsycINFO Database Record (c) 2016 APA, all
  17480. rights reserved)",
  17481. journal = "Psychol. Rev.",
  17482. publisher = "psycnet.apa.org",
  17483. volume = 113,
  17484. number = 4,
  17485. pages = "917--927",
  17486. month = oct,
  17487. year = 2006
  17488. }
  17489. @ARTICLE{Botvinick2008-sh,
  17490. title = "Hierarchical models of behavior and prefrontal function",
  17491. author = "Botvinick, Matthew M",
  17492. abstract = "The recognition of hierarchical structure in human behavior was
  17493. one of the founding insights of the cognitive revolution.
  17494. Despite decades of research, however, the computational
  17495. mechanisms underlying hierarchically organized behavior are
  17496. still not fully understood. Recent findings from behavioral and
  17497. neuroscientific research have fueled a resurgence of interest in
  17498. the problem, inspiring a new generation of computational models.
  17499. In addition to developing some classic proposals, these models
  17500. also break fresh ground, teasing apart different forms of
  17501. hierarchical structure, placing a new focus on the issue of
  17502. learning and addressing recent findings concerning the
  17503. representation of behavioral hierarchies within the prefrontal
  17504. cortex. In addition to offering explanations for some key
  17505. aspects of behavior and functional neuroanatomy, the latest
  17506. models also pose new questions for empirical research.",
  17507. journal = "Trends Cogn. Sci.",
  17508. publisher = "Elsevier",
  17509. volume = 12,
  17510. number = 5,
  17511. pages = "201--208",
  17512. month = may,
  17513. year = 2008,
  17514. language = "en"
  17515. }
  17516. @ARTICLE{Dayan1995-ai,
  17517. title = "The Helmholtz machine",
  17518. author = "Dayan, P and Hinton, G E and Neal, R M and Zemel, R S",
  17519. abstract = "Discovering the structure inherent in a set of patterns is a
  17520. fundamental aim of statistical inference or learning. One
  17521. fruitful approach is to build a parameterized stochastic
  17522. generative model, independent draws from which are likely to
  17523. produce the patterns. For all but the simplest generative
  17524. models, each pattern can be generated in exponentially many
  17525. ways. It is thus intractable to adjust the parameters to
  17526. maximize the probability of the observed patterns. We describe a
  17527. way of finessing this combinatorial explosion by maximizing an
  17528. easily computed lower bound on the probability of the
  17529. observations. Our method can be viewed as a form of hierarchical
  17530. self-supervised learning that may relate to the function of
  17531. bottom-up and top-down cortical processing pathways.",
  17532. journal = "Neural Comput.",
  17533. publisher = "MIT Press",
  17534. volume = 7,
  17535. number = 5,
  17536. pages = "889--904",
  17537. month = sep,
  17538. year = 1995,
  17539. language = "en"
  17540. }
  17541. @ARTICLE{Gatto2018-oi,
  17542. title = "Locomotion Control: Brainstem Circuits Satisfy the Need for Speed",
  17543. author = "Gatto, Graziana and Goulding, Martyn",
  17544. abstract = "Three new and closely complementary studies have defined the
  17545. architecture of the circuits underlying the descending control of
  17546. locomotion, identifying neurons that drive fast motor responses
  17547. and those that seem to be specialised for exploratory behaviors.",
  17548. journal = "Curr. Biol.",
  17549. volume = 28,
  17550. number = 6,
  17551. pages = "R256--R259",
  17552. month = mar,
  17553. year = 2018,
  17554. keywords = "Locomotion",
  17555. language = "en"
  17556. }
  17557. @ARTICLE{Corrado2007-ds,
  17558. title = "Understanding neural coding through the model-based analysis of
  17559. decision making",
  17560. author = "Corrado, Greg and Doya, Kenji",
  17561. abstract = "The study of decision making poses new methodological challenges
  17562. for systems neuroscience. Whereas our traditional approach linked
  17563. neural activity to external variables that the experimenter
  17564. directly observed and manipulated, many of the key elements that
  17565. contribute to decisions are internal to the decider. Variables
  17566. such as subjective value or subjective probability may be
  17567. influenced by experimental conditions and manipulations but can
  17568. neither be directly measured nor precisely controlled. Pioneering
  17569. work on the neural basis of decision circumvented this difficulty
  17570. by studying behavior in static conditions, in which knowledge of
  17571. the average state of these quantities was sufficient. More
  17572. recently, a new wave of studies has confronted the conundrum of
  17573. internal decision variables more directly by leveraging
  17574. quantitative behavioral models. When these behavioral models are
  17575. successful in predicting a subject's choice, the model's internal
  17576. variables may serve as proxies for the unobservable decision
  17577. variables that actually drive behavior. This new methodology has
  17578. allowed researchers to localize neural subsystems that encode
  17579. hidden decision variables related to free choice and to study
  17580. these variables under dynamic conditions.",
  17581. journal = "J. Neurosci.",
  17582. volume = 27,
  17583. number = 31,
  17584. pages = "8178--8180",
  17585. month = aug,
  17586. year = 2007,
  17587. language = "en"
  17588. }
  17589. @ARTICLE{Hikosaka2010-ql,
  17590. title = "The habenula: from stress evasion to value-based decision-making",
  17591. author = "Hikosaka, Okihide",
  17592. abstract = "Surviving in a world with hidden rewards and dangers requires
  17593. choosing the appropriate behaviours. Recent discoveries indicate
  17594. that the habenula plays a prominent part in such behavioural
  17595. choice through its effects on neuromodulator systems, in
  17596. particular the dopamine and serotonin systems. By inhibiting
  17597. dopamine-releasing neurons, habenula activation leads to the
  17598. suppression of motor behaviour when an animal fails to obtain a
  17599. reward or anticipates an aversive outcome. Moreover, the habenula
  17600. is involved in behavioural responses to pain, stress, anxiety,
  17601. sleep and reward, and its dysfunction is associated with
  17602. depression, schizophrenia and drug-induced psychosis. As a highly
  17603. conserved structure in the brain, the habenula provides a
  17604. fundamental mechanism for both survival and decision-making.",
  17605. journal = "Nat. Rev. Neurosci.",
  17606. volume = 11,
  17607. number = 7,
  17608. pages = "503--513",
  17609. month = jul,
  17610. year = 2010,
  17611. language = "en"
  17612. }
  17613. @ARTICLE{Balleine2007-qv,
  17614. title = "The role of the dorsal striatum in reward and decision-making",
  17615. author = "Balleine, Bernard W and Delgado, Mauricio R and Hikosaka, Okihide",
  17616. abstract = "Although the involvement in the striatum in the refinement and
  17617. control of motor movement has long been recognized, recent
  17618. description of discrete frontal corticobasal ganglia networks in
  17619. a range of species has focused attention on the role particularly
  17620. of the dorsal striatum in executive functions. Current evidence
  17621. suggests that the dorsal striatum contributes directly to
  17622. decision-making, especially to action selection and initiation,
  17623. through the integration of sensorimotor, cognitive, and
  17624. motivational/emotional information within specific
  17625. corticostriatal circuits involving discrete regions of striatum.
  17626. We review key evidence from recent studies in rodent, nonhuman
  17627. primate, and human subjects.",
  17628. journal = "J. Neurosci.",
  17629. volume = 27,
  17630. number = 31,
  17631. pages = "8161--8165",
  17632. month = aug,
  17633. year = 2007,
  17634. language = "en"
  17635. }
  17636. @ARTICLE{Hanks2017-mu,
  17637. title = "Perceptual Decision Making in Rodents, Monkeys, and Humans",
  17638. author = "Hanks, Timothy D and Summerfield, Christopher",
  17639. abstract = "Perceptual decision making is the process by which animals
  17640. detect, discriminate, and categorize information from the senses.
  17641. Over the past two decades, understanding how perceptual decisions
  17642. are made has become a central theme in the neurosciences.
  17643. Exceptional progress has been made by recording from single
  17644. neurons in the cortex of the macaque monkey and using
  17645. computational models from mathematical psychology to relate these
  17646. neural data to behavior. More recently, however, the range of
  17647. available techniques and paradigms has dramatically broadened,
  17648. and researchers have begun to harness new approaches to explore
  17649. how rodents and humans make perceptual decisions. The results
  17650. have illustrated some striking convergences with findings from
  17651. the monkey, but also raised new questions and provided new
  17652. theoretical insights. In this review, we summarize key findings,
  17653. and highlight open challenges, for understanding perceptual
  17654. decision making in rodents, monkeys, and humans.",
  17655. journal = "Neuron",
  17656. volume = 93,
  17657. number = 1,
  17658. pages = "15--31",
  17659. month = jan,
  17660. year = 2017,
  17661. keywords = "confidence; decision making; functional neuroimaging; human;
  17662. non-human primate; parietal cortex; psychophysics; rodent;
  17663. single-cell recordings",
  17664. language = "en"
  17665. }
  17666. @ARTICLE{Dalley2004-ta,
  17667. title = "Prefrontal executive and cognitive functions in rodents: neural
  17668. and neurochemical substrates",
  17669. author = "Dalley, Jeffrey W and Cardinal, Rudolf N and Robbins, Trevor W",
  17670. abstract = "The prefrontal cortex has been implicated in a variety of
  17671. cognitive and executive processes, including working memory,
  17672. decision-making, inhibitory response control, attentional
  17673. set-shifting and the temporal integration of voluntary behaviour.
  17674. This article reviews current progress in our understanding of the
  17675. rodent prefrontal cortex, especially evidence for functional
  17676. divergence of the anatomically distinct sub-regions of the rat
  17677. prefrontal cortex. Recent findings suggest clear distinctions
  17678. between the dorsal (precentral and anterior cingulate) and
  17679. ventral (prelimbic, infralimbic and medial orbital) sub-divisions
  17680. of the medial prefrontal cortex, and between the orbitofrontal
  17681. cortex (ventral orbital, ventrolateral orbital, dorsal and
  17682. ventral agranular cortices) and the adjacent medial wall of the
  17683. prefrontal cortex. The dorso-medial prefrontal cortex is
  17684. implicated in memory for motor responses, including response
  17685. selection, and the temporal processing of information. Ventral
  17686. regions of the medial prefrontal cortex are implicated in
  17687. interrelated 'supervisory' attentional functions, including
  17688. attention to stimulus features and task contingencies (or
  17689. action-outcome rules), attentional set-shifting, and behavioural
  17690. flexibility. The orbitofrontal cortex is implicated in
  17691. lower-order discriminations, including reversal of
  17692. stimulus-reward associations (reversal learning), and choice
  17693. involving delayed reinforcement. It is anticipated that a greater
  17694. understanding of the prefrontal cortex will come from using tasks
  17695. that load specific cognitive and executive processes, in parallel
  17696. with discovering new ways of manipulating the different
  17697. sub-regions and neuromodulatory systems of the prefrontal cortex.",
  17698. journal = "Neurosci. Biobehav. Rev.",
  17699. volume = 28,
  17700. number = 7,
  17701. pages = "771--784",
  17702. month = nov,
  17703. year = 2004,
  17704. language = "en"
  17705. }
  17706. @ARTICLE{Coutlee2012-hy,
  17707. title = "The functional neuroanatomy of decision making: prefrontal
  17708. control of thought and action",
  17709. author = "Coutlee, Christopher G and Huettel, Scott A",
  17710. abstract = "Humans exhibit a remarkable capacity for flexible thought and
  17711. action. Despite changing internal needs and external context,
  17712. individuals maintain stable goals and pursue purposeful action.
  17713. Functional neuroimaging research examining the neural
  17714. underpinnings of such behavioral flexibility has progressed
  17715. within several distinct traditions, as evident in the largely
  17716. separate literatures on ``cognitive control'' and on ``decision
  17717. making.'' Both topics investigate the formulation of desires and
  17718. intentions, the integration of knowledge and context, and the
  17719. resolution of conflict and uncertainty. Additionally, each
  17720. recognizes the fundamental role of the prefrontal cortex in
  17721. supporting flexible selection of behavior. But despite this
  17722. notable overlap, neuroimaging studies in cognitive control and
  17723. decision making have exerted only limited influence on each
  17724. other, in part due to differences in their theoretical and
  17725. experimental groundings. Additionally, the precise organization
  17726. of control processing within prefrontal cortex has remained
  17727. unclear, fostering an acceptance of vague descriptions of
  17728. decision making in terms of canonical cognitive control functions
  17729. such as ``inhibition'' or ``self-control.'' We suggest a unifying
  17730. role for models of the hierarchical organization of action
  17731. selection within prefrontal cortex. These models provide an
  17732. important conceptual link between decision-making phenomena and
  17733. cognitive-control processes, potentially facilitating
  17734. cross-fertilization between these topics.",
  17735. journal = "Brain Res.",
  17736. volume = 1428,
  17737. pages = "3--12",
  17738. month = jan,
  17739. year = 2012,
  17740. language = "en"
  17741. }
  17742. @ARTICLE{Clark2004-oz,
  17743. title = "The neuropsychology of ventral prefrontal cortex: decision-making
  17744. and reversal learning",
  17745. author = "Clark, L and Cools, R and Robbins, T W",
  17746. abstract = "Converging evidence from human lesion, animal lesion, and human
  17747. functional neuroimaging studies implicates overlapping neural
  17748. circuitry in ventral prefrontal cortex in decision-making and
  17749. reversal learning. The ascending 5-HT and dopamine
  17750. neurotransmitter systems have a modulatory role in both
  17751. processes. There is accumulating evidence that measures of
  17752. decision-making and reversal learning may be useful as functional
  17753. markers of ventral prefrontal cortex integrity in psychiatric and
  17754. neurological disorders. Whilst existing measures of
  17755. decision-making may have superior sensitivity, reversal learning
  17756. may offer superior selectivity, particularly within prefrontal
  17757. cortex. Effective decision-making on existing measures requires
  17758. the ability to adapt behaviour on the basis of changes in
  17759. emotional significance, and this may underlie the shared neural
  17760. circuitry with reversal learning.",
  17761. journal = "Brain Cogn.",
  17762. volume = 55,
  17763. number = 1,
  17764. pages = "41--53",
  17765. month = jun,
  17766. year = 2004,
  17767. language = "en"
  17768. }
  17769. @ARTICLE{Doya2008-qf,
  17770. title = "Modulators of decision making",
  17771. author = "Doya, Kenji",
  17772. abstract = "Human and animal decisions are modulated by a variety of
  17773. environmental and intrinsic contexts. Here I consider
  17774. computational factors that can affect decision making and review
  17775. anatomical structures and neurochemical systems that are related
  17776. to contextual modulation of decision making. Expectation of a
  17777. high reward can motivate a subject to go for an action despite a
  17778. large cost, a decision that is influenced by dopamine in the
  17779. anterior cingulate cortex. Uncertainty of action outcomes can
  17780. promote risk taking and exploratory choices, in which
  17781. norepinephrine and the orbitofrontal cortex appear to be
  17782. involved. Predictable environments should facilitate
  17783. consideration of longer-delayed rewards, which depends on
  17784. serotonin in the dorsal striatum and dorsal prefrontal cortex.
  17785. This article aims to sort out factors that affect the process of
  17786. decision making from the viewpoint of reinforcement learning
  17787. theory and to bridge between such computational needs and their
  17788. neurophysiological substrates.",
  17789. journal = "Nat. Neurosci.",
  17790. volume = 11,
  17791. number = 4,
  17792. pages = "410--416",
  17793. month = apr,
  17794. year = 2008,
  17795. language = "en"
  17796. }
  17797. @ARTICLE{Frank2006-an,
  17798. title = "Hold your horses: a dynamic computational role for the
  17799. subthalamic nucleus in decision making",
  17800. author = "Frank, Michael J",
  17801. abstract = "The basal ganglia (BG) coordinate decision making processes by
  17802. facilitating adaptive frontal motor commands while suppressing
  17803. others. In previous work, neural network simulations accounted
  17804. for response selection deficits associated with BG dopamine
  17805. depletion in Parkinson's disease. Novel predictions from this
  17806. model have been subsequently confirmed in Parkinson patients and
  17807. in healthy participants under pharmacological challenge.
  17808. Nevertheless, one clear limitation of that model is in its
  17809. omission of the subthalamic nucleus (STN), a key BG structure
  17810. that participates in both motor and cognitive processes. The
  17811. present model incorporates the STN and shows that by modulating
  17812. when a response is executed, the STN reduces premature responding
  17813. and therefore has substantial effects on which response is
  17814. ultimately selected, particularly when there are multiple
  17815. competing responses. Increased cortical response conflict leads
  17816. to dynamic adjustments in response thresholds via
  17817. cortico-subthalamic-pallidal pathways. The model accurately
  17818. captures the dynamics of activity in various BG areas during
  17819. response selection. Simulated dopamine depletion results in
  17820. emergent oscillatory activity in BG structures, which has been
  17821. linked with Parkinson's tremor. Finally, the model accounts for
  17822. the beneficial effects of STN lesions on these oscillations, but
  17823. suggests that this benefit may come at the expense of impaired
  17824. decision making.",
  17825. journal = "Neural Netw.",
  17826. volume = 19,
  17827. number = 8,
  17828. pages = "1120--1136",
  17829. month = oct,
  17830. year = 2006,
  17831. language = "en"
  17832. }
  17833. % The entry below contains non-ASCII chars that could not be converted
  17834. % to a LaTeX equivalent.
  17835. @ARTICLE{Olson2017-vi,
  17836. title = "Subiculum neurons map the current axis of travel",
  17837. author = "Olson, Jacob M and Tongprasearth, Kanyanat and Nitz, Douglas A",
  17838. abstract = "Flexible navigation demands knowledge of boundaries, routes and
  17839. their relationships. Within a multi-path environment, a
  17840. subpopulation of subiculum neurons robustly encoded the axis of
  17841. travel. The firing of axis-tuned neurons peaked bimodally, at
  17842. head orientations 180° apart. Environmental manipulations showed
  17843. these neurons to be anchored to environmental boundaries but to
  17844. lack axis tuning in an open arena. Axis-tuned neurons thus
  17845. provide a powerful mechanism for mapping relationships between
  17846. routes and the larger environmental context.",
  17847. journal = "Nat. Neurosci.",
  17848. volume = 20,
  17849. number = 2,
  17850. pages = "170--172",
  17851. month = feb,
  17852. year = 2017,
  17853. language = "en"
  17854. }
  17855. % The entry below contains non-ASCII chars that could not be converted
  17856. % to a LaTeX equivalent.
  17857. @UNPUBLISHED{Lee2019-qv,
  17858. title = "The statistical structure of the hippocampal code for space as a
  17859. function of time, context, and value",
  17860. author = "Lee, Jae Sung and Briguglio, John and Romani, Sandro and Lee,
  17861. Albert K",
  17862. abstract = "Hippocampal activity represents many behaviorally important
  17863. variables, including context, an animal9s location within a given
  17864. environmental context, time, and reward. Here we used
  17865. longitudinal calcium imaging in mice, multiple large virtual
  17866. environments, and differing reward contingencies to derive a
  17867. unified probabilistic model of hippocampal CA1 representations
  17868. centered on a single feature − the field propensity. Each cell9s
  17869. propensity governs how many place fields it has per unit space,
  17870. predicts its reward−related activity, and is preserved across
  17871. distinct environments and over months. The propensity is broadly
  17872. distributed−with many low, and some very high, propensity cells
  17873. −and thus strongly shapes hippocampal representations. The result
  17874. is a range of spatial codes, from sparse to dense. Propensity
  17875. varied ~10−fold between adjacent cells in a salt-and-pepper
  17876. fashion, indicating substantial functional differences within a
  17877. presumed cell type. The stability of each cell9s propensity
  17878. across conditions suggests this fundamental property has
  17879. anatomical, transcriptional, and/or developmental origins.",
  17880. journal = "bioRxiv",
  17881. pages = "615203",
  17882. month = apr,
  17883. year = 2019,
  17884. language = "en"
  17885. }
  17886. @UNPUBLISHED{Pisupati2019-mc,
  17887. title = "Lapses in perceptual judgments reflect exploration",
  17888. author = "Pisupati, Sashank and Chartarifsky-Lynn, Lital and Khanal, Anup
  17889. and Churchland, Anne K",
  17890. abstract = "During perceptual decision making, subjects often display a
  17891. constant rate of errors independent of evidence strength,
  17892. referred to as lapses. Their proper treatment is crucial for
  17893. accurate estimation of perceptual parameters, however they are
  17894. often treated as a nuisance arising from motor errors or
  17895. inattention. Here, we propose that lapses can instead reflect a
  17896. dynamic form of exploration. We demonstrate that perceptual
  17897. uncertainty modulates the probability of lapses both across and
  17898. within modalities on a multisensory discrimination task in rats.
  17899. These effects cannot be accounted for by inattention or motor
  17900. error, however they are concisely explained by uncertainty-guided
  17901. exploration. We confirm the predictions of the exploration model
  17902. by showing that changing the magnitude or probability of reward
  17903. associated with one of the decisions selectively affects the
  17904. lapses associated with that decision in uncertain conditions,
  17905. while leaving sure-bet decisions unchanged, as predicted by the
  17906. model. Finally, we demonstrate that muscimol inactivations of
  17907. secondary motor cortex and posterior striatum affect lapses
  17908. asymmetrically across modalities. The inactivations can be
  17909. captured by a devaluation of actions corresponding to the
  17910. inactivated side, and do not affect sure-bet decisions. Together,
  17911. our results suggest that far from being a nuisance, lapses are
  17912. informative about subjects9 action values, and deficits thereof,
  17913. during perceptual decisions.",
  17914. journal = "bioRxiv",
  17915. pages = "613828",
  17916. month = apr,
  17917. year = 2019,
  17918. language = "en"
  17919. }
  17920. @ARTICLE{De_Cheveigne2019-ab,
  17921. title = "Filters: When, Why, and How (Not) to Use Them",
  17922. author = "de Cheveign{\'e}, Alain and Nelken, Israel",
  17923. abstract = "Filters are commonly used to reduce noise and improve data
  17924. quality. Filter theory is part of a scientist's training, yet the
  17925. impact of filters on interpreting data is not always fully
  17926. appreciated. This paper reviews the issue and explains what a
  17927. filter is, what problems are to be expected when using them, how
  17928. to choose the right filter, and how to avoid filtering by using
  17929. alternative tools. Time-frequency analysis shares some of the
  17930. same problems that filters have, particularly in the case of
  17931. wavelet transforms. We recommend reporting filter characteristics
  17932. with sufficient details, including a plot of the impulse or step
  17933. response as an inset.",
  17934. journal = "Neuron",
  17935. volume = 102,
  17936. number = 2,
  17937. pages = "280--293",
  17938. month = apr,
  17939. year = 2019,
  17940. keywords = "Fourier analysis; causality; distortions; filter; impulse
  17941. response; oscillations; ringing; time-frequency representation",
  17942. language = "en"
  17943. }
  17944. @ARTICLE{Berridge2004-eo,
  17945. title = "Motivation concepts in behavioral neuroscience",
  17946. author = "Berridge, Kent C",
  17947. abstract = "Concepts of motivation are vital to progress in behavioral
  17948. neuroscience. Motivational concepts help us to understand what
  17949. limbic brain systems are chiefly evolved to do, i.e., to mediate
  17950. psychological processes that guide real behavior. This article
  17951. evaluates some major motivation concepts that have historic
  17952. importance or have influenced the interpretation of behavioral
  17953. neuroscience research. These concepts include homeostasis,
  17954. setpoints and settling points, intervening variables, hydraulic
  17955. drives, drive reduction, appetitive and consummatory behavior,
  17956. opponent processes, hedonic reactions, incentive motivation,
  17957. drive centers, dedicated drive neurons (and drive neuropeptides
  17958. and receptors), neural hierarchies, and new concepts from
  17959. affective neuroscience such as allostasis, cognitive incentives,
  17960. and reward 'liking' versus 'wanting'.",
  17961. journal = "Physiol. Behav.",
  17962. publisher = "Elsevier",
  17963. volume = 81,
  17964. number = 2,
  17965. pages = "179--209",
  17966. month = apr,
  17967. year = 2004,
  17968. language = "en"
  17969. }
  17970. @ARTICLE{Gottlieb2018-jh,
  17971. title = "Towards a neuroscience of active sampling and curiosity",
  17972. author = "Gottlieb, Jacqueline and Oudeyer, Pierre-Yves",
  17973. abstract = "In natural behaviour, animals actively interrogate their
  17974. environments using endogenously generated 'question-and-answer'
  17975. strategies. However, in laboratory settings participants
  17976. typically engage with externally imposed stimuli and tasks, and
  17977. the mechanisms of active sampling remain poorly understood. We
  17978. review a nascent neuroscientific literature that examines
  17979. active-sampling policies and their relation to attention and
  17980. curiosity. We distinguish between information sampling, in which
  17981. organisms reduce uncertainty relevant to a familiar task, and
  17982. information search, in which they investigate in an open-ended
  17983. fashion to discover new tasks. We review evidence that both
  17984. sampling and search depend on individual preferences over
  17985. cognitive states, including attitudes towards uncertainty,
  17986. learning progress and types of information. We propose that,
  17987. although these preferences are non-instrumental and can on
  17988. occasion interfere with external goals, they are important
  17989. heuristics that allow organisms to cope with the high complexity
  17990. of both sampling and search, and generate curiosity-driven
  17991. investigations in large, open environments in which rewards are
  17992. sparse and ex ante unknown.",
  17993. journal = "Nat. Rev. Neurosci.",
  17994. publisher = "nature.com",
  17995. volume = 19,
  17996. number = 12,
  17997. pages = "758--770",
  17998. month = dec,
  17999. year = 2018,
  18000. language = "en"
  18001. }
  18002. @ARTICLE{Dayan2008-jq,
  18003. title = "Decision theory, reinforcement learning, and the brain",
  18004. author = "Dayan, Peter and Daw, Nathaniel D",
  18005. abstract = "Decision making is a core competence for animals and humans
  18006. acting and surviving in environments they only partially
  18007. comprehend, gaining rewards and punishments for their troubles.
  18008. Decision-theoretic concepts permeate experiments and
  18009. computational models in ethology, psychology, and neuroscience.
  18010. Here, we review a well-known, coherent Bayesian approach to
  18011. decision making, showing how it unifies issues in Markovian
  18012. decision problems, signal detection psychophysics, sequential
  18013. sampling, and optimal exploration and discuss paradigmatic
  18014. psychological and neural examples of each problem. We discuss
  18015. computational issues concerning what subjects know about their
  18016. task and how ambitious they are in seeking optimal solutions; we
  18017. address algorithmic topics concerning model-based and model-free
  18018. methods for making choices; and we highlight key aspects of the
  18019. neural implementation of decision making.",
  18020. journal = "Cogn. Affect. Behav. Neurosci.",
  18021. publisher = "Springer",
  18022. volume = 8,
  18023. number = 4,
  18024. pages = "429--453",
  18025. month = dec,
  18026. year = 2008,
  18027. language = "en"
  18028. }
  18029. @ARTICLE{Jiang2019-in,
  18030. title = "{Short-Term} Influence of Recent Trial History on Perceptual
  18031. Choice Changes with Stimulus Strength",
  18032. author = "Jiang, Weiqian and Liu, Jing and Zhang, Dinghong and Xie, Taorong
  18033. and Yao, Haishan",
  18034. abstract = "Perceptual decisions, especially for difficult stimuli, can be
  18035. influenced by choices and outcomes in previous trials. However,
  18036. it is not well understood how stimulus strength modulates the
  18037. temporal characteristics as well as the magnitude of trial
  18038. history influence. We addressed this question using a contrast
  18039. detection task in freely moving mice. We found that, at lower as
  18040. compared to higher stimulus contrast, the current choice of the
  18041. mice was more influenced by choices and outcomes in the past
  18042. trials and the influence emerged from a longer history. To
  18043. examine the neural basis of stimulus strength-dependent history
  18044. influence, we recorded from the secondary motor cortex (M2), a
  18045. prefrontal region that plays an important role in cue-guided
  18046. actions and memory-guided behaviors. We found that more M2
  18047. neurons conveyed information about choices on the past two trials
  18048. at lower than at higher contrast. Furthermore, history-trial
  18049. activity in M2 was important for decoding upcoming choice at low
  18050. contrast. Thus, trial history influence of perceptual choice is
  18051. adaptive to the strength of sensory evidence, which may be
  18052. important for action selection in a dynamic environment.",
  18053. journal = "Neuroscience",
  18054. month = apr,
  18055. year = 2019,
  18056. keywords = "choice history; contrast; decision making; rodent; secondary
  18057. motor cortex",
  18058. language = "en"
  18059. }
  18060. @ARTICLE{Lintz2019-sx,
  18061. title = "Spatial Representations in the Superior Colliculus Are Modulated
  18062. by Competition among Targets",
  18063. author = "Lintz, Mario J and Essig, Jaclyn and Zylberberg, Joel and Felsen,
  18064. Gidon",
  18065. abstract = "Selecting and moving to spatial targets are critical components
  18066. of goal-directed behavior, yet their neural bases are not well
  18067. understood. The superior colliculus (SC) is thought to contain a
  18068. topographic map of contralateral space in which the activity of
  18069. specific neuronal populations corresponds to particular spatial
  18070. locations. However, these spatial representations are modulated
  18071. by several decision-related variables, suggesting that they
  18072. reflect information beyond simply the location of an upcoming
  18073. movement. Here, we examine the extent to which these
  18074. representations arise from competitive spatial choice. We
  18075. recorded SC activity in male mice performing a behavioral task
  18076. requiring orienting movements to targets for a water reward in
  18077. two contexts. In ``competitive'' trials, either the left or right
  18078. target could be rewarded, depending on which stimulus was
  18079. presented at the central port. In ``noncompetitive'' trials, the
  18080. same target (e.g., left) was rewarded throughout an entire block.
  18081. While both trial types required orienting movements to the same
  18082. spatial targets, only in competitive trials do targets compete
  18083. for selection. We found that in competitive trials, pre-movement
  18084. SC activity predicted movement to contralateral targets, as
  18085. expected. However, in noncompetitive trials, some neurons lost
  18086. their spatial selectivity and in others activity predicted
  18087. movement to ipsilateral targets. Consistent with these findings,
  18088. unilateral optogenetic inactivation of pre-movement SC activity
  18089. ipsiversively biased competitive, but not noncompetitive, trials.
  18090. Incorporating these results into an attractor model of SC
  18091. activity points to distinct pathways for orienting movements
  18092. under competitive and noncompetitive conditions, with the SC
  18093. specifically required for selecting among multiple potential
  18094. targets.",
  18095. journal = "Neuroscience",
  18096. month = apr,
  18097. year = 2019,
  18098. keywords = "Superior colliculus; decision making; freely-moving mice; target
  18099. selection",
  18100. language = "en"
  18101. }
  18102. @ARTICLE{Churchland2008-av,
  18103. title = "Decision-making with multiple alternatives",
  18104. author = "Churchland, Anne K and Kiani, Roozbeh and Shadlen, Michael N",
  18105. abstract = "Simple perceptual tasks have laid the groundwork for
  18106. understanding the neurobiology of decision-making. Here, we
  18107. examined this foundation to explain how decision-making
  18108. circuitry adjusts in the face of a more difficult task. We
  18109. measured behavioral and physiological responses of monkeys on a
  18110. two- and four-choice direction-discrimination decision task. For
  18111. both tasks, firing rates in the lateral intraparietal area
  18112. appeared to reflect the accumulation of evidence for or against
  18113. each choice. Evidence accumulation began at a lower firing rate
  18114. for the four-choice task, but reached a common level by the end
  18115. of the decision process. The larger excursion suggests that the
  18116. subjects required more evidence before making a choice.
  18117. Furthermore, on both tasks, we observed a time-dependent rise in
  18118. firing rates that may impose a deadline for deciding. These
  18119. physiological observations constitute an effective strategy for
  18120. handling increased task difficulty. The differences appear to
  18121. explain subjects' accuracy and reaction times.",
  18122. journal = "Nat. Neurosci.",
  18123. publisher = "nature.com",
  18124. volume = 11,
  18125. number = 6,
  18126. pages = "693--702",
  18127. month = jun,
  18128. year = 2008,
  18129. language = "en"
  18130. }
  18131. @ARTICLE{Louie2014-lo,
  18132. title = "Dynamic divisive normalization predicts time-varying value
  18133. coding in decision-related circuits",
  18134. author = "Louie, Kenway and LoFaro, Thomas and Webb, Ryan and Glimcher,
  18135. Paul W",
  18136. abstract = "Normalization is a widespread neural computation, mediating
  18137. divisive gain control in sensory processing and implementing a
  18138. context-dependent value code in decision-related frontal and
  18139. parietal cortices. Although decision-making is a dynamic process
  18140. with complex temporal characteristics, most models of
  18141. normalization are time-independent and little is known about the
  18142. dynamic interaction of normalization and choice. Here, we show
  18143. that a simple differential equation model of normalization
  18144. explains the characteristic phasic-sustained pattern of cortical
  18145. decision activity and predicts specific normalization dynamics:
  18146. value coding during initial transients, time-varying value
  18147. modulation, and delayed onset of contextual information.
  18148. Empirically, we observe these predicted dynamics in
  18149. saccade-related neurons in monkey lateral intraparietal cortex.
  18150. Furthermore, such models naturally incorporate a time-weighted
  18151. average of past activity, implementing an intrinsic
  18152. reference-dependence in value coding. These results suggest that
  18153. a single network mechanism can explain both transient and
  18154. sustained decision activity, emphasizing the importance of a
  18155. dynamic view of normalization in neural coding.",
  18156. journal = "J. Neurosci.",
  18157. publisher = "Soc Neuroscience",
  18158. volume = 34,
  18159. number = 48,
  18160. pages = "16046--16057",
  18161. month = nov,
  18162. year = 2014,
  18163. keywords = "computational modeling; decision-making; divisive normalization;
  18164. dynamical system; reward",
  18165. language = "en"
  18166. }
  18167. @ARTICLE{Schacter_Daniel_L2007-rc,
  18168. title = "The cognitive neuroscience of constructive memory: remembering
  18169. the past and imagining the future",
  18170. author = "{Schacter Daniel L} and {Addis Donna Rose}",
  18171. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  18172. publisher = "Royal Society",
  18173. volume = 362,
  18174. number = 1481,
  18175. pages = "773--786",
  18176. month = may,
  18177. year = 2007
  18178. }
  18179. @ARTICLE{Khamassi2005-rl,
  18180. title = "{Actor--Critic} Models of Reinforcement Learning in the Basal
  18181. Ganglia: From Natural to Artificial Rats",
  18182. author = "Khamassi, Mehdi and Lach{\`e}ze, Lo{\"\i}c and Girard,
  18183. Beno{\^\i}t and Berthoz, Alain and Guillot, Agn{\`e}s",
  18184. abstract = "Since 1995, numerous Actor?Critic architectures for
  18185. reinforcement learning have been proposed as models of
  18186. dopamine-like reinforcement learning mechanisms in the rat?s
  18187. basal ganglia. However, these models were usually tested in
  18188. different tasks, and it is then difficult to compare their
  18189. efficiency for an autonomous animat. We present here the
  18190. comparison of four architectures in an animat as it per forms
  18191. the same reward-seeking task. This will illustrate the
  18192. consequences of different hypotheses about the management of
  18193. different Actor sub-modules and Critic units, and their more or
  18194. less autono mously determined coordination. We show that the
  18195. classical method of coordination of modules by mixture of
  18196. experts, depending on each module?s performance, did not allow
  18197. solving our task. Then we address the question of which
  18198. principle should be applied efficiently to combine these units.
  18199. Improve ments for Critic modeling and accuracy of Actor?Critic
  18200. models for a natural task are finally discussed in the
  18201. perspective of our Psikharpax project?an artificial rat having
  18202. to survive autonomously in unpre dictable environments.",
  18203. journal = "Adapt. Behav.",
  18204. publisher = "SAGE Publications Ltd STM",
  18205. volume = 13,
  18206. number = 2,
  18207. pages = "131--148",
  18208. month = jun,
  18209. year = 2005
  18210. }
  18211. @ARTICLE{Joel2002-xo,
  18212. title = "Actor--critic models of the basal ganglia: new anatomical and
  18213. computational perspectives",
  18214. author = "Joel, Daphna and Niv, Yael and Ruppin, Eytan",
  18215. abstract = "A large number of computational models of information processing
  18216. in the basal ganglia have been developed in recent years.
  18217. Prominent in these are actor--critic models of basal ganglia
  18218. functioning, which build on the strong resemblance between
  18219. dopamine neuron activity and the temporal difference prediction
  18220. error signal in the critic, and between dopamine-dependent
  18221. long-term synaptic plasticity in the striatum and learning
  18222. guided by a prediction error signal in the actor. We selectively
  18223. review several actor--critic models of the basal ganglia with an
  18224. emphasis on two important aspects: the way in which models of
  18225. the critic reproduce the temporal dynamics of dopamine firing,
  18226. and the extent to which models of the actor take into account
  18227. known basal ganglia anatomy and physiology. To complement the
  18228. efforts to relate basal ganglia mechanisms to reinforcement
  18229. learning (RL), we introduce an alternative approach to modeling
  18230. a critic network, which uses Evolutionary Computation techniques
  18231. to `evolve' an optimal RL mechanism, and relate the evolved
  18232. mechanism to the basic model of the critic. We conclude our
  18233. discussion of models of the critic by a critical discussion of
  18234. the anatomical plausibility of implementations of a critic in
  18235. basal ganglia circuitry, and conclude that such implementations
  18236. build on assumptions that are inconsistent with the known
  18237. anatomy of the basal ganglia. We return to the actor component
  18238. of the actor--critic model, which is usually modeled at the
  18239. striatal level with very little detail. We describe an
  18240. alternative model of the basal ganglia which takes into account
  18241. several important, and previously neglected, anatomical and
  18242. physiological characteristics of basal ganglia--thalamocortical
  18243. connectivity and suggests that the basal ganglia performs
  18244. reinforcement-biased dimensionality reduction of cortical
  18245. inputs. We further suggest that since such selective encoding
  18246. may bias the representation at the level of the frontal cortex
  18247. towards the selection of rewarded plans and actions, the
  18248. reinforcement-driven dimensionality reduction framework may
  18249. serve as a basis for basal ganglia actor models. We conclude
  18250. with a short discussion of the dual role of the dopamine signal
  18251. in RL and in behavioral switching.",
  18252. journal = "Neural Netw.",
  18253. publisher = "Elsevier",
  18254. volume = 15,
  18255. number = 4,
  18256. pages = "535--547",
  18257. month = jun,
  18258. year = 2002,
  18259. keywords = "Basal ganglia; Dopamine; Reinforcement learning; Actor--critic;
  18260. Dimensionality reduction; Evolutionary computation; Behavioral
  18261. switching; Striosomes/patches"
  18262. }
  18263. @ARTICLE{Lee2014-ji,
  18264. title = "Neural computations underlying arbitration between model-based
  18265. and model-free learning",
  18266. author = "Lee, Sang Wan and Shimojo, Shinsuke and O'Doherty, John P",
  18267. abstract = "There is accumulating neural evidence to support the existence
  18268. of two distinct systems for guiding action selection, a
  18269. deliberative ``model-based'' and a reflexive ``model-free''
  18270. system. However, little is known about how the brain determines
  18271. which of these systems controls behavior at one moment in time.
  18272. We provide evidence for an arbitration mechanism that allocates
  18273. the degree of control over behavior by model-based and
  18274. model-free systems as a function of the reliability of their
  18275. respective predictions. We show that the inferior lateral
  18276. prefrontal and frontopolar cortex encode both reliability
  18277. signals and the output of a comparison between those signals,
  18278. implicating these regions in the arbitration process. Moreover,
  18279. connectivity between these regions and model-free valuation
  18280. areas is negatively modulated by the degree of model-based
  18281. control in the arbitrator, suggesting that arbitration may work
  18282. through modulation of the model-free valuation system when the
  18283. arbitrator deems that the model-based system should drive
  18284. behavior.",
  18285. journal = "Neuron",
  18286. publisher = "Elsevier",
  18287. volume = 81,
  18288. number = 3,
  18289. pages = "687--699",
  18290. month = feb,
  18291. year = 2014,
  18292. language = "en"
  18293. }
  18294. @ARTICLE{Papale2016-ml,
  18295. title = "Interplay between Hippocampal {Sharp-Wave-Ripple} Events and
  18296. Vicarious Trial and Error Behaviors in Decision Making",
  18297. author = "Papale, Andrew E and Zielinski, Mark C and Frank, Loren M and
  18298. Jadhav, Shantanu P and Redish, A David",
  18299. abstract = "Current theories posit that memories encoded during experiences
  18300. are subsequently consolidated into longer-term storage.
  18301. Hippocampal sharp-wave-ripple (SWR) events have been linked to
  18302. this consolidation process during sleep, but SWRs also occur
  18303. during awake immobility, where their role remains unclear. We
  18304. report that awake SWR rates at the reward site are inversely
  18305. related to the prevalence of vicarious trial and error (VTE)
  18306. behaviors, thought to be involved in deliberation processes. SWR
  18307. rates were diminished immediately after VTE behaviors and an
  18308. increase in the rate of SWR events at the reward site predicted
  18309. a decrease in subsequent VTE behaviors at the choice point.
  18310. Furthermore, SWR disruptions increased VTE behaviors. These
  18311. results suggest an inverse relationship between SWRs and VTE
  18312. behaviors and suggest that awake SWRs and associated planning
  18313. and memory consolidation mechanisms are engaged specifically in
  18314. the context of higher levels of behavioral certainty.",
  18315. journal = "Neuron",
  18316. publisher = "Elsevier",
  18317. volume = 92,
  18318. number = 5,
  18319. pages = "975--982",
  18320. month = dec,
  18321. year = 2016,
  18322. language = "en"
  18323. }
  18324. @ARTICLE{Van_der_Meer2010-ed,
  18325. title = "Expectancies in decision making, reinforcement learning, and
  18326. ventral striatum",
  18327. author = "van der Meer, Matthijs A A and Redish, A David",
  18328. abstract = "Decisions can arise in different ways, such as from a gut
  18329. feeling, doing what worked last time, or planful deliberation.
  18330. Different decision-making systems are dissociable behaviorally,
  18331. map onto distinct brain systems, and have different
  18332. computational demands. For instance, ``model-free'' decision
  18333. strategies use prediction errors to estimate scalar action
  18334. values from previous experience, while ``model-based''
  18335. strategies leverage internal forward models to generate and
  18336. evaluate potentially rich outcome expectancies. Animal learning
  18337. studies indicate that expectancies may arise from different
  18338. sources, including not only forward models but also Pavlovian
  18339. associations, and the flexibility with which such
  18340. representations impact behavior may depend on how they are
  18341. generated. In the light of these considerations, we review the
  18342. results of van der Meer and Redish (2009a), who found that
  18343. ventral striatal neurons that respond to reward delivery can
  18344. also be activated at other points, notably at a decision point
  18345. where hippocampal forward representations were also observed.
  18346. These data suggest the possibility that ventral striatal reward
  18347. representations contribute to model-based expectancies used in
  18348. deliberative decision making.",
  18349. journal = "Front. Neurosci.",
  18350. publisher = "frontiersin.org",
  18351. volume = 4,
  18352. pages = "6",
  18353. month = may,
  18354. year = 2010,
  18355. keywords = "Pavlovian-instrumental transfer; actor--critic; planning;
  18356. reinforcement learning; reward",
  18357. language = "en"
  18358. }
  18359. @ARTICLE{Alexander1990-rz,
  18360. title = "Functional architecture of basal ganglia circuits: neural
  18361. substrates of parallel processing",
  18362. author = "Alexander, G E and Crutcher, M D",
  18363. abstract = "Concepts of basal ganglia organization have changed markedly
  18364. over the past decade, due to significant advances in our
  18365. understanding of the anatomy, physiology and pharmacology of
  18366. these structures. Independent evidence from each of these fields
  18367. has reinforced a growing perception that the functional
  18368. architecture of the basal ganglia is essentially parallel in
  18369. nature, regardless of the perspective from which these
  18370. structures are viewed. This represents a significant departure
  18371. from earlier concepts of basal ganglia organization, which
  18372. generally emphasized the serial aspects of their connectivity.
  18373. Current evidence suggests that the basal ganglia are organized
  18374. into several structurally and functionally distinct 'circuits'
  18375. that link cortex, basal ganglia and thalamus, with each circuit
  18376. focused on a different portion of the frontal lobe. In this
  18377. review, Garrett Alexander and Michael Crutcher, using the basal
  18378. ganglia 'motor' circuit as the principal example, discuss recent
  18379. evidence indicating that a parallel functional architecture may
  18380. also be characteristic of the organization within each
  18381. individual circuit.",
  18382. journal = "Trends Neurosci.",
  18383. publisher = "Elsevier",
  18384. volume = 13,
  18385. number = 7,
  18386. pages = "266--271",
  18387. month = jul,
  18388. year = 1990,
  18389. language = "en"
  18390. }
  18391. @ARTICLE{Dickinson1994-td,
  18392. title = "Motivational control of goal-directed action",
  18393. author = "Dickinson, Anthony and Balleine, Bernard",
  18394. abstract = "The control of goal-directed, instrumental actions by primary
  18395. motivational states, such as hunger and thirst, is mediated by
  18396. two processes. The first is engaged by the Pavlovian association
  18397. between contextual or discriminative stimuli and the outcome or
  18398. reinforcer presented during instrumental training. Such stimuli
  18399. exert a motivational influence on instrumental performance that
  18400. depends upon the relevance of the associated outcome to the
  18401. current motivational state of the agent. Moreover, the
  18402. motivational effects of these stimuli operate in the absence of
  18403. prior experience with the outcome under the relevant
  18404. motivational state. The second, instrumental, process is
  18405. mediated by knowledge of the contingency between the action and
  18406. its outcome and controls the value assigned to this outcome. In
  18407. contrast to the Pavlovian process, motivational states do not
  18408. influence the instrumental process directly; rather, the agent
  18409. has to learn about the value of an outcome in a given
  18410. motivational state by exposure to it while in that state. This
  18411. incentive learning is similar in certain respects to the
  18412. acquisition of ``cathexes'' envisaged by Tolman (1949a, 1949b).",
  18413. journal = "Anim. Learn. Behav.",
  18414. publisher = "Springer",
  18415. volume = 22,
  18416. number = 1,
  18417. pages = "1--18",
  18418. month = mar,
  18419. year = 1994
  18420. }
  18421. % The entry below contains non-ASCII chars that could not be converted
  18422. % to a LaTeX equivalent.
  18423. @ARTICLE{Dickinson1985-pt,
  18424. title = "Actions and habits: the development of behavioural autonomy",
  18425. author = "Dickinson, Anthony",
  18426. abstract = "The study of animal behaviour has been dominated by two general
  18427. models. According to the mechanistic stimulus-response model, a
  18428. particular behaviour is either an innate or an acquired habit
  18429. which is simply triggered by the appropriate stimulus. By
  18430. contrast, the teleological model argues that, at least, some
  18431. activities are purposive actions controlled by the current value
  18432. of their goals through knowledge about the instrumental
  18433. relations between the actions and their consequences. The type
  18434. of control over any particular behaviour can …",
  18435. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  18436. publisher = "The Royal Society London",
  18437. volume = 308,
  18438. number = 1135,
  18439. pages = "67--78",
  18440. year = 1985
  18441. }
  18442. @ARTICLE{Dayan1993-rr,
  18443. title = "Improving Generalization for Temporal Difference Learning: The
  18444. Successor Representation",
  18445. author = "Dayan, Peter",
  18446. abstract = "Estimation of returns over time, the focus of temporal
  18447. difference (TD) algorithms, imposes particular constraints on
  18448. good function approximators or representations. Appropriate
  18449. generalization between states is determined by how similar their
  18450. successors are, and representations should follow suit. This
  18451. paper shows how TD machinery can be used to learn such
  18452. representations, and illustrates, using a navigation task, the
  18453. appropriately distributed nature of the result.",
  18454. journal = "Neural Comput.",
  18455. publisher = "MIT Press",
  18456. volume = 5,
  18457. number = 4,
  18458. pages = "613--624",
  18459. month = jul,
  18460. year = 1993
  18461. }
  18462. % The entry below contains non-ASCII chars that could not be converted
  18463. % to a LaTeX equivalent.
  18464. @ARTICLE{Ramachandran2007-tz,
  18465. title = "Bayesian Inverse Reinforcement Learning",
  18466. author = "Ramachandran, D and Amir, E",
  18467. abstract = "Abstract Inverse Reinforcement Learning (IRL) is the problem of
  18468. learning the reward function underlying a Markov Decision
  18469. Process given the dynamics of the system and the behaviour of an
  18470. expert. IRL is motivated by situations where knowledge of the
  18471. rewards is a goal by …",
  18472. journal = "IJCAI",
  18473. publisher = "aaai.org",
  18474. year = 2007
  18475. }
  18476. @ARTICLE{noauthor_undated-wn,
  18477. title = "icml00-irl.pdf"
  18478. }
  18479. @INCOLLECTION{Choi2011-ua,
  18480. title = "{MAP} Inference for Bayesian Inverse Reinforcement Learning",
  18481. booktitle = "Advances in Neural Information Processing Systems 24",
  18482. author = "Choi, Jaedeug and Kim, Kee-Eung",
  18483. editor = "Shawe-Taylor, J and Zemel, R S and Bartlett, P L and Pereira, F
  18484. and Weinberger, K Q",
  18485. publisher = "Curran Associates, Inc.",
  18486. pages = "1989--1997",
  18487. year = 2011
  18488. }
  18489. @ARTICLE{Burton2015-am,
  18490. title = "From ventral-medial to dorsal-lateral striatum: neural correlates
  18491. of reward-guided decision-making",
  18492. author = "Burton, Amanda C and Nakamura, Kae and Roesch, Matthew R",
  18493. abstract = "The striatum is critical for reward-guided and habitual behavior.
  18494. Anatomical and interference studies suggest a functional
  18495. heterogeneity within striatum. Medial regions, such as nucleus
  18496. accumbens core and dorsal medial striatum play roles in
  18497. goal-directed behavior, while dorsal lateral striatum is critical
  18498. for control of habitual action. Subdivisions of striatum are
  18499. topographically connected with different cortical and subcortical
  18500. structures forming channels that carry information related to
  18501. limbic, associative, and sensorimotor functions. Here, we
  18502. describe data showing that as one progresses from ventral-medial
  18503. to dorsal-lateral striatum, there is a shift from more prominent
  18504. value encoding to activity more closely related to associative
  18505. and motor aspects of decision-making. In addition, we will
  18506. describe data suggesting that striatal circuits work in parallel
  18507. to control behavior and that regions within striatum can
  18508. compensate for each other when functions are disrupted.",
  18509. journal = "Neurobiol. Learn. Mem.",
  18510. volume = 117,
  18511. pages = "51--59",
  18512. month = jan,
  18513. year = 2015,
  18514. keywords = "Goal; Habit; Monkey; Nucleus accumbens; Rat; Single unit;
  18515. Striatum; Value",
  18516. language = "en"
  18517. }
  18518. @ARTICLE{Guazzelli1998-ub,
  18519. title = "Affordances. Motivations, and the World Graph Theory",
  18520. author = "Guazzelli, Alex and Bota, Mihail and Corbacho, Fernando J and
  18521. Arbib, Michael A",
  18522. abstract = "O'Keefe and Nadel (1978) distinguish two paradigms for
  18523. navigation, the ``locale system'' for map-based navigation and
  18524. the ``taxon (behavioral orientation) system'' for route
  18525. navigation. This article models the taxon system, the map-based
  18526. system, and their interaction, and argues that the map-based
  18527. system involves the interaction of hippocampus and other
  18528. systems.We relate taxes to the notion of an affordance. Just as
  18529. a rat may have basic taxes for approaching food or avoiding a
  18530. bright light, so does it have a wider repertoire of affordances
  18531. for possible actions associated with immediate sensing of its
  18532. environment. We propose that affordances are extracted by the
  18533. rat posterior parietal cortex, which guides action selection by
  18534. the premotor cortex and is influenced also by hypothalamic drive
  18535. information.The taxon-affordances model (TAM) for taxon-based
  18536. determination of movement direction is based on models of frog
  18537. detour behavior, with expectations of future reward implemented
  18538. using reinforcement learning. The specification of the direction
  18539. of movement is refined by current affordances and motivational
  18540. information to yield an appropriate course of action.The world
  18541. graph (WG) theory expands the idea of a map by developing the
  18542. hypothesis that cognitive and motivational states interact. This
  18543. article describes an implementation of this theory, the WG
  18544. model. The integrated TAM-WG model then allows us to explain
  18545. data on the behavior of rats with and without fornix lesions,
  18546. which disconnect the hippocampus from other neural systems.",
  18547. journal = "Adapt. Behav.",
  18548. publisher = "SAGE Publications Ltd STM",
  18549. volume = 6,
  18550. number = "3-4",
  18551. pages = "435--471",
  18552. month = jan,
  18553. year = 1998
  18554. }
  18555. @ARTICLE{Chersi2013-el,
  18556. title = "Mental imagery in the navigation domain: a computational model
  18557. of sensory-motor simulation mechanisms",
  18558. author = "Chersi, Fabian and Donnarumma, Francesco and Pezzulo, Giovanni",
  18559. abstract = "Recent experimental evidence indicates that animals can use
  18560. mental simulation to make decisions about the actions to take
  18561. during goal-directed navigation. The principal brain areas found
  18562. to be active during this process are the hippocampus, the
  18563. ventral striatum and the sensory-motor cortex. In this paper, we
  18564. present a computational model that includes biological aspects
  18565. of this circuit and explains mechanistically how it may be used
  18566. to imagine and evaluate future events. Its most salient
  18567. characteristic is that choices about actions are made by
  18568. simulating movements and their sensory effects using the same
  18569. brain areas that are active during overt execution. More
  18570. precisely, the simulation of an action (e.g., walking) creates a
  18571. new sensory pattern that is evaluated in the same way as real
  18572. inputs. The model is validated in a navigation task in which a
  18573. simulated rat is placed in a complex maze. We show that
  18574. hippocampal and striatal cells are activated to simulate paths,
  18575. to retrieve their estimated value and to make decisions. We link
  18576. these results with a general framework that sees the brain as a
  18577. predictive device that can ?detach? itself from the here-and-now
  18578. of current perception using mechanisms such as episodic
  18579. memories, motor and visual imagery.",
  18580. journal = "Adapt. Behav.",
  18581. publisher = "SAGE Publications Ltd STM",
  18582. volume = 21,
  18583. number = 4,
  18584. pages = "251--262",
  18585. month = aug,
  18586. year = 2013
  18587. }
  18588. @ARTICLE{Lynn2018-pf,
  18589. title = "The physics of brain network structure, function, and
  18590. control",
  18591. author = "Lynn, Christopher W and Bassett, Danielle S",
  18592. abstract = "The brain is a complex organ characterized by heterogeneous
  18593. patterns of structural connections supporting unparalleled
  18594. feats of cognition and a wide range of behaviors. New
  18595. noninvasive imaging techniques now allow these patterns to
  18596. be carefully and comprehensively mapped in individual humans
  18597. and animals. Yet, it remains a fundamental challenge to
  18598. understand how the brain's structural wiring supports
  18599. cognitive processes, with major implications for the
  18600. personalized treatment of mental health disorders. Here, we
  18601. review recent efforts to meet this challenge that draw on
  18602. intuitions, models, and theories from physics, spanning the
  18603. domains of statistical mechanics, information theory, and
  18604. dynamical systems and control. We begin by considering the
  18605. organizing principles of brain network architecture
  18606. instantiated in structural wiring under constraints of
  18607. symmetry, spatial embedding, and energy minimization. We
  18608. next consider models of brain network function that
  18609. stipulate how neural activity propagates along these
  18610. structural connections, producing the long-range
  18611. interactions and collective dynamics that support a rich
  18612. repertoire of system functions. Finally, we consider
  18613. perturbative experiments and models for brain network
  18614. control, which leverage the physics of signal transmission
  18615. along structural wires to infer intrinsic control processes
  18616. that support goal-directed behavior and to inform
  18617. stimulation-based therapies for neurological disease and
  18618. psychiatric disorders. Throughout, we highlight several open
  18619. questions in the physics of brain network structure,
  18620. function, and control that will require creative efforts
  18621. from physicists willing to brave the complexities of living
  18622. matter.",
  18623. month = sep,
  18624. year = 2018,
  18625. archivePrefix = "arXiv",
  18626. primaryClass = "q-bio.NC",
  18627. eprint = "1809.06441"
  18628. }
  18629. @UNPUBLISHED{Chalk2019-yt,
  18630. title = "Inferring the function performed by a recurrent neural network",
  18631. author = "Chalk, Matthew and Tkacik, Gasper and Marre, Olivier",
  18632. abstract = "A central goal in systems neuroscience is to understand the
  18633. functions performed by neural circuits. Previous top-down models
  18634. addressed this question by comparing the behaviour of an ideal
  18635. model circuit, optimised to perform a given function, with neural
  18636. recordings. However, this requires guessing in advance what
  18637. function is being performed, which may not be possible for many
  18638. neural systems. Here, we propose an alternative approach that
  18639. uses recorded neural responses to directly infer the function
  18640. performed by a neural network. We assume that the goal of the
  18641. network can be expressed via a reward function, which describes
  18642. how desirable each state of the network is for carrying out a
  18643. given objective. This allows us to frame the problem of
  18644. optimising each neuron9s responses by viewing neurons as agents
  18645. in a reinforcement learning (RL) paradigm; likewise the problem
  18646. of inferring the reward function from the observed dynamics can
  18647. be treated using inverse RL. Our framework encompasses previous
  18648. influential theories of neural coding, such as efficient coding
  18649. and attractor network models, as special cases, given specific
  18650. choices of reward function. Finally, we can use the reward
  18651. function inferred from recorded neural responses to make testable
  18652. predictions about how the network dynamics will adapt depending
  18653. on contextual changes, such as cell death and/or varying input
  18654. statistics, so as to carry out the same underlying function with
  18655. different constraints.",
  18656. journal = "bioRxiv",
  18657. pages = "598086",
  18658. month = apr,
  18659. year = 2019,
  18660. language = "en"
  18661. }
  18662. @ARTICLE{Janabi-Sharifi2000-tj,
  18663. title = "Discrete-time adaptive windowing for velocity estimation",
  18664. author = "Janabi-Sharifi, F and Hayward, V and -. J. Chen, C",
  18665. abstract = "We present methods for velocity estimation from discrete and
  18666. quantized position samples using adaptive windowing. Previous
  18667. methods necessitate trade-offs between noise reduction, control
  18668. delay, estimate accuracy, reliability, computational load,
  18669. transient preservation, and difficulties with tuning. In
  18670. contrast, a first-order adaptive windowing method is shown to be
  18671. optimal in the sense that it minimizes the velocity error
  18672. variance while maximizes the accuracy of the estimates, requiring
  18673. no tradeoff. Variants of this method are also discussed. The
  18674. effectiveness of the proposed technique is verified in simulation
  18675. and by experiments on the control of a haptic device.",
  18676. journal = "IEEE Trans. Control Syst. Technol.",
  18677. volume = 8,
  18678. number = 6,
  18679. pages = "1003--1009",
  18680. month = nov,
  18681. year = 2000,
  18682. keywords = "haptic interfaces;velocity control;adaptive estimation;discrete
  18683. time systems;filtering theory;optimisation;adaptive
  18684. windowing;velocity estimation;discrete-time systems;control
  18685. delay;haptic interface;optimisation;Delay estimation;Finite
  18686. impulse response filter;Noise reduction;Haptic
  18687. interfaces;Velocity control;Filtering;Finite difference
  18688. methods;Intelligent robots;Machine intelligence;Size control"
  18689. }
  18690. @ARTICLE{Wang2015-ic,
  18691. title = "Covert rapid action-memory simulation ({CRAMS)}: A hypothesis of
  18692. hippocampal--prefrontal interactions for adaptive behavior",
  18693. author = "Wang, Jane X and Cohen, Neal J and Voss, Joel L",
  18694. abstract = "Effective choices generally require memory, yet little is known
  18695. regarding the cognitive or neural mechanisms that allow memory
  18696. to influence choices. We outline a new framework proposing that
  18697. covert memory processing of hippocampus interacts with
  18698. action-generation processing of prefrontal cortex in order to
  18699. arrive at optimal, memory-guided choices. Covert, rapid
  18700. action-memory simulation (CRAMS) is proposed here as a framework
  18701. for understanding cognitive and/or behavioral choices, whereby
  18702. prefrontal--hippocampal interactions quickly provide multiple
  18703. simulations of potential outcomes used to evaluate the set of
  18704. possible choices. We hypothesize that this CRAMS process is
  18705. automatic, obligatory, and covert, meaning that many cycles of
  18706. action-memory simulation occur in response to choice conflict
  18707. without an individual's necessary intention and generally
  18708. without awareness of the simulations, leading to adaptive
  18709. behavior with little perceived effort. CRAMS is thus distinct
  18710. from influential proposals that adaptive memory-based behavior
  18711. in humans requires consciously experienced memory-based
  18712. construction of possible future scenarios and deliberate
  18713. decisions among possible future constructions. CRAMS provides an
  18714. account of why hippocampus has been shown to make critical
  18715. contributions to the short-term control of behavior, and it
  18716. motivates several new experimental approaches and hypotheses
  18717. that could be used to better understand the ubiquitous role of
  18718. prefrontal--hippocampal interactions in situations that require
  18719. adaptively using memory to guide choices. Importantly, this
  18720. framework provides a perspective that allows for testing
  18721. decision-making mechanisms in a manner that translates well
  18722. across human and nonhuman animal model systems.",
  18723. journal = "Neurobiol. Learn. Mem.",
  18724. publisher = "Elsevier",
  18725. volume = 117,
  18726. pages = "22--33",
  18727. month = jan,
  18728. year = 2015,
  18729. keywords = "Learning; Memory; Decision-making; Hippocampus; Prefrontal
  18730. cortex; Simulation; Imagination; Adaptive function"
  18731. }
  18732. @ARTICLE{Preston2013-qk,
  18733. title = "Interplay of hippocampus and prefrontal cortex in memory",
  18734. author = "Preston, Alison R and Eichenbaum, Howard",
  18735. abstract = "Recent studies on the hippocampus and the prefrontal cortex have
  18736. considerably advanced our understanding of the distinct roles of
  18737. these brain areas in the encoding and retrieval of memories, and
  18738. of how they interact in the prolonged process by which new
  18739. memories are consolidated into our permanent storehouse of
  18740. knowledge. These studies have led to a new model of how the
  18741. hippocampus forms and replays memories and how the prefrontal
  18742. cortex engages representations of the meaningful contexts in
  18743. which related memories occur, as well as how these areas
  18744. interact during memory retrieval. Furthermore, they have
  18745. provided new insights into how interactions between the
  18746. hippocampus and prefrontal cortex support the assimilation of
  18747. new memories into pre-existing networks of knowledge, called
  18748. schemas, and how schemas are modified in this process as the
  18749. foundation of memory consolidation.",
  18750. journal = "Curr. Biol.",
  18751. publisher = "Elsevier",
  18752. volume = 23,
  18753. number = 17,
  18754. pages = "R764--73",
  18755. month = sep,
  18756. year = 2013,
  18757. language = "en"
  18758. }
  18759. @ARTICLE{Schmidt2019-gu,
  18760. title = "Disrupting the medial Prefrontal Cortex Alters Hippocampal
  18761. Sequences during Deliberative {Decision-Making}",
  18762. author = "Schmidt, Brandy and Duin, Anneke A and Redish, A David",
  18763. abstract = "Current theories of deliberative decision-making suggest that
  18764. deliberative decisions arise from imagined simulations that
  18765. require interactions between the prefrontal cortex and
  18766. hippocampus. In rodent navigation experiments, hippocampal theta
  18767. sequences advance from the location of the rat ahead to the
  18768. subsequent goal. In order to examine the role of the medial
  18769. prefrontal cortex (mPFC) on the hippocampus, we disrupted the
  18770. mPFC with DREADDs (Designer Receptors Exclusively Activated by
  18771. Designer Drugs). Using the Restaurant Row foraging task, we found
  18772. that mPFC disruption resulted in decreased vicarious trial and
  18773. error (VTE) behavior, reduced the number of theta sequences, and
  18774. impaired theta sequences in hippocampus. mPFC disruption led to
  18775. larger changes in the initiation of the hippocampal theta
  18776. sequences that represent the current location of the rat rather
  18777. than to the later portions that represent the future outcomes.
  18778. These data suggest that the mPFC likely provides an important
  18779. component to the initiation of deliberative sequences, and
  18780. provides support for an episodic-future thinking, working memory
  18781. interpretation of deliberation.",
  18782. journal = "J. Neurophysiol.",
  18783. month = mar,
  18784. year = 2019,
  18785. keywords = "hippocampus; place cell; prelimbic cortex; theta; vicarious trial
  18786. and error",
  18787. language = "en"
  18788. }
  18789. @ARTICLE{Knill2004-uu,
  18790. title = "The Bayesian brain: the role of uncertainty in neural coding and
  18791. computation",
  18792. author = "Knill, David C and Pouget, Alexandre",
  18793. abstract = "To use sensory information efficiently to make judgments and
  18794. guide action in the world, the brain must represent and use
  18795. information about uncertainty in its computations for perception
  18796. and action. Bayesian methods have proven successful in building
  18797. computational theories for perception and sensorimotor control,
  18798. and psychophysics is providing a growing body of evidence that
  18799. human perceptual computations are ``Bayes' optimal''. This leads
  18800. to the ``Bayesian coding hypothesis'': that the brain represents
  18801. sensory information probabilistically, in the form of probability
  18802. distributions. Several computational schemes have recently been
  18803. proposed for how this might be achieved in populations of
  18804. neurons. Neurophysiological data on the hypothesis, however, is
  18805. almost non-existent. A major challenge for neuroscientists is to
  18806. test these ideas experimentally, and so determine whether and how
  18807. neurons code information about sensory uncertainty.",
  18808. journal = "Trends Neurosci.",
  18809. volume = 27,
  18810. number = 12,
  18811. pages = "712--719",
  18812. month = dec,
  18813. year = 2004,
  18814. language = "en"
  18815. }
  18816. % The entry below contains non-ASCII chars that could not be converted
  18817. % to a LaTeX equivalent.
  18818. @ARTICLE{Friston2012-rd,
  18819. title = "The history of the future of the Bayesian brain",
  18820. author = "Friston, Karl",
  18821. abstract = "The slight perversion of the original title of this piece (The
  18822. Future of the Bayesian Brain ) reflects my attempt to write
  18823. prospectively about 'Science and Stories' over the past 20
  18824. years. I will meet this challenge by dealing with the future and
  18825. then turning to its history. The future of …",
  18826. journal = "Neuroimage",
  18827. publisher = "Elsevier",
  18828. volume = 62,
  18829. number = 2,
  18830. pages = "1230--1233",
  18831. year = 2012
  18832. }
  18833. @ARTICLE{Havenith2019-st,
  18834. title = "The {Virtual-Environment-Foraging} Task enables rapid training
  18835. and single-trial metrics of rule acquisition and reversal in
  18836. head-fixed mice",
  18837. author = "Havenith, Martha N and Zijderveld, Peter M and van Heukelum,
  18838. Sabrina and Abghari, Shaghayegh and Tiesinga, Paul and Glennon,
  18839. Jeffrey C",
  18840. abstract = "Behavioural flexibility is an essential survival skill, yet our
  18841. understanding of its neuronal substrates is still limited. While
  18842. mouse research offers unique tools to dissect the neuronal
  18843. circuits involved, the measurement of flexible behaviour in mice
  18844. often suffers from long training times, poor experimental
  18845. control, and temporally imprecise binary (hit/miss) performance
  18846. readouts. Here we present a virtual-environment task for mice
  18847. that tackles these limitations. It offers fast training of
  18848. vision-based rule reversals (~100 trials per reversal) with full
  18849. stimulus control and continuous behavioural readouts. By
  18850. generating multiple non-binary performance metrics per trial, it
  18851. provides single-trial estimates not only of response accuracy and
  18852. speed, but also of underlying processes like choice certainty and
  18853. alertness (discussed in detail in a companion paper). Based on
  18854. these metrics, we show that mice can predict new task rules long
  18855. before they are able to execute them, and that this delay varies
  18856. across animals. We also provide and validate single-trial
  18857. estimates of whether an error was committed with or without
  18858. awareness of the task rule. By tracking in unprecedented detail
  18859. the cognitive dynamics underlying flexible behaviour, this task
  18860. enables new investigations into the neuronal interactions that
  18861. shape behavioural flexibility moment by moment.",
  18862. journal = "Sci. Rep.",
  18863. volume = 9,
  18864. number = 1,
  18865. pages = "4790",
  18866. month = mar,
  18867. year = 2019,
  18868. language = "en"
  18869. }
  18870. @UNPUBLISHED{Russek2017-vw,
  18871. title = "Predictive representations can link model-based reinforcement
  18872. learning to model-free mechanisms",
  18873. author = "Russek, Evan M and Momennejad, Ida and Botvinick, Matthew M and
  18874. Gershman, Samuel J and Daw, Nathaniel D",
  18875. abstract = "Humans and animals are capable of evaluating actions by
  18876. considering their long-run future rewards through a process
  18877. described using model-based reinforcement learning (RL)
  18878. algorithms. The mechanisms by which neural circuits perform the
  18879. computations prescribed by model-based RL remain largely unknown;
  18880. however, multiple lines of evidence suggest that neural circuits
  18881. supporting model-based behavior are structurally homologous to
  18882. and overlapping with those thought to carry out model-free
  18883. temporal difference (TD) learning. Here, we lay out a family of
  18884. approaches by which model-based computation may be built upon a
  18885. core of TD learning. The foundation of this framework is the
  18886. successor representation, a predictive state representation that,
  18887. when combined with TD learning of value predictions, can produce
  18888. a subset of the behaviors associated with model-based learning at
  18889. a fraction of the computational cost. Using simulations, we
  18890. delineate the precise behavioral capabilities enabled by
  18891. evaluating actions using this approach, and compare them to those
  18892. demonstrated by biological organisms. We then introduce two new
  18893. algorithms that build upon the successor representation while
  18894. progressively mitigating its limitations. Because this framework
  18895. can account for the full range of observed putatively model-based
  18896. behaviors while still utilizing a core TD framework, we suggest
  18897. that it represents a neurally plausible family of mechanisms for
  18898. model-based evaluation.",
  18899. journal = "bioRxiv",
  18900. pages = "083857",
  18901. month = aug,
  18902. year = 2017,
  18903. language = "en"
  18904. }
  18905. @ARTICLE{Sul2010-js,
  18906. title = "Distinct roles of rodent orbitofrontal and medial prefrontal
  18907. cortex in decision making",
  18908. author = "Sul, Jung Hoon and Kim, Hoseok and Huh, Namjung and Lee, Daeyeol
  18909. and Jung, Min Whan",
  18910. abstract = "We investigated how different subregions of rodent prefrontal
  18911. cortex contribute to value-based decision making, by comparing
  18912. neural signals related to animal's choice, its outcome, and
  18913. action value in orbitofrontal cortex (OFC) and medial prefrontal
  18914. cortex (mPFC) of rats performing a dynamic two-armed bandit
  18915. task. Neural signals for upcoming action selection arose in the
  18916. mPFC, including the anterior cingulate cortex, only immediately
  18917. before the behavioral manifestation of animal's choice,
  18918. suggesting that rodent prefrontal cortex is not involved in
  18919. advanced action planning. Both OFC and mPFC conveyed signals
  18920. related to the animal's past choices and their outcomes over
  18921. multiple trials, but neural signals for chosen value and reward
  18922. prediction error were more prevalent in the OFC. Our results
  18923. suggest that rodent OFC and mPFC serve distinct roles in
  18924. value-based decision making and that the OFC plays a prominent
  18925. role in updating the values of outcomes expected from chosen
  18926. actions.",
  18927. journal = "Neuron",
  18928. publisher = "Elsevier",
  18929. volume = 66,
  18930. number = 3,
  18931. pages = "449--460",
  18932. month = may,
  18933. year = 2010,
  18934. language = "en"
  18935. }
  18936. @ARTICLE{Ito2011-yi,
  18937. title = "Multiple representations and algorithms for reinforcement
  18938. learning in the cortico-basal ganglia circuit",
  18939. author = "Ito, Makoto and Doya, Kenji",
  18940. abstract = "Accumulating evidence shows that the neural network of the
  18941. cerebral cortex and the basal ganglia is critically involved in
  18942. reinforcement learning. Recent studies found functional
  18943. heterogeneity within the cortico-basal ganglia circuit,
  18944. especially in its ventromedial to dorsolateral axis. Here we
  18945. review computational issues in reinforcement learning and
  18946. propose a working hypothesis on how multiple reinforcement
  18947. learning algorithms are implemented in the cortico-basal ganglia
  18948. circuit using different representations of states, values, and
  18949. actions.",
  18950. journal = "Curr. Opin. Neurobiol.",
  18951. publisher = "Elsevier",
  18952. volume = 21,
  18953. number = 3,
  18954. pages = "368--373",
  18955. month = jun,
  18956. year = 2011,
  18957. language = "en"
  18958. }
  18959. @ARTICLE{Solway2012-bf,
  18960. title = "Goal-directed decision making as probabilistic inference: a
  18961. computational framework and potential neural correlates",
  18962. author = "Solway, Alec and Botvinick, Matthew M",
  18963. abstract = "Recent work has given rise to the view that reward-based
  18964. decision making is governed by two key controllers: a habit
  18965. system, which stores stimulus-response associations shaped by
  18966. past reward, and a goal-oriented system that selects actions
  18967. based on their anticipated outcomes. The current literature
  18968. provides a rich body of computational theory addressing habit
  18969. formation, centering on temporal-difference learning mechanisms.
  18970. Less progress has been made toward formalizing the processes
  18971. involved in goal-directed decision making. We draw on recent
  18972. work in cognitive neuroscience, animal conditioning, cognitive
  18973. and developmental psychology, and machine learning to outline a
  18974. new theory of goal-directed decision making. Our basic proposal
  18975. is that the brain, within an identifiable network of cortical
  18976. and subcortical structures, implements a probabilistic
  18977. generative model of reward, and that goal-directed decision
  18978. making is effected through Bayesian inversion of this model. We
  18979. present a set of simulations implementing the account, which
  18980. address benchmark behavioral and neuroscientific findings, and
  18981. give rise to a set of testable predictions. We also discuss the
  18982. relationship between the proposed framework and other models of
  18983. decision making, including recent models of perceptual choice,
  18984. to which our theory bears a direct connection.",
  18985. journal = "Psychol. Rev.",
  18986. publisher = "psycnet.apa.org",
  18987. volume = 119,
  18988. number = 1,
  18989. pages = "120--154",
  18990. month = jan,
  18991. year = 2012,
  18992. language = "en"
  18993. }
  18994. @ARTICLE{Huys2012-my,
  18995. title = "Bonsai trees in your head: how the pavlovian system sculpts
  18996. goal-directed choices by pruning decision trees",
  18997. author = "Huys, Quentin J M and Eshel, Neir and O'Nions, Elizabeth and
  18998. Sheridan, Luke and Dayan, Peter and Roiser, Jonathan P",
  18999. abstract = "When planning a series of actions, it is usually infeasible to
  19000. consider all potential future sequences; instead, one must prune
  19001. the decision tree. Provably optimal pruning is, however, still
  19002. computationally ruinous and the specific approximations humans
  19003. employ remain unknown. We designed a new sequential
  19004. reinforcement-based task and showed that human subjects adopted
  19005. a simple pruning strategy: during mental evaluation of a
  19006. sequence of choices, they curtailed any further evaluation of a
  19007. sequence as soon as they encountered a large loss. This pruning
  19008. strategy was Pavlovian: it was reflexively evoked by large
  19009. losses and persisted even when overwhelmingly counterproductive.
  19010. It was also evident above and beyond loss aversion. We found
  19011. that the tendency towards Pavlovian pruning was selectively
  19012. predicted by the degree to which subjects exhibited sub-clinical
  19013. mood disturbance, in accordance with theories that ascribe
  19014. Pavlovian behavioural inhibition, via serotonin, a role in mood
  19015. disorders. We conclude that Pavlovian behavioural inhibition
  19016. shapes highly flexible, goal-directed choices in a manner that
  19017. may be important for theories of decision-making in mood
  19018. disorders.",
  19019. journal = "PLoS Comput. Biol.",
  19020. publisher = "journals.plos.org",
  19021. volume = 8,
  19022. number = 3,
  19023. pages = "e1002410",
  19024. month = mar,
  19025. year = 2012,
  19026. language = "en"
  19027. }
  19028. @ARTICLE{Collins2012-sn,
  19029. title = "How much of reinforcement learning is working memory, not
  19030. reinforcement learning? A behavioral, computational, and
  19031. neurogenetic analysis",
  19032. author = "Collins, Anne G E and Frank, Michael J",
  19033. abstract = "Instrumental learning involves corticostriatal circuitry and the
  19034. dopaminergic system. This system is typically modeled in the
  19035. reinforcement learning (RL) framework by incrementally
  19036. accumulating reward values of states and actions. However, human
  19037. learning also implicates prefrontal cortical mechanisms involved
  19038. in higher level cognitive functions. The interaction of these
  19039. systems remains poorly understood, and models of human behavior
  19040. often ignore working memory (WM) and therefore incorrectly
  19041. assign behavioral variance to the RL system. Here we designed a
  19042. task that highlights the profound entanglement of these two
  19043. processes, even in simple learning problems. By systematically
  19044. varying the size of the learning problem and delay between
  19045. stimulus repetitions, we separately extracted WM-specific
  19046. effects of load and delay on learning. We propose a new
  19047. computational model that accounts for the dynamic integration of
  19048. RL and WM processes observed in subjects' behavior.
  19049. Incorporating capacity-limited WM into the model allowed us to
  19050. capture behavioral variance that could not be captured in a pure
  19051. RL framework even if we (implausibly) allowed separate RL
  19052. systems for each set size. The WM component also allowed for a
  19053. more reasonable estimation of a single RL process. Finally, we
  19054. report effects of two genetic polymorphisms having relative
  19055. specificity for prefrontal and basal ganglia functions. Whereas
  19056. the COMT gene coding for catechol-O-methyl transferase
  19057. selectively influenced model estimates of WM capacity, the GPR6
  19058. gene coding for G-protein-coupled receptor 6 influenced the RL
  19059. learning rate. Thus, this study allowed us to specify distinct
  19060. influences of the high-level and low-level cognitive functions
  19061. on instrumental learning, beyond the possibilities offered by
  19062. simple RL models.",
  19063. journal = "Eur. J. Neurosci.",
  19064. publisher = "Wiley Online Library",
  19065. volume = 35,
  19066. number = 7,
  19067. pages = "1024--1035",
  19068. month = apr,
  19069. year = 2012,
  19070. language = "en"
  19071. }
  19072. @ARTICLE{Ruediger2012-lx,
  19073. title = "Goal-oriented searching mediated by ventral hippocampus early in
  19074. trial-and-error learning",
  19075. author = "Ruediger, Sarah and Spirig, Dominique and Donato, Flavio and
  19076. Caroni, Pico",
  19077. abstract = "Most behavioral learning in biology is trial and error, but how
  19078. these learning processes are influenced by individual brain
  19079. systems is poorly understood. Here we show that
  19080. ventral-to-dorsal hippocampal subdivisions have specific and
  19081. sequential functions in trial-and-error maze navigation, with
  19082. ventral hippocampus (vH) mediating early task-specific
  19083. goal-oriented searching. Although performance and strategy
  19084. deployment progressed continuously at the population level,
  19085. individual mice showed discrete learning phases, each
  19086. characterized by particular search habits. Transitions in
  19087. learning phases reflected feedforward inhibitory connectivity
  19088. (FFI) growth occurring sequentially in ventral, then
  19089. intermediate, then dorsal hippocampal subdivisions. FFI growth
  19090. at vH occurred abruptly upon behavioral learning of goal-task
  19091. relationships. vH lesions or the absence of vH FFI growth
  19092. delayed early learning and disrupted performance consistency.
  19093. Intermediate hippocampus lesions impaired intermediate place
  19094. learning, whereas dorsal hippocampus lesions specifically
  19095. disrupted late spatial learning. Trial-and-error navigational
  19096. learning processes in naive mice thus involve a stereotype
  19097. sequence of increasingly precise subtasks learned through
  19098. distinct hippocampal subdivisions. Because of its unique
  19099. connectivity, vH may relate specific goals to internal states in
  19100. learning under healthy and pathological conditions.",
  19101. journal = "Nat. Neurosci.",
  19102. publisher = "nature.com",
  19103. volume = 15,
  19104. number = 11,
  19105. pages = "1563--1571",
  19106. month = nov,
  19107. year = 2012,
  19108. language = "en"
  19109. }
  19110. @ARTICLE{Glascher2010-ne,
  19111. title = "States versus rewards: dissociable neural prediction error
  19112. signals underlying model-based and model-free reinforcement
  19113. learning",
  19114. author = "Gl{\"a}scher, Jan and Daw, Nathaniel and Dayan, Peter and
  19115. O'Doherty, John P",
  19116. abstract = "Reinforcement learning (RL) uses sequential experience with
  19117. situations (``states'') and outcomes to assess actions. Whereas
  19118. model-free RL uses this experience directly, in the form of a
  19119. reward prediction error (RPE), model-based RL uses it
  19120. indirectly, building a model of the state transition and outcome
  19121. structure of the environment, and evaluating actions by
  19122. searching this model. A state prediction error (SPE) plays a
  19123. central role, reporting discrepancies between the current model
  19124. and the observed state transitions. Using functional magnetic
  19125. resonance imaging in humans solving a probabilistic Markov
  19126. decision task, we found the neural signature of an SPE in the
  19127. intraparietal sulcus and lateral prefrontal cortex, in addition
  19128. to the previously well-characterized RPE in the ventral
  19129. striatum. This finding supports the existence of two unique
  19130. forms of learning signal in humans, which may form the basis of
  19131. distinct computational strategies for guiding behavior.",
  19132. journal = "Neuron",
  19133. publisher = "Elsevier",
  19134. volume = 66,
  19135. number = 4,
  19136. pages = "585--595",
  19137. month = may,
  19138. year = 2010,
  19139. language = "en"
  19140. }
  19141. % The entry below contains non-ASCII chars that could not be converted
  19142. % to a LaTeX equivalent.
  19143. @ARTICLE{Tolman1930-yp,
  19144. title = "Introduction and removal of reward, and maze performance in rats",
  19145. author = "Tolman, E C and Honzik, C H",
  19146. abstract = "Two groups of rats, one which ran the maze with reward, and one
  19147. which ran the maze without reward, were tested to determine the
  19148. influence upon the learning curve of a sudden removal, or a
  19149. sudden introduction, of a food reward. The maze was a 14-unit
  19150. T-maze. Reliability coefficients ranged from. 876 to. 965. When
  19151. the food reward was removed from the maze the error scores and
  19152. time scores of the rewarded rats showed a large increase. When
  19153. reward was introduced into the maze the non-rewarded rats showed
  19154. a large decrease …",
  19155. journal = "Univ. Calif. Publ. Zool.",
  19156. publisher = "psycnet.apa.org",
  19157. year = 1930
  19158. }
  19159. @ARTICLE{Doya2002-bl,
  19160. title = "Multiple model-based reinforcement learning",
  19161. author = "Doya, Kenji and Samejima, Kazuyuki and Katagiri, Ken-Ichi and
  19162. Kawato, Mitsuo",
  19163. abstract = "We propose a modular reinforcement learning architecture for
  19164. nonlinear, nonstationary control tasks, which we call multiple
  19165. model-based reinforcement learning (MMRL). The basic idea is to
  19166. decompose a complex task into multiple domains in space and time
  19167. based on the predictability of the environmental dynamics. The
  19168. system is composed of multiple modules, each of which consists
  19169. of a state prediction model and a reinforcement learning
  19170. controller. The ``responsibility signal,'' which is given by the
  19171. softmax function of the prediction errors, is used to weight the
  19172. outputs of multiple modules, as well as to gate the learning of
  19173. the prediction models and the reinforcement learning
  19174. controllers. We formulate MMRL for both discrete-time,
  19175. finite-state case and continuous-time, continuous-state case.
  19176. The performance of MMRL was demonstrated for discrete case in a
  19177. nonstationary hunting task in a grid world and for continuous
  19178. case in a nonlinear, nonstationary control task of swinging up a
  19179. pendulum with variable physical parameters.",
  19180. journal = "Neural Comput.",
  19181. publisher = "MIT Press",
  19182. volume = 14,
  19183. number = 6,
  19184. pages = "1347--1369",
  19185. month = jun,
  19186. year = 2002,
  19187. language = "en"
  19188. }
  19189. @ARTICLE{Dayan2009-zu,
  19190. title = "Goal-directed control and its antipodes",
  19191. author = "Dayan, Peter",
  19192. abstract = "In instrumental conditioning, there is a rather precise
  19193. definition of goal-directed control, and therefore an acute
  19194. boundary between it and the somewhat more amorphous category
  19195. comprising its opposites. Here, we review this division in terms
  19196. of the various distinctions that accompany it in the fields of
  19197. reinforcement learning and cognitive architectures, considering
  19198. issues such as declarative and procedural control, the effect of
  19199. prior distributions over environments, the neural substrates
  19200. involved, and the differing views about the relative rationality
  19201. of the various forms of control. Our overall aim is to reconnect
  19202. some presently far-flung relations.",
  19203. journal = "Neural Netw.",
  19204. publisher = "Elsevier",
  19205. volume = 22,
  19206. number = 3,
  19207. pages = "213--219",
  19208. month = apr,
  19209. year = 2009,
  19210. language = "en"
  19211. }
  19212. @ARTICLE{Balleine1998-ur,
  19213. title = "Goal-directed instrumental action: contingency and incentive
  19214. learning and their cortical substrates",
  19215. author = "Balleine, B W and Dickinson, A",
  19216. abstract = "Instrumental behaviour is controlled by two systems: a
  19217. stimulus-response habit mechanism and a goal-directed process
  19218. that involves two forms of learning. The first is learning about
  19219. the instrumental contingency between the response and reward,
  19220. whereas the second consists of the acquisition of incentive
  19221. value by the reward. Evidence for contingency learning comes
  19222. from studies of reward devaluation and from demonstrations that
  19223. instrumental performance is sensitive not only the probability
  19224. of contiguous reward but also to the probability of unpaired
  19225. rewards. The process of incentive learning is evident in the
  19226. acquisition of control over performance by primary motivational
  19227. states. Preliminary lesion studies of the rat suggest that the
  19228. prelimbic area of prefrontal cortex plays a role in the
  19229. contingency learning, whereas the incentive learning for food
  19230. rewards involves the insular cortex.",
  19231. journal = "Neuropharmacology",
  19232. publisher = "Elsevier",
  19233. volume = 37,
  19234. number = "4-5",
  19235. pages = "407--419",
  19236. month = apr,
  19237. year = 1998,
  19238. language = "en"
  19239. }
  19240. @ARTICLE{Doll2009-ki,
  19241. title = "Instructional control of reinforcement learning: a behavioral
  19242. and neurocomputational investigation",
  19243. author = "Doll, Bradley B and Jacobs, W Jake and Sanfey, Alan G and Frank,
  19244. Michael J",
  19245. abstract = "Humans learn how to behave directly through environmental
  19246. experience and indirectly through rules and instructions.
  19247. Behavior analytic research has shown that instructions can
  19248. control behavior, even when such behavior leads to sub-optimal
  19249. outcomes (Hayes, S. (Ed.). 1989. Rule-governed behavior:
  19250. cognition, contingencies, and instructional control. Plenum
  19251. Press.). Here we examine the control of behavior through
  19252. instructions in a reinforcement learning task known to depend on
  19253. striatal dopaminergic function. Participants selected between
  19254. probabilistically reinforced stimuli, and were (incorrectly)
  19255. told that a specific stimulus had the highest (or lowest)
  19256. reinforcement probability. Despite experience to the contrary,
  19257. instructions drove choice behavior. We present neural network
  19258. simulations that capture the interactions between
  19259. instruction-driven and reinforcement-driven behavior via two
  19260. potential neural circuits: one in which the striatum is
  19261. inaccurately trained by instruction representations coming from
  19262. prefrontal cortex/hippocampus (PFC/HC), and another in which the
  19263. striatum learns the environmentally based reinforcement
  19264. contingencies, but is ``overridden'' at decision output. Both
  19265. models capture the core behavioral phenomena but, because they
  19266. differ fundamentally on what is learned, make distinct
  19267. predictions for subsequent behavioral and neuroimaging
  19268. experiments. Finally, we attempt to distinguish between the
  19269. proposed computational mechanisms governing instructed behavior
  19270. by fitting a series of abstract ``Q-learning'' and Bayesian
  19271. models to subject data. The best-fitting model supports one of
  19272. the neural models, suggesting the existence of a ``confirmation
  19273. bias'' in which the PFC/HC system trains the reinforcement
  19274. system by amplifying outcomes that are consistent with
  19275. instructions while diminishing inconsistent outcomes.",
  19276. journal = "Brain Res.",
  19277. publisher = "Elsevier",
  19278. volume = 1299,
  19279. pages = "74--94",
  19280. month = nov,
  19281. year = 2009,
  19282. language = "en"
  19283. }
  19284. @ARTICLE{Otto2013-yk,
  19285. title = "The curse of planning: dissecting multiple
  19286. reinforcement-learning systems by taxing the central executive",
  19287. author = "Otto, A Ross and Gershman, Samuel J and Markman, Arthur B and
  19288. Daw, Nathaniel D",
  19289. abstract = "A number of accounts of human and animal behavior posit the
  19290. operation of parallel and competing valuation systems in the
  19291. control of choice behavior. In these accounts, a flexible but
  19292. computationally expensive model-based reinforcement-learning
  19293. system has been contrasted with a less flexible but more
  19294. efficient model-free reinforcement-learning system. The factors
  19295. governing which system controls behavior-and under what
  19296. circumstances-are still unclear. Following the hypothesis that
  19297. model-based reinforcement learning requires cognitive resources,
  19298. we demonstrated that having human decision makers perform a
  19299. demanding secondary task engenders increased reliance on a
  19300. model-free reinforcement-learning strategy. Further, we showed
  19301. that, across trials, people negotiate the trade-off between the
  19302. two systems dynamically as a function of concurrent
  19303. executive-function demands, and people's choice latencies
  19304. reflect the computational expenses of the strategy they employ.
  19305. These results demonstrate that competition between multiple
  19306. learning systems can be controlled on a trial-by-trial basis by
  19307. modulating the availability of cognitive resources.",
  19308. journal = "Psychol. Sci.",
  19309. publisher = "journals.sagepub.com",
  19310. volume = 24,
  19311. number = 5,
  19312. pages = "751--761",
  19313. month = may,
  19314. year = 2013,
  19315. keywords = "cognitive neuroscience; decision making",
  19316. language = "en"
  19317. }
  19318. @ARTICLE{Wolpert1996-cd,
  19319. title = "The Lack of A Priori Distinctions Between Learning Algorithms",
  19320. author = "Wolpert, David H",
  19321. abstract = "This is the first of two papers that use off-training set (OTS)
  19322. error to investigate the assumption-free relationship between
  19323. learning algorithms. This first paper discusses the senses in
  19324. which there are no a priori distinctions between learning
  19325. algorithms. (The second paper discusses the senses in which
  19326. there are such distinctions.) In this first paper it is shown,
  19327. loosely speaking, that for any two algorithms A and B, there are
  19328. ?as many? targets (or priors over targets) for which A has lower
  19329. expected OTS error than B as vice versa, for loss functions like
  19330. zero-one loss. In particular, this is true if A is
  19331. cross-validation and B is ?anti-cross-validation? (choose the
  19332. learning algorithm with largest cross-validation error). This
  19333. paper ends with a discussion of the implications of these
  19334. results for computational learning theory. It is shown that one
  19335. cannot say: if empirical misclassification rate is low, the
  19336. Vapnik-Chervonenkis dimension of your generalizer is small, and
  19337. the training set is large, then with high probability your OTS
  19338. error is small. Other implications for ?membership queries?
  19339. algorithms and ?punting? algorithms are also discussed.",
  19340. journal = "Neural Comput.",
  19341. publisher = "MIT Press",
  19342. volume = 8,
  19343. number = 7,
  19344. pages = "1341--1390",
  19345. month = oct,
  19346. year = 1996
  19347. }
  19348. @ARTICLE{Evans2019-tq,
  19349. title = "Cognitive Control of Escape Behaviour",
  19350. author = "Evans, Dominic A and Stempel, A Vanessa and Vale, Ruben and
  19351. Branco, Tiago",
  19352. abstract = "When faced with potential predators, animals instinctively decide
  19353. whether there is a threat they should escape from, and also when,
  19354. how, and where to take evasive action. While escape is often
  19355. viewed in classical ethology as an action that is released upon
  19356. presentation of specific stimuli, successful and adaptive escape
  19357. behaviour relies on integrating information from sensory systems,
  19358. stored knowledge, and internal states. From a neuroscience
  19359. perspective, escape is an incredibly rich model that provides
  19360. opportunities for investigating processes such as perceptual and
  19361. value-based decision-making, or action selection, in an
  19362. ethological setting. We review recent research from laboratory
  19363. and field studies that explore, at the behavioural and
  19364. mechanistic levels, how elements from multiple information
  19365. streams are integrated to generate flexible escape behaviour.",
  19366. journal = "Trends Cogn. Sci.",
  19367. volume = 23,
  19368. number = 4,
  19369. pages = "334--348",
  19370. month = apr,
  19371. year = 2019,
  19372. keywords = "behavioural flexibility; defence; instinctive decisions; threat",
  19373. language = "en"
  19374. }
  19375. @UNPUBLISHED{Zador2019-fj,
  19376. title = "A Critique of Pure Learning: What Artificial Neural Networks can
  19377. Learn from Animal Brains",
  19378. author = "Zador, Anthony",
  19379. abstract = "Over the last decade, artificial neural networks (ANNs), have
  19380. undergone a revolution, catalyzed in large part by better tools
  19381. for supervised learning. However, training such networks requires
  19382. enormous data sets of labeled examples, whereas young animals
  19383. (including humans) typically learn with few or no labeled
  19384. examples. This stark contrast with biological learning has led
  19385. many in the ANN community posit that instead of supervised
  19386. paradigms, animals must rely instead primarily on unsupervised
  19387. learning, leading the search for better unsupervised algorithms.
  19388. Here we argue that much of an animal9s behavioral repertoire is
  19389. not the result of clever learning algorithms--supervised or
  19390. unsupervised--but arises instead from behavior programs already
  19391. present at birth. These programs arise through evolution, are
  19392. encoded in the genome, and emerge as a consequence of wiring up
  19393. the brain. Specifically, animals are born with highly structured
  19394. brain connectivity, which enables them learn very rapidly.
  19395. Recognizing the importance of the highly structured connectivity
  19396. suggests a path toward building ANNs capable of rapid learning.",
  19397. journal = "bioRxiv",
  19398. pages = "582643",
  19399. month = mar,
  19400. year = 2019,
  19401. language = "en"
  19402. }
  19403. @ARTICLE{Ahrlund-Richter2019-ff,
  19404. title = "A whole-brain atlas of monosynaptic input targeting four
  19405. different cell types in the medial prefrontal cortex of the mouse",
  19406. author = "{\"A}hrlund-Richter, Sofie and Xuan, Yang and van Lunteren,
  19407. Josina Anna and Kim, Hoseok and Ortiz, Cantin and Pollak Dorocic,
  19408. Iskra and Meletis, Konstantinos and Carl{\'e}n, Marie",
  19409. abstract = "The local and long-range connectivity of cortical neurons are
  19410. considered instrumental to the functional repertoire of the
  19411. cortical region in which they reside. In cortical networks,
  19412. distinct cell types build local circuit structures enabling
  19413. computational operations. Computations in the medial prefrontal
  19414. cortex (mPFC) are thought to be central to cognitive operation,
  19415. including decision-making and memory. We used a retrograde
  19416. trans-synaptic rabies virus system to generate brain-wide maps of
  19417. the input to excitatory neurons as well as three inhibitory
  19418. interneuron subtypes in the mPFC. On the global scale the input
  19419. patterns were found to be mainly cell type independent, with
  19420. quantitative differences in key brain regions, including the
  19421. basal forebrain. Mapping of the local mPFC network revealed high
  19422. connectivity between the different subtypes of interneurons. The
  19423. connectivity mapping gives insight into the information that the
  19424. mPFC processes and the structural architecture underlying the
  19425. mPFC's unique functions.",
  19426. journal = "Nat. Neurosci.",
  19427. month = mar,
  19428. year = 2019,
  19429. language = "en"
  19430. }
  19431. @ARTICLE{Saxena2019-it,
  19432. title = "Towards the neural population doctrine",
  19433. author = "Saxena, Shreya and Cunningham, John P",
  19434. abstract = "Across neuroscience, large-scale data recording and
  19435. population-level analysis methods have experienced explosive
  19436. growth. While the underlying hardware and computational
  19437. techniques have been well reviewed, we focus here on the novel
  19438. science that these technologies have enabled. We detail four
  19439. areas of the field where the joint analysis of neural populations
  19440. has significantly furthered our understanding of computation in
  19441. the brain: correlated variability, decoding, neural dynamics, and
  19442. artificial neural networks. Together, these findings suggest an
  19443. exciting trend towards a new era where neural populations are
  19444. understood to be the essential unit of computation in many brain
  19445. regions, a classic idea that has been given new life.",
  19446. journal = "Curr. Opin. Neurobiol.",
  19447. volume = 55,
  19448. pages = "103--111",
  19449. month = mar,
  19450. year = 2019,
  19451. language = "en"
  19452. }
  19453. @ARTICLE{Gershman2018-gf,
  19454. title = "The Successor Representation: Its Computational Logic and Neural
  19455. Substrates",
  19456. author = "Gershman, Samuel J",
  19457. abstract = "Reinforcement learning is the process by which an agent learns to
  19458. predict long-term future reward. We now understand a great deal
  19459. about the brain's reinforcement learning algorithms, but we know
  19460. considerably less about the representations of states and actions
  19461. over which these algorithms operate. A useful starting point is
  19462. asking what kinds of representations we would want the brain to
  19463. have, given the constraints on its computational architecture.
  19464. Following this logic leads to the idea of the successor
  19465. representation, which encodes states of the environment in terms
  19466. of their predictive relationships with other states. Recent
  19467. behavioral and neural studies have provided evidence for the
  19468. successor representation, and computational studies have explored
  19469. ways to extend the original idea. This paper reviews progress on
  19470. these fronts, organizing them within a broader framework for
  19471. understanding how the brain negotiates tradeoffs between
  19472. efficiency and flexibility for reinforcement learning.",
  19473. journal = "J. Neurosci.",
  19474. volume = 38,
  19475. number = 33,
  19476. pages = "7193--7200",
  19477. month = aug,
  19478. year = 2018,
  19479. keywords = "cognitive map; dopamine; hippocampus; reinforcement learning;
  19480. reward",
  19481. language = "en"
  19482. }
  19483. @ARTICLE{De_Martino2013-rw,
  19484. title = "Confidence in value-based choice",
  19485. author = "De Martino, Benedetto and Fleming, Stephen M and Garrett, Neil
  19486. and Dolan, Raymond J",
  19487. abstract = "Decisions are never perfect, with confidence in one's choices
  19488. fluctuating over time. How subjective confidence and valuation of
  19489. choice options interact at the level of brain and behavior is
  19490. unknown. Using a dynamic model of the decision process, we show
  19491. that confidence reflects the evolution of a decision variable
  19492. over time, explaining the observed relation between confidence,
  19493. value, accuracy and reaction time. As predicted by our dynamic
  19494. model, we show that a functional magnetic resonance imaging
  19495. signal in human ventromedial prefrontal cortex (vmPFC) reflects
  19496. both value comparison and confidence in the value comparison
  19497. process. Crucially, individuals varied in how they related
  19498. confidence to accuracy, allowing us to show that this
  19499. introspective ability is predicted by a measure of functional
  19500. connectivity between vmPFC and rostrolateral prefrontal cortex.
  19501. Our findings provide a mechanistic link between noise in value
  19502. comparison and metacognitive awareness of choice, enabling us
  19503. both to want and to express knowledge of what we want.",
  19504. journal = "Nat. Neurosci.",
  19505. volume = 16,
  19506. number = 1,
  19507. pages = "105--110",
  19508. month = jan,
  19509. year = 2013,
  19510. language = "en"
  19511. }
  19512. @ARTICLE{Cortese2018-hj,
  19513. title = "The neural and cognitive architecture for learning from a
  19514. small sample",
  19515. author = "Cortese, Aurelio and De Martino, Benedetto and Kawato,
  19516. Mitsuo",
  19517. abstract = "Artificial intelligence algorithms are capable of fantastic
  19518. exploits, yet they are still grossly inefficient compared
  19519. with the brain's ability to learn from few exemplars or
  19520. solve problems that have not been explicitly defined. What
  19521. is the secret that the evolution of human intelligence has
  19522. unlocked? Generalization is one answer, but there is more to
  19523. it. The brain does not directly solve difficult problems, it
  19524. is able to recast them into new and more tractable problems.
  19525. Here we propose a model whereby higher cognitive functions
  19526. profoundly interact with reinforcement learning to
  19527. drastically reduce the degrees of freedom of the search
  19528. space, simplifying complex problems and fostering more
  19529. efficient learning.",
  19530. month = oct,
  19531. year = 2018,
  19532. archivePrefix = "arXiv",
  19533. primaryClass = "q-bio.NC",
  19534. eprint = "1810.02476"
  19535. }
  19536. @ARTICLE{Burgess2007-hn,
  19537. title = "An oscillatory interference model of grid cell firing",
  19538. author = "Burgess, Neil and Barry, Caswell and O'Keefe, John",
  19539. abstract = "We expand upon our proposal that the oscillatory interference
  19540. mechanism proposed for the phase precession effect in place cells
  19541. underlies the grid-like firing pattern of dorsomedial entorhinal
  19542. grid cells (O'Keefe and Burgess (2005) Hippocampus 15:853-866).
  19543. The original one-dimensional interference model is generalized to
  19544. an appropriate two-dimensional mechanism. Specifically, dendritic
  19545. subunits of layer II medial entorhinal stellate cells provide
  19546. multiple linear interference patterns along different directions,
  19547. with their product determining the firing of the cell. Connection
  19548. of appropriate speed- and direction-dependent inputs onto
  19549. dendritic subunits could result from an unsupervised learning
  19550. rule which maximizes postsynaptic firing (e.g. competitive
  19551. learning). These inputs cause the intrinsic oscillation of
  19552. subunit membrane potential to increase above theta frequency by
  19553. an amount proportional to the animal's speed of running in the
  19554. ``preferred'' direction. The phase difference between this
  19555. oscillation and a somatic input at theta-frequency essentially
  19556. integrates velocity so that the interference of the two
  19557. oscillations reflects distance traveled in the preferred
  19558. direction. The overall grid pattern is maintained in
  19559. environmental location by phase reset of the grid cell by place
  19560. cells receiving sensory input from the environment, and
  19561. environmental boundaries in particular. We also outline possible
  19562. variations on the basic model, including the generation of
  19563. grid-like firing via the interaction of multiple cells rather
  19564. than via multiple dendritic subunits. Predictions of the
  19565. interference model are given for the frequency composition of EEG
  19566. power spectra and temporal autocorrelograms of grid cell firing
  19567. as functions of the speed and direction of running and the
  19568. novelty of the environment.",
  19569. journal = "Hippocampus",
  19570. volume = 17,
  19571. number = 9,
  19572. pages = "801--812",
  19573. year = 2007,
  19574. language = "en"
  19575. }
  19576. @ARTICLE{OKeefe1996-rj,
  19577. title = "Geometric determinants of the place fields of hippocampal neurons",
  19578. author = "O'Keefe, J and Burgess, N",
  19579. abstract = "The human hippocampus has been implicated in memory, in
  19580. particular episodic or declarative memory. In rats, hippocampal
  19581. lesions cause selective spatial deficits, and hippocampal complex
  19582. spike cells (place cells) exhibit spatially localized firing,
  19583. suggesting a role in spatial memory, although broader functions
  19584. have also been suggested. Here we report the identification of
  19585. the environmental features controlling the location and shape of
  19586. the receptive fields (place fields) of the place cells. This was
  19587. done by recording from the same cell in four rectangular boxes
  19588. that differed solely in the length of one or both sides. Most of
  19589. our results are explained by a model in which the place field is
  19590. formed by the summation of gaussian tuning curves, each oriented
  19591. perpendicular to a box wall and peaked at a fixed distance from
  19592. it.",
  19593. journal = "Nature",
  19594. volume = 381,
  19595. number = 6581,
  19596. pages = "425--428",
  19597. month = may,
  19598. year = 1996,
  19599. language = "en"
  19600. }
  19601. @ARTICLE{Akrami2018-xc,
  19602. title = "Posterior parietal cortex represents sensory history and mediates
  19603. its effects on behaviour",
  19604. author = "Akrami, Athena and Kopec, Charles D and Diamond, Mathew E and
  19605. Brody, Carlos D",
  19606. abstract = "Many models of cognition and of neural computations posit the use
  19607. and estimation of prior stimulus statistics: it has long been
  19608. known that working memory and perception are strongly impacted by
  19609. previous sensory experience, even when that sensory history is
  19610. not relevant to the current task at hand. Nevertheless, the
  19611. neural mechanisms and regions of the brain that are necessary for
  19612. computing and using such prior experience are unknown. Here we
  19613. report that the posterior parietal cortex (PPC) is a critical
  19614. locus for the representation and use of prior stimulus
  19615. information. We trained rats in an auditory parametric working
  19616. memory task, and found that they displayed substantial and
  19617. readily quantifiable behavioural effects of sensory-stimulus
  19618. history, similar to those observed in humans and monkeys. Earlier
  19619. proposals that the PPC supports working memory predict that
  19620. optogenetic silencing of this region would impair behaviour in
  19621. our working memory task. Contrary to this prediction, we found
  19622. that silencing the PPC significantly improved performance.
  19623. Quantitative analyses of behaviour revealed that this improvement
  19624. was due to the selective reduction of the effects of prior
  19625. sensory stimuli. Electrophysiological recordings showed that PPC
  19626. neurons carried far more information about the sensory stimuli of
  19627. previous trials than about the stimuli of the current trial.
  19628. Furthermore, for a given rat, the more information about previous
  19629. trial sensory history in the neural firing rates of the PPC, the
  19630. greater the behavioural effect of sensory history, suggesting a
  19631. tight link between behaviour and PPC representations of stimulus
  19632. history. Our results indicate that the PPC is a central component
  19633. in the processing of sensory-stimulus history, and could enable
  19634. further neurobiological investigation of long-standing questions
  19635. regarding how perception and working memory are affected by prior
  19636. sensory information.",
  19637. journal = "Nature",
  19638. volume = 554,
  19639. number = 7692,
  19640. pages = "368--372",
  19641. month = feb,
  19642. year = 2018,
  19643. language = "en"
  19644. }
  19645. @UNPUBLISHED{Duan2018-yh,
  19646. title = "Collicular circuits for flexible sensorimotor routing",
  19647. author = "Duan, Chunyu A and Pagan, Marino and Piet, Alex T and Kopec,
  19648. Charles D and Akrami, Athena and Riordan, Alexander J and Erlich,
  19649. Jeffrey C and Brody, Carlos D",
  19650. abstract = "Flexible and fast sensorimotor routing, based on relevant
  19651. environmental context, is a central component of executive
  19652. control, with prefrontal cortex (PFC) thought of as playing a
  19653. critical role and the midbrain superior colliculus (SC) more
  19654. traditionally viewed as the output of cortical flexible routing.
  19655. Here, using a rat task in which subjects switch rapidly between
  19656. task contexts that demand changes in sensorimotor mappings, we
  19657. report that silencing of the SC during a delay period, during
  19658. which task context is encoded in SC activity, impaired choice
  19659. accuracy. But inactivations during the subsequent choice period,
  19660. during which the subject selects their motor response, did not.
  19661. Furthermore, a defined subset of SC neurons encoded task context
  19662. more strongly than PFC neurons, and encoded the subject9s motor
  19663. output choice faster than PFC neurons or other SC neurons. These
  19664. data suggest cognitive and decision-making roles for the SC. We
  19665. used computational methods to identify different SC circuit
  19666. architectures that could account for these results. We found
  19667. numerous, highly varied SC model circuits that matched our
  19668. experimental data, including circuits without inhibitory
  19669. connections between units representing opposite decision outputs.
  19670. But all successful model circuits had inhibitory connections
  19671. between units on the same side of the brain representing opposite
  19672. contexts. This anatomical feature appears to be a key
  19673. experimental prediction for models in which the SC plays a
  19674. decision-making role during executive control.",
  19675. journal = "bioRxiv",
  19676. pages = "245613",
  19677. month = jan,
  19678. year = 2018,
  19679. language = "en"
  19680. }
  19681. @ARTICLE{noauthor_undated-cn,
  19682. }
  19683. @ARTICLE{Erdem2012-xv,
  19684. title = "A goal-directed spatial navigation model using forward
  19685. trajectory planning based on grid cells",
  19686. author = "Erdem, U{\u g}ur M and Hasselmo, Michael",
  19687. abstract = "A goal-directed navigation model is proposed based on forward
  19688. linear look-ahead probe of trajectories in a network of head
  19689. direction cells, grid cells, place cells and prefrontal cortex
  19690. (PFC) cells. The model allows selection of new goal-directed
  19691. trajectories. In a novel environment, the virtual rat
  19692. incrementally creates a map composed of place cells and PFC
  19693. cells by random exploration. After exploration, the rat
  19694. retrieves memory of the goal location, picks its next movement
  19695. direction by forward linear look-ahead probe of trajectories in
  19696. several candidate directions while stationary in one location,
  19697. and finds the one activating PFC cells with the highest reward
  19698. signal. Each probe direction involves activation of a static
  19699. pattern of head direction cells to drive an interference model
  19700. of grid cells to update their phases in a specific direction.
  19701. The updating of grid cell spiking drives place cells along the
  19702. probed look-ahead trajectory similar to the forward replay
  19703. during waking seen in place cell recordings. Directions are
  19704. probed until the look-ahead trajectory activates the reward
  19705. signal and the corresponding direction is used to guide
  19706. goal-finding behavior. We report simulation results in several
  19707. mazes with and without barriers. Navigation with barriers
  19708. requires a PFC map topology based on the temporal vicinity of
  19709. visited place cells and a reward signal diffusion process. The
  19710. interaction of the forward linear look-ahead trajectory probes
  19711. with the reward diffusion allows discovery of never-before
  19712. experienced shortcuts towards a goal location.",
  19713. journal = "Eur. J. Neurosci.",
  19714. publisher = "Wiley Online Library",
  19715. volume = 35,
  19716. number = 6,
  19717. pages = "916--931",
  19718. month = mar,
  19719. year = 2012,
  19720. language = "en"
  19721. }
  19722. @ARTICLE{Vikbladh2019-qe,
  19723. title = "Hippocampal Contributions to {Model-Based} Planning and Spatial
  19724. Memory",
  19725. author = "Vikbladh, Oliver M and Meager, Michael R and King, John and
  19726. Blackmon, Karen and Devinsky, Orrin and Shohamy, Daphna and
  19727. Burgess, Neil and Daw, Nathaniel D",
  19728. abstract = "SummaryLittle is known about the neural mechanisms that allow
  19729. humans and animals to plan actions using knowledge of task
  19730. contingencies. Emerging theories hypothesize that it involves
  19731. the same hippocampal mechanisms that support self-localization
  19732. and memory for locations. Yet limited direct evidence supports
  19733. the link between planning and the hippocampal place map. We
  19734. addressed this by investigating model-based planning and place
  19735. memory in healthy controls and epilepsy patients treated using
  19736. unilateral anterior temporal lobectomy with hippocampal
  19737. resection. Both functions were impaired in the patient group.
  19738. Specifically, the planning impairment was related to right
  19739. hippocampal lesion size, controlling for overall lesion size.
  19740. Furthermore, although planning and boundary-driven place memory
  19741. covaried in the control group, this relationship was attenuated
  19742. in patients, consistent with both functions relying on the same
  19743. structure in the healthy brain. These findings clarify both the
  19744. neural mechanism of model-based planning and the scope of
  19745. hippocampal contributions to behavior.",
  19746. journal = "Neuron",
  19747. publisher = "Elsevier",
  19748. volume = 0,
  19749. number = 0,
  19750. month = mar,
  19751. year = 2019,
  19752. keywords = "decision-making; model-based; reinforcement learning; planning;
  19753. spatial; memory; human; hippocampus; anterior temporal lobe;
  19754. lesion",
  19755. language = "en"
  19756. }
  19757. @ARTICLE{Alvernhe2008-dx,
  19758. title = "Different {CA1} and {CA3} representations of novel routes in a
  19759. shortcut situation",
  19760. author = "Alvernhe, Alice and Van Cauter, Tiffany and Save, Etienne and
  19761. Poucet, Bruno",
  19762. abstract = "Place cells are hippocampal neurons whose discharge is strongly
  19763. related to a rat's location in its environment. The existence of
  19764. place cells has led to the proposal that they are part of an
  19765. integrated neural system dedicated to spatial navigation. To
  19766. further understand the relationships between place cell firing
  19767. and spatial problem solving, we examined the discharge of CA1
  19768. and CA3 place cells as rats were exposed to a shortcut in a
  19769. runway maze. On specific sessions, a wall section of the maze
  19770. was removed so as to open a shorter novel route within the
  19771. otherwise familiar maze. We found that the discharge of both CA1
  19772. and CA3 cells was strongly affected in the vicinity of the
  19773. shortcut region but was much less affected farther away. In
  19774. addition, CA3 fields away from the shortcut were more altered
  19775. than CA1 fields. Thus, place cell firing appears to reflect more
  19776. than just the animal's spatial location and may provide
  19777. additional information about possible motions, or routes, within
  19778. the environment. This kinematic representation appears to be
  19779. spatially more extended in CA3 than in CA1, suggesting
  19780. interesting computational differences between the two
  19781. subregions.",
  19782. journal = "J. Neurosci.",
  19783. publisher = "Soc Neuroscience",
  19784. volume = 28,
  19785. number = 29,
  19786. pages = "7324--7333",
  19787. month = jul,
  19788. year = 2008,
  19789. language = "en"
  19790. }
  19791. @ARTICLE{Alvernhe2011-st,
  19792. title = "Local remapping of place cell firing in the Tolman detour task",
  19793. author = "Alvernhe, Alice and Save, Etienne and Poucet, Bruno",
  19794. abstract = "The existence of place cells, whose discharge is strongly
  19795. related to a rat's location in its environment, has led to the
  19796. proposal that they form part of an integrated neural system
  19797. dedicated to spatial navigation. It has been suggested that this
  19798. system could represent space as a cognitive map, which is
  19799. flexibly used by animals to plan new shortcuts or efficient
  19800. detours. To further understand the relationships between
  19801. hippocampal place cell firing and cognitive maps, we examined
  19802. the discharge of place cells as rats were exposed to a
  19803. Tolman-type detour problem. In specific sessions, a transparent
  19804. barrier was placed onto the maze so as to block the shortest
  19805. central path between the two rewarded end locations of a
  19806. familiar three-way maze. We found that rats rapidly and
  19807. consistently chose the shortest alternative detour. Furthermore,
  19808. both CA1 and CA3 place cells that had a field in the vicinity of
  19809. the barrier displayed local remapping. In contrast, neither CA1
  19810. nor CA3 cells that had a field away from the barrier were
  19811. affected. This finding, at odds with our previous report of
  19812. altered CA3 discharge for distant fields in a shortcut task,
  19813. suggests that the availability of a novel path and the blocking
  19814. of a familiar path are not equivalent and could lead to
  19815. different responses of the CA3 place cell population. Together,
  19816. the two studies point to a specific role of CA3 in the
  19817. representation of spatial connectivity and sequences.",
  19818. journal = "Eur. J. Neurosci.",
  19819. publisher = "Wiley Online Library",
  19820. volume = 33,
  19821. number = 9,
  19822. pages = "1696--1705",
  19823. month = may,
  19824. year = 2011,
  19825. language = "en"
  19826. }
  19827. @ARTICLE{Simon2011-ky,
  19828. title = "Neural correlates of forward planning in a spatial decision task
  19829. in humans",
  19830. author = "Simon, Dylan Alexander and Daw, Nathaniel D",
  19831. abstract = "Although reinforcement learning (RL) theories have been
  19832. influential in characterizing the mechanisms for reward-guided
  19833. choice in the brain, the predominant temporal difference (TD)
  19834. algorithm cannot explain many flexible or goal-directed actions
  19835. that have been demonstrated behaviorally. We investigate such
  19836. actions by contrasting an RL algorithm that is model based, in
  19837. that it relies on learning a map or model of the task and
  19838. planning within it, to traditional model-free TD learning. To
  19839. distinguish these approaches in humans, we used functional
  19840. magnetic resonance imaging in a continuous spatial navigation
  19841. task, in which frequent changes to the layout of the maze forced
  19842. subjects continually to relearn their favored routes, thereby
  19843. exposing the RL mechanisms used. We sought evidence for the
  19844. neural substrates of such mechanisms by comparing choice
  19845. behavior and blood oxygen level-dependent (BOLD) signals to
  19846. decision variables extracted from simulations of either
  19847. algorithm. Both choices and value-related BOLD signals in
  19848. striatum, although most often associated with TD learning, were
  19849. better explained by the model-based theory. Furthermore,
  19850. predecessor quantities for the model-based value computation
  19851. were correlated with BOLD signals in the medial temporal lobe
  19852. and frontal cortex. These results point to a significant
  19853. extension of both the computational and anatomical substrates
  19854. for RL in the brain.",
  19855. journal = "J. Neurosci.",
  19856. publisher = "Soc Neuroscience",
  19857. volume = 31,
  19858. number = 14,
  19859. pages = "5526--5539",
  19860. month = apr,
  19861. year = 2011,
  19862. language = "en"
  19863. }
  19864. @ARTICLE{Gustafson2011-ua,
  19865. title = "Grid cells, place cells, and geodesic generalization for spatial
  19866. reinforcement learning",
  19867. author = "Gustafson, Nicholas J and Daw, Nathaniel D",
  19868. abstract = "Reinforcement learning (RL) provides an influential
  19869. characterization of the brain's mechanisms for learning to make
  19870. advantageous choices. An important problem, though, is how
  19871. complex tasks can be represented in a way that enables efficient
  19872. learning. We consider this problem through the lens of spatial
  19873. navigation, examining how two of the brain's location
  19874. representations--hippocampal place cells and entorhinal grid
  19875. cells--are adapted to serve as basis functions for approximating
  19876. value over space for RL. Although much previous work has focused
  19877. on these systems' roles in combining upstream sensory cues to
  19878. track location, revisiting these representations with a focus on
  19879. how they support this downstream decision function offers
  19880. complementary insights into their characteristics. Rather than
  19881. localization, the key problem in learning is generalization
  19882. between past and present situations, which may not match
  19883. perfectly. Accordingly, although neural populations collectively
  19884. offer a precise representation of position, our simulations of
  19885. navigational tasks verify the suggestion that RL gains efficiency
  19886. from the more diffuse tuning of individual neurons, which allows
  19887. learning about rewards to generalize over longer distances given
  19888. fewer training experiences. However, work on generalization in RL
  19889. suggests the underlying representation should respect the
  19890. environment's layout. In particular, although it is often assumed
  19891. that neurons track location in Euclidean coordinates (that a
  19892. place cell's activity declines ``as the crow flies'' away from
  19893. its peak), the relevant metric for value is geodesic: the
  19894. distance along a path, around any obstacles. We formalize this
  19895. intuition and present simulations showing how Euclidean, but not
  19896. geodesic, representations can interfere with RL by generalizing
  19897. inappropriately across barriers. Our proposal that place and grid
  19898. responses should be modulated by geodesic distances suggests
  19899. novel predictions about how obstacles should affect spatial
  19900. firing fields, which provides a new viewpoint on data concerning
  19901. both spatial codes.",
  19902. journal = "PLoS Comput. Biol.",
  19903. volume = 7,
  19904. number = 10,
  19905. pages = "e1002235",
  19906. month = oct,
  19907. year = 2011,
  19908. language = "en"
  19909. }
  19910. @INCOLLECTION{Lengyel2008-xy,
  19911. title = "Hippocampal Contributions to Control: The Third Way",
  19912. booktitle = "Advances in Neural Information Processing Systems 20",
  19913. author = "Lengyel, M{\'a}t{\'e} and Dayan, Peter",
  19914. editor = "Platt, J C and Koller, D and Singer, Y and Roweis, S T",
  19915. publisher = "Curran Associates, Inc.",
  19916. pages = "889--896",
  19917. year = 2008
  19918. }
  19919. @ARTICLE{Bennett1996-wj,
  19920. title = "Do animals have cognitive maps?",
  19921. author = "Bennett, A T",
  19922. abstract = "Drawing on studies of humans, rodents, birds and arthropods, I
  19923. show that 'cognitive maps' have been used to describe a wide
  19924. variety of spatial concepts. There are, however, two main
  19925. definitions. One, sensu Tolman, O'Keefe and Nadel, is that a
  19926. cognitive map is a powerful memory of landmarks which allows
  19927. novel short-cutting to occur. The other, sensu Gallistel, is that
  19928. a cognitive map is any representation of space held by an animal.
  19929. Other definitions with quite different meanings are also
  19930. summarised. I argue that no animal has been conclusively shown to
  19931. have a cognitive map, sensu Tolman, O'Keefe and Nadel, because
  19932. simpler explanations of the crucial novel short-cutting results
  19933. are invariably possible. Owing to the repeated inability of
  19934. experimenters to eliminate these simpler explanations over at
  19935. least 15 years, and the confusion caused by the numerous
  19936. contradictory definitions of a cognitive map, I argue that the
  19937. cognitive map is no longer a useful hypothesis for elucidating
  19938. the spatial behaviour of animals and that use of the term should
  19939. be avoided.",
  19940. journal = "J. Exp. Biol.",
  19941. volume = 199,
  19942. number = "Pt 1",
  19943. pages = "219--224",
  19944. month = jan,
  19945. year = 1996,
  19946. language = "en"
  19947. }
  19948. @UNPUBLISHED{Alvarez2019-gb,
  19949. title = "Modeling decision-making under uncertainty: a direct comparison
  19950. study between human and mouse gambling data",
  19951. author = "Alvarez, Lidia Cabeza and Giustiniani, Julie and Chabin, Thibault
  19952. and Ramadan, Bahrie and Joucla, Coralie and Nicolier, Magali and
  19953. Pazart, Lionel and Haffen, Emmanuel and Fellmann, Dominique and
  19954. Gabriel, Damien and Peterschmitt, Yvan",
  19955. abstract = "Decision-making is a conserved evolutionary process enabling to
  19956. choose one option among several alternatives, and relying on
  19957. reward and cognitive control systems. The Iowa Gambling Task
  19958. allows to assess human decision-making under uncertainty by
  19959. presenting four cards decks with various cost-benefit
  19960. probabilities. Participants seek to maximize their monetary gains
  19961. by developing long-term optimal choice strategies. Animal
  19962. versions have been adapted with nutritional rewards but
  19963. interspecies data comparisons are still scarce. Our study
  19964. directly compared physiological decision-making performances
  19965. between humans and wild-type C57BL/6 mice. Human subjects
  19966. fulfilled an electronic Iowa Gambling Task version while mice
  19967. performed a maze-based adaptation with four arms baited in a
  19968. probabilistic way. Our data show closely matching performances
  19969. among species with similar patterns of choice behaviors.
  19970. Moreover, both populations clustered into good, intermediate, and
  19971. poor decision-making categories with similar proportions.
  19972. Remarkably, mice good decision-makers behaved as humans of the
  19973. same category, but slight differences among species have been
  19974. evidenced for the other two subpopulations. Overall, our direct
  19975. comparative study confirms the good face validity of the rodent
  19976. gambling task. Extended behavioral characterization and
  19977. pathological animal models should help strengthen its construct
  19978. validity and disentangle determinants of decision-making in
  19979. animals and humans.",
  19980. journal = "bioRxiv",
  19981. pages = "570499",
  19982. month = mar,
  19983. year = 2019,
  19984. language = "en"
  19985. }
  19986. @ARTICLE{Spiers2015-fh,
  19987. title = "Solving the detour problem in navigation: a model of prefrontal
  19988. and hippocampal interactions",
  19989. author = "Spiers, Hugo J and Gilbert, Sam J",
  19990. abstract = "Adapting behavior to accommodate changes in the environment is an
  19991. important function of the nervous system. A universal problem for
  19992. motile animals is the discovery that a learned route is blocked
  19993. and a detour is required. Given the substantial neuroscience
  19994. research on spatial navigation and decision-making it is
  19995. surprising that so little is known about how the brain solves the
  19996. detour problem. Here we review the limited number of relevant
  19997. functional neuroimaging, single unit recording and lesion
  19998. studies. We find that while the prefrontal cortex (PFC)
  19999. consistently responds to detours, the hippocampus does not.
  20000. Recent evidence suggests the hippocampus tracks information about
  20001. the future path distance to the goal. Based on this evidence we
  20002. postulate a conceptual model in which: Lateral PFC provides a
  20003. prediction error signal about the change in the path, frontopolar
  20004. and superior PFC support the re-formulation of the route plan as
  20005. a novel subgoal and the hippocampus simulates the new path. More
  20006. data will be required to validate this model and understand (1)
  20007. how the system processes the different options; and (2) deals
  20008. with situations where a new path becomes available (i.e.,
  20009. shortcuts).",
  20010. journal = "Front. Hum. Neurosci.",
  20011. volume = 9,
  20012. pages = "125",
  20013. month = mar,
  20014. year = 2015,
  20015. keywords = "artificial intelligence; goals; hippocampus; place cells;
  20016. planning; prediction error; reinforcement learning; virtual
  20017. reality",
  20018. language = "en"
  20019. }
  20020. @ARTICLE{Pezzulo2013-jy,
  20021. title = "The mixed instrumental controller: using value of information to
  20022. combine habitual choice and mental simulation",
  20023. author = "Pezzulo, Giovanni and Rigoli, Francesco and Chersi, Fabian",
  20024. abstract = "Instrumental behavior depends on both goal-directed and habitual
  20025. mechanisms of choice. Normative views cast these mechanisms in
  20026. terms of model-free and model-based methods of reinforcement
  20027. learning, respectively. An influential proposal hypothesizes that
  20028. model-free and model-based mechanisms coexist and compete in the
  20029. brain according to their relative uncertainty. In this paper we
  20030. propose a novel view in which a single Mixed Instrumental
  20031. Controller produces both goal-directed and habitual behavior by
  20032. flexibly balancing and combining model-based and model-free
  20033. computations. The Mixed Instrumental Controller performs a
  20034. cost-benefits analysis to decide whether to chose an action
  20035. immediately based on the available ``cached'' value of actions
  20036. (linked to model-free mechanisms) or to improve value estimation
  20037. by mentally simulating the expected outcome values (linked to
  20038. model-based mechanisms). Since mental simulation entails
  20039. cognitive effort and increases the reward delay, it is activated
  20040. only when the associated ``Value of Information'' exceeds its
  20041. costs. The model proposes a method to compute the Value of
  20042. Information, based on the uncertainty of action values and on the
  20043. distance of alternative cached action values. Overall, the model
  20044. by default chooses on the basis of lighter model-free estimates,
  20045. and integrates them with costly model-based predictions only when
  20046. useful. Mental simulation uses a sampling method to produce
  20047. reward expectancies, which are used to update the cached value of
  20048. one or more actions; in turn, this updated value is used for the
  20049. choice. The key predictions of the model are tested in different
  20050. settings of a double T-maze scenario. Results are discussed in
  20051. relation with neurobiological evidence on the hippocampus -
  20052. ventral striatum circuit in rodents, which has been linked to
  20053. goal-directed spatial navigation.",
  20054. journal = "Front. Psychol.",
  20055. volume = 4,
  20056. pages = "92",
  20057. month = mar,
  20058. year = 2013,
  20059. keywords = "exploration-exploitation; forward sweeps; goal-directed
  20060. decision-making; hippocampus; model-based reinforcement learning;
  20061. value of information; ventral striatum",
  20062. language = "en"
  20063. }
  20064. @ARTICLE{Lansink2012-ha,
  20065. title = "Reward cues in space: commonalities and differences in neural
  20066. coding by hippocampal and ventral striatal ensembles",
  20067. author = "Lansink, Carien S and Jackson, Jadin C and Lankelma, Jan V and
  20068. Ito, Rutsuko and Robbins, Trevor W and Everitt, Barry J and
  20069. Pennartz, Cyriel M A",
  20070. abstract = "Forming place-reward associations critically depends on the
  20071. integrity of the hippocampal-ventral striatal system. The ventral
  20072. striatum (VS) receives a strong hippocampal input conveying
  20073. spatial-contextual information, but it is unclear how this
  20074. structure integrates this information to invigorate
  20075. reward-directed behavior. Neuronal ensembles in rat hippocampus
  20076. (HC) and VS were simultaneously recorded during a conditioning
  20077. task in which navigation depended on path integration. In
  20078. contrast to HC, ventral striatal neurons showed low spatial
  20079. selectivity, but rather coded behavioral task phases toward
  20080. reaching goal sites. Outcome-predicting cues induced a remapping
  20081. of firing patterns in the HC, consistent with its role in
  20082. episodic memory. VS remapped in conjunction with the HC,
  20083. indicating that remapping can take place in multiple brain
  20084. regions engaged in the same task. Subsets of ventral striatal
  20085. neurons showed a ``flip'' from high activity when cue lights were
  20086. illuminated to low activity in intertrial intervals, or vice
  20087. versa. The cues induced an increase in spatial information
  20088. transmission and sparsity in both structures. These effects were
  20089. paralleled by an enhanced temporal specificity of ensemble coding
  20090. and a more accurate reconstruction of the animal's position from
  20091. population firing patterns. Altogether, the results reveal strong
  20092. differences in spatial processing between hippocampal area CA1
  20093. and VS, but indicate similarities in how discrete cues impact on
  20094. this processing.",
  20095. journal = "J. Neurosci.",
  20096. volume = 32,
  20097. number = 36,
  20098. pages = "12444--12459",
  20099. month = sep,
  20100. year = 2012,
  20101. language = "en"
  20102. }
  20103. @ARTICLE{Pezzulo2014-ph,
  20104. title = "Internally generated sequences in learning and executing
  20105. goal-directed behavior",
  20106. author = "Pezzulo, Giovanni and van der Meer, Matthijs A A and Lansink,
  20107. Carien S and Pennartz, Cyriel M A",
  20108. abstract = "A network of brain structures including hippocampus (HC),
  20109. prefrontal cortex, and striatum controls goal-directed behavior
  20110. and decision making. However, the neural mechanisms underlying
  20111. these functions are unknown. Here, we review the role of
  20112. 'internally generated sequences': structured, multi-neuron firing
  20113. patterns in the network that are not confined to signaling the
  20114. current state or location of an agent, but are generated on the
  20115. basis of internal brain dynamics. Neurophysiological studies
  20116. suggest that such sequences fulfill functions in memory
  20117. consolidation, augmentation of representations, internal
  20118. simulation, and recombination of acquired information. Using
  20119. computational modeling, we propose that internally generated
  20120. sequences may be productively considered a component of
  20121. goal-directed decision systems, implementing a sampling-based
  20122. inference engine that optimizes goal acquisition at multiple
  20123. timescales of on-line choice, action control, and learning.",
  20124. journal = "Trends Cogn. Sci.",
  20125. volume = 18,
  20126. number = 12,
  20127. pages = "647--657",
  20128. month = dec,
  20129. year = 2014,
  20130. keywords = "decision making; forward sweep; generative models; hippocampus;
  20131. inference; prospection; reinforcement learning; replay; spatial
  20132. navigation; theta rhythm; ventral striatum",
  20133. language = "en"
  20134. }
  20135. @ARTICLE{Daw2014-nr,
  20136. title = "The algorithmic anatomy of model-based evaluation",
  20137. author = "Daw, Nathaniel D and Dayan, Peter",
  20138. abstract = "Despite many debates in the first half of the twentieth century,
  20139. it is now largely a truism that humans and other animals build
  20140. models of their environments and use them for prediction and
  20141. control. However, model-based (MB) reasoning presents severe
  20142. computational challenges. Alternative, computationally simpler,
  20143. model-free (MF) schemes have been suggested in the reinforcement
  20144. learning literature, and have afforded influential accounts of
  20145. behavioural and neural data. Here, we study the realization of
  20146. MB calculations, and the ways that this might be woven together
  20147. with MF values and evaluation methods. There are as yet mostly
  20148. only hints in the literature as to the resulting tapestry, so we
  20149. offer more preview than review.",
  20150. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  20151. publisher = "royalsocietypublishing.org",
  20152. volume = 369,
  20153. number = 1655,
  20154. month = nov,
  20155. year = 2014,
  20156. keywords = "Monte Carlo tree search; model-based reasoning; model-free
  20157. reasoning; orbitofrontal cortex; reinforcement learning;
  20158. striatum",
  20159. language = "en"
  20160. }
  20161. @ARTICLE{Dolan2013-rb,
  20162. title = "Goals and habits in the brain",
  20163. author = "Dolan, Ray J and Dayan, Peter",
  20164. abstract = "An enduring and richly elaborated dichotomy in cognitive
  20165. neuroscience is that of reflective versus reflexive decision
  20166. making and choice. Other literatures refer to the two ends of
  20167. what is likely to be a spectrum with terms such as goal-directed
  20168. versus habitual, model-based versus model-free or prospective
  20169. versus retrospective. One of the most rigorous traditions of
  20170. experimental work in the field started with studies in rodents
  20171. and graduated via human versions and enrichments of those
  20172. experiments to a current state in which new paradigms are probing
  20173. and challenging the very heart of the distinction. We review four
  20174. generations of work in this tradition and provide pointers to the
  20175. forefront of the field's fifth generation.",
  20176. journal = "Neuron",
  20177. volume = 80,
  20178. number = 2,
  20179. pages = "312--325",
  20180. month = oct,
  20181. year = 2013,
  20182. language = "en"
  20183. }
  20184. @ARTICLE{Nitz2012-rm,
  20185. title = "Spaces within spaces: rat parietal cortex neurons register
  20186. position across three reference frames",
  20187. author = "Nitz, Douglas A",
  20188. abstract = "We recorded parietal cortex neurons as rats traversed squared
  20189. spiral tracks. Spatial firing patterns distinguished the
  20190. behaviorally identical track segments composing loops, yet
  20191. recurred with increasing or decreasing amplitude across the five
  20192. loops composing the full track. These results indicate that
  20193. parietal cortex neurons simultaneously respond to spatial
  20194. relationships in multiple external reference frames, a phenomenon
  20195. that may reflect a neural mechanism for relating parts to a
  20196. whole.",
  20197. journal = "Nat. Neurosci.",
  20198. volume = 15,
  20199. number = 10,
  20200. pages = "1365--1367",
  20201. month = oct,
  20202. year = 2012,
  20203. keywords = "Locomotion",
  20204. language = "en"
  20205. }
  20206. @ARTICLE{noauthor_undated-uu,
  20207. title = "Prefrontal cortex creates novel navigation sequences from
  20208. hippocampal place-cell replay with spatial reward propagation"
  20209. }
  20210. @ARTICLE{Hok2007-zu,
  20211. title = "Goal-related activity in hippocampal place cells",
  20212. author = "Hok, Vincent and Lenck-Santini, Pierre-Pascal and Roux,
  20213. S{\'e}bastien and Save, Etienne and Muller, Robert U and Poucet,
  20214. Bruno",
  20215. abstract = "Place cells are hippocampal neurons whose discharge is strongly
  20216. related to a rat's location in its environment. The existence of
  20217. place cells has led to the proposal that they are part of an
  20218. integrated neural system dedicated to spatial navigation, an idea
  20219. supported by the discovery of strong relationships between place
  20220. cell activity and spatial problem solving. To further understand
  20221. such relationships, we examined the discharge of place cells
  20222. recorded while rats solved a place navigation task. We report
  20223. that, in addition to having widely distributed firing fields,
  20224. place cells also discharge selectively while the hungry rat waits
  20225. in an unmarked goal location to release a food pellet. Such
  20226. firing is not duplicated in other locations outside the main
  20227. firing field even when the rat's behavior is constrained to be
  20228. extremely similar to the behavior at the goal. We therefore
  20229. propose that place cells provide both a geometric representation
  20230. of the current environment and a reflection of the rat's
  20231. expectancy that it is located correctly at the goal. This on-line
  20232. feedback about a critical aspect of navigational performance is
  20233. proposed to be signaled by the synchronous activity of the large
  20234. fraction of place cells active at the goal. In combination with
  20235. other (prefrontal) cells that provide coarse encoding of goal
  20236. location, hippocampal place cells may therefore participate in a
  20237. neural network allowing the rat to plan accurate trajectories in
  20238. space.",
  20239. journal = "J. Neurosci.",
  20240. volume = 27,
  20241. number = 3,
  20242. pages = "472--482",
  20243. month = jan,
  20244. year = 2007,
  20245. language = "en"
  20246. }
  20247. @ARTICLE{Botvinick2012-ez,
  20248. title = "Hierarchical reinforcement learning and decision making",
  20249. author = "Botvinick, Matthew Michael",
  20250. abstract = "The hierarchical structure of human and animal behavior has been
  20251. of critical interest in neuroscience for many years. Yet
  20252. understanding the neural processes that give rise to such
  20253. structure remains an open challenge. In recent research, a new
  20254. perspective on hierarchical behavior has begun to take shape,
  20255. inspired by ideas from machine learning, and in particular the
  20256. framework of hierarchical reinforcement learning. Hierarchical
  20257. reinforcement learning builds on traditional reinforcement
  20258. learning mechanisms, extending them to accommodate temporally
  20259. extended behaviors or subroutines. The resulting computational
  20260. paradigm has begun to influence both theoretical and empirical
  20261. work in neuroscience, conceptually aligning the study of
  20262. hierarchical behavior with research on other aspects of learning
  20263. and decision making, and giving rise to some thought-provoking
  20264. new findings.",
  20265. journal = "Curr. Opin. Neurobiol.",
  20266. volume = 22,
  20267. number = 6,
  20268. pages = "956--962",
  20269. month = dec,
  20270. year = 2012,
  20271. language = "en"
  20272. }
  20273. @ARTICLE{Barto2003-bz,
  20274. title = "Recent Advances in Hierarchical Reinforcement Learning",
  20275. author = "Barto, Andrew G and Mahadevan, Sridhar",
  20276. abstract = "Reinforcement learning is bedeviled by the curse of
  20277. dimensionality: the number of parameters to be learned grows
  20278. exponentially with the size of any compact encoding of a state.
  20279. Recent attempts to combat the curse of dimensionality have
  20280. turned to principled ways of exploiting temporal abstraction,
  20281. where decisions are not required at each step, but rather invoke
  20282. the execution of temporally-extended activities which follow
  20283. their own policies until termination. This leads naturally to
  20284. hierarchical control architectures and associated learning
  20285. algorithms. We review several approaches to temporal abstraction
  20286. and hierarchical organization that machine learning researchers
  20287. have recently developed. Common to these approaches is a
  20288. reliance on the theory of semi-Markov decision processes, which
  20289. we emphasize in our review. We then discuss extensions of these
  20290. ideas to concurrent activities, multiagent coordination, and
  20291. hierarchical memory for addressing partial observability.
  20292. Concluding remarks address open challenges facing the further
  20293. development of reinforcement learning in a hierarchical setting.",
  20294. journal = "Discrete Event Dyn. Syst.: Theory Appl.",
  20295. publisher = "Springer",
  20296. volume = 13,
  20297. number = 1,
  20298. pages = "41--77",
  20299. month = jan,
  20300. year = 2003
  20301. }
  20302. @ARTICLE{Dosovitskiy2016-au,
  20303. title = "Learning to Act by Predicting the Future",
  20304. author = "Dosovitskiy, Alexey and Koltun, Vladlen",
  20305. abstract = "We present an approach to sensorimotor control in immersive
  20306. environments. Our approach utilizes a high-dimensional
  20307. sensory stream and a lower-dimensional measurement stream.
  20308. The cotemporal structure of these streams provides a rich
  20309. supervisory signal, which enables training a sensorimotor
  20310. control model by interacting with the environment. The model
  20311. is trained using supervised learning techniques, but without
  20312. extraneous supervision. It learns to act based on raw
  20313. sensory input from a complex three-dimensional environment.
  20314. The presented formulation enables learning without a fixed
  20315. goal at training time, and pursuing dynamically changing
  20316. goals at test time. We conduct extensive experiments in
  20317. three-dimensional simulations based on the classical
  20318. first-person game Doom. The results demonstrate that the
  20319. presented approach outperforms sophisticated prior
  20320. formulations, particularly on challenging tasks. The results
  20321. also show that trained models successfully generalize across
  20322. environments and goals. A model trained using the presented
  20323. approach won the Full Deathmatch track of the Visual Doom AI
  20324. Competition, which was held in previously unseen
  20325. environments.",
  20326. month = nov,
  20327. year = 2016,
  20328. archivePrefix = "arXiv",
  20329. primaryClass = "cs.LG",
  20330. eprint = "1611.01779"
  20331. }
  20332. @ARTICLE{Lei2018-ih,
  20333. title = "Dynamic Path Planning of Unknown Environment Based on Deep
  20334. Reinforcement Learning",
  20335. author = "Lei, Xiaoyun and Zhang, Zhian and Dong, Peifang",
  20336. abstract = "PDF | Dynamic path planning of unknown environment has always
  20337. been a challenge for mobile robots. In this paper, we apply
  20338. double Q-network (DDQN) deep reinforcement learning proposed by
  20339. DeepMind in 2016 to dynamic path planning of unknown environment.
  20340. The reward and punishment...",
  20341. journal = "Journal of Robotics",
  20342. volume = 2018,
  20343. number = 12,
  20344. pages = "1--10",
  20345. month = sep,
  20346. year = 2018
  20347. }
  20348. @INPROCEEDINGS{Mannucci2016-ja,
  20349. title = "A hierarchical maze navigation algorithm with Reinforcement
  20350. Learning and mapping",
  20351. booktitle = "2016 {IEEE} Symposium Series on Computational Intelligence
  20352. ({SSCI})",
  20353. author = "Mannucci, T and van Kampen, E",
  20354. abstract = "Goal-finding in an unknown maze is a challenging problem for a
  20355. Reinforcement Learning agent, because the corresponding state
  20356. space can be large if not intractable, and the agent does not
  20357. usually have a model of the environment. Hierarchical
  20358. Reinforcement Learning has been shown in the past to improve
  20359. tractability and learning time of complex problems, as well as
  20360. facilitate learning a coherent transition model for the
  20361. environment. Nonetheless, considerable time is still needed to
  20362. learn the transition model, so that initially the agent can
  20363. perform poorly by getting trapped into dead ends and colliding
  20364. with obstacles. This paper proposes a strategy for maze
  20365. exploration that, by means of sequential tasking and off-line
  20366. training on an abstract environment, provides the agent with a
  20367. minimal level of performance from the very beginning of
  20368. exploration. In particular, this approach allows to prevent
  20369. collisions with obstacles, thus enforcing a safety restraint on
  20370. the agent.",
  20371. pages = "1--8",
  20372. month = dec,
  20373. year = 2016,
  20374. keywords = "learning (artificial intelligence);safety restraint;off-line
  20375. training;sequential tasking;maze exploration;reinforcement
  20376. learning;hierarchical maze navigation
  20377. algorithm;Trajectory;Navigation;Robot sensing systems;Learning
  20378. (artificial intelligence);Training;Sonar"
  20379. }
  20380. @ARTICLE{Oess2017-fp,
  20381. title = "A Computational Model for Spatial Navigation Based on Reference
  20382. Frames in the Hippocampus, Retrosplenial Cortex, and Posterior
  20383. Parietal Cortex",
  20384. author = "Oess, Timo and Krichmar, Jeffrey L and R{\"o}hrbein, Florian",
  20385. abstract = "Behavioral studies for humans, monkeys, and rats have shown that,
  20386. while traversing an environment, these mammals tend to use
  20387. different frames of reference and frequently switch between them.
  20388. These frames represent allocentric, egocentric, or route-centric
  20389. views of the environment. However, combinations of either of them
  20390. are often deployed. Neurophysiological studies on rats have
  20391. indicated that the hippocampus, the retrosplenial cortex, and the
  20392. posterior parietal cortex contribute to the formation of these
  20393. frames and mediate the transformation between those. In this
  20394. paper, we construct a computational model of the posterior
  20395. parietal cortex and the retrosplenial cortex for spatial
  20396. navigation. We demonstrate how the transformation of reference
  20397. frames could be realized in the brain and suggest how different
  20398. brain areas might use these reference frames to form navigational
  20399. strategies and predict under what conditions an animal might use
  20400. a specific type of reference frame. Our simulated navigation
  20401. experiments demonstrate that the model's results closely resemble
  20402. behavioral findings in humans and rats. These results suggest
  20403. that navigation strategies may depend on the animal's reliance in
  20404. a particular reference frame and shows how low confidence in a
  20405. reference frame can lead to fluid adaptation and deployment of
  20406. alternative navigation strategies. Because of its flexibility,
  20407. our biologically inspired navigation system may be applied to
  20408. autonomous robots.",
  20409. journal = "Front. Neurorobot.",
  20410. volume = 11,
  20411. pages = "4",
  20412. month = feb,
  20413. year = 2017,
  20414. keywords = "computational model; frames of reference; hippocampus; posterior
  20415. parietal cortex; retrosplenial cortex; spatial navigation",
  20416. language = "en"
  20417. }
  20418. % The entry below contains non-ASCII chars that could not be converted
  20419. % to a LaTeX equivalent.
  20420. @ARTICLE{Darekar2018-xd,
  20421. title = "Modeling spatial navigation in the presence of dynamic obstacles:
  20422. a differential games approach",
  20423. author = "Darekar, Anuja and Goussev, Valery and McFadyen, Bradford J and
  20424. Lamontagne, Anouk and Fung, Joyce",
  20425. abstract = "Obstacle circumvention strategies can be shaped by the dynamic
  20426. interaction of an individual (evader) and an obstacle (pursuer).
  20427. We have developed a mathematical model with predictive and
  20428. emergent components, using experimental data from seven healthy
  20429. young adults walking toward a target while avoiding collision
  20430. with a stationary or moving obstacle (approaching head-on, or
  20431. diagonally 30° left or right) in a virtual environment. Two
  20432. linear properties from the predictive component enable the evader
  20433. to predict the minimum distance between itself and the obstacle
  20434. at all times, including the future intersection of trajectories.
  20435. The emergent component uses the classical differential games
  20436. model to solve for an optimal circumvention while reaching the
  20437. target, wherein the locomotor strategy is influenced by the
  20438. obstacle, target, and the evader velocity. Both model components
  20439. were fitted to a different set of experimental data obtained from
  20440. five poststroke and healthy participants to derive the minimum
  20441. predicted distance (predictive component) and obstacle influence
  20442. dimensions (emergent component) during circumvention. Minimum
  20443. predicted distance between evader and pursuer was kept constant
  20444. when the evader was closest to the obstacle in all participants.
  20445. Obstacle influence dimensions varied depending on obstacle
  20446. approach condition and preferred side of circumvention,
  20447. reflecting differences in locomotor strategies between poststroke
  20448. and healthy individuals. Additionally, important associations
  20449. between model outputs and observed experimental outcomes were
  20450. found. The model, supported by experimental data, suggests that
  20451. both predictive and emergent processes can shape obstacle
  20452. circumvention strategies in healthy and poststroke individuals.
  20453. NEW \& NOTEWORTHY Obstacle circumvention during goal-directed
  20454. locomotion is modeled with a new mathematical approach comprising
  20455. both predictive and emergent elements. The major novelty is using
  20456. differential games solutions to illustrate the dynamic
  20457. interactions between the individual as an evader and the
  20458. approaching obstacle as a pursuer. The model is supported by
  20459. experimental evidence that explains the behavior along the
  20460. continuum of locomotor adaptation displayed by healthy subjects
  20461. and individuals with stroke.",
  20462. journal = "J. Neurophysiol.",
  20463. volume = 119,
  20464. number = 3,
  20465. pages = "990--1004",
  20466. month = mar,
  20467. year = 2018,
  20468. keywords = "collision avoidance; locomotion; modeling; obstacle
  20469. circumvention; stroke;Locomotion",
  20470. language = "en"
  20471. }
  20472. @ARTICLE{Panov2018-xf,
  20473. title = "Grid Path Planning with Deep Reinforcement Learning: Preliminary
  20474. Results",
  20475. author = "Panov, Aleksandr I and Yakovlev, Konstantin S and Suvorov, Roman",
  20476. abstract = "Single-shot grid-based path finding is an important problem with
  20477. the applications in robotics, video games etc. Typically in AI
  20478. community heuristic search methods (based on A* and its
  20479. variations) are used to solve it. In this work we present the
  20480. results of preliminary studies on how neural networks can be
  20481. utilized to path planning on square grids, e.g. how well they can
  20482. cope with path finding tasks by themselves within the well-known
  20483. reinforcement problem statement. Conducted experiments show that
  20484. the agent using neural Q-learning algorithm robustly learns to
  20485. achieve the goal on small maps and demonstrate promising results
  20486. on the maps have ben never seen by him before.",
  20487. journal = "Procedia Comput. Sci.",
  20488. volume = 123,
  20489. pages = "347--353",
  20490. month = jan,
  20491. year = 2018,
  20492. keywords = "path planning; reinforcement learning; neural networks;
  20493. Q-learning; convolution networks; Q-network"
  20494. }
  20495. @ARTICLE{Schmidhuber2015-ym,
  20496. title = "On Learning to Think: Algorithmic Information Theory for
  20497. Novel Combinations of Reinforcement Learning Controllers and
  20498. Recurrent Neural World Models",
  20499. author = "Schmidhuber, Juergen",
  20500. abstract = "This paper addresses the general problem of reinforcement
  20501. learning (RL) in partially observable environments. In 2013,
  20502. our large RL recurrent neural networks (RNNs) learned from
  20503. scratch to drive simulated cars from high-dimensional video
  20504. input. However, real brains are more powerful in many ways.
  20505. In particular, they learn a predictive model of their
  20506. initially unknown environment, and somehow use it for
  20507. abstract (e.g., hierarchical) planning and reasoning. Guided
  20508. by algorithmic information theory, we describe RNN-based AIs
  20509. (RNNAIs) designed to do the same. Such an RNNAI can be
  20510. trained on never-ending sequences of tasks, some of them
  20511. provided by the user, others invented by the RNNAI itself in
  20512. a curious, playful fashion, to improve its RNN-based world
  20513. model. Unlike our previous model-building RNN-based RL
  20514. machines dating back to 1990, the RNNAI learns to actively
  20515. query its model for abstract reasoning and planning and
  20516. decision making, essentially ``learning to think.'' The
  20517. basic ideas of this report can be applied to many other
  20518. cases where one RNN-like system exploits the algorithmic
  20519. information content of another. They are taken from a grant
  20520. proposal submitted in Fall 2014, and also explain concepts
  20521. such as ``mirror neurons.'' Experimental results will be
  20522. described in separate papers.",
  20523. month = nov,
  20524. year = 2015,
  20525. archivePrefix = "arXiv",
  20526. primaryClass = "cs.AI",
  20527. eprint = "1511.09249"
  20528. }
  20529. @ARTICLE{Arulkumaran2017-bm,
  20530. title = "A Brief Survey of Deep Reinforcement Learning",
  20531. author = "Arulkumaran, Kai and Deisenroth, Marc Peter and Brundage,
  20532. Miles and Bharath, Anil Anthony",
  20533. abstract = "Deep reinforcement learning is poised to revolutionise the
  20534. field of AI and represents a step towards building
  20535. autonomous systems with a higher level understanding of the
  20536. visual world. Currently, deep learning is enabling
  20537. reinforcement learning to scale to problems that were
  20538. previously intractable, such as learning to play video games
  20539. directly from pixels. Deep reinforcement learning algorithms
  20540. are also applied to robotics, allowing control policies for
  20541. robots to be learned directly from camera inputs in the real
  20542. world. In this survey, we begin with an introduction to the
  20543. general field of reinforcement learning, then progress to
  20544. the main streams of value-based and policy-based methods.
  20545. Our survey will cover central algorithms in deep
  20546. reinforcement learning, including the deep $Q$-network,
  20547. trust region policy optimisation, and asynchronous advantage
  20548. actor-critic. In parallel, we highlight the unique
  20549. advantages of deep neural networks, focusing on visual
  20550. understanding via reinforcement learning. To conclude, we
  20551. describe several current areas of research within the field.",
  20552. month = aug,
  20553. year = 2017,
  20554. archivePrefix = "arXiv",
  20555. primaryClass = "cs.LG",
  20556. eprint = "1708.05866"
  20557. }
  20558. @ARTICLE{Madl2015-wd,
  20559. title = "Computational cognitive models of spatial memory in navigation
  20560. space: a review",
  20561. author = "Madl, Tamas and Chen, Ke and Montaldi, Daniela and Trappl, Robert",
  20562. abstract = "Spatial memory refers to the part of the memory system that
  20563. encodes, stores, recognizes and recalls spatial information about
  20564. the environment and the agent's orientation within it. Such
  20565. information is required to be able to navigate to goal locations,
  20566. and is vitally important for any embodied agent, or model
  20567. thereof, for reaching goals in a spatially extended environment.
  20568. In this paper, a number of computationally implemented cognitive
  20569. models of spatial memory are reviewed and compared. Three
  20570. categories of models are considered: symbolic models, neural
  20571. network models, and models that are part of a systems-level
  20572. cognitive architecture. Representative models from each category
  20573. are described and compared in a number of dimensions along which
  20574. simulation models can differ (level of modeling, types of
  20575. representation, structural accuracy, generality and abstraction,
  20576. environment complexity), including their possible mapping to the
  20577. underlying neural substrate. Neural mappings are rarely
  20578. explicated in the context of behaviorally validated models, but
  20579. they could be useful to cognitive modeling research by providing
  20580. a new approach for investigating a model's plausibility. Finally,
  20581. suggested experimental neuroscience methods are described for
  20582. verifying the biological plausibility of computational cognitive
  20583. models of spatial memory, and open questions for the field of
  20584. spatial memory modeling are outlined.",
  20585. journal = "Neural Netw.",
  20586. volume = 65,
  20587. pages = "18--43",
  20588. month = may,
  20589. year = 2015,
  20590. keywords = "Computational cognitive modeling; Spatial memory models",
  20591. language = "en"
  20592. }
  20593. @ARTICLE{Ha2018-sa,
  20594. title = "World Models",
  20595. author = "Ha, David and Schmidhuber, J{\"u}rgen",
  20596. abstract = "We explore building generative neural network models of popular
  20597. reinforcement learning environments. Our world model can be
  20598. trained quickly in an unsupervised manner to learn a compressed
  20599. spatial and temporal representation of the environment. By using
  20600. features extracted from the world model as inputs to an agent,
  20601. we can train a very compact and simple policy that can solve the
  20602. required task. We can even train our agent entirely inside of
  20603. its own hallucinated dream generated by its world model, and
  20604. transfer this policy back into the actual environment. An
  20605. interactive version of this article is available at
  20606. worldmodels.github.io.",
  20607. publisher = "Zenodo",
  20608. year = 2018
  20609. }
  20610. @ARTICLE{Stoianov2018-og,
  20611. title = "Model-based spatial navigation in the hippocampus-ventral
  20612. striatum circuit: A computational analysis",
  20613. author = "Stoianov, Ivilin Peev and Pennartz, Cyriel M A and Lansink,
  20614. Carien S and Pezzulo, Giovani",
  20615. abstract = "While the neurobiology of simple and habitual choices is
  20616. relatively well known, our current understanding of goal-directed
  20617. choices and planning in the brain is still limited. Theoretical
  20618. work suggests that goal-directed computations can be productively
  20619. associated to model-based (reinforcement learning) computations,
  20620. yet a detailed mapping between computational processes and
  20621. neuronal circuits remains to be fully established. Here we report
  20622. a computational analysis that aligns Bayesian nonparametrics and
  20623. model-based reinforcement learning (MB-RL) to the functioning of
  20624. the hippocampus (HC) and the ventral striatum (vStr)-a neuronal
  20625. circuit that increasingly recognized to be an appropriate model
  20626. system to understand goal-directed (spatial) decisions and
  20627. planning mechanisms in the brain. We test the MB-RL agent in a
  20628. contextual conditioning task that depends on intact hippocampus
  20629. and ventral striatal (shell) function and show that it solves the
  20630. task while showing key behavioral and neuronal signatures of the
  20631. HC-vStr circuit. Our simulations also explore the benefits of
  20632. biological forms of look-ahead prediction (forward sweeps) during
  20633. both learning and control. This article thus contributes to fill
  20634. the gap between our current understanding of computational
  20635. algorithms and biological realizations of (model-based)
  20636. reinforcement learning.",
  20637. journal = "PLoS Comput. Biol.",
  20638. volume = 14,
  20639. number = 9,
  20640. pages = "e1006316",
  20641. month = sep,
  20642. year = 2018,
  20643. language = "en"
  20644. }
  20645. @ARTICLE{Frank2000-xi,
  20646. title = "Trajectory encoding in the hippocampus and entorhinal cortex",
  20647. author = "Frank, L M and Brown, E N and Wilson, M",
  20648. abstract = "We recorded from single neurons in the hippocampus and entorhinal
  20649. cortex (EC) of rats to investigate the role of these structures
  20650. in navigation and memory representation. Our results revealed two
  20651. novel phenomena: first, many cells in CA1 and the EC fired at
  20652. significantly different rates when the animal was in the same
  20653. position depending on where the animal had come from or where it
  20654. was going. Second, cells in deep layers of the EC, the targets of
  20655. hippocampal outputs, appeared to represent the similarities
  20656. between locations on spatially distinct trajectories through the
  20657. environment. Our findings suggest that the hippocampus represents
  20658. the animal's position in the context of a trajectory through
  20659. space and that the EC represents regularities across different
  20660. trajectories that could allow for generalization across
  20661. experiences.",
  20662. journal = "Neuron",
  20663. volume = 27,
  20664. number = 1,
  20665. pages = "169--178",
  20666. month = jul,
  20667. year = 2000,
  20668. keywords = "Non-programmatic",
  20669. language = "en"
  20670. }
  20671. @ARTICLE{Trullier1997-gt,
  20672. title = "Biologically based artificial navigation systems: review and
  20673. prospects",
  20674. author = "Trullier, O and Wiener, S I and Berthoz, A and Meyer, J A",
  20675. abstract = "Diverse theories of animal navigation aim at explaining how to
  20676. determine and maintain a course from one place to another in the
  20677. environment, although each presents a particular perspective with
  20678. its own terminologies. These vocabularies sometimes overlap, but
  20679. unfortunately with different meanings. This paper attempts to
  20680. define precisely the existing concepts and terminologies, so as
  20681. to describe comprehensively the different theories and models
  20682. within the same unifying framework. We present navigation
  20683. strategies within a four-level hierarchical framework based upon
  20684. levels of complexity of required processing (Guidance, Place
  20685. recognition-triggered Response, Topological navigation, Metric
  20686. navigation). This classification is based upon what information
  20687. is perceived, represented and processed. It contrasts with common
  20688. distinctions based upon the availability of certain sensors or
  20689. cues and rather stresses the information structure and content of
  20690. central processors. We then review computational models of animal
  20691. navigation, i.e. of animats. These are introduced along with the
  20692. underlying conceptual basis in biological data drawn from
  20693. behavioral and physiological experiments, with emphasis on
  20694. theories of ``spatial cognitive maps''. The goal is to aid in
  20695. deriving algorithms based upon insights into these processes,
  20696. algorithms that can be useful both for psychobiologists and
  20697. roboticists. The main observation is, however, that despite the
  20698. fact that all reviewed models claim to have biological
  20699. inspiration and that some of them explicitly use ``Cognitive
  20700. Map''-like mechanisms, they correspond to different levels of our
  20701. proposed hierarchy and that none of them exhibits the main
  20702. capabilities of real ``Cognitive Maps''--in Tolman's sense--that
  20703. is, a robust capacity for detour and shortcut behaviors.",
  20704. journal = "Prog. Neurobiol.",
  20705. volume = 51,
  20706. number = 5,
  20707. pages = "483--544",
  20708. month = apr,
  20709. year = 1997,
  20710. language = "en"
  20711. }
  20712. @ARTICLE{Olafsdottir2018-on,
  20713. title = "The Role of Hippocampal Replay in Memory and Planning",
  20714. author = "{\'O}lafsd{\'o}ttir, H Freyja and Bush, Daniel and Barry, Caswell",
  20715. abstract = "The mammalian hippocampus is important for normal memory
  20716. function, particularly memory for places and events. Place cells,
  20717. neurons within the hippocampus that have spatial receptive
  20718. fields, represent information about an animal's position. During
  20719. periods of rest, but also during active task engagement, place
  20720. cells spontaneously recapitulate past trajectories. Such 'replay'
  20721. has been proposed as a mechanism necessary for a range of
  20722. neurobiological functions, including systems memory
  20723. consolidation, recall and spatial working memory, navigational
  20724. planning, and reinforcement learning. Focusing mainly, but not
  20725. exclusively, on work conducted in rodents, we describe the
  20726. methodologies used to analyse replay and review evidence for its
  20727. putative roles. We identify outstanding questions as well as
  20728. apparent inconsistencies in existing data, making suggestions as
  20729. to how these might be resolved. In particular, we find support
  20730. for the involvement of replay in disparate processes, including
  20731. the maintenance of hippocampal memories and decision making. We
  20732. propose that the function of replay changes dynamically according
  20733. to task demands placed on an organism and its current level of
  20734. arousal.",
  20735. journal = "Curr. Biol.",
  20736. volume = 28,
  20737. number = 1,
  20738. pages = "R37--R50",
  20739. month = jan,
  20740. year = 2018,
  20741. language = "en"
  20742. }
  20743. @ARTICLE{Stella2019-ap,
  20744. title = "Hippocampal Reactivation of Random Trajectories Resembling
  20745. Brownian Diffusion",
  20746. author = "Stella, Federico and Baracskay, Peter and O'Neill, Joseph and
  20747. Csicsvari, Jozsef",
  20748. abstract = "SummaryHippocampal activity patterns representing movement
  20749. trajectories are reactivated in immobility and sleep periods, a
  20750. process associated with memory recall, consolidation, and
  20751. decision making. It is thought that only fixed, behaviorally
  20752. relevant patterns can be reactivated, which are stored across
  20753. hippocampal synaptic connections. To test whether some
  20754. generalized rules govern reactivation, we examined trajectory
  20755. reactivation following non-stereotypical exploration of familiar
  20756. open-field environments. We found that random trajectories of
  20757. varying lengths and timescales were reactivated, resembling that
  20758. of Brownian motion of particles. The animals' behavioral
  20759. trajectory did not follow Brownian diffusion demonstrating that
  20760. the exact behavioral experience is not reactivated. Therefore,
  20761. hippocampal circuits are able to generate random trajectories of
  20762. any recently active map by following diffusion dynamics. This
  20763. ability of hippocampal circuits to generate representations of
  20764. all behavioral outcome combinations, experienced or not, may
  20765. underlie a wide variety of hippocampal-dependent cognitive
  20766. functions such as learning, generalization, and planning.",
  20767. journal = "Neuron",
  20768. publisher = "Elsevier",
  20769. volume = 0,
  20770. number = 0,
  20771. month = feb,
  20772. year = 2019,
  20773. language = "en"
  20774. }
  20775. @BOOK{Pfeifer2006-sa,
  20776. title = "How the Body Shapes the Way We Think: A New View of Intelligence",
  20777. author = "Pfeifer, Rolf and Bongard, Josh",
  20778. abstract = "An exploration of embodied intelligence and its implications
  20779. points toward a theory of intelligence in general; with case
  20780. studies of intelligent systems in ubiquitous computing, business
  20781. and management, human memory, and robotics.How could the body
  20782. influence our thinking when it seems obvious that the brain
  20783. controls the body? In How the Body Shapes the Way We Think, Rolf
  20784. Pfeifer and Josh Bongard demonstrate that thought is not
  20785. independent of the body but is tightly constrained, and at the
  20786. same time enabled, by it. They argue that the kinds of thoughts
  20787. we are capable of have their foundation in our embodiment---in
  20788. our morphology and the material properties of our bodies.This
  20789. crucial notion of embodiment underlies fundamental changes in
  20790. the field of artificial intelligence over the past two decades,
  20791. and Pfeifer and Bongard use the basic methodology of artificial
  20792. intelligence---``understanding by building''---to describe their
  20793. insights. If we understand how to design and build intelligent
  20794. systems, they reason, we will better understand intelligence in
  20795. general. In accessible, nontechnical language, and using many
  20796. examples, they introduce the basic concepts by building on
  20797. recent developments in robotics, biology, neuroscience, and
  20798. psychology to outline a possible theory of intelligence. They
  20799. illustrate applications of such a theory in ubiquitous
  20800. computing, business and management, and the psychology of human
  20801. memory. Embodied intelligence, as described by Pfeifer and
  20802. Bongard, has important implications for our understanding of
  20803. both natural and artificial intelligence.",
  20804. publisher = "MIT Press",
  20805. month = oct,
  20806. year = 2006,
  20807. keywords = "books;Books",
  20808. language = "en"
  20809. }
  20810. @ARTICLE{Stephenson-Jones2016-wq,
  20811. title = "A basal ganglia circuit for evaluating action outcomes",
  20812. author = "Stephenson-Jones, Marcus and Yu, Kai and Ahrens, Sandra and
  20813. Tucciarone, Jason M and van Huijstee, Aile N and Mejia, Luis A
  20814. and Penzo, Mario A and Tai, Lung-Hao and Wilbrecht, Linda and Li,
  20815. Bo",
  20816. abstract = "The basal ganglia, a group of subcortical nuclei, play a crucial
  20817. role in decision-making by selecting actions and evaluating their
  20818. outcomes. While much is known about the function of the basal
  20819. ganglia circuitry in selection, how these nuclei contribute to
  20820. outcome evaluation is less clear. Here we show that neurons in
  20821. the habenula-projecting globus pallidus (GPh) in mice are
  20822. essential for evaluating action outcomes and are regulated by a
  20823. specific set of inputs from the basal ganglia. We find in a
  20824. classical conditioning task that individual mouse GPh neurons
  20825. bidirectionally encode whether an outcome is better or worse than
  20826. expected. Mimicking these evaluation signals with optogenetic
  20827. inhibition or excitation is sufficient to reinforce or discourage
  20828. actions in a decision-making task. Moreover, cell-type-specific
  20829. synaptic manipulations reveal that the inhibitory and excitatory
  20830. inputs to the GPh are necessary for mice to appropriately
  20831. evaluate positive and negative feedback, respectively. Finally,
  20832. using rabies-virus-assisted monosynaptic tracing, we show that
  20833. the GPh is embedded in a basal ganglia circuit wherein it
  20834. receives inhibitory input from both striosomal and matrix
  20835. compartments of the striatum, and excitatory input from the
  20836. 'limbic' regions of the subthalamic nucleus. Our results provide
  20837. evidence that information about the selection and evaluation of
  20838. actions is channelled through distinct sets of basal ganglia
  20839. circuits, with the GPh representing a key locus in which
  20840. information of opposing valence is integrated to determine
  20841. whether action outcomes are better or worse than expected.",
  20842. journal = "Nature",
  20843. volume = 539,
  20844. number = 7628,
  20845. pages = "289--293",
  20846. month = nov,
  20847. year = 2016,
  20848. language = "en"
  20849. }
  20850. @ARTICLE{Rolls2004-wu,
  20851. title = "The functions of the orbitofrontal cortex",
  20852. author = "Rolls, Edmund T",
  20853. abstract = "The orbitofrontal cortex contains the secondary taste cortex, in
  20854. which the reward value of taste is represented. It also contains
  20855. the secondary and tertiary olfactory cortical areas, in which
  20856. information about the identity and also about the reward value of
  20857. odours is represented. The orbitofrontal cortex also receives
  20858. information about the sight of objects from the temporal lobe
  20859. cortical visual areas, and neurons in it learn and reverse the
  20860. visual stimulus to which they respond when the association of the
  20861. visual stimulus with a primary reinforcing stimulus (such as
  20862. taste) is reversed. This is an example of stimulus-reinforcement
  20863. association learning, and is a type of stimulus-stimulus
  20864. association learning. More generally, the stimulus might be a
  20865. visual or olfactory stimulus, and the primary (unlearned)
  20866. positive or negative reinforcer a taste or touch. A somatosensory
  20867. input is revealed by neurons that respond to the texture of food
  20868. in the mouth, including a population that responds to the mouth
  20869. feel of fat. In complementary neuroimaging studies in humans, it
  20870. is being found that areas of the orbitofrontal cortex are
  20871. activated by pleasant touch, by painful touch, by taste, by
  20872. smell, and by more abstract reinforcers such as winning or losing
  20873. money. Damage to the orbitofrontal cortex can impair the learning
  20874. and reversal of stimulus-reinforcement associations, and thus the
  20875. correction of behavioural responses when there are no longer
  20876. appropriate because previous reinforcement contingencies change.
  20877. The information which reaches the orbitofrontal cortex for these
  20878. functions includes information about faces, and damage to the
  20879. orbitofrontal cortex can impair face (and voice) expression
  20880. identification. This evidence thus shows that the orbitofrontal
  20881. cortex is involved in decoding and representing some primary
  20882. reinforcers such as taste and touch; in learning and reversing
  20883. associations of visual and other stimuli to these primary
  20884. reinforcers; and in controlling and correcting reward-related and
  20885. punishment-related behavior, and thus in emotion. The approach
  20886. described here is aimed at providing a fundamental understanding
  20887. of how the orbitofrontal cortex actually functions, and thus in
  20888. how it is involved in motivational behavior such as feeding and
  20889. drinking, in emotional behavior, and in social behavior.",
  20890. journal = "Brain Cogn.",
  20891. volume = 55,
  20892. number = 1,
  20893. pages = "11--29",
  20894. month = jun,
  20895. year = 2004,
  20896. language = "en"
  20897. }
  20898. @ARTICLE{Ernst2005-xn,
  20899. title = "Neurobiology of decision making: a selective review from a
  20900. neurocognitive and clinical perspective",
  20901. author = "Ernst, Monique and Paulus, Martin P",
  20902. abstract = "We present a temporal map of key processes that occur during
  20903. decision making, which consists of three stages: 1) formation of
  20904. preferences among options, 2) selection and execution of an
  20905. action, and 3) experience or evaluation of an outcome. This
  20906. framework can be used to integrate findings of traditional choice
  20907. psychology, neuropsychology, brain lesion studies, and functional
  20908. neuroimaging. Decision making is distributed across various brain
  20909. centers, which are differentially active across these stages of
  20910. decision making. This approach can be used to follow
  20911. developmental trajectories of the different stages of decision
  20912. making and to identify unique deficits associated with distinct
  20913. psychiatric disorders.",
  20914. journal = "Biol. Psychiatry",
  20915. volume = 58,
  20916. number = 8,
  20917. pages = "597--604",
  20918. month = oct,
  20919. year = 2005,
  20920. language = "en"
  20921. }
  20922. @ARTICLE{Wallis2011-jp,
  20923. title = "Cross-species studies of orbitofrontal cortex and value-based
  20924. decision-making",
  20925. author = "Wallis, Jonathan D",
  20926. abstract = "Recent work has emphasized the role that orbitofrontal cortex
  20927. (OFC) has in value-based decision-making. However, it is also
  20928. clear that a number of discrepancies have arisen when comparing
  20929. the findings from animal models to those from humans. Here, we
  20930. examine several possibilities that might explain these
  20931. discrepancies, including anatomical difference between species,
  20932. the behavioral tasks used to probe decision-making and the
  20933. methodologies used to assess neural function. Understanding how
  20934. these differences affect the interpretation of experimental
  20935. results will help us to better integrate future results from
  20936. animal models. This will enable us to fully realize the benefits
  20937. of using multiple approaches to understand OFC function.",
  20938. journal = "Nat. Neurosci.",
  20939. volume = 15,
  20940. number = 1,
  20941. pages = "13--19",
  20942. month = nov,
  20943. year = 2011,
  20944. language = "en"
  20945. }
  20946. @ARTICLE{Jones2012-oh,
  20947. title = "Orbitofrontal cortex supports behavior and learning using
  20948. inferred but not cached values",
  20949. author = "Jones, Joshua L and Esber, Guillem R and McDannald, Michael A and
  20950. Gruber, Aaron J and Hernandez, Alex and Mirenzi, Aaron and
  20951. Schoenbaum, Geoffrey",
  20952. abstract = "Computational and learning theory models propose that behavioral
  20953. control reflects value that is both cached (computed and stored
  20954. during previous experience) and inferred (estimated on the fly on
  20955. the basis of knowledge of the causal structure of the
  20956. environment). The latter is thought to depend on the
  20957. orbitofrontal cortex. Yet some accounts propose that the
  20958. orbitofrontal cortex contributes to behavior by signaling
  20959. ``economic'' value, regardless of the associative basis of the
  20960. information. We found that the orbitofrontal cortex is critical
  20961. for both value-based behavior and learning when value must be
  20962. inferred but not when a cached value is sufficient. The
  20963. orbitofrontal cortex is thus fundamental for accessing
  20964. model-based representations of the environment to compute value
  20965. rather than for signaling value per se.",
  20966. journal = "Science",
  20967. volume = 338,
  20968. number = 6109,
  20969. pages = "953--956",
  20970. month = nov,
  20971. year = 2012,
  20972. language = "en"
  20973. }
  20974. @ARTICLE{Gershman2014-ll,
  20975. title = "Retrospective revaluation in sequential decision making: a tale
  20976. of two systems",
  20977. author = "Gershman, Samuel J and Markman, Arthur B and Otto, A Ross",
  20978. abstract = "Recent computational theories of decision making in humans and
  20979. animals have portrayed 2 systems locked in a battle for control
  20980. of behavior. One system--variously termed model-free or
  20981. habitual--favors actions that have previously led to reward,
  20982. whereas a second--called the model-based or goal-directed
  20983. system--favors actions that causally lead to reward according to
  20984. the agent's internal model of the environment. Some evidence
  20985. suggests that control can be shifted between these systems using
  20986. neural or behavioral manipulations, but other evidence suggests
  20987. that the systems are more intertwined than a competitive account
  20988. would imply. In 4 behavioral experiments, using a retrospective
  20989. revaluation design and a cognitive load manipulation, we show
  20990. that human decisions are more consistent with a cooperative
  20991. architecture in which the model-free system controls behavior,
  20992. whereas the model-based system trains the model-free system by
  20993. replaying and simulating experience.",
  20994. journal = "J. Exp. Psychol. Gen.",
  20995. volume = 143,
  20996. number = 1,
  20997. pages = "182--194",
  20998. month = feb,
  20999. year = 2014,
  21000. language = "en"
  21001. }
  21002. @ARTICLE{Richard_R_Sutton2015-ot,
  21003. title = "Reinforcement Learning - An introduction",
  21004. author = "Richard R. Sutton, Andrew G Burton",
  21005. journal = "The MIT Press",
  21006. year = 2015,
  21007. keywords = "books;Books"
  21008. }
  21009. @ARTICLE{Akam2014-ig,
  21010. title = "When brains flip coins",
  21011. author = "Akam, Thomas and Costa, Rui M",
  21012. abstract = "In a recent study in the journal Cell, Tervo et al. (2014) show
  21013. that animals can implement stochastic choice policies in
  21014. environments unfavorable to predictive strategies. The shift
  21015. toward stochastic behavior was driven by noradrenergic signaling
  21016. in the anterior cingulate cortex.",
  21017. journal = "Neuron",
  21018. volume = 84,
  21019. number = 1,
  21020. pages = "9--11",
  21021. month = oct,
  21022. year = 2014,
  21023. language = "en"
  21024. }
  21025. @ARTICLE{Tervo2014-uw,
  21026. title = "Behavioral variability through stochastic choice and its gating
  21027. by anterior cingulate cortex",
  21028. author = "Tervo, Dougal G R and Proskurin, Mikhail and Manakov, Maxim and
  21029. Kabra, Mayank and Vollmer, Alison and Branson, Kristin and
  21030. Karpova, Alla Y",
  21031. abstract = "Behavioral choices that ignore prior experience promote
  21032. exploration and unpredictability but are seemingly at odds with
  21033. the brain's tendency to use experience to optimize behavioral
  21034. choice. Indeed, when faced with virtual competitors, primates
  21035. resort to strategic counter prediction rather than to stochastic
  21036. choice. Here, we show that rats also use history- and model-based
  21037. strategies when faced with similar competitors but can switch to
  21038. a ``stochastic'' mode when challenged with a competitor that they
  21039. cannot defeat by counter prediction. In this mode, outcomes
  21040. associated with an animal's actions are ignored, and normal
  21041. engagement of anterior cingulate cortex (ACC) is suppressed.
  21042. Using circuit perturbations in transgenic rats, we demonstrate
  21043. that switching between strategic and stochastic behavioral modes
  21044. is controlled by locus coeruleus input into ACC. Our findings
  21045. suggest that, under conditions of uncertainty about environmental
  21046. rules, changes in noradrenergic input alter ACC output and
  21047. prevent erroneous beliefs from guiding decisions, thus enabling
  21048. behavioral variation. PAPERCLIP:",
  21049. journal = "Cell",
  21050. volume = 159,
  21051. number = 1,
  21052. pages = "21--32",
  21053. month = sep,
  21054. year = 2014,
  21055. language = "en"
  21056. }
  21057. @ARTICLE{Koechlin2016-yb,
  21058. title = "Prefrontal executive function and adaptive behavior in complex
  21059. environments",
  21060. author = "Koechlin, Etienne",
  21061. abstract = "The prefrontal cortex (PFC) subserves higher cognitive abilities
  21062. such as planning, reasoning and creativity. Here we review recent
  21063. findings from both empirical and theoretical studies providing
  21064. new insights about these cognitive abilities and their neural
  21065. underpinnings in the PFC as overcoming key adaptive limitations
  21066. in reinforcement learning. We outline a unified theoretical
  21067. framework describing the PFC function as implementing an
  21068. algorithmic solution approximating statistically optimal, but
  21069. computationally intractable, adaptive processes. The resulting
  21070. PFC functional architecture combines learning, planning,
  21071. reasoning and creativity processes for balancing exploitation and
  21072. exploration behaviors and optimizing behavioral adaptations in
  21073. uncertain, variable and open-ended environments.",
  21074. journal = "Curr. Opin. Neurobiol.",
  21075. volume = 37,
  21076. pages = "1--6",
  21077. month = apr,
  21078. year = 2016,
  21079. language = "en"
  21080. }
  21081. @UNPUBLISHED{Zhang2019-bo,
  21082. title = "A Sequence Learning Model for Decision Making in the Brain",
  21083. author = "Zhang, Zhewei and Cheng, Huzi and Lin, Zhongqiao and Yang,
  21084. Tianming",
  21085. abstract = "Decision making is often modelled as a competition between
  21086. options. Currently, a great number of popular models to explain
  21087. the accuracy and speed in decision making are based on variations
  21088. of drift diffusion models (DDM), in which the options compete by
  21089. accumulating evidence toward decision bounds. Attractor-based
  21090. recurrent neural networks have been proposed to explain the
  21091. underlying neural mechanism. Yet, it is questionable that either
  21092. the DDM or attractor network is the brain9s general solution for
  21093. decision making. Here, we propose an alternative recurrent neural
  21094. network modeling approach based on gated recurrent units and
  21095. sequence learning. Our network model is trained to learn the
  21096. statistical structure of temporal sequences of sensory events,
  21097. action events, and reward events. We demonstrate its learning
  21098. with a reaction-time version of the weather prediction task
  21099. previously studied in monkey experiments, in which both the
  21100. animals9 behavior and the neuronal responses were consistent with
  21101. the DDM. The network model9s performance is able to reflect the
  21102. accuracy and reaction time pattern of the animals9 choice
  21103. behavior. The analyses of the unit responses in the network
  21104. reveal that they match important experimental findings. Notably,
  21105. we find units encoding the accumulated evidence and the urgency
  21106. signal. We further identify two groups of units based on their
  21107. connection weights to the choice output units. Simulated lesions
  21108. of each group of units produce doubly-dissociable effects on the
  21109. network9s choice and reaction time behavior. Graph analyses
  21110. reveal that these two groups of units belong to one highly
  21111. inter-connected sub-network. Finally, we show that the network is
  21112. capable of making predictions consistent with the predictive
  21113. coding and Bayesian inference framework. Our work offers
  21114. experimentally testable predictions of how decision making is
  21115. achieved in the brain. It provides an approach that may piece
  21116. together experimental findings of decision making, reinforcement
  21117. learning, and predictive coding. In particular, it suggests that
  21118. the DDM may be a manifestation of a more general computational
  21119. mechanism in the brain.",
  21120. journal = "bioRxiv",
  21121. pages = "555862",
  21122. month = feb,
  21123. year = 2019,
  21124. language = "en"
  21125. }
  21126. @ARTICLE{Amemiya2018-ym,
  21127. title = "Hippocampal {Theta-Gamma} Coupling Reflects {State-Dependent}
  21128. Information Processing in Decision Making",
  21129. author = "Amemiya, Seiichiro and Redish, A David",
  21130. abstract = "During decision making, hippocampal activity encodes information
  21131. sometimes about present and sometimes about potential future
  21132. plans. The mechanisms underlying this transition remain unknown.
  21133. Building on the evidence that gamma oscillations at different
  21134. frequencies (low gamma [LG], 30-55 Hz; high gamma [HG], 60-90 Hz;
  21135. and epsilon, 100-140 Hz) reflect inputs from different circuits,
  21136. we identified how changes in those frequencies reflect different
  21137. information-processing states. Using a unique noradrenergic
  21138. manipulation by clonidine, which shifted both neural
  21139. representations and gamma states, we found that future
  21140. representations depended on gamma components. These changes were
  21141. identifiable on each cycle of theta as asymmetries in the theta
  21142. cycle, which arose from changes within the ratio of LG and HG
  21143. power and the underlying phases of those gamma rhythms within the
  21144. theta cycle. These changes in asymmetry of the theta cycle
  21145. reflected changes in representations of present and future on
  21146. each theta cycle.",
  21147. journal = "Cell Rep.",
  21148. volume = 22,
  21149. number = 12,
  21150. pages = "3328--3338",
  21151. month = mar,
  21152. year = 2018,
  21153. keywords = "decision making; gamma; hippocampus; local field potential;
  21154. noradrenaline; norepinephrine; place cell; theta; vicarious trial
  21155. and error",
  21156. language = "en"
  21157. }
  21158. @ARTICLE{Sweis2018-ep,
  21159. title = "Mice learn to avoid regret",
  21160. author = "Sweis, Brian M and Thomas, Mark J and Redish, A David",
  21161. abstract = "Regret can be defined as the subjective experience of recognizing
  21162. that one has made a mistake and that a better alternative could
  21163. have been selected. The experience of regret is thought to carry
  21164. negative utility. This typically takes two distinct forms:
  21165. augmenting immediate postregret valuations to make up for losses,
  21166. and augmenting long-term changes in decision-making strategies to
  21167. avoid future instances of regret altogether. While the short-term
  21168. changes in valuation have been studied in human psychology,
  21169. economics, neuroscience, and even recently in nonhuman-primate
  21170. and rodent neurophysiology, the latter long-term process has
  21171. received far less attention, with no reports of regret avoidance
  21172. in nonhuman decision-making paradigms. We trained 31 mice in a
  21173. novel variant of the Restaurant Row economic decision-making
  21174. task, in which mice make decisions of whether to spend time from
  21175. a limited budget to achieve food rewards of varying costs
  21176. (delays). Importantly, we tested mice longitudinally for 70
  21177. consecutive days, during which the task provided their only
  21178. source of food. Thus, decision strategies were interdependent
  21179. across both trials and days. We separated principal commitment
  21180. decisions from secondary reevaluation decisions across space and
  21181. time and found evidence for regret-like behaviors following
  21182. change-of-mind decisions that corrected prior economically
  21183. disadvantageous choices. Immediately following change-of-mind
  21184. events, subsequent decisions appeared to make up for lost effort
  21185. by altering willingness to wait, decision speed, and pellet
  21186. consumption speed, consistent with past reports of regret in
  21187. rodents. As mice were exposed to an increasingly reward-scarce
  21188. environment, we found they adapted and refined distinct economic
  21189. decision-making strategies over the course of weeks to maximize
  21190. reinforcement rate. However, we also found that even without
  21191. changes in reinforcement rate, mice transitioned from an early
  21192. strategy rooted in foraging to a strategy rooted in deliberation
  21193. and planning that prevented future regret-inducing change-of-mind
  21194. episodes from occurring. These data suggest that mice are
  21195. learning to avoid future regret, independent of and separate from
  21196. reinforcement rate maximization.",
  21197. journal = "PLoS Biol.",
  21198. volume = 16,
  21199. number = 6,
  21200. pages = "e2005853",
  21201. month = jun,
  21202. year = 2018,
  21203. language = "en"
  21204. }
  21205. @ARTICLE{Knierim2009-ze,
  21206. title = "Imagining the possibilities: ripples, routes, and reactivation",
  21207. author = "Knierim, James J",
  21208. abstract = "Hippocampal place cells fire selectively when a rat occupies a
  21209. particular location. Under certain conditions, the cells briefly
  21210. represent trajectories along locations away from the rat's
  21211. current location. New results lend important insight into this
  21212. phenomenon and demonstrate spatiotemporally coherent, cognitive
  21213. representations that are independent of current sensory input.",
  21214. journal = "Neuron",
  21215. volume = 63,
  21216. number = 4,
  21217. pages = "421--423",
  21218. month = aug,
  21219. year = 2009,
  21220. language = "en"
  21221. }
  21222. @ARTICLE{Deshmukh2013-nc,
  21223. title = "Influence of local objects on hippocampal representations:
  21224. Landmark vectors and memory",
  21225. author = "Deshmukh, Sachin S and Knierim, James J",
  21226. abstract = "The hippocampus is thought to represent nonspatial information in
  21227. the context of spatial information. An animal can derive both
  21228. spatial information as well as nonspatial information from the
  21229. objects (landmarks) it encounters as it moves around in an
  21230. environment. In this article, correlates of both object-derived
  21231. spatial as well as nonspatial information in the hippocampus of
  21232. rats foraging in the presence of objects are demonstrated. A new
  21233. form of CA1 place cells, called landmark-vector cells, that
  21234. encode spatial locations as a vector relationship to local
  21235. landmarks is described. Such landmark vector relationships can be
  21236. dynamically encoded. Of the 26 CA1 neurons that developed new
  21237. fields in the course of a day's recording sessions, in eight
  21238. cases, the new fields were located at a similar distance and
  21239. direction from a landmark as the initial field was located
  21240. relative to a different landmark. In addition, object-location
  21241. memory in the hippocampus is also described. When objects were
  21242. removed from an environment or moved to new locations, a small
  21243. number of neurons in CA1 and CA3 increased firing at the
  21244. locations where the objects used to be. In some neurons, this
  21245. increase occurred only in one location, indicating object + place
  21246. conjunctive memory; in other neurons, the increase in firing was
  21247. seen at multiple locations where an object used to be. Taken
  21248. together, these results demonstrate that the spatially restricted
  21249. firing of hippocampal neurons encode multiple types of
  21250. information regarding the relationship between an animal's
  21251. location and the location of objects in its environment.",
  21252. journal = "Hippocampus",
  21253. volume = 23,
  21254. number = 4,
  21255. pages = "253--267",
  21256. month = apr,
  21257. year = 2013,
  21258. keywords = "boundary vector cell; hippocampus; landmark; memory; objects",
  21259. language = "en"
  21260. }
  21261. @ARTICLE{Knierim_James_J2014-wd,
  21262. title = "Functional correlates of the lateral and medial entorhinal
  21263. cortex: objects, path integration and local--global reference
  21264. frames",
  21265. author = "{Knierim James J.} and {Neunuebel Joshua P.} and {Deshmukh
  21266. Sachin S.}",
  21267. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  21268. publisher = "Royal Society",
  21269. volume = 369,
  21270. number = 1635,
  21271. pages = "20130369",
  21272. month = feb,
  21273. year = 2014
  21274. }
  21275. @ARTICLE{Knierim2016-cx,
  21276. title = "Tracking the flow of hippocampal computation: Pattern separation,
  21277. pattern completion, and attractor dynamics",
  21278. author = "Knierim, James J and Neunuebel, Joshua P",
  21279. abstract = "Classic computational theories of the mnemonic functions of the
  21280. hippocampus ascribe the processes of pattern separation to the
  21281. dentate gyrus (DG) and pattern completion to the CA3 region.
  21282. Until the last decade, the large majority of single-unit studies
  21283. of the hippocampus in behaving animals were from the CA1 region.
  21284. The lack of data from the DG, CA3, and the entorhinal inputs to
  21285. the hippocampus severely hampered the ability to test these
  21286. theories with neurophysiological techniques. The past ten years
  21287. have seen a major increase in the recordings from the CA3 region
  21288. and the medial entorhinal cortex (MEC), with an increasing (but
  21289. still limited) number of experiments from the lateral entorhinal
  21290. cortex (LEC) and DG. This paper reviews a series of studies in a
  21291. local-global cue mismatch (double-rotation) experiment in which
  21292. recordings were made from cells in the anterior thalamus, MEC,
  21293. LEC, DG, CA3, and CA1 regions. Compared to the standard cue
  21294. environment, the change in the DG representation of the
  21295. cue-mismatch environment was greater than the changes in its
  21296. entorhinal inputs, providing support for the theory of pattern
  21297. separation in the DG. In contrast, the change in the CA3
  21298. representation of the cue-mismatch environment was less than the
  21299. changes in its entorhinal and DG inputs, providing support for a
  21300. pattern completion/error correction function of CA3. The results
  21301. are interpreted in terms of continuous attractor network models
  21302. of the hippocampus and the relationship of these models to
  21303. pattern separation and pattern completion theories. Whereas DG
  21304. may perform an automatic pattern separation function, the
  21305. attractor dynamics of CA3 allow it to perform a pattern
  21306. separation or pattern completion function, depending on the
  21307. nature of its inputs and the relative strength of the internal
  21308. attractor dynamics.",
  21309. journal = "Neurobiol. Learn. Mem.",
  21310. volume = 129,
  21311. pages = "38--49",
  21312. month = mar,
  21313. year = 2016,
  21314. keywords = "Attractors; CA3; Dentate gyrus; Pattern completion; Pattern
  21315. separation; Place cells",
  21316. language = "en"
  21317. }
  21318. @ARTICLE{Davidson2009-cn,
  21319. title = "Hippocampal replay of extended experience",
  21320. author = "Davidson, Thomas J and Kloosterman, Fabian and Wilson, Matthew A",
  21321. abstract = "During pauses in exploration, ensembles of place cells in the
  21322. rat hippocampus re-express firing sequences corresponding to
  21323. recent spatial experience. Such ``replay'' co-occurs with ripple
  21324. events: short-lasting (approximately 50-120 ms), high-frequency
  21325. (approximately 200 Hz) oscillations that are associated with
  21326. increased hippocampal-cortical communication. In previous
  21327. studies, rats exploring small environments showed replay
  21328. anchored to the rat's current location and compressed in time
  21329. into a single ripple event. Here, we show, using a neural
  21330. decoding approach, that firing sequences corresponding to long
  21331. runs through a large environment are replayed with high fidelity
  21332. and that such replay can begin at remote locations on the track.
  21333. Extended replay proceeds at a characteristic virtual speed of
  21334. approximately 8 m/s and remains coherent across trains of ripple
  21335. events. These results suggest that extended replay is composed
  21336. of chains of shorter subsequences, which may reflect a strategy
  21337. for the storage and flexible expression of memories of prolonged
  21338. experience.",
  21339. journal = "Neuron",
  21340. publisher = "Elsevier",
  21341. volume = 63,
  21342. number = 4,
  21343. pages = "497--507",
  21344. month = aug,
  21345. year = 2009,
  21346. language = "en"
  21347. }
  21348. @ARTICLE{Foster2000-uu,
  21349. title = "A model of hippocampally dependent navigation, using the
  21350. temporal difference learning rule",
  21351. author = "Foster, D J and Morris, R G and Dayan, P",
  21352. abstract = "This paper presents a model of how hippocampal place cells might
  21353. be used for spatial navigation in two watermaze tasks: the
  21354. standard reference memory task and a delayed matching-to-place
  21355. task. In the reference memory task, the escape platform occupies
  21356. a single location and rats gradually learn relatively direct
  21357. paths to the goal over the course of days, in each of which they
  21358. perform a fixed number of trials. In the delayed
  21359. matching-to-place task, the escape platform occupies a novel
  21360. location on each day, and rats gradually acquire one-trial
  21361. learning, i.e., direct paths on the second trial of each day.
  21362. The model uses a local, incremental, and statistically efficient
  21363. connectionist algorithm called temporal difference learning in
  21364. two distinct components. The first is a reinforcement-based
  21365. ``actor-critic'' network that is a general model of classical
  21366. and instrumental conditioning. In this case, it is applied to
  21367. navigation, using place cells to provide information about
  21368. state. By itself, the actor-critic can learn the reference
  21369. memory task, but this learning is inflexible to changes to the
  21370. platform location. We argue that one-trial learning in the
  21371. delayed matching-to-place task demands a goal-independent
  21372. representation of space. This is provided by the second
  21373. component of the model: a network that uses temporal difference
  21374. learning and self-motion information to acquire consistent
  21375. spatial coordinates in the environment. Each component of the
  21376. model is necessary at a different stage of the task; the
  21377. actor-critic provides a way of transferring control to the
  21378. component that performs best. The model successfully captures
  21379. gradual acquisition in both tasks, and, in particular, the
  21380. ultimate development of one-trial learning in the delayed
  21381. matching-to-place task. Place cells report a form of stable,
  21382. allocentric information that is well-suited to the various kinds
  21383. of learning in the model.",
  21384. journal = "Hippocampus",
  21385. publisher = "Wiley Online Library",
  21386. volume = 10,
  21387. number = 1,
  21388. pages = "1--16",
  21389. year = 2000,
  21390. language = "en"
  21391. }
  21392. @ARTICLE{Jacob2019-iv,
  21393. title = "Path integration maintains spatial periodicity of grid cell
  21394. firing in a {1D} circular track",
  21395. author = "Jacob, Pierre-Yves and Capitano, Fabrizio and Poucet, Bruno and
  21396. Save, Etienne and Sargolini, Francesca",
  21397. abstract = "Entorhinal grid cells are thought to provide a 2D spatial metric
  21398. of the environment. In this study we demonstrate that in a
  21399. familiar 1D circular track (i.e., a continuous space) grid cells
  21400. display a novel 1D equidistant firing pattern based on
  21401. integrated distance rather than travelled distance or time. In
  21402. addition, field spacing is increased compared to a 2D open
  21403. field, probably due to a reduced access to the visual cue in the
  21404. track. This metrical modification is accompanied by a change in
  21405. LFP theta oscillations, but no change in intrinsic grid cell
  21406. rhythmicity, or firing activity of entorhinal speed and
  21407. head-direction cells. These results suggest that in a 1D
  21408. circular space grid cell spatial selectivity is shaped by path
  21409. integration processes, while grid scale relies on external
  21410. information.",
  21411. journal = "Nat. Commun.",
  21412. publisher = "Nature Publishing Group",
  21413. volume = 10,
  21414. number = 1,
  21415. pages = "840",
  21416. month = feb,
  21417. year = 2019,
  21418. language = "en"
  21419. }
  21420. @ARTICLE{Minderer2019-kz,
  21421. title = "The Spatial Structure of Neural Encoding in Mouse Posterior
  21422. Cortex during Navigation",
  21423. author = "Minderer, Matthias and Brown, Kristen D and Harvey, Christopher D",
  21424. abstract = "Navigation engages many cortical areas, including visual,
  21425. parietal, and retrosplenial cortices. These regions have been
  21426. mapped anatomically and with sensory stimuli and studied
  21427. individually during behavior. Here, we investigated how
  21428. behaviorally driven neural activity is distributed and combined
  21429. across these regions. We performed dense sampling of
  21430. single-neuron activity across the mouse posterior cortex and
  21431. developed unbiased methods to relate neural activity to behavior
  21432. and anatomical space. Most parts of the posterior cortex encoded
  21433. most behavior-related features. However, the relative strength
  21434. with which features were encoded varied across space. Therefore,
  21435. the posterior cortex could be divided into discriminable areas
  21436. based solely on behaviorally relevant neural activity, revealing
  21437. functional structure in association regions. Multimodal
  21438. representations combining sensory and movement signals were
  21439. strongest in posterior parietal cortex, where gradients of
  21440. single-feature representations spatially overlapped. We propose
  21441. that encoding of behavioral features is not constrained by
  21442. retinotopic borders and instead varies smoothly over space within
  21443. association regions.",
  21444. journal = "Neuron",
  21445. month = feb,
  21446. year = 2019,
  21447. keywords = "calcium imaging; cortical architecture; mouse cortex; navigation;
  21448. optogenetics; parietal cortex; virtual reality; visual cortex",
  21449. language = "en"
  21450. }
  21451. @ARTICLE{Stachenfeld2017-yp,
  21452. title = "The hippocampus as a predictive map",
  21453. author = "Stachenfeld, Kimberly L and Botvinick, Matthew M and Gershman,
  21454. Samuel J",
  21455. abstract = "A cognitive map has long been the dominant metaphor for
  21456. hippocampal function, embracing the idea that place cells encode
  21457. a geometric representation of space. However, evidence for
  21458. predictive coding, reward sensitivity and policy dependence in
  21459. place cells suggests that the representation is not purely
  21460. spatial. We approach this puzzle from a reinforcement learning
  21461. perspective: what kind of spatial representation is most useful
  21462. for maximizing future reward? We show that the answer takes the
  21463. form of a predictive representation. This representation captures
  21464. many aspects of place cell responses that fall outside the
  21465. traditional view of a cognitive map. Furthermore, we argue that
  21466. entorhinal grid cells encode a low-dimensionality basis set for
  21467. the predictive representation, useful for suppressing noise in
  21468. predictions and extracting multiscale structure for hierarchical
  21469. planning.",
  21470. journal = "Nat. Neurosci.",
  21471. volume = 20,
  21472. number = 11,
  21473. pages = "1643--1653",
  21474. month = nov,
  21475. year = 2017,
  21476. language = "en"
  21477. }
  21478. @ARTICLE{Foster2012-lf,
  21479. title = "Sequence learning and the role of the hippocampus in rodent
  21480. navigation",
  21481. author = "Foster, David J and Knierim, James J",
  21482. abstract = "The hippocampus has long been associated with navigation and
  21483. spatial representations, but it has been difficult to link
  21484. directly the neurophysiological correlates of hippocampal place
  21485. cells with navigational planning and action. In recent years,
  21486. large-scale population recordings of place cells have revealed
  21487. that spatial sequences are stored and activated in ways that may
  21488. support navigational strategies. Plasticity mechanisms allow the
  21489. hippocampus to store learned sequences of locations that may
  21490. allow predictions of future locations based on past experience.
  21491. These sequences can also be activated during navigational
  21492. behavior in ways that may allow the animal to learn trajectories
  21493. toward goals. Task-dependent alterations in place cell firing
  21494. patterns may reflect the operation of the hippocampus in
  21495. associating locations with navigationally relevant decision
  21496. variables.",
  21497. journal = "Curr. Opin. Neurobiol.",
  21498. publisher = "Elsevier",
  21499. volume = 22,
  21500. number = 2,
  21501. pages = "294--300",
  21502. month = apr,
  21503. year = 2012,
  21504. language = "en"
  21505. }
  21506. @ARTICLE{Whitlock2008-wa,
  21507. title = "Navigating from hippocampus to parietal cortex",
  21508. author = "Whitlock, Jonathan R and Sutherland, Robert J and Witter, Menno
  21509. P and Moser, May-Britt and Moser, Edvard I",
  21510. abstract = "The navigational system of the mammalian cortex comprises a
  21511. number of interacting brain regions. Grid cells in the medial
  21512. entorhinal cortex and place cells in the hippocampus are thought
  21513. to participate in the formation of a dynamic representation of
  21514. the animal's current location, and these cells are presumably
  21515. critical for storing the representation in memory. To traverse
  21516. the environment, animals must be able to translate coordinate
  21517. information from spatial maps in the entorhinal cortex and
  21518. hippocampus into body-centered representations that can be used
  21519. to direct locomotion. How this is done remains an enigma. We
  21520. propose that the posterior parietal cortex is critical for this
  21521. transformation.",
  21522. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  21523. publisher = "National Acad Sciences",
  21524. volume = 105,
  21525. number = 39,
  21526. pages = "14755--14762",
  21527. month = sep,
  21528. year = 2008,
  21529. language = "en"
  21530. }
  21531. @ARTICLE{Chersi2015-qr,
  21532. title = "The Cognitive Architecture of Spatial Navigation: Hippocampal and
  21533. Striatal Contributions",
  21534. author = "Chersi, Fabian and Burgess, Neil",
  21535. abstract = "Spatial navigation can serve as a model system in cognitive
  21536. neuroscience, in which specific neural representations, learning
  21537. rules, and control strategies can be inferred from the vast
  21538. experimental literature that exists across many species,
  21539. including humans. Here, we review this literature, focusing on
  21540. the contributions of hippocampal and striatal systems, and
  21541. attempt to outline a minimal cognitive architecture that is
  21542. consistent with the experimental literature and that synthesizes
  21543. previous related computational modeling. The resulting
  21544. architecture includes striatal reinforcement learning based on
  21545. egocentric representations of sensory states and actions,
  21546. incidental Hebbian association of sensory information with
  21547. allocentric state representations in the hippocampus, and
  21548. arbitration of the outputs of both systems based on
  21549. confidence/uncertainty in medial prefrontal cortex. We discuss
  21550. the relationship between this architecture and learning in
  21551. model-free and model-based systems, episodic memory, imagery, and
  21552. planning, including some open questions and directions for
  21553. further experiments.",
  21554. journal = "Neuron",
  21555. volume = 88,
  21556. number = 1,
  21557. pages = "64--77",
  21558. month = oct,
  21559. year = 2015,
  21560. language = "en"
  21561. }
  21562. @ARTICLE{Maaswinkel1999-yb,
  21563. title = "Homing with locale, taxon, and dead reckoning strategies by
  21564. foraging rats: sensory hierarchy in spatial navigation",
  21565. author = "Maaswinkel, H and Whishaw, I Q",
  21566. abstract = "Studies on foraging rats suggest that they can use visual,
  21567. olfactory, and self-movement cues for spatial guidance, but their
  21568. relative reliance on these different cues is not well understood.
  21569. In the present study, rats left a hidden refuge to search for a
  21570. large food pellet located somewhere on a circular table, and the
  21571. accuracy with which they returned to the refuge with the food
  21572. pellet was measured. Cue use was manipulated by administering
  21573. probe trials from novel locations, blindfolding, moving the home
  21574. cage relative to the table, rotating the table and using
  21575. combinations of these manipulations. When visual cues were
  21576. available and a consistent starting location used, a visual
  21577. strategy dominated performance. When blindfolded, the rats used
  21578. olfactory cues from the surface of the table and from the
  21579. starting hole. When olfactory stimuli were made uninformative, by
  21580. changing the starting hole and rotating the table, the rats still
  21581. homed accurately, suggesting they used self-movement cues. In a
  21582. number of cue combinations, in which cues gave conflicting
  21583. information, performance degraded. The results suggest that rats
  21584. display a hierarchical preference in using visual, olfactory and
  21585. self-movement cues while at the same time being able to reaffirm
  21586. or switch between various cue combinations. The results are
  21587. discussed in relation to ideas concerning the neural basis of
  21588. spatial navigation.",
  21589. journal = "Behav. Brain Res.",
  21590. volume = 99,
  21591. number = 2,
  21592. pages = "143--152",
  21593. month = mar,
  21594. year = 1999,
  21595. language = "en"
  21596. }
  21597. @ARTICLE{Cheung2007-hd,
  21598. title = "Animal navigation: the difficulty of moving in a straight line",
  21599. author = "Cheung, Allen and Zhang, Shaowu and Stricker, Christian and
  21600. Srinivasan, Mandyam V",
  21601. abstract = "In principle, there are two strategies for navigating a straight
  21602. course. One is to use an external directional reference and
  21603. continually reorienting with reference to it, while the other is
  21604. to infer body rotations from internal sensory information only.
  21605. We show here that, while the first strategy will enable an
  21606. animal or mobile agent to move arbitrarily far away from its
  21607. starting point, the second strategy will not do so, even after
  21608. an infinite number of steps. Thus, an external directional
  21609. reference-some form of compass-is indispensable for ensuring
  21610. progress away from home. This limitation must place significant
  21611. constraints on the evolution of biological navigation systems.
  21612. Some specific examples are discussed. An important corollary
  21613. arising from the analysis of compassless navigation is that the
  21614. maximum expected displacement represents a robust measure of the
  21615. straightness of a path.",
  21616. journal = "Biol. Cybern.",
  21617. publisher = "Springer",
  21618. volume = 97,
  21619. number = 1,
  21620. pages = "47--61",
  21621. month = jul,
  21622. year = 2007,
  21623. language = "en"
  21624. }
  21625. @ARTICLE{Zhang2014-nb,
  21626. title = "Spatial representations of place cells in darkness are supported
  21627. by path integration and border information",
  21628. author = "Zhang, Sijie and Sch{\"o}nfeld, Fabian and Wiskott, Laurenz and
  21629. Manahan-Vaughan, Denise",
  21630. abstract = "Effective spatial navigation is enabled by reliable reference
  21631. cues that derive from sensory information from the external
  21632. environment, as well as from internal sources such as the
  21633. vestibular system. The integration of information from these
  21634. sources enables dead reckoning in the form of path integration.
  21635. Navigation in the dark is associated with the accumulation of
  21636. errors in terms of perception of allocentric position and this
  21637. may relate to error accumulation in path integration. We
  21638. assessed this by recording from place cells in the dark under
  21639. circumstances where spatial sensory cues were suppressed.
  21640. Spatial information content, spatial coherence, place field
  21641. size, and peak and infield firing rates decreased whereas
  21642. sparsity increased following exploration in the dark compared to
  21643. the light. Nonetheless it was observed that place field
  21644. stability in darkness was sustained by border information in a
  21645. subset of place cells. To examine the impact of encountering the
  21646. environment's border on navigation, we analyzed the trajectory
  21647. and spiking data gathered during navigation in the dark. Our
  21648. data suggest that although error accumulation in path
  21649. integration drives place field drift in darkness, under
  21650. circumstances where border contact is possible, this information
  21651. is integrated to enable retention of spatial representations.",
  21652. journal = "Front. Behav. Neurosci.",
  21653. publisher = "frontiersin.org",
  21654. volume = 8,
  21655. pages = "222",
  21656. month = jun,
  21657. year = 2014,
  21658. keywords = "CA1; hippocampus; place cells; sensory",
  21659. language = "en"
  21660. }
  21661. @ARTICLE{Battaglia2004-ef,
  21662. title = "Local sensory cues and place cell directionality: additional
  21663. evidence of prospective coding in the hippocampus",
  21664. author = "Battaglia, Francesco P and Sutherland, Gary R and McNaughton,
  21665. Bruce L",
  21666. abstract = "In tasks involving goal-directed, stereotyped trajectories on
  21667. uniform tracks, the spatially selective activity of hippocampal
  21668. principal cells depends on the animal's direction of motion.
  21669. Principal cell ensemble activity while the rat moves in opposite
  21670. directions through a given location is typically uncorrelated.
  21671. It is shown here, with data from three experiments, that
  21672. multimodal, local sensory cues can change the directional
  21673. properties of CA1 pyramidal cells, inducing bidirectionality in
  21674. a significant proportion of place cells. For a majority of these
  21675. bidirectional place cells, place field centers in the two
  21676. directions of motion were displaced relative to one another, as
  21677. would be the case if the cells were representing a position in
  21678. space approximately 5-10 cm ahead of the rat or if place cells
  21679. were subject to strong accommodation or inhibition in the latter
  21680. half of their input fields. However, place field density was not
  21681. affected by the presence of local cues, but in the experimental
  21682. condition with the most salient sensory cues, the CA1 population
  21683. vectors in the ``cue-rich'' condition were sparser and changed
  21684. more quickly in space than in the ``cue-poor'' condition. These
  21685. results suggest that ``view-invariant'' object representations
  21686. are projected to the hippocampus from lower cortical areas and
  21687. can have the effect of increasing the correlation of the
  21688. hippocampal input vectors in the two directions, hence
  21689. decreasing the orthogonality of hippocampal output.",
  21690. journal = "J. Neurosci.",
  21691. publisher = "Soc Neuroscience",
  21692. volume = 24,
  21693. number = 19,
  21694. pages = "4541--4550",
  21695. month = may,
  21696. year = 2004,
  21697. language = "en"
  21698. }
  21699. % The entry below contains non-ASCII chars that could not be converted
  21700. % to a LaTeX equivalent.
  21701. @ARTICLE{Doya1999-jy,
  21702. title = "What are the computations of the cerebellum, the basal ganglia
  21703. and the cerebral cortex?",
  21704. author = "Doya, K",
  21705. abstract = "The classical notion that the cerebellum and the basal ganglia
  21706. are dedicated to motor control is under dispute given increasing
  21707. evidence of their involvement in non-motor functions. Is it then
  21708. impossible to characterize the functions of the cerebellum, the
  21709. basal ganglia and the cerebral cortex in a simplistic manner?
  21710. This paper presents a novel view that their computational roles
  21711. can be characterized not by asking what are the ``goals'' of
  21712. their computation, such as motor or sensory, but by asking what
  21713. are the ``methods'' of their …",
  21714. journal = "Neural Netw.",
  21715. publisher = "Elsevier",
  21716. year = 1999
  21717. }
  21718. @ARTICLE{Killcross2003-ij,
  21719. title = "Coordination of actions and habits in the medial prefrontal
  21720. cortex of rats",
  21721. author = "Killcross, Simon and Coutureau, Etienne",
  21722. abstract = "As animals learn novel behavioural responses, performance is
  21723. maintained by two dissociable influences. Initial responding is
  21724. goal-directed and under voluntary control, but overtraining of
  21725. the same response routine leads to behavioural autonomy and the
  21726. development of habits that are no longer voluntary or
  21727. goal-directed. Rats normally show goal-directed performance
  21728. after limited training, indexed by sensitivity to changes in the
  21729. value of reward, but this sensitivity to goal value is lost with
  21730. extended training. Rats with selective lesions of the prelimbic
  21731. medial prefrontal cortex showed no sensitivity to goal value
  21732. after either limited or extended training, whereas rats with
  21733. lesions of the infralimbic region of the medial prefrontal
  21734. cortex showed the opposite pattern of deficit, a marked
  21735. sensitivity to goal value after both limited and extended
  21736. training. This double-dissociation suggests that the prelimbic
  21737. region is responsible for voluntary response performance and the
  21738. infralimbic cortex mediates the incremental ability of extended
  21739. training to override this goal-directed behaviour.",
  21740. journal = "Cereb. Cortex",
  21741. publisher = "academic.oup.com",
  21742. volume = 13,
  21743. number = 4,
  21744. pages = "400--408",
  21745. month = apr,
  21746. year = 2003,
  21747. language = "en"
  21748. }
  21749. @ARTICLE{Doya2000-wf,
  21750. title = "Complementary roles of basal ganglia and cerebellum in learning
  21751. and motor control",
  21752. author = "Doya, K",
  21753. abstract = "The classical notion that the basal ganglia and the cerebellum
  21754. are dedicated to motor control has been challenged by the
  21755. accumulation of evidence revealing their involvement in
  21756. non-motor, cognitive functions. From a computational viewpoint,
  21757. it has been suggested that the cerebellum, the basal ganglia, and
  21758. the cerebral cortex are specialized for different types of
  21759. learning: namely, supervised learning, reinforcement learning and
  21760. unsupervised learning, respectively. This idea of
  21761. learning-oriented specialization is helpful in understanding the
  21762. complementary roles of the basal ganglia and the cerebellum in
  21763. motor control and cognitive functions.",
  21764. journal = "Curr. Opin. Neurobiol.",
  21765. volume = 10,
  21766. number = 6,
  21767. pages = "732--739",
  21768. month = dec,
  21769. year = 2000,
  21770. language = "en"
  21771. }
  21772. @ARTICLE{Botvinick2009-wz,
  21773. title = "Hierarchically organized behavior and its neural foundations: a
  21774. reinforcement learning perspective",
  21775. author = "Botvinick, Matthew M and Niv, Yael and Barto, Andrew C",
  21776. abstract = "Research on human and animal behavior has long emphasized its
  21777. hierarchical structure-the divisibility of ongoing behavior into
  21778. discrete tasks, which are comprised of subtask sequences, which
  21779. in turn are built of simple actions. The hierarchical structure
  21780. of behavior has also been of enduring interest within
  21781. neuroscience, where it has been widely considered to reflect
  21782. prefrontal cortical functions. In this paper, we reexamine
  21783. behavioral hierarchy and its neural substrates from the point of
  21784. view of recent developments in computational reinforcement
  21785. learning. Specifically, we consider a set of approaches known
  21786. collectively as hierarchical reinforcement learning, which extend
  21787. the reinforcement learning paradigm by allowing the learning
  21788. agent to aggregate actions into reusable subroutines or skills. A
  21789. close look at the components of hierarchical reinforcement
  21790. learning suggests how they might map onto neural structures, in
  21791. particular regions within the dorsolateral and orbital prefrontal
  21792. cortex. It also suggests specific ways in which hierarchical
  21793. reinforcement learning might provide a complement to existing
  21794. psychological models of hierarchically structured behavior. A
  21795. particularly important question that hierarchical reinforcement
  21796. learning brings to the fore is that of how learning identifies
  21797. new action routines that are likely to provide useful building
  21798. blocks in solving a wide range of future problems. Here and at
  21799. many other points, hierarchical reinforcement learning offers an
  21800. appealing framework for investigating the computational and
  21801. neural underpinnings of hierarchically structured behavior.",
  21802. journal = "Cognition",
  21803. volume = 113,
  21804. number = 3,
  21805. pages = "262--280",
  21806. month = dec,
  21807. year = 2009,
  21808. language = "en"
  21809. }
  21810. @ARTICLE{Dezfouli2013-ik,
  21811. title = "Actions, action sequences and habits: evidence that goal-directed
  21812. and habitual action control are hierarchically organized",
  21813. author = "Dezfouli, Amir and Balleine, Bernard W",
  21814. abstract = "Behavioral evidence suggests that instrumental conditioning is
  21815. governed by two forms of action control: a goal-directed and a
  21816. habit learning process. Model-based reinforcement learning (RL)
  21817. has been argued to underlie the goal-directed process; however,
  21818. the way in which it interacts with habits and the structure of
  21819. the habitual process has remained unclear. According to a flat
  21820. architecture, the habitual process corresponds to model-free RL,
  21821. and its interaction with the goal-directed process is coordinated
  21822. by an external arbitration mechanism. Alternatively, the
  21823. interaction between these systems has recently been argued to be
  21824. hierarchical, such that the formation of action sequences
  21825. underlies habit learning and a goal-directed process selects
  21826. between goal-directed actions and habitual sequences of actions
  21827. to reach the goal. Here we used a two-stage decision-making task
  21828. to test predictions from these accounts. The hierarchical account
  21829. predicts that, because they are tied to each other as an action
  21830. sequence, selecting a habitual action in the first stage will be
  21831. followed by a habitual action in the second stage, whereas the
  21832. flat account predicts that the statuses of the first and second
  21833. stage actions are independent of each other. We found, based on
  21834. subjects' choices and reaction times, that human subjects
  21835. combined single actions to build action sequences and that the
  21836. formation of such action sequences was sufficient to explain
  21837. habitual actions. Furthermore, based on Bayesian model
  21838. comparison, a family of hierarchical RL models, assuming a
  21839. hierarchical interaction between habit and goal-directed
  21840. processes, provided a better fit of the subjects' behavior than a
  21841. family of flat models. Although these findings do not rule out
  21842. all possible model-free accounts of instrumental conditioning,
  21843. they do show such accounts are not necessary to explain habitual
  21844. actions and provide a new basis for understanding how
  21845. goal-directed and habitual action control interact.",
  21846. journal = "PLoS Comput. Biol.",
  21847. volume = 9,
  21848. number = 12,
  21849. pages = "e1003364",
  21850. month = dec,
  21851. year = 2013,
  21852. language = "en"
  21853. }
  21854. @ARTICLE{Pennartz2011-jd,
  21855. title = "The hippocampal-striatal axis in learning, prediction and
  21856. goal-directed behavior",
  21857. author = "Pennartz, C M A and Ito, R and Verschure, P F M J and Battaglia,
  21858. F P and Robbins, T W",
  21859. abstract = "The hippocampal formation and striatum subserve declarative and
  21860. procedural memory, respectively. However, experimental evidence
  21861. suggests that the ventral striatum, as opposed to the dorsal
  21862. striatum, does not lend itself to being part of either system.
  21863. Instead, it may constitute a system integrating inputs from the
  21864. amygdala, prefrontal cortex and hippocampus to generate
  21865. motivational, outcome-predicting signals that invigorate
  21866. goal-directed behaviors. Inspired by reinforcement learning
  21867. models, we suggest an alternative scheme for computational
  21868. functions of the striatum. Dorsal and ventral striatum are
  21869. proposed to compute outcome predictions largely in parallel,
  21870. using different types of information as input. The nature of the
  21871. inputs to striatum is furthermore combinatorial, and the
  21872. specificity of predictions transcends the level of scalar value
  21873. signals, incorporating episodic information.",
  21874. journal = "Trends Neurosci.",
  21875. volume = 34,
  21876. number = 10,
  21877. pages = "548--559",
  21878. month = oct,
  21879. year = 2011,
  21880. language = "en"
  21881. }
  21882. @ARTICLE{Verschure_Paul_F_M_J2014-hg,
  21883. title = "The why, what, where, when and how of goal-directed choice:
  21884. neuronal and computational principles",
  21885. author = "{Verschure Paul F. M. J.} and {Pennartz Cyriel M. A.} and
  21886. {Pezzulo Giovanni}",
  21887. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  21888. publisher = "Royal Society",
  21889. volume = 369,
  21890. number = 1655,
  21891. pages = "20130483",
  21892. month = nov,
  21893. year = 2014
  21894. }
  21895. @ARTICLE{Balleine2015-wp,
  21896. title = "Hierarchical control of goal-directed action in the
  21897. cortical--basal ganglia network",
  21898. author = "Balleine, Bernard W and Dezfouli, Amir and Ito, Makato and Doya,
  21899. Kenji",
  21900. abstract = "Goal-directed control depends on constructing a model of the
  21901. world that maps actions onto specific outcomes, allowing choice
  21902. to remain adaptive when the values of outcomes change. In complex
  21903. environments, however, such models can become computationally
  21904. unwieldy. One solution to this problem is to develop a
  21905. hierarchical control structure within which more complex, or
  21906. abstract, actions are built from simpler ones. Here we review
  21907. findings suggesting that the acquisition, evaluation and
  21908. execution of goal-directed actions accords well with predictions
  21909. from hierarchical models. We describe recent evidence that
  21910. hierarchical action control is implemented in a series of
  21911. feedback loops integrating secondary motor areas with the basal
  21912. ganglia and describe how such a structure not only overcomes
  21913. issues of dimensionality, but also helps to explain the formation
  21914. of actions sequences, action chunking and the relationship
  21915. between goal-directed actions and habits.",
  21916. journal = "Current Opinion in Behavioral Sciences",
  21917. volume = 5,
  21918. pages = "1--7",
  21919. month = oct,
  21920. year = 2015
  21921. }
  21922. @ARTICLE{Mobbs2015-ag,
  21923. title = "Neuroethological studies of fear, anxiety, and risky
  21924. decision-making in rodents and humans",
  21925. author = "Mobbs, Dean and Kim, Jeansok J",
  21926. abstract = "Prey are relentlessly faced with a series of survival problems to
  21927. solve. One enduring problem is predation, where the prey's
  21928. answers rely on the complex interaction between actions
  21929. cultivated during its life course and defense reactions passed
  21930. down by descendants. To understand the proximate neural responses
  21931. to analogous threats, affective neuroscientists have favored
  21932. well-controlled associative learning paradigms, yet researchers
  21933. are now creating semi-realistic environments that examine the
  21934. dynamic flow of decision-making and escape calculations that
  21935. mimic the prey's real world choices. In the context of research
  21936. from the field of ethology and behavioral ecology, we review some
  21937. of the recent literature in rodent and human neuroscience and
  21938. discuss how these studies have the potential to provide new
  21939. insights into the behavioral expression, computations, and the
  21940. neural circuits that underlie healthy and pathological fear and
  21941. anxiety.",
  21942. journal = "Curr Opin Behav Sci",
  21943. volume = 5,
  21944. pages = "8--15",
  21945. month = oct,
  21946. year = 2015,
  21947. language = "en"
  21948. }
  21949. @ARTICLE{Ito2016-tm,
  21950. title = "The role of the hippocampus in approach-avoidance conflict
  21951. decision-making: Evidence from rodent and human studies",
  21952. author = "Ito, Rutsuko and Lee, Andy C H",
  21953. abstract = "The hippocampus (HPC) has been traditionally considered to
  21954. subserve mnemonic processing and spatial cognition. Over the past
  21955. decade, however, there has been increasing interest in its
  21956. contributions to processes beyond these two domains. One question
  21957. is whether the HPC plays an important role in decision-making
  21958. under conditions of high approach-avoidance conflict, a scenario
  21959. that arises when a goal stimulus is simultaneously associated
  21960. with reward and punishment. This idea has its origins in rodent
  21961. work conducted in the 1950s and 1960s, and has recently
  21962. experienced a resurgence of interest in the literature. In this
  21963. review, we will first provide an overview of classic rodent
  21964. lesion data that first suggested a role for the HPC in
  21965. approach-avoidance conflict processing and then proceed to
  21966. describe a wide range of more recent evidence from studies
  21967. conducted in rodents and humans. We will demonstrate that there
  21968. is substantial, converging cross-species evidence to support the
  21969. idea that the HPC, in particular the ventral (in
  21970. rodents)/anterior (in humans) portion, contributes to
  21971. approach-avoidance conflict decision making. Furthermore, we
  21972. suggest that the seemingly disparate functions of the HPC (e.g.
  21973. memory, spatial cognition, conflict processing) need not be
  21974. mutually exclusive.",
  21975. journal = "Behav. Brain Res.",
  21976. volume = 313,
  21977. pages = "345--357",
  21978. month = oct,
  21979. year = 2016,
  21980. keywords = "Approach-avoidance; Conflict; Decision-making; Functional
  21981. neuroimaging; Hippocampus; Human; Lesion; Long axis; Memory;
  21982. Rodent; Septotemporal axis; Spatial cognition",
  21983. language = "en"
  21984. }
  21985. @ARTICLE{Viard2011-oi,
  21986. title = "Anterior hippocampus and goal-directed spatial decision making",
  21987. author = "Viard, Armelle and Doeller, Christian F and Hartley, Tom and
  21988. Bird, Chris M and Burgess, Neil",
  21989. abstract = "Planning spatial paths through our environment is an important
  21990. part of everyday life and is supported by a neural system
  21991. including the hippocampus and prefrontal cortex. Here we
  21992. investigated the precise functional roles of the components of
  21993. this system in humans by using fMRI as participants performed a
  21994. simple goal-directed route-planning task. Participants had to
  21995. choose the shorter of two routes to a goal in a visual scene that
  21996. might contain a barrier blocking the most direct route, requiring
  21997. a detour, or might be obscured by a curtain, requiring memory for
  21998. the scene. The participant's start position was varied to
  21999. parametrically manipulate their proximity to the goal and the
  22000. difference in length of the two routes. Activity in medial
  22001. prefrontal cortex, precuneus, and left posterior parietal cortex
  22002. was associated with detour planning, regardless of difficulty,
  22003. whereas activity in parahippocampal gyrus was associated with
  22004. remembering the spatial layout of the visual scene. Activity in
  22005. bilateral anterior hippocampal formation showed a strong increase
  22006. the closer the start position was to the goal, together with
  22007. medial prefrontal, medial and posterior parietal cortices. Our
  22008. results are consistent with computational models in which goal
  22009. proximity is used to guide subsequent navigation and with the
  22010. association of anterior hippocampal areas with nonspatial
  22011. functions such as arousal and reward expectancy. They illustrate
  22012. how spatial and nonspatial functions combine within the anterior
  22013. hippocampus, and how these functions interact with
  22014. parahippocampal, parietal, and prefrontal areas in decision
  22015. making and mnemonic function.",
  22016. journal = "J. Neurosci.",
  22017. volume = 31,
  22018. number = 12,
  22019. pages = "4613--4621",
  22020. month = mar,
  22021. year = 2011,
  22022. language = "en"
  22023. }
  22024. @ARTICLE{Seymour2008-vs,
  22025. title = "Emotion, decision making, and the amygdala",
  22026. author = "Seymour, Ben and Dolan, Ray",
  22027. abstract = "Emotion plays a critical role in many contemporary accounts of
  22028. decision making, but exactly what underlies its influence and how
  22029. this is mediated in the brain remain far from clear. Here, we
  22030. review behavioral studies that suggest that Pavlovian processes
  22031. can exert an important influence over choice and may account for
  22032. many effects that have traditionally been attributed to emotion.
  22033. We illustrate how recent experiments cast light on the underlying
  22034. structure of Pavlovian control and argue that generally this
  22035. influence makes good computational sense. Corresponding
  22036. neuroscientific data from both animals and humans implicate a
  22037. central role for the amygdala through interactions with other
  22038. brain areas. This yields a neurobiological account of emotion in
  22039. which it may operate, often covertly, to optimize rather than
  22040. corrupt economic choice.",
  22041. journal = "Neuron",
  22042. volume = 58,
  22043. number = 5,
  22044. pages = "662--671",
  22045. month = jun,
  22046. year = 2008,
  22047. language = "en"
  22048. }
  22049. % The entry below contains non-ASCII chars that could not be converted
  22050. % to a LaTeX equivalent.
  22051. @ARTICLE{Payne1988-pk,
  22052. title = "Adaptive strategy selection in decision making",
  22053. author = "Payne, John W and Bettman, James R and Johnson, Eric J",
  22054. abstract = "The role of effort and accuracy in the adaptive use of decision
  22055. processes is examined. A computer simulation using the concept
  22056. of elementary information processes identified heuristic choice
  22057. strategies that approximate the accuracy of normative procedures
  22058. while saving substantial effort. However, no single heuristic
  22059. did well across all task and context conditions. Of particular
  22060. interest was the finding that under time constraints, several
  22061. heuristics were more accurate than a truncated normative
  22062. procedure. Using a process …",
  22063. journal = "J. Exp. Psychol. Learn. Mem. Cogn.",
  22064. publisher = "American Psychological Association",
  22065. volume = 14,
  22066. number = 3,
  22067. pages = "534",
  22068. year = 1988
  22069. }
  22070. @ARTICLE{Dayan2012-jd,
  22071. title = "How to set the switches on this thing",
  22072. author = "Dayan, Peter",
  22073. abstract = "Reinforcement learning (RL) has become a dominant computational
  22074. paradigm for modeling psychological and neural aspects of
  22075. affectively charged decision-making tasks. RL is normally
  22076. construed in terms of the interaction between a subject and its
  22077. environment, with the former emitting actions, and the latter
  22078. providing stimuli, and appetitive and aversive reinforcement.
  22079. However, there is recent emphasis on redrawing the boundary
  22080. between the two, with the organism constructing its own notion of
  22081. reward, punishment and state, and with internal actions, such as
  22082. the gating of working memory, being treated on an equal footing
  22083. with external manipulation of the environment. We review recent
  22084. work in this area, focusing on cognitive control.",
  22085. journal = "Curr. Opin. Neurobiol.",
  22086. volume = 22,
  22087. number = 6,
  22088. pages = "1068--1074",
  22089. month = dec,
  22090. year = 2012,
  22091. language = "en"
  22092. }
  22093. @ARTICLE{Balleine2010-es,
  22094. title = "Human and rodent homologies in action control: corticostriatal
  22095. determinants of goal-directed and habitual action",
  22096. author = "Balleine, Bernard W and O'Doherty, John P",
  22097. abstract = "Recent behavioral studies in both humans and rodents have found
  22098. evidence that performance in decision-making tasks depends on two
  22099. different learning processes; one encoding the relationship
  22100. between actions and their consequences and a second involving the
  22101. formation of stimulus-response associations. These learning
  22102. processes are thought to govern goal-directed and habitual
  22103. actions, respectively, and have been found to depend on
  22104. homologous corticostriatal networks in these species. Thus,
  22105. recent research using comparable behavioral tasks in both humans
  22106. and rats has implicated homologous regions of cortex (medial
  22107. prefrontal cortex/medial orbital cortex in humans and prelimbic
  22108. cortex in rats) and of dorsal striatum (anterior caudate in
  22109. humans and dorsomedial striatum in rats) in goal-directed action
  22110. and in the control of habitual actions (posterior lateral putamen
  22111. in humans and dorsolateral striatum in rats). These learning
  22112. processes have been argued to be antagonistic or competing
  22113. because their control over performance appears to be all or none.
  22114. Nevertheless, evidence has started to accumulate suggesting that
  22115. they may at times compete and at others cooperate in the
  22116. selection and subsequent evaluation of actions necessary for
  22117. normal choice performance. It appears likely that cooperation or
  22118. competition between these sources of action control depends not
  22119. only on local interactions in dorsal striatum but also on the
  22120. cortico-basal ganglia network within which the striatum is
  22121. embedded and that mediates the integration of learning with basic
  22122. motivational and emotional processes. The neural basis of the
  22123. integration of learning and motivation in choice and
  22124. decision-making is still controversial and we review some recent
  22125. hypotheses relating to this issue.",
  22126. journal = "Neuropsychopharmacology",
  22127. volume = 35,
  22128. number = 1,
  22129. pages = "48--69",
  22130. month = jan,
  22131. year = 2010,
  22132. language = "en"
  22133. }
  22134. @ARTICLE{Arnsten2009-ow,
  22135. title = "Stress signalling pathways that impair prefrontal cortex
  22136. structure and function",
  22137. author = "Arnsten, Amy F T",
  22138. abstract = "The prefrontal cortex (PFC) - the most evolved brain region -
  22139. subserves our highest-order cognitive abilities. However, it is
  22140. also the brain region that is most sensitive to the detrimental
  22141. effects of stress exposure. Even quite mild acute uncontrollable
  22142. stress can cause a rapid and dramatic loss of prefrontal
  22143. cognitive abilities, and more prolonged stress exposure causes
  22144. architectural changes in prefrontal dendrites. Recent research
  22145. has begun to reveal the intracellular signalling pathways that
  22146. mediate the effects of stress on the PFC. This research has
  22147. provided clues as to why genetic or environmental insults that
  22148. disinhibit stress signalling pathways can lead to symptoms of
  22149. profound prefrontal cortical dysfunction in mental illness.",
  22150. journal = "Nat. Rev. Neurosci.",
  22151. volume = 10,
  22152. number = 6,
  22153. pages = "410--422",
  22154. month = jun,
  22155. year = 2009,
  22156. language = "en"
  22157. }
  22158. @ARTICLE{Schwabe2013-ka,
  22159. title = "Stress and multiple memory systems: from 'thinking' to 'doing'",
  22160. author = "Schwabe, Lars and Wolf, Oliver T",
  22161. abstract = "Although it has been known for decades that stress influences
  22162. memory performance, it was only recently shown that stress may
  22163. alter the contribution of multiple, anatomically and functionally
  22164. distinct memory systems to behavior. Here, we review recent
  22165. animal and human studies demonstrating that stress promotes a
  22166. shift from flexible 'cognitive' to rather rigid 'habit' memory
  22167. systems and discuss, based on recent neuroimaging data in humans,
  22168. the underlying brain mechanisms. We argue that, despite being
  22169. generally adaptive, this stress-induced shift towards 'habit'
  22170. memory may, in vulnerable individuals, be a risk factor for
  22171. psychopathology.",
  22172. journal = "Trends Cogn. Sci.",
  22173. volume = 17,
  22174. number = 2,
  22175. pages = "60--68",
  22176. month = feb,
  22177. year = 2013,
  22178. language = "en"
  22179. }
  22180. @ARTICLE{Wirz2018-zd,
  22181. title = "Habits under stress: mechanistic insights across different types
  22182. of learning",
  22183. author = "Wirz, Lisa and Bogdanov, Mario and Schwabe, Lars",
  22184. abstract = "Learning can be controlled by reflective, `cognitive' or
  22185. reflexive, `habitual' systems. An essential question is what
  22186. factors determine which system governs behavior. Here we review
  22187. recent evidence from navigation, classification, and instrumental
  22188. learning, demonstrating that stressful events induce a shift from
  22189. cognitive to habitual control of learning. We propose that this
  22190. shift, mediated by noradrenaline and glucocorticoids acting
  22191. through mineralocorticoid receptors, is orchestrated by the
  22192. amygdala. Although generally adaptive for coping with acute
  22193. stress, the bias toward habits comes at the cost of reduced
  22194. flexibility of learning and may ultimately contribute to
  22195. stress-related psychopathologies.",
  22196. journal = "Current Opinion in Behavioral Sciences",
  22197. volume = 20,
  22198. pages = "9--16",
  22199. month = apr,
  22200. year = 2018
  22201. }
  22202. @ARTICLE{Corbit2018-ep,
  22203. title = "Understanding the balance between goal-directed and habitual
  22204. behavioral control",
  22205. author = ". Corbit, Laura H",
  22206. abstract = "Decisions can be reached in different ways. Sometimes they
  22207. involve careful consideration of the expected outcome of our
  22208. behavior. Other times, behavior is generated more automatically
  22209. if a particular response has been repeatedly successful in the
  22210. past. I review animal research into goal-directed and habit
  22211. learning including common training paradigms and studies
  22212. investigating the neural substrates of actions and habits.
  22213. Further, I summarize the wide range of factors (e.g., drugs,
  22214. stress, diet) that promote habitual control. Since habitual
  22215. control is prevalent across a range of neuropsychiatric disorders
  22216. it is important to be able to accurately identify when behavior
  22217. is habitual and better understanding of the behavioral and neural
  22218. determinants of habitual control may enable behavior change when
  22219. needed.",
  22220. journal = "Current Opinion in Behavioral Sciences",
  22221. volume = 20,
  22222. pages = "161--168",
  22223. month = apr,
  22224. year = 2018
  22225. }
  22226. @ARTICLE{Schwabe2011-ko,
  22227. title = "Stress-induced modulation of instrumental behavior: from
  22228. goal-directed to habitual control of action",
  22229. author = "Schwabe, Lars and Wolf, Oliver T",
  22230. abstract = "Actions that are directed at achieving pleasant or avoiding
  22231. unpleasant states are referred to as instrumental. The
  22232. acquisition of instrumental actions can be controlled by two
  22233. anatomically and functionally distinct processes: a goal-directed
  22234. process that is based on the prefrontal cortex and dorsomedial
  22235. striatum and encodes the causal relationship between an action
  22236. and the motivational value of the outcome and a dorsolateral
  22237. striatum-based habit process that learns associations between
  22238. actions and antecedent stimuli. Here, we review recent research
  22239. showing that stress modulates the control of instrumental action
  22240. in a manner that favors habitual over goal-directed action. At
  22241. the neuroendocrine level, this stress-induced shift towards habit
  22242. action requires the concerted action of glucocorticoids and
  22243. noradrenergic arousal and is most likely accompanied by opposite
  22244. functional changes in the corticostriatal circuits underlying
  22245. goal-directed and habitual actions. Although generally adaptive,
  22246. these changes in the control of instrumental action under stress
  22247. may promote dysfunctional behaviors and the development of
  22248. psychiatric disorders such as addiction.",
  22249. journal = "Behav. Brain Res.",
  22250. volume = 219,
  22251. number = 2,
  22252. pages = "321--328",
  22253. month = jun,
  22254. year = 2011,
  22255. language = "en"
  22256. }
  22257. @ARTICLE{Keinan1987-ci,
  22258. title = "Decision making under stress: scanning of alternatives under
  22259. controllable and uncontrollable threats",
  22260. author = "Keinan, G",
  22261. abstract = "This study tested the proposition that deficient decision making
  22262. under stress is due, to a significant extent, to the individual's
  22263. failure to fulfill adequately an elementary requirement of the
  22264. decision-making process, that is, the systematic consideration of
  22265. all relevant alternatives. One hundred one undergraduate students
  22266. (59 women and 42 men), aged 20-40, served as subjects in this
  22267. experiment. They were requested to solve decision problems, using
  22268. an interactive computer paradigm, while being exposed to
  22269. controllable stress, uncontrollable stress, or no stress at all.
  22270. There was no time constraint for the performance of the task. The
  22271. controllability of the stressor was found to have no effect on
  22272. the participants' performance. However, those who were exposed to
  22273. either controllable or uncontrollable stress showed a
  22274. significantly stronger tendency to offer solutions before all
  22275. available alternatives had been considered and to scan their
  22276. alternatives in a nonsystematic fashion than did participants who
  22277. were not exposed to stress. In addition, patterns of alternative
  22278. scanning were found to be correlated with the correctness of
  22279. solutions to decision problems.",
  22280. journal = "J. Pers. Soc. Psychol.",
  22281. volume = 52,
  22282. number = 3,
  22283. pages = "639--644",
  22284. month = mar,
  22285. year = 1987,
  22286. language = "en"
  22287. }
  22288. @ARTICLE{Boureau2015-eo,
  22289. title = "Deciding How To Decide: {Self-Control} and {Meta-Decision} Making",
  22290. author = "Boureau, Y-Lan and Sokol-Hessner, Peter and Daw, Nathaniel D",
  22291. abstract = "Many different situations related to self control involve
  22292. competition between two routes to decisions: default and frugal
  22293. versus more resource-intensive. Examples include habits versus
  22294. deliberative decisions, fatigue versus cognitive effort, and
  22295. Pavlovian versus instrumental decision making. We propose that
  22296. these situations are linked by a strikingly similar core dilemma,
  22297. pitting the opportunity costs of monopolizing shared resources
  22298. such as executive functions for some time, against the
  22299. possibility of obtaining a better outcome. We offer a unifying
  22300. normative perspective on this underlying rational
  22301. meta-optimization, review how this may tie together recent
  22302. advances in many separate areas, and connect several independent
  22303. models. Finally, we suggest that the crucial mechanisms and
  22304. meta-decision variables may be shared across domains.",
  22305. journal = "Trends Cogn. Sci.",
  22306. volume = 19,
  22307. number = 11,
  22308. pages = "700--710",
  22309. month = nov,
  22310. year = 2015,
  22311. language = "en"
  22312. }
  22313. @ARTICLE{Keramati2011-oj,
  22314. title = "Speed/accuracy trade-off between the habitual and the
  22315. goal-directed processes",
  22316. author = "Keramati, Mehdi and Dezfouli, Amir and Piray, Payam",
  22317. abstract = "Instrumental responses are hypothesized to be of two kinds:
  22318. habitual and goal-directed, mediated by the sensorimotor and the
  22319. associative cortico-basal ganglia circuits, respectively. The
  22320. existence of the two heterogeneous associative learning
  22321. mechanisms can be hypothesized to arise from the comparative
  22322. advantages that they have at different stages of learning. In
  22323. this paper, we assume that the goal-directed system is
  22324. behaviourally flexible, but slow in choice selection. The
  22325. habitual system, in contrast, is fast in responding, but
  22326. inflexible in adapting its behavioural strategy to new
  22327. conditions. Based on these assumptions and using the
  22328. computational theory of reinforcement learning, we propose a
  22329. normative model for arbitration between the two processes that
  22330. makes an approximately optimal balance between search-time and
  22331. accuracy in decision making. Behaviourally, the model can explain
  22332. experimental evidence on behavioural sensitivity to outcome at
  22333. the early stages of learning, but insensitivity at the later
  22334. stages. It also explains that when two choices with equal
  22335. incentive values are available concurrently, the behaviour
  22336. remains outcome-sensitive, even after extensive training.
  22337. Moreover, the model can explain choice reaction time variations
  22338. during the course of learning, as well as the experimental
  22339. observation that as the number of choices increases, the reaction
  22340. time also increases. Neurobiologically, by assuming that phasic
  22341. and tonic activities of midbrain dopamine neurons carry the
  22342. reward prediction error and the average reward signals used by
  22343. the model, respectively, the model predicts that whereas phasic
  22344. dopamine indirectly affects behaviour through reinforcing
  22345. stimulus-response associations, tonic dopamine can directly
  22346. affect behaviour through manipulating the competition between the
  22347. habitual and the goal-directed systems and thus, affect reaction
  22348. time.",
  22349. journal = "PLoS Comput. Biol.",
  22350. volume = 7,
  22351. number = 5,
  22352. pages = "e1002055",
  22353. month = may,
  22354. year = 2011,
  22355. language = "en"
  22356. }
  22357. @ARTICLE{Yu2016-ns,
  22358. title = "Stress potentiates decision biases: A stress induced
  22359. deliberation-to-intuition ({SIDI}) model",
  22360. author = "Yu, Rongjun",
  22361. abstract = "Humans often make decisions in stressful situations, for example
  22362. when the stakes are high and the potential consequences severe,
  22363. or when the clock is ticking and the task demand is overwhelming.
  22364. In response, a whole train of biological responses to stress has
  22365. evolved to allow organisms to make a fight-or-flight response.
  22366. When under stress, fast and effortless heuristics may dominate
  22367. over slow and demanding deliberation in making decisions under
  22368. uncertainty. Here, I review evidence from behavioral studies and
  22369. neuroimaging research on decision making under stress and propose
  22370. that stress elicits a switch from an analytic reasoning system to
  22371. intuitive processes, and predict that this switch is associated
  22372. with diminished activity in the prefrontal executive control
  22373. regions and exaggerated activity in subcortical reactive emotion
  22374. brain areas. Previous studies have shown that when stressed,
  22375. individuals tend to make more habitual responses than
  22376. goal-directed choices, be less likely to adjust their initial
  22377. judgment, and rely more on gut feelings in social situations. It
  22378. is possible that stress influences the arbitration between the
  22379. emotion responses in subcortical regions and deliberative
  22380. processes in the prefrontal cortex, so that final decisions are
  22381. based on unexamined innate responses. Future research may further
  22382. test this 'stress induced deliberation-to-intuition' (SIDI) model
  22383. and examine its underlying neural mechanisms.",
  22384. journal = "Neurobiol Stress",
  22385. volume = 3,
  22386. pages = "83--95",
  22387. month = jun,
  22388. year = 2016,
  22389. keywords = "Cortisol; Decision making; Stress",
  22390. language = "en"
  22391. }
  22392. @ARTICLE{Frank2009-nv,
  22393. title = "Multiple Systems in Decision Making: A Neurocomputational
  22394. Perspective",
  22395. author = "Frank, Michael J and Cohen, Michael X and Sanfey, Alan G",
  22396. abstract = "Various psychological models posit the existence of two systems
  22397. that contribute to decision making. The first system is
  22398. bottom-up, automatic, intuitive, emotional, and implicit, while
  22399. the second system is top-down, controlled, deliberative, and
  22400. explicit. It has become increasingly evident that this dichotomy
  22401. is both too simplistic and too vague. Here we consider insights
  22402. gained from a different approach, one that considers the
  22403. multiple computational demands of the decision-making system in
  22404. the context of neural mechanisms specialized to accomplish some
  22405. of that system's more basic functions. The use of explicit
  22406. computational models has led to (a) identification of core
  22407. trade-offs imposed by a single-system solution to cognitive
  22408. problems that are solved by having multiple neural systems, and
  22409. (b) novel predictions that can be tested empirically and that
  22410. serve to further refine the models.",
  22411. journal = "Curr. Dir. Psychol. Sci.",
  22412. publisher = "SAGE Publications Inc",
  22413. volume = 18,
  22414. number = 2,
  22415. pages = "73--77",
  22416. month = apr,
  22417. year = 2009
  22418. }
  22419. @ARTICLE{Hasselmo2007-mt,
  22420. title = "Arc length coding by interference of theta frequency oscillations
  22421. may underlie context-dependent hippocampal unit data and episodic
  22422. memory function",
  22423. author = "Hasselmo, Michael E",
  22424. abstract = "Many memory models focus on encoding of sequences by excitatory
  22425. recurrent synapses in region CA3 of the hippocampus. However,
  22426. data and modeling suggest an alternate mechanism for encoding of
  22427. sequences in which interference between theta frequency
  22428. oscillations encodes the position within a sequence based on
  22429. spatial arc length or time. Arc length can be coded by an
  22430. oscillatory interference model that accounts for many features of
  22431. the context-dependent firing properties of hippocampal neurons
  22432. observed during performance of spatial memory tasks. In
  22433. continuous spatial alternation, many neurons fire selectively
  22434. depending on the direction of prior or future response (left or
  22435. right). In contrast, in delayed non-match to position, most
  22436. neurons fire selectively for task phase (sample vs. choice), with
  22437. less selectivity for left versus right. These seemingly disparate
  22438. results are effectively simulated by the same model, based on
  22439. mechanisms similar to a model of grid cell firing in entorhinal
  22440. cortex. The model also simulates forward shifting of firing over
  22441. trials. Adding effects of persistent firing with reset at reward
  22442. locations addresses changes in context-dependent firing with
  22443. different task designs. Arc length coding could contribute to
  22444. episodic encoding of trajectories as sequences of states and
  22445. actions.",
  22446. journal = "Learn. Mem.",
  22447. volume = 14,
  22448. number = 11,
  22449. pages = "782--794",
  22450. month = nov,
  22451. year = 2007,
  22452. language = "en"
  22453. }
  22454. @ARTICLE{Savelli2019-fx,
  22455. title = "Origin and role of path integration in the cognitive
  22456. representations of the hippocampus: computational insights into
  22457. open questions",
  22458. author = "Savelli, Francesco and Knierim, James J",
  22459. abstract = "ABSTRACT Path integration is a straightforward concept with
  22460. varied connotations that are important to different disciplines
  22461. concerned with navigation, such as ethology, cognitive science,
  22462. robotics and neuroscience. In studying the hippocampal
  22463. formation, it is fruitful to think of path integration as a
  22464. computation that transforms a sense of motion into a sense of
  22465. location, continuously integrated with landmark perception.
  22466. Here, we review experimental evidence that path integration is
  22467. intimately involved in fundamental properties of place cells and
  22468. other spatial cells that are thought to support a cognitive
  22469. abstraction of space in this brain system. We discuss hypotheses
  22470. about the anatomical and computational origin of path
  22471. integration in the well-characterized circuits of the rodent
  22472. limbic system. We highlight how computational frameworks for
  22473. map-building in robotics and cognitive science alike suggest an
  22474. essential role for path integration in the creation of a new map
  22475. in unfamiliar territory, and how this very role can help us make
  22476. sense of differences in neurophysiological data from novel
  22477. versus familiar and small versus large environments. Similar
  22478. computational principles could be at work when the hippocampus
  22479. builds certain non-spatial representations, such as time
  22480. intervals or trajectories defined in a sensory stimulus space.",
  22481. journal = "J. Exp. Biol.",
  22482. publisher = "The Company of Biologists Ltd",
  22483. volume = 222,
  22484. number = "Suppl 1",
  22485. pages = "jeb188912",
  22486. month = feb,
  22487. year = 2019,
  22488. language = "en"
  22489. }
  22490. @UNPUBLISHED{Rubin2019-ag,
  22491. title = "Revealing neural correlates of behavior without behavioral
  22492. measurements",
  22493. author = "Rubin, Alon and Sheintuch, Liron and Brande-Eilat, Noa and
  22494. Pinchasof, Or and Rechavi, Yoav and Geva, Nitzan and Ziv, Yaniv",
  22495. abstract = "Measuring neuronal tuning curves has been instrumental for many
  22496. discoveries in neuroscience but requires a-priori assumptions
  22497. regarding the identity of the encoded variables. We applied
  22498. unsupervised learning to large-scale neuronal recordings in
  22499. behaving mice from circuits involved in spatial cognition, and
  22500. uncovered a highly-organized internal structure of ensemble
  22501. activity patterns. This emergent structure allowed defining for
  22502. each neuron an 9internal tuning-curve9 that characterizes its
  22503. activity relative to the network activity, rather than relative
  22504. to any pre-defined external variable -revealing place-tuning in
  22505. the hippocampus and head-direction tuning in the thalamus and
  22506. postsubiculum, without relying on measurements of place or
  22507. head-direction. Similar investigation in prefrontal cortex
  22508. revealed schematic representations of distances and actions, and
  22509. exposed a previously unknown variable, the 9trajectory-phase9.
  22510. The structure of ensemble activity patterns was conserved across
  22511. mice, allowing using one animal9s data to decode another animal9s
  22512. behavior. Thus, the internal structure of neuronal activity
  22513. itself enables reconstructing internal representations and
  22514. discovering new behavioral variables hidden within a neural code.",
  22515. journal = "bioRxiv",
  22516. pages = "540195",
  22517. month = feb,
  22518. year = 2019,
  22519. language = "en"
  22520. }
  22521. @ARTICLE{Rabinowitz2018-ur,
  22522. title = "Machine Theory of Mind",
  22523. author = "Rabinowitz, Neil C and Perbet, Frank and Francis Song, H and
  22524. Zhang, Chiyuan and Ali Eslami, S M and Botvinick, Matthew",
  22525. abstract = "Theory of mind (ToM; Premack \& Woodruff, 1978) broadly
  22526. refers to humans' ability to represent the mental states of
  22527. others, including their desires, beliefs, and intentions. We
  22528. propose to train a machine to build such models too. We
  22529. design a Theory of Mind neural network -- a ToMnet -- which
  22530. uses meta-learning to build models of the agents it
  22531. encounters, from observations of their behaviour alone.
  22532. Through this process, it acquires a strong prior model for
  22533. agents' behaviour, as well as the ability to bootstrap to
  22534. richer predictions about agents' characteristics and mental
  22535. states using only a small number of behavioural
  22536. observations. We apply the ToMnet to agents behaving in
  22537. simple gridworld environments, showing that it learns to
  22538. model random, algorithmic, and deep reinforcement learning
  22539. agents from varied populations, and that it passes classic
  22540. ToM tasks such as the ``Sally-Anne'' test (Wimmer \& Perner,
  22541. 1983; Baron-Cohen et al., 1985) of recognising that others
  22542. can hold false beliefs about the world. We argue that this
  22543. system -- which autonomously learns how to model other
  22544. agents in its world -- is an important step forward for
  22545. developing multi-agent AI systems, for building
  22546. intermediating technology for machine-human interaction, and
  22547. for advancing the progress on interpretable AI.",
  22548. month = feb,
  22549. year = 2018,
  22550. archivePrefix = "arXiv",
  22551. primaryClass = "cs.AI",
  22552. eprint = "1802.07740"
  22553. }
  22554. @UNPUBLISHED{Brette2018-gj,
  22555. title = "Is coding a relevant metaphor for the brain?",
  22556. author = "Brette, Romain",
  22557. abstract = "``Neural coding'' is a popular metaphor in neuroscience, where
  22558. objective properties of the world are communicated to the brain
  22559. in the form of spikes. Here I argue that this metaphor is often
  22560. inappropriate and misleading. First, when neurons are said to
  22561. encode experimental parameters, the neural code depends on
  22562. experimental details that are not carried by the coding variable.
  22563. Thus, the representational power of neural codes is much more
  22564. limited than generally implied. Second, neural codes carry
  22565. information only by reference to things with known meaning. In
  22566. contrast, perceptual systems must build information from
  22567. relations between sensory signals and actions, forming a
  22568. structured internal model. Neural codes are inadequate for this
  22569. purpose because they are unstructured. Third, coding variables
  22570. are observables tied to the temporality of experiments, while
  22571. spikes are timed actions that mediate coupling in a distributed
  22572. dynamical system. The coding metaphor tries to fit the dynamic,
  22573. circular and distributed causal structure of the brain into a
  22574. linear chain of transformations between observables, but the two
  22575. causal structures are incongruent. I conclude that the neural
  22576. coding metaphor cannot provide a basis for theories of brain
  22577. function, because it is incompatible with both the causal
  22578. structure of the brain and the informational requirements of
  22579. cognition.",
  22580. journal = "bioRxiv",
  22581. pages = "168237",
  22582. month = jul,
  22583. year = 2018,
  22584. language = "en"
  22585. }
  22586. @ARTICLE{Lisman2009-tn,
  22587. title = "Prediction, sequences and the hippocampus",
  22588. author = "Lisman, John and Redish, A D",
  22589. abstract = "Recordings of rat hippocampal place cells have provided
  22590. information about how the hippocampus retrieves memory
  22591. sequences. One line of evidence has to do with phase precession,
  22592. a process organized by theta and gamma oscillations. This
  22593. precession can be interpreted as the cued prediction of the
  22594. sequence of upcoming positions. In support of this
  22595. interpretation, experiments in two-dimensional environments and
  22596. on a cue-rich linear track demonstrate that many cells represent
  22597. a position ahead of the animal and that this position is the
  22598. same irrespective of which direction the rat is coming from.
  22599. Other lines of investigation have demonstrated that such
  22600. predictive processes also occur in the non-spatial domain and
  22601. that retrieval can be internally or externally cued. The
  22602. mechanism of sequence retrieval and the usefulness of this
  22603. retrieval to guide behaviour are discussed.",
  22604. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  22605. publisher = "rstb.royalsocietypublishing.org",
  22606. volume = 364,
  22607. number = 1521,
  22608. pages = "1193--1201",
  22609. month = may,
  22610. year = 2009,
  22611. language = "en"
  22612. }
  22613. @ARTICLE{Ji2008-cy,
  22614. title = "Firing rate dynamics in the hippocampus induced by trajectory
  22615. learning",
  22616. author = "Ji, Daoyun and Wilson, Matthew A",
  22617. abstract = "The hippocampus is essential for spatial navigation, which may
  22618. involve sequential learning. However, how the hippocampus
  22619. encodes new sequences in familiar environments is unknown. To
  22620. study the impact of novel spatial sequences on the activity of
  22621. hippocampal neurons, we monitored hippocampal ensembles while
  22622. rats learned to switch from two familiar trajectories to a new
  22623. one in a familiar environment. Here, we show that this novel
  22624. spatial experience induces two types of changes in firing rates,
  22625. but not locations of hippocampal place cells. First, place-cell
  22626. firing rates on the two familiar trajectories start to change
  22627. before the actual behavioral switch to the new trajectory.
  22628. Second, repeated exposure on the new trajectory is associated
  22629. with an increased dependence of place-cell firing rates on
  22630. immediate past locations. The result suggests that sequence
  22631. encoding in the hippocampus may involve integration of
  22632. information about the recent past into current state.",
  22633. journal = "J. Neurosci.",
  22634. publisher = "Soc Neuroscience",
  22635. volume = 28,
  22636. number = 18,
  22637. pages = "4679--4689",
  22638. month = apr,
  22639. year = 2008,
  22640. language = "en"
  22641. }
  22642. @ARTICLE{Eichenbaum2009-jb,
  22643. title = "The neurobiology of memory based predictions",
  22644. author = "Eichenbaum, Howard and Fortin, Norbert J",
  22645. abstract = "Recent findings indicate that, in humans, the hippocampal memory
  22646. system is involved in the capacity to imagine the future as well
  22647. as remember the past. Other studies have suggested that animals
  22648. may also have the capacity to recall the past and plan for the
  22649. future. Here, we will consider data that bridge between these
  22650. sets of findings by assessing the role of the hippocampus in
  22651. memory and prediction in rats. We will argue that animals have
  22652. the capacity for recollection and that the hippocampus plays a
  22653. central and selective role in binding information in the service
  22654. of recollective memory. Then we will consider examples of
  22655. transitive inference, a paradigm that requires the integration
  22656. of overlapping memories and flexible use of the resulting
  22657. relational memory networks for generating predictions in novel
  22658. situations. Our data show that animals have the capacity for
  22659. transitive inference and that the hippocampus plays a central
  22660. role in the ability to predict outcomes of events that have not
  22661. yet occurred.",
  22662. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  22663. publisher = "royalsocietypublishing.org",
  22664. volume = 364,
  22665. number = 1521,
  22666. pages = "1183--1191",
  22667. month = may,
  22668. year = 2009,
  22669. language = "en"
  22670. }
  22671. @ARTICLE{Eichenbaum2000-rr,
  22672. title = "A cortical--hippocampal system for declarative memory",
  22673. author = "Eichenbaum, Howard",
  22674. abstract = "Recent neurobiological studies have begun to reveal the
  22675. cognitive and neural coding mechanisms that underlie declarative
  22676. memory --- our ability to recollect everyday events and factual
  22677. knowledge. These studies indicate that the critical circuitry
  22678. involves bidirectional connections between the neocortex, the
  22679. parahippocampal region and the hippocampus. Each of these areas
  22680. makes a unique contribution to memory processing. Widespread
  22681. high-order neocortical areas provide dedicated processors for
  22682. perceptual, motor or cognitive information that is influenced by
  22683. other components of the system. The parahippocampal region
  22684. mediates convergence of this information and extends the
  22685. persistence of neocortical memory representations. The
  22686. hippocampus encodes the sequences of places and events that
  22687. compose episodic memories, and links them together through their
  22688. common elements. Here I describe how these mechanisms work
  22689. together to create and re-create fully networked representations
  22690. of previous experiences and knowledge about the world.",
  22691. journal = "Nat. Rev. Neurosci.",
  22692. publisher = "Macmillan Magazines Ltd.",
  22693. volume = 1,
  22694. pages = "41",
  22695. month = oct,
  22696. year = 2000
  22697. }
  22698. @ARTICLE{Moser2008-af,
  22699. title = "Place cells, grid cells, and the brain's spatial representation
  22700. system",
  22701. author = "Moser, Edvard I and Kropff, Emilio and Moser, May-Britt",
  22702. abstract = "More than three decades of research have demonstrated a role for
  22703. hippocampal place cells in representation of the spatial
  22704. environment in the brain. New studies have shown that place
  22705. cells are part of a broader circuit for dynamic representation
  22706. of self-location. A key component of this network is the
  22707. entorhinal grid cells, which, by virtue of their tessellating
  22708. firing fields, may provide the elements of a path
  22709. integration-based neural map. Here we review how place cells and
  22710. grid cells may form the basis for quantitative spatiotemporal
  22711. representation of places, routes, and associated experiences
  22712. during behavior and in memory. Because these cell types have
  22713. some of the most conspicuous behavioral correlates among neurons
  22714. in nonsensory cortical systems, and because their spatial firing
  22715. structure reflects computations internally in the system,
  22716. studies of entorhinal-hippocampal representations may offer
  22717. considerable insight into general principles of cortical network
  22718. dynamics.",
  22719. journal = "Annu. Rev. Neurosci.",
  22720. publisher = "annualreviews.org",
  22721. volume = 31,
  22722. pages = "69--89",
  22723. year = 2008,
  22724. language = "en"
  22725. }
  22726. @ARTICLE{Squire2004-hb,
  22727. title = "The medial temporal lobe",
  22728. author = "Squire, Larry R and Stark, Craig E L and Clark, Robert E",
  22729. abstract = "The medial temporal lobe includes a system of anatomically
  22730. related structures that are essential for declarative memory
  22731. (conscious memory for facts and events). The system consists of
  22732. the hippocampal region (CA fields, dentate gyrus, and subicular
  22733. complex) and the adjacent perirhinal, entorhinal, and
  22734. parahippocampal cortices. Here, we review findings from humans,
  22735. monkeys, and rodents that illuminate the function of these
  22736. structures. Our analysis draws on studies of human memory
  22737. impairment and animal models of memory impairment, as well as
  22738. neurophysiological and neuroimaging data, to show that this
  22739. system (a) is principally concerned with memory, (b) operates
  22740. with neocortex to establish and maintain long-term memory, and
  22741. (c) ultimately, through a process of consolidation, becomes
  22742. independent of long-term memory, though questions remain about
  22743. the role of perirhinal and parahippocampal cortices in this
  22744. process and about spatial memory in rodents. Data from
  22745. neurophysiology, neuroimaging, and neuroanatomy point to a
  22746. division of labor within the medial temporal lobe. However, the
  22747. available data do not support simple dichotomies between the
  22748. functions of the hippocampus and the adjacent medial temporal
  22749. cortex, such as associative versus nonassociative memory,
  22750. episodic versus semantic memory, and recollection versus
  22751. familiarity.",
  22752. journal = "Annu. Rev. Neurosci.",
  22753. volume = 27,
  22754. pages = "279--306",
  22755. year = 2004,
  22756. language = "en"
  22757. }
  22758. @ARTICLE{Kloosterman2014-ki,
  22759. title = "Bayesian decoding using unsorted spikes in the rat hippocampus",
  22760. author = "Kloosterman, Fabian and Layton, Stuart P and Chen, Zhe and
  22761. Wilson, Matthew A",
  22762. abstract = "A fundamental task in neuroscience is to understand how neural
  22763. ensembles represent information. Population decoding is a useful
  22764. tool to extract information from neuronal populations based on
  22765. the ensemble spiking activity. We propose a novel Bayesian
  22766. decoding paradigm to decode unsorted spikes in the rat
  22767. hippocampus. Our approach uses a direct mapping between spike
  22768. waveform features and covariates of interest and avoids
  22769. accumulation of spike sorting errors. Our decoding paradigm is
  22770. nonparametric, encoding model-free for representing stimuli, and
  22771. extracts information from all available spikes and their
  22772. waveform features. We apply the proposed Bayesian decoding
  22773. algorithm to a position reconstruction task for freely behaving
  22774. rats based on tetrode recordings of rat hippocampal neuronal
  22775. activity. Our detailed decoding analyses demonstrate that our
  22776. approach is efficient and better utilizes the available
  22777. information in the nonsortable hash than the standard
  22778. sorting-based decoding algorithm. Our approach can be adapted to
  22779. an online encoding/decoding framework for applications that
  22780. require real-time decoding, such as brain-machine interfaces.",
  22781. journal = "J. Neurophysiol.",
  22782. publisher = "physiology.org",
  22783. volume = 111,
  22784. number = 1,
  22785. pages = "217--227",
  22786. month = jan,
  22787. year = 2014,
  22788. keywords = "kernel density estimation; neural decoding; population codes;
  22789. spatial-temporal Poisson process; spike sorting",
  22790. language = "en"
  22791. }
  22792. @ARTICLE{Wikenheiser2015-pj,
  22793. title = "Hippocampal theta sequences reflect current goals",
  22794. author = "Wikenheiser, Andrew M and Redish, A David",
  22795. abstract = "Hippocampal information processing is discretized by
  22796. oscillations, and the ensemble activity of place cells is
  22797. organized into temporal sequences bounded by theta cycles. Theta
  22798. sequences represent time-compressed trajectories through space.
  22799. Their forward-directed nature makes them an intuitive candidate
  22800. mechanism for planning future trajectories, but their connection
  22801. to goal-directed behavior remains unclear. As rats performed a
  22802. value-guided decision-making task, the extent to which theta
  22803. sequences projected ahead of the animal's current location
  22804. varied on a moment-by-moment basis depending on the rat's goals.
  22805. Look-ahead extended farther on journeys to distant goals than on
  22806. journeys to more proximal goals and was predictive of the
  22807. animal's destination. On arrival at goals, however, look-ahead
  22808. was similar regardless of where the animal began its journey
  22809. from. Together, these results provide evidence that hippocampal
  22810. theta sequences contain information related to goals or
  22811. intentions, pointing toward a potential spatial basis for
  22812. planning.",
  22813. journal = "Nat. Neurosci.",
  22814. publisher = "nature.com",
  22815. volume = 18,
  22816. number = 2,
  22817. pages = "289--294",
  22818. month = feb,
  22819. year = 2015,
  22820. language = "en"
  22821. }
  22822. % The entry below contains non-ASCII chars that could not be converted
  22823. % to a LaTeX equivalent.
  22824. @ARTICLE{Buckner2010-by,
  22825. title = "The role of the hippocampus in prediction and imagination",
  22826. author = "Buckner, Randy L",
  22827. abstract = "Traditionally, the hippocampal system has been studied in
  22828. relation to the goal of retrieving memories about the past.
  22829. Recent work in humans and rodents suggests that the hippocampal
  22830. system may be better understood as a system that facilitates
  22831. predictions about upcoming events. The hippocampus and
  22832. associated cortical structures are active when people envision
  22833. future events, and damage that includes the hippocampal region
  22834. impairs this ability. In rats, hippocampal ensembles preplay and
  22835. replay event sequences in the …",
  22836. journal = "Annu. Rev. Psychol.",
  22837. publisher = "Annual Reviews",
  22838. volume = 61,
  22839. pages = "27--48",
  22840. year = 2010
  22841. }
  22842. @ARTICLE{Mullally2014-cr,
  22843. title = "Memory, Imagination, and Predicting the Future: A Common Brain
  22844. Mechanism?",
  22845. author = "Mullally, Sin{\'e}ad L and Maguire, Eleanor A",
  22846. abstract = "On the face of it, memory, imagination, and prediction seem to
  22847. be distinct cognitive functions. However, metacognitive,
  22848. cognitive, neuropsychological, and neuroimaging evidence is
  22849. emerging that they are not, suggesting intimate links in their
  22850. underlying processes. Here, we explore these empirical findings
  22851. and the evolving theoretical frameworks that seek to explain how
  22852. a common neural system supports our recollection of times past,
  22853. imagination, and our attempts to predict the future.",
  22854. journal = "Neuroscientist",
  22855. publisher = "journals.sagepub.com",
  22856. volume = 20,
  22857. number = 3,
  22858. pages = "220--234",
  22859. month = jun,
  22860. year = 2014,
  22861. keywords = "amnesia; episodic memory; fMRI; future; hippocampus;
  22862. imagination; navigation; neuropsychology; prediction;
  22863. prospection; scene construction; scenes; simulation",
  22864. language = "en"
  22865. }
  22866. @ARTICLE{Smallwood2015-wd,
  22867. title = "The science of mind wandering: empirically navigating the stream
  22868. of consciousness",
  22869. author = "Smallwood, Jonathan and Schooler, Jonathan W",
  22870. abstract = "Conscious experience is fluid; it rarely remains on one topic
  22871. for an extended period without deviation. Its dynamic nature is
  22872. illustrated by the experience of mind wandering, in which
  22873. attention switches from a current task to unrelated thoughts and
  22874. feelings. Studies exploring the phenomenology of mind wandering
  22875. highlight the importance of its content and relation to
  22876. meta-cognition in determining its functional outcomes.
  22877. Examination of the information-processing demands of the
  22878. mind-wandering state suggests that it involves perceptual
  22879. decoupling to escape the constraints of the moment, its content
  22880. arises from episodic and affective processes, and its regulation
  22881. relies on executive control. Mind wandering also involves a
  22882. complex balance of costs and benefits: Its association with
  22883. various kinds of error underlines its cost, whereas its
  22884. relationship to creativity and future planning suggest its
  22885. potential value. Although essential to the stream of
  22886. consciousness, various strategies may minimize the downsides of
  22887. mind wandering while maintaining its productive aspects.",
  22888. journal = "Annu. Rev. Psychol.",
  22889. publisher = "annualreviews.org",
  22890. volume = 66,
  22891. pages = "487--518",
  22892. month = jan,
  22893. year = 2015,
  22894. keywords = "default mode network; mental time travel; meta-awareness; mind
  22895. wandering; perceptual decoupling; self-generated thought",
  22896. language = "en"
  22897. }
  22898. @ARTICLE{Voermans2004-ir,
  22899. title = "Interaction between the human hippocampus and the caudate
  22900. nucleus during route recognition",
  22901. author = "Voermans, Nicol C and Petersson, Karl Magnus and Daudey, Leonie
  22902. and Weber, Bernd and Van Spaendonck, Karel P and Kremer,
  22903. Hubertus P H and Fern{\'a}ndez, Guill{\'e}n",
  22904. abstract = "Navigation through familiar environments can rely upon distinct
  22905. neural representations that are related to different memory
  22906. systems with either the hippocampus or the caudate nucleus at
  22907. their core. However, it is a fundamental question whether and
  22908. how these systems interact during route recognition. To address
  22909. this issue, we combined a functional neuroimaging approach with
  22910. a naturally occurring, well-controlled human model of caudate
  22911. nucleus dysfunction (i.e., preclinical and early-stage
  22912. Huntington's disease). Our results reveal a noncompetitive
  22913. interaction so that the hippocampus compensates for gradual
  22914. caudate nucleus dysfunction with a gradual activity increase,
  22915. maintaining normal behavior. Furthermore, we revealed an
  22916. interaction between medial temporal and caudate activity in
  22917. healthy subjects, which was adaptively modified in Huntington
  22918. patients to allow compensatory hippocampal processing. Thus, the
  22919. two memory systems contribute in a noncompetitive, cooperative
  22920. manner to route recognition, which enables the hippocampus to
  22921. compensate seamlessly for the functional degradation of the
  22922. caudate nucleus.",
  22923. journal = "Neuron",
  22924. publisher = "Elsevier",
  22925. volume = 43,
  22926. number = 3,
  22927. pages = "427--435",
  22928. month = aug,
  22929. year = 2004,
  22930. language = "en"
  22931. }
  22932. @ARTICLE{Lee2011-an,
  22933. title = "Using hierarchical Bayesian methods to examine the tools of
  22934. decision-making",
  22935. author = "Lee, Michael D and Newell, Ben R",
  22936. abstract = "Author(s): Lee, Michael D.; Newell, Ben R. | Abstract:
  22937. Hierarchical Bayesian methods offer a principled and
  22938. comprehensive way to relate psychological models to data. Here
  22939. we use them to model the patterns of information search,
  22940. stopping and deciding in a simulated binary comparison judgment
  22941. task. The simulation involves 20 subjects making 100 forced
  22942. choice comparisons about the relative magnitudes of two objects
  22943. (which of two German cities has more inhabitants). Two
  22944. worked-examples show how hierarchical models can be developed to
  22945. account for and explain the diversity of both search and
  22946. stopping rules seen across the simulated individuals. We discuss
  22947. how the results provide insight into current debates in the
  22948. literature on heuristic decision making and argue that they
  22949. demonstrate the power and flexibility of hierarchical Bayesian
  22950. methods in modeling human decision-making.",
  22951. publisher = "escholarship.org",
  22952. volume = 6,
  22953. number = 8,
  22954. pages = "832--842",
  22955. month = dec,
  22956. year = 2011,
  22957. keywords = "Social and Behavioral Sciences"
  22958. }
  22959. @ARTICLE{Zatka-Haas_undated-ob,
  22960. title = "Distinct contributions of mouse cortical areas to visual
  22961. discrimination",
  22962. author = "Zatka-Haas, Peter and Steinmetz, Nicholas A and Carandini, Matteo
  22963. and Harris, Kenneth D"
  22964. }
  22965. % The entry below contains non-ASCII chars that could not be converted
  22966. % to a LaTeX equivalent.
  22967. @ARTICLE{Penny2012-xo,
  22968. title = "Bayesian models of brain and behaviour",
  22969. author = "Penny, William",
  22970. abstract = "… A readjust- ment at the next level in the hierarchy may
  22971. increase … have multiple hidden variables, for example,
  22972. representing different levels of abstraction in cortical
  22973. hierarchies , and multiple … the applicability of the approach,
  22974. but it is highly efficient for many hierarchical models [21] …",
  22975. journal = "ISRN Biomathematics",
  22976. publisher = "Hindawi Publishing Corporation",
  22977. volume = 2012,
  22978. year = 2012
  22979. }
  22980. @UNPUBLISHED{Bernardi2018-ma,
  22981. title = "The geometry of abstraction in hippocampus and prefrontal cortex",
  22982. author = "Bernardi, Silvia and Benna, Marcus K and Rigotti, Mattia and
  22983. Munuera, Jerome and Fusi, Stefano and Salzman, Daniel",
  22984. abstract = "Abstraction can be defined as a cognitive process that identifies
  22985. common features - abstract variables, or concepts - shared by
  22986. many examples. Such conceptual knowledge enables subjects to
  22987. generalize upon encountering new examples, an ability that
  22988. supports inferential reasoning and cognitive flexibility. To
  22989. confer the ability to generalize, the brain must represent
  22990. variables in a particular `abstract9 format. Here we show how to
  22991. construct neural representations that encode multiple variables
  22992. in an abstract format simultaneously, and we characterize their
  22993. geometry. Neural representations conforming to this geometry were
  22994. observed in dorsolateral pre-frontal cortex, anterior cingulate
  22995. cortex and the hippocampus in monkeys performing a serial
  22996. reversal-learning task. Similar representations are observed in a
  22997. simulated multi-layer neural network trained with
  22998. back-propagation. These findings provide a novel framework for
  22999. characterizing how different brain areas represent abstract
  23000. variables that are critical for flexible conceptual
  23001. generalization.",
  23002. journal = "bioRxiv",
  23003. pages = "408633",
  23004. month = dec,
  23005. year = 2018,
  23006. language = "en"
  23007. }
  23008. @ARTICLE{Lee2012-pe,
  23009. title = "Neural basis of reinforcement learning and decision making",
  23010. author = "Lee, Daeyeol and Seo, Hyojung and Jung, Min Whan",
  23011. abstract = "Reinforcement learning is an adaptive process in which an animal
  23012. utilizes its previous experience to improve the outcomes of
  23013. future choices. Computational theories of reinforcement learning
  23014. play a central role in the newly emerging areas of neuroeconomics
  23015. and decision neuroscience. In this framework, actions are chosen
  23016. according to their value functions, which describe how much
  23017. future reward is expected from each action. Value functions can
  23018. be adjusted not only through reward and penalty, but also by the
  23019. animal's knowledge of its current environment. Studies have
  23020. revealed that a large proportion of the brain is involved in
  23021. representing and updating value functions and using them to
  23022. choose an action. However, how the nature of a behavioral task
  23023. affects the neural mechanisms of reinforcement learning remains
  23024. incompletely understood. Future studies should uncover the
  23025. principles by which different computational elements of
  23026. reinforcement learning are dynamically coordinated across the
  23027. entire brain.",
  23028. journal = "Annu. Rev. Neurosci.",
  23029. volume = 35,
  23030. pages = "287--308",
  23031. month = mar,
  23032. year = 2012,
  23033. language = "en"
  23034. }
  23035. @ARTICLE{Groman2018-tb,
  23036. title = "Neurochemical and behavioral dissections of decision-making in a
  23037. rodent multi-stage task",
  23038. author = "Groman, Stephanie M and Massi, Bart and Mathias, Samuel R and
  23039. Curry, Daniel W and Lee, Daeyeol and Taylor, Jane R",
  23040. abstract = "Flexible decision-making in dynamic environments requires both
  23041. retrospective appraisal of reinforced actions and prospective
  23042. reasoning about the consequences of actions. These complementary
  23043. reinforcement-learning systems can be characterized
  23044. computationally with model-free and model-based algorithms, but
  23045. how these processes interact at a neurobehavioral level in normal
  23046. and pathological states is unknown. Here, we developed a
  23047. translationally analogous multi-stage decision-making task to
  23048. independently quantify model-free and model-based behavioral
  23049. mechanisms in rats. We provide the first direct evidence that
  23050. male rats, similar to humans, use both model-free and model-based
  23051. learning when making value-based choices in the multi-stage
  23052. decision-making task and provide novel analytic approaches for
  23053. independently quantifying these reinforcement-learning
  23054. strategies. Furthermore, we report that ex vivo dopamine tone in
  23055. the ventral striatum and orbitofrontal cortex correlate with
  23056. model-based, but not model-free, strategies indicating that the
  23057. biological mechanisms mediating decision-making in the
  23058. multi-stage task are conserved in rats and humans. This new
  23059. multi-stage task provides a unique behavioral platform for
  23060. conducting systems level analyses of decision-making in normal
  23061. and pathological states.Significance statementDecision-making is
  23062. influenced by both a retrospective ``model free'' system and a
  23063. prospective ``model based'' system in humans, but the
  23064. biobehavioral mechanisms mediating these learning systems in
  23065. normal and disease states are unknown. Here, we describe a
  23066. translationally analogous multi-stage decision-making task to
  23067. provide a behavioral platform for conducting neuroscience studies
  23068. of decision-making in rats. We provide the first evidence that
  23069. choice behavior in rats is influenced by model-free and
  23070. model-based systems and demonstrate that model-based, but not
  23071. model-free, learning is associated with cortico-striatal dopamine
  23072. tone. This novel behavioral paradigm has the potential to yield
  23073. critical insights into the mechanisms mediating decision-making
  23074. alterations in mental disorders.",
  23075. journal = "J. Neurosci.",
  23076. month = nov,
  23077. year = 2018,
  23078. language = "en"
  23079. }
  23080. @UNPUBLISHED{Clawson2019-kk,
  23081. title = "Computing Hubs in the Hippocampus and Cortex",
  23082. author = "Clawson, Wesley P and Vicente, Ana F and Bernard, Christophe and
  23083. Battaglia, Demian and Quilichini, Pascale P",
  23084. abstract = "Neural computation, which relies on the active storage and
  23085. sharing of information, occurs within large neuron networks in
  23086. the highly dynamic context of varying brain states. Whether such
  23087. functions are performed by specific subsets of neurons and
  23088. whether they occur in specific dynamical regimes remain poorly
  23089. understood. Using high density recordings in the hippocampus,
  23090. medial entorhinal and medial prefrontal cortex of the rat, we
  23091. identify computing microstates, or discreet epochs, in which
  23092. specific computing hub neurons perform well defined storage and
  23093. sharing operations in a brain state-dependent manner. We retrieve
  23094. a multiplicity of distinct computing microstates within each
  23095. global brain state, such as REM and nonREM sleep. Half of
  23096. recorded neurons act as computing hubs in at least one
  23097. microstate, suggesting that functional roles are not firmly
  23098. hardwired but dynamically reassigned at the second timescale. We
  23099. identify sequences of microstates whose temporal organization is
  23100. dynamic and stands between order and disorder. We propose that
  23101. global brain states constrain the language of neuronal
  23102. computations by regulating the syntactic complexity of these
  23103. microstate sequences.",
  23104. journal = "bioRxiv",
  23105. pages = "513424",
  23106. month = jan,
  23107. year = 2019,
  23108. language = "en"
  23109. }
  23110. @ARTICLE{Crochet2019-lk,
  23111. title = "Neural Circuits for {Goal-Directed} Sensorimotor Transformations",
  23112. author = "Crochet, Sylvain and Lee, Seung-Hee and Petersen, Carl C H",
  23113. abstract = "Precisely wired neuronal circuits process sensory information in
  23114. a learning- and context-dependent manner in order to govern
  23115. behavior. Simple sensory decision-making tasks in rodents are now
  23116. beginning to reveal the contributions of distinct cell types and
  23117. brain regions participating in the conversion of sensory
  23118. information into learned goal-directed motor output. Task
  23119. learning is accompanied by target-specific routing of sensory
  23120. information to specific downstream cortical regions, with
  23121. higher-order cortical regions such as the posterior parietal
  23122. cortex, medial prefrontal cortex, and hippocampus appearing to
  23123. play important roles in learning- and context-dependent
  23124. processing of sensory input. An important challenge for future
  23125. research is to connect cell-type-specific activity in these brain
  23126. regions with motor neurons responsible for action initiation.",
  23127. journal = "Trends Neurosci.",
  23128. volume = 42,
  23129. number = 1,
  23130. pages = "66--77",
  23131. month = jan,
  23132. year = 2019,
  23133. keywords = "decision-making; neocortex; neuronal cell-types; sensory
  23134. perception",
  23135. language = "en"
  23136. }
  23137. @MISC{John_OKeefe_undated-cu,
  23138. title = "The hippocampus as a cognitive map",
  23139. author = "John O'Keefe, Lynn Nadel",
  23140. keywords = "books;Books"
  23141. }
  23142. @ARTICLE{McNaughton2006-by,
  23143. title = "Path integration and the neural basis of the 'cognitive map'",
  23144. author = "McNaughton, Bruce L and Battaglia, Francesco P and Jensen, Ole
  23145. and Moser, Edvard I and Moser, May-Britt",
  23146. abstract = "The hippocampal formation can encode relative spatial location,
  23147. without reference to external cues, by the integration of linear
  23148. and angular self-motion (path integration). Theoretical studies,
  23149. in conjunction with recent empirical discoveries, suggest that
  23150. the medial entorhinal cortex (MEC) might perform some of the
  23151. essential underlying computations by means of a unique, periodic
  23152. synaptic matrix that could be self-organized in early development
  23153. through a simple, symmetry-breaking operation. The scale at which
  23154. space is represented increases systematically along the
  23155. dorsoventral axis in both the hippocampus and the MEC, apparently
  23156. because of systematic variation in the gain of a movement-speed
  23157. signal. Convergence of spatially periodic input at multiple
  23158. scales, from so-called grid cells in the entorhinal cortex, might
  23159. result in non-periodic spatial firing patterns (place fields) in
  23160. the hippocampus.",
  23161. journal = "Nat. Rev. Neurosci.",
  23162. volume = 7,
  23163. number = 8,
  23164. pages = "663--678",
  23165. month = aug,
  23166. year = 2006,
  23167. language = "en"
  23168. }
  23169. @ARTICLE{Pereira2016-eq,
  23170. title = "Is there anybody out there? Neural circuits of threat detection
  23171. in vertebrates",
  23172. author = "Pereira, Ana G and Moita, Marta A",
  23173. abstract = "Avoiding or escaping a predator is arguably one of the most
  23174. important functions of a prey's brain, hence of most animals'
  23175. brains. Studies on fear conditioning have greatly advanced our
  23176. understanding of the circuits that regulate learned defensive
  23177. behaviours. However, animals possess a multitude of threat
  23178. detection mechanisms, from hardwired circuits that ensure innate
  23179. responses to predator cues, to the use of social information.
  23180. Surprisingly, only more recently have these circuits captured the
  23181. attention of a wider range of researchers working on different
  23182. species and behavioural paradigms. These have shed new light into
  23183. the mechanisms of threat detection revealing conservation of the
  23184. kinds of cues animals use and of its underlying detection
  23185. circuits across vertebrates. As most of these studies focus on
  23186. single cues, we argue for the need to study multisensory
  23187. integration, a process that we believe is determinant for the
  23188. prey's defence responses.",
  23189. journal = "Curr. Opin. Neurobiol.",
  23190. volume = 41,
  23191. pages = "179--187",
  23192. month = dec,
  23193. year = 2016,
  23194. language = "en"
  23195. }
  23196. @ARTICLE{Stott2014-ac,
  23197. title = "A functional difference in information processing between
  23198. orbitofrontal cortex and ventral striatum during decision-making
  23199. behaviour",
  23200. author = "Stott, Jeffrey J and Redish, A David",
  23201. abstract = "Both orbitofrontal cortex (OFC) and ventral striatum (vStr) have
  23202. been identified as key structures that represent information
  23203. about value in decision-making tasks. However, the dynamics of
  23204. how this information is processed are not yet understood. We
  23205. recorded ensembles of cells from OFC and vStr in rats engaged in
  23206. the spatial adjusting delay-discounting task, a decision-making
  23207. task that involves a trade-off between delay to and magnitude of
  23208. reward. Ventral striatal neural activity signalled information
  23209. about reward before the rat's decision, whereas such
  23210. reward-related signals were absent in OFC until after the animal
  23211. had committed to its decision. These data support models in which
  23212. vStr is directly involved in action selection, but OFC processes
  23213. decision-related information afterwards that can be used to
  23214. compare the predicted and actual consequences of behaviour.",
  23215. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  23216. volume = 369,
  23217. number = 1655,
  23218. month = nov,
  23219. year = 2014,
  23220. keywords = "decision-making; neuroeconomics; nucleus accumbens; orbitofrontal
  23221. cortex; vicarious trial and error",
  23222. language = "en"
  23223. }
  23224. @ARTICLE{Jones2005-ye,
  23225. title = "Theta rhythms coordinate hippocampal-prefrontal interactions in a
  23226. spatial memory task",
  23227. author = "Jones, Matthew W and Wilson, Matthew A",
  23228. abstract = "Decision-making requires the coordinated activity of diverse
  23229. brain structures. For example, in maze-based tasks, the
  23230. prefrontal cortex must integrate spatial information encoded in
  23231. the hippocampus with mnemonic information concerning route and
  23232. task rules in order to direct behavior appropriately. Using
  23233. simultaneous tetrode recordings from CA1 of the rat hippocampus
  23234. and medial prefrontal cortex, we show that correlated firing in
  23235. the two structures is selectively enhanced during behavior that
  23236. recruits spatial working memory, allowing the integration of
  23237. hippocampal spatial information into a broader, decision-making
  23238. network. The increased correlations are paralleled by enhanced
  23239. coupling of the two structures in the 4- to 12-Hz theta-frequency
  23240. range. Thus the coordination of theta rhythms may constitute a
  23241. general mechanism through which the relative timing of disparate
  23242. neural activities can be controlled, allowing specialized brain
  23243. structures to both encode information independently and to
  23244. interact selectively according to current behavioral demands.",
  23245. journal = "PLoS Biol.",
  23246. volume = 3,
  23247. number = 12,
  23248. pages = "e402",
  23249. month = dec,
  23250. year = 2005,
  23251. language = "en"
  23252. }
  23253. @ARTICLE{Powell2014-iy,
  23254. title = "Complex neural codes in rat prelimbic cortex are stable across
  23255. days on a spatial decision task",
  23256. author = "Powell, Nathaniel J and Redish, A David",
  23257. abstract = "The rodent prelimbic cortex has been shown to play an important
  23258. role in cognitive processing, and has been implicated in encoding
  23259. many different parameters relevant to solving decision-making
  23260. tasks. However, it is not known how the prelimbic cortex
  23261. represents all these disparate variables, and if they are
  23262. simultaneously represented when the task requires it. In order to
  23263. investigate this question, we trained rats to run the Multiple-T
  23264. Left Right Alternate (MT-LRA) task and recorded multi-unit
  23265. ensembles from their prelimbic regions. Significant populations
  23266. of cells in the prelimbic cortex represented the strategy
  23267. controlling reward receipt on a given lap, whether the animal
  23268. chose to go right or left on a given lap, and whether the animal
  23269. made a correct decision or an error on a given lap. These
  23270. populations overlapped in the cells recorded, with several cells
  23271. demonstrating differential firing to all three variables. The
  23272. spatial and strategic firing patterns of individual prelimbic
  23273. cells were highly conserved across several days of running this
  23274. task, indicating that each cell encoded the same information
  23275. across days.",
  23276. journal = "Front. Behav. Neurosci.",
  23277. volume = 8,
  23278. pages = "120",
  23279. month = apr,
  23280. year = 2014,
  23281. keywords = "animal; behavior; decision-making; neural ensemble data;
  23282. prefrontal cortex (PFC); prelimbic cortex; rats; tetrode
  23283. recording",
  23284. language = "en"
  23285. }
  23286. @ARTICLE{Amemiya2016-fb,
  23287. title = "Manipulating Decisiveness in Decision Making: Effects of
  23288. Clonidine on Hippocampal Search Strategies",
  23289. author = "Amemiya, Seiichiro and Redish, A David",
  23290. abstract = "Decisiveness is the ability to commit to a decision quickly and
  23291. efficiently; in contrast, indecision entails the repeated
  23292. consideration of multiple alternative possibilities. In humans,
  23293. the $\alpha$2-adrenergic receptor agonist clonidine increases
  23294. decisiveness in tasks that require planning through unknown
  23295. neural mechanisms. In rats, indecision is manifested as
  23296. reorienting behaviors at choice points (vicarious trial and error
  23297. [VTE]), during which hippocampal representations alternate
  23298. between prospective options. To determine whether the increase in
  23299. decisiveness driven by clonidine also entails changes in
  23300. hippocampal search processes, we compared the effect of clonidine
  23301. on spatial representations in hippocampal neural ensembles as
  23302. rats passed through a T-shaped decision point. Consistent with
  23303. previous experiments, hippocampal representations reflected both
  23304. chosen and unchosen paths during VTE events under saline control
  23305. conditions. Also, consistent with previous experiments,
  23306. hippocampal representations reflected the chosen path more than
  23307. the unchosen path when the animal did not show VTE at the choice
  23308. point. Injection of clonidine suppressed the spatial
  23309. representation of the unchosen path at the choice point on VTE
  23310. laps and hastened the differentiation of spatial representations
  23311. of the chosen path from the unchosen path on non-VTE laps to
  23312. appear before reaching the choice point. These results suggest
  23313. that the decisiveness seen under clonidine is due to limited
  23314. exploration of potential options in hippocampus, and suggest
  23315. novel roles for noradrenaline as a modulator of the hippocampal
  23316. search processes. Significance statement: Clonidine, an
  23317. $\alpha$2-adrenergic receptor agonist, which decreases the level
  23318. of noradrenaline in vivo, has an interesting effect in humans and
  23319. other animals: it makes them more decisive. However, the
  23320. mechanisms by which clonidine makes them more decisive remain
  23321. unknown. Researchers have speculated that clonidine limits the
  23322. amount of mental search that subjects do when planning options.
  23323. We test this hypothesis by measuring the mental search strategy
  23324. in rats through hippocampal recordings. We find that clonidine
  23325. limits the options searched by rats, suggesting that
  23326. noradrenaline also plays a role in balancing exploration and
  23327. exploitation in internally simulated behaviors, similar to its
  23328. role in balancing exploration and exploitation in external
  23329. behaviors.",
  23330. journal = "J. Neurosci.",
  23331. volume = 36,
  23332. number = 3,
  23333. pages = "814--827",
  23334. month = jan,
  23335. year = 2016,
  23336. keywords = "VTE; hippocampus; noradrenaline; norepinphrine; place field;
  23337. vicarious trial and error",
  23338. language = "en"
  23339. }
  23340. @ARTICLE{Schmitzer-Torbert2004-tj,
  23341. title = "Neuronal activity in the rodent dorsal striatum in sequential
  23342. navigation: separation of spatial and reward responses on the
  23343. multiple {T} task",
  23344. author = "Schmitzer-Torbert, Neil and Redish, A David",
  23345. abstract = "The striatum plays an important role in ``habitual'' learning and
  23346. memory and has been hypothesized to implement a
  23347. reinforcement-learning algorithm to select actions to perform
  23348. given the current sensory input. Many experimental approaches to
  23349. striatal activity have made use of temporally structured tasks,
  23350. which imply that the striatal representation is temporal. To test
  23351. this assumption, we recorded neurons in the dorsal striatum of
  23352. rats running a sequential navigation task: the multiple T maze.
  23353. Rats navigated a sequence of four T maze turns to receive food
  23354. rewards delivered in two locations. The responses of neurons that
  23355. fired phasically were examined. Task-responsive phasic neurons
  23356. were active as rats ran on the maze (maze-responsive) or during
  23357. reward receipt (reward-responsive). Neither maze- nor
  23358. reward-responsive neurons encoded simple motor commands:
  23359. maze-responses were not well correlated with the shape of the
  23360. rat's path and most reward-responsive neurons did not fire at
  23361. similar rates at both food-delivery sites. Maze-responsive
  23362. neurons were active at one or more locations on the maze, but
  23363. these responses did not cluster at spatial landmarks such as
  23364. turns. Across sessions the activity of maze-responsive neurons
  23365. was highly correlated when rats ran the same maze. Maze-responses
  23366. encoded the location of the rat on the maze and imply a spatial
  23367. representation in the striatum in a task with prominent spatial
  23368. demands. Maze-responsive and reward-responsive neurons were two
  23369. separate populations, suggesting a divergence in striatal
  23370. information processing of navigation and reward.",
  23371. journal = "J. Neurophysiol.",
  23372. volume = 91,
  23373. number = 5,
  23374. pages = "2259--2272",
  23375. month = may,
  23376. year = 2004,
  23377. language = "en"
  23378. }
  23379. @ARTICLE{Johnson2007-br,
  23380. title = "Neural ensembles in {CA3} transiently encode paths forward of the
  23381. animal at a decision point",
  23382. author = "Johnson, Adam and Redish, A David",
  23383. abstract = "Neural ensembles were recorded from the CA3 region of rats
  23384. running on T-based decision tasks. Examination of neural
  23385. representations of space at fast time scales revealed a transient
  23386. but repeatable phenomenon as rats made a decision: the location
  23387. reconstructed from the neural ensemble swept forward, first down
  23388. one path and then the other. Estimated representations were
  23389. coherent and preferentially swept ahead of the animal rather than
  23390. behind the animal, implying it represented future possibilities
  23391. rather than recently traveled paths. Similar phenomena occurred
  23392. at other important decisions (such as in recovery from an error).
  23393. Local field potentials from these sites contained pronounced
  23394. theta and gamma frequencies, but no sharp wave frequencies.
  23395. Forward-shifted spatial representations were influenced by task
  23396. demands and experience. These data suggest that the hippocampus
  23397. does not represent space as a passive computation, but rather
  23398. that hippocampal spatial processing is an active process likely
  23399. regulated by cognitive mechanisms.",
  23400. journal = "J. Neurosci.",
  23401. volume = 27,
  23402. number = 45,
  23403. pages = "12176--12189",
  23404. month = nov,
  23405. year = 2007,
  23406. language = "en"
  23407. }
  23408. @ARTICLE{Steiner2012-xp,
  23409. title = "The road not taken: neural correlates of decision making in
  23410. orbitofrontal cortex",
  23411. author = "Steiner, Adam P and Redish, A David",
  23412. abstract = "Empirical research links human orbitofrontal cortex (OFC) to the
  23413. evaluation of outcomes during decision making and the
  23414. representation of alternative (better) outcomes after failures.
  23415. When faced with a difficult decision, rats sometimes pause and
  23416. turn back-and-forth toward goals, until finally orienting toward
  23417. the chosen direction. Neural representations of reward in rodent
  23418. OFC increased immediately following each reorientation, implying
  23419. a transient representation of the expected outcome following
  23420. self-initiated decisions. Upon reaching reward locations and
  23421. finding no reward (having made an error), OFC representations of
  23422. reward decreased locally indicating a disappointment signal that
  23423. then switched to represent the unrewarded, non-local,
  23424. would-have-been rewarded site. These results illustrate that
  23425. following a decision to act, neural ensembles in OFC represent
  23426. reward, and upon the realization of an error, represent the
  23427. reward that could have been.",
  23428. journal = "Front. Neurosci.",
  23429. volume = 6,
  23430. pages = "131",
  23431. month = sep,
  23432. year = 2012,
  23433. keywords = "counterfactual; covert representation of reward; multiple T;
  23434. orbitofrontal cortex; regret; vicarious trial and error",
  23435. language = "en"
  23436. }
  23437. @ARTICLE{Tolman_undated-yh,
  23438. title = "Cognitive Critique",
  23439. author = "Tolman, Revisiting and Ries, His Theo"
  23440. }
  23441. @ARTICLE{Redish2016-wf,
  23442. title = "The Computational Complexity of Valuation and Motivational Forces
  23443. in {Decision-Making} Processes",
  23444. author = "Redish, A David and Schultheiss, Nathan W and Carter, Evan C",
  23445. abstract = "The concept of value is fundamental to most theories of
  23446. motivation and decision making. However, value has to be measured
  23447. experimentally. Different methods of measuring value produce
  23448. incompatible valuation hierarchies. Taking the agent's
  23449. perspective (rather than the experimenter's), we interpret the
  23450. different valuation measurement methods as accessing different
  23451. decision-making systems and show how these different systems
  23452. depend on different information processing algorithms. This
  23453. identifies the translation from these multiple decision-making
  23454. systems into a single action taken by a given agent as one of the
  23455. most important open questions in decision making today. We
  23456. conclude by looking at how these different valuation measures
  23457. accessing different decision-making systems can be used to
  23458. understand and treat decision dysfunction such as in addiction.",
  23459. journal = "Curr. Top. Behav. Neurosci.",
  23460. volume = 27,
  23461. pages = "313--333",
  23462. year = 2016,
  23463. keywords = "Decision-Making; Multiple Decision Theory; Neuroeconomonics;
  23464. Valuation",
  23465. language = "en"
  23466. }
  23467. @ARTICLE{Phelps2014-qu,
  23468. title = "Emotion and decision making: multiple modulatory neural circuits",
  23469. author = "Phelps, Elizabeth A and Lempert, Karolina M and Sokol-Hessner,
  23470. Peter",
  23471. abstract = "Although the prevalent view of emotion and decision making is
  23472. derived from the notion that there are dual systems of emotion
  23473. and reason, a modulatory relationship more accurately reflects
  23474. the current research in affective neuroscience and
  23475. neuroeconomics. Studies show two potential mechanisms for
  23476. affect's modulation of the computation of subjective value and
  23477. decisions. Incidental affective states may carry over to the
  23478. assessment of subjective value and the decision, and emotional
  23479. reactions to the choice may be incorporated into the value
  23480. calculation. In addition, this modulatory relationship is
  23481. reciprocal: Changing emotion can change choices. This research
  23482. suggests that the neural mechanisms mediating the relation
  23483. between affect and choice vary depending on which affective
  23484. component is engaged and which decision variables are assessed.
  23485. We suggest that a detailed and nuanced understanding of emotion
  23486. and decision making requires characterizing the multiple
  23487. modulatory neural circuits underlying the different means by
  23488. which emotion and affect can influence choices.",
  23489. journal = "Annu. Rev. Neurosci.",
  23490. volume = 37,
  23491. pages = "263--287",
  23492. month = may,
  23493. year = 2014,
  23494. keywords = "amygdala; insular cortex; mood; orbitofrontal cortex; stress;
  23495. striatum",
  23496. language = "en"
  23497. }
  23498. @ARTICLE{Geva-Sagiv2015-cl,
  23499. title = "Spatial cognition in bats and rats: from sensory acquisition to
  23500. multiscale maps and navigation",
  23501. author = "Geva-Sagiv, Maya and Las, Liora and Yovel, Yossi and Ulanovsky,
  23502. Nachum",
  23503. abstract = "Spatial orientation and navigation rely on the acquisition of
  23504. several types of sensory information. This information is then
  23505. transformed into a neural code for space in the hippocampal
  23506. formation through the activity of place cells, grid cells and
  23507. head-direction cells. These spatial representations, in turn, are
  23508. thought to guide long-range navigation. But how the
  23509. representations encoded by these different cell types are
  23510. integrated in the brain to form a neural 'map and compass' is
  23511. largely unknown. Here, we discuss this problem in the context of
  23512. spatial navigation by bats and rats. We review the experimental
  23513. findings and theoretical models that provide insight into the
  23514. mechanisms that link sensory systems to spatial representations
  23515. and to large-scale natural navigation.",
  23516. journal = "Nat. Rev. Neurosci.",
  23517. volume = 16,
  23518. number = 2,
  23519. pages = "94--108",
  23520. month = feb,
  23521. year = 2015,
  23522. language = "en"
  23523. }
  23524. @ARTICLE{Johnson2009-bx,
  23525. title = "Looking for cognition in the structure within the noise",
  23526. author = "Johnson, Adam and Fenton, Andr{\'e} A and Kentros, Cliff and
  23527. Redish, A David",
  23528. abstract = "Neural activity in the mammalian CNS is determined by both
  23529. observable processes, such as sensory stimuli or motor output,
  23530. and covert, internal cognitive processes that cannot be directly
  23531. observed. We propose methods to identify these cognitive
  23532. processes by examining the covert structure within the apparent
  23533. 'noise' in spike trains. Contemporary analyses of neural codes
  23534. include encoding (tuning curves derived from spike trains and
  23535. behavioral, sensory or motor variables), decoding (reconstructing
  23536. behavioral, sensory or motor variables from spike trains and
  23537. hypothesized tuning curves) and generative models (predicting the
  23538. spike trains from hypothesized encoding models and decoded
  23539. variables). We review examples of each of these processes in
  23540. hippocampal activity, and propose a general methodology to
  23541. examine cognitive processes via the identification of dynamic
  23542. changes in covert variables.",
  23543. journal = "Trends Cogn. Sci.",
  23544. volume = 13,
  23545. number = 2,
  23546. pages = "55--64",
  23547. month = feb,
  23548. year = 2009,
  23549. language = "en"
  23550. }
  23551. @ARTICLE{Mysore2011-fr,
  23552. title = "The role of a midbrain network in competitive stimulus selection",
  23553. author = "Mysore, Shreesh P and Knudsen, Eric I",
  23554. abstract = "A midbrain network interacts with the well-known frontoparietal
  23555. forebrain network to select stimuli for gaze and spatial
  23556. attention. The midbrain network, containing the superior
  23557. colliculus (SC; optic tectum, OT, in non-mammalian vertebrates)
  23558. and the isthmic nuclei, helps evaluate the relative priorities of
  23559. competing stimuli and encodes them in a topographic map of space.
  23560. Behavioral experiments in monkeys demonstrate an essential
  23561. contribution of the SC to stimulus selection when the relative
  23562. priorities of competing stimuli are similar. Neurophysiological
  23563. results from the owl OT demonstrate a neural correlate of this
  23564. essential contribution of the SC/OT. The multi-layered,
  23565. spatiotopic organization of the midbrain network lends itself to
  23566. the analysis and modeling of the mechanisms underlying stimulus
  23567. selection for gaze and spatial attention.",
  23568. journal = "Curr. Opin. Neurobiol.",
  23569. volume = 21,
  23570. number = 4,
  23571. pages = "653--660",
  23572. month = aug,
  23573. year = 2011,
  23574. language = "en"
  23575. }
  23576. @ARTICLE{Cisek2010-ke,
  23577. title = "Neural mechanisms for interacting with a world full of action
  23578. choices",
  23579. author = "Cisek, Paul and Kalaska, John F",
  23580. abstract = "The neural bases of behavior are often discussed in terms of
  23581. perceptual, cognitive, and motor stages, defined within an
  23582. information processing framework that was originally inspired by
  23583. models of human abstract problem solving. Here, we review a
  23584. growing body of neurophysiological data that is difficult to
  23585. reconcile with this influential theoretical perspective. As an
  23586. alternative foundation for interpreting neural data, we consider
  23587. frameworks borrowed from ethology, which emphasize the kinds of
  23588. real-time interactive behaviors that animals have engaged in for
  23589. millions of years. In particular, we discuss an
  23590. ethologically-inspired view of interactive behavior as
  23591. simultaneous processes that specify potential motor actions and
  23592. select between them. We review how recent neurophysiological data
  23593. from diverse cortical and subcortical regions appear more
  23594. compatible with this parallel view than with the classical view
  23595. of serial information processing stages.",
  23596. journal = "Annu. Rev. Neurosci.",
  23597. volume = 33,
  23598. pages = "269--298",
  23599. year = 2010,
  23600. language = "en"
  23601. }
  23602. @ARTICLE{Koene2003-ep,
  23603. title = "Modeling goal-directed spatial navigation in the rat based on
  23604. physiological data from the hippocampal formation",
  23605. author = "Koene, Randal A and Gorchetchnikov, Anatoli and Cannon, Robert C
  23606. and Hasselmo, Michael E",
  23607. abstract = "We investigated the importance of hippocampal theta oscillations
  23608. and the significance of phase differences of theta modulation in
  23609. the cortical regions that are involved in goal-directed spatial
  23610. navigation. Our models used representations of entorhinal cortex
  23611. layer III (ECIII), hippocampus and prefrontal cortex (PFC) to
  23612. guide movements of a virtual rat in a virtual environment. The
  23613. model encoded representations of the environment through
  23614. long-term potentiation of excitatory recurrent connections
  23615. between sequentially spiking place cells in ECIII and CA3. This
  23616. encoding required buffering of place cell activity, which was
  23617. achieved by a short-term memory (STM) in EC that was regulated by
  23618. theta modulation and allowed synchronized reactivation with
  23619. encoding phases in ECIII and CA3. Inhibition at a specific theta
  23620. phase deactivated the oldest item in the buffer when new input
  23621. was presented to a full STM buffer. A 180 degrees phase
  23622. difference separated retrieval and encoding in ECIII and CA3,
  23623. which enabled us to simulate data on theta phase precession of
  23624. place cells. Retrieval of known paths was elicited in ECIII by
  23625. input at the retrieval phase from PFC working memory for goal
  23626. location, requiring strict theta phase relationships with PFC.
  23627. Known locations adjacent to the virtual rat were retrieved in
  23628. CA3. Together, input from ECIII and CA3 activated predictive
  23629. spiking in cells in CA1 for the next desired place on a shortest
  23630. path to a goal. Consistent with data, place cell activity in CA1
  23631. and CA3 showed smaller place fields than in ECIII.",
  23632. journal = "Neural Netw.",
  23633. volume = 16,
  23634. number = "5-6",
  23635. pages = "577--584",
  23636. month = jun,
  23637. year = 2003,
  23638. language = "en"
  23639. }
  23640. @ARTICLE{Van_Strien2009-yf,
  23641. title = "The anatomy of memory: an interactive overview of the
  23642. parahippocampal-hippocampal network",
  23643. author = "van Strien, N M and Cappaert, N L M and Witter, M P",
  23644. abstract = "Converging evidence suggests that each parahippocampal and
  23645. hippocampal subregion contributes uniquely to the encoding,
  23646. consolidation and retrieval of declarative memories, but their
  23647. precise roles remain elusive. Current functional thinking does
  23648. not fully incorporate the intricately connected networks that
  23649. link these subregions, owing to their organizational complexity;
  23650. however, such detailed anatomical knowledge is of pivotal
  23651. importance for comprehending the unique functional contribution
  23652. of each subregion. We have therefore developed an interactive
  23653. diagram with the aim to display all of the currently known
  23654. anatomical connections of the rat parahippocampal-hippocampal
  23655. network. In this Review, we integrate the existing anatomical
  23656. knowledge into a concise description of this network and discuss
  23657. the functional implications of some relatively underexposed
  23658. connections.",
  23659. journal = "Nat. Rev. Neurosci.",
  23660. volume = 10,
  23661. number = 4,
  23662. pages = "272--282",
  23663. month = apr,
  23664. year = 2009,
  23665. language = "en"
  23666. }
  23667. @ARTICLE{Oh2014-ws,
  23668. title = "A mesoscale connectome of the mouse brain",
  23669. author = "Oh, Seung Wook and Harris, Julie A and Ng, Lydia and Winslow,
  23670. Brent and Cain, Nicholas and Mihalas, Stefan and Wang, Quanxin
  23671. and Lau, Chris and Kuan, Leonard and Henry, Alex M and Mortrud,
  23672. Marty T and Ouellette, Benjamin and Nguyen, Thuc Nghi and
  23673. Sorensen, Staci A and Slaughterbeck, Clifford R and Wakeman,
  23674. Wayne and Li, Yang and Feng, David and Ho, Anh and Nicholas, Eric
  23675. and Hirokawa, Karla E and Bohn, Phillip and Joines, Kevin M and
  23676. Peng, Hanchuan and Hawrylycz, Michael J and Phillips, John W and
  23677. Hohmann, John G and Wohnoutka, Paul and Gerfen, Charles R and
  23678. Koch, Christof and Bernard, Amy and Dang, Chinh and Jones, Allan
  23679. R and Zeng, Hongkui",
  23680. abstract = "Comprehensive knowledge of the brain's wiring diagram is
  23681. fundamental for understanding how the nervous system processes
  23682. information at both local and global scales. However, with the
  23683. singular exception of the C. elegans microscale connectome, there
  23684. are no complete connectivity data sets in other species. Here we
  23685. report a brain-wide, cellular-level, mesoscale connectome for the
  23686. mouse. The Allen Mouse Brain Connectivity Atlas uses enhanced
  23687. green fluorescent protein (EGFP)-expressing adeno-associated
  23688. viral vectors to trace axonal projections from defined regions
  23689. and cell types, and high-throughput serial two-photon tomography
  23690. to image the EGFP-labelled axons throughout the brain. This
  23691. systematic and standardized approach allows spatial registration
  23692. of individual experiments into a common three dimensional (3D)
  23693. reference space, resulting in a whole-brain connectivity matrix.
  23694. A computational model yields insights into connectional strength
  23695. distribution, symmetry and other network properties. Virtual
  23696. tractography illustrates 3D topography among interconnected
  23697. regions. Cortico-thalamic pathway analysis demonstrates
  23698. segregation and integration of parallel pathways. The Allen Mouse
  23699. Brain Connectivity Atlas is a freely available, foundational
  23700. resource for structural and functional investigations into the
  23701. neural circuits that support behavioural and cognitive processes
  23702. in health and disease.",
  23703. journal = "Nature",
  23704. volume = 508,
  23705. number = 7495,
  23706. pages = "207--214",
  23707. month = apr,
  23708. year = 2014,
  23709. language = "en"
  23710. }
  23711. @ARTICLE{Sofroniew2014-mh,
  23712. title = "Natural whisker-guided behavior by head-fixed mice in tactile
  23713. virtual reality",
  23714. author = "Sofroniew, Nicholas J and Cohen, Jeremy D and Lee, Albert K and
  23715. Svoboda, Karel",
  23716. abstract = "During many natural behaviors the relevant sensory stimuli and
  23717. motor outputs are difficult to quantify. Furthermore, the high
  23718. dimensionality of the space of possible stimuli and movements
  23719. compounds the problem of experimental control. Head fixation
  23720. facilitates stimulus control and movement tracking, and can be
  23721. combined with techniques for recording and manipulating neural
  23722. activity. However, head-fixed mouse behaviors are typically
  23723. trained through extensive instrumental conditioning. Here we
  23724. present a whisker-based, tactile virtual reality system for
  23725. head-fixed mice running on a spherical treadmill. Head-fixed mice
  23726. displayed natural movements, including running and rhythmic
  23727. whisking at 16 Hz. Whisking was centered on a set point that
  23728. changed in concert with running so that more protracted whisking
  23729. was correlated with faster running. During turning, whiskers
  23730. moved in an asymmetric manner, with more retracted whisker
  23731. positions in the turn direction and protracted whisker movements
  23732. on the other side. Under some conditions, whisker movements were
  23733. phase-coupled to strides. We simulated a virtual reality tactile
  23734. corridor, consisting of two moveable walls controlled in a
  23735. closed-loop by running speed and direction. Mice used their
  23736. whiskers to track the walls of the winding corridor without
  23737. training. Whisker curvature changes, which cause forces in the
  23738. sensory follicles at the base of the whiskers, were tightly
  23739. coupled to distance from the walls. Our behavioral system allows
  23740. for precise control of sensorimotor variables during natural
  23741. tactile navigation.",
  23742. journal = "J. Neurosci.",
  23743. volume = 34,
  23744. number = 29,
  23745. pages = "9537--9550",
  23746. month = jul,
  23747. year = 2014,
  23748. language = "en"
  23749. }
  23750. @ARTICLE{Churchland2012-kw,
  23751. title = "New advances in understanding decisions among multiple
  23752. alternatives",
  23753. author = "Churchland, Anne K and Ditterich, Jochen",
  23754. abstract = "Experimental studies of decision-making have put a strong
  23755. emphasis on choices between two alternatives. However, real-life
  23756. decisions often involve multiple alternatives. This article
  23757. provides an overview of theoretical frameworks that have been
  23758. proposed to account for behavioral data from both economic and
  23759. perceptual multialternative decision-making. We further review
  23760. recent neurophysiological data collected in conjunction with
  23761. decision-making behavior. These neural recordings provide
  23762. constraints on putative models of the decision mechanism. For
  23763. example, the time course of inhibition provides insight into how
  23764. the competition between alternatives is mediated. Furthermore,
  23765. whereas decision-related neural activity seems to reach a common
  23766. threshold at the end of the decision period, the starting point
  23767. tends to depend systematically on the number of alternatives. We
  23768. discuss candidate mechanisms that could drive the reduction in
  23769. firing rates on decisions among multiple alternatives.",
  23770. journal = "Curr. Opin. Neurobiol.",
  23771. volume = 22,
  23772. number = 6,
  23773. pages = "920--926",
  23774. month = dec,
  23775. year = 2012,
  23776. language = "en"
  23777. }
  23778. @ARTICLE{Cain2012-tf,
  23779. title = "Computational models of decision making: integration, stability,
  23780. and noise",
  23781. author = "Cain, Nicholas and Shea-Brown, Eric",
  23782. abstract = "Decision making demands the accumulation of sensory evidence over
  23783. time. Questions remain about how this occurs, but recent years
  23784. have seen progress on several fronts. The first concerns when
  23785. optimal accumulation of evidence coincides with the simplest
  23786. method of accumulating neural activity: summation over time. The
  23787. second involves what computations the brain might perform when
  23788. summation is difficult due to imprecision in neural circuits or
  23789. is suboptimal due to uncertainty or variability in how evidence
  23790. arrives. Finally, the third concerns sources of noise in decision
  23791. circuits. Empirical studies have better constrained the extent of
  23792. this noise, and modeling work is helping to clarify its possible
  23793. origins.",
  23794. journal = "Curr. Opin. Neurobiol.",
  23795. volume = 22,
  23796. number = 6,
  23797. pages = "1047--1053",
  23798. month = dec,
  23799. year = 2012,
  23800. language = "en"
  23801. }
  23802. @ARTICLE{Wang2012-ac,
  23803. title = "Neural dynamics and circuit mechanisms of decision-making",
  23804. author = "Wang, Xiao-Jing",
  23805. abstract = "In this review, I briefly summarize current neurobiological
  23806. studies of decision-making that bear on two general themes. The
  23807. first focuses on the nature of neural representation and dynamics
  23808. in a decision circuit. Experimental and computational results
  23809. suggest that ramping-to-threshold in the temporal domain and
  23810. trajectory of population activity in the state space represent a
  23811. duality of perspectives on a decision process. Moreover, a
  23812. decision circuit can display several different dynamical regimes,
  23813. such as the ramping mode and the jumping mode with distinct
  23814. defining properties. The second is concerned with the
  23815. relationship between biologically-based mechanistic models and
  23816. normative-type models. A fruitful interplay between experiments
  23817. and these models at different levels of abstraction have enabled
  23818. investigators to pose increasingly refined questions and gain new
  23819. insights into the neural basis of decision-making. In particular,
  23820. recent work on multi-alternative decisions suggests that
  23821. deviations from rational models of choice behavior can be
  23822. explained by established neural mechanisms.",
  23823. journal = "Curr. Opin. Neurobiol.",
  23824. volume = 22,
  23825. number = 6,
  23826. pages = "1039--1046",
  23827. month = dec,
  23828. year = 2012
  23829. }
  23830. @ARTICLE{Adams2012-mw,
  23831. title = "Neuroethology of decision-making",
  23832. author = "Adams, Geoffrey K and Watson, Karli K and Pearson, John and
  23833. Platt, Michael L",
  23834. abstract = "A neuroethological approach to decision-making considers the
  23835. effect of evolutionary pressures on neural circuits mediating
  23836. choice. In this view, decision systems are expected to enhance
  23837. fitness with respect to the local environment, and particularly
  23838. efficient solutions to specific problems should be conserved,
  23839. expanded, and repurposed to solve other problems. Here, we
  23840. discuss basic prerequisites for a variety of decision systems
  23841. from this viewpoint. We focus on two of the best-studied and most
  23842. widely represented decision problems. First, we examine patch
  23843. leaving, a prototype of environmentally based switching between
  23844. action patterns. Second, we consider social information seeking,
  23845. a process resembling foraging with search costs. We argue that
  23846. while the specific neural solutions to these problems sometimes
  23847. differ across species, both the problems themselves and the
  23848. algorithms instantiated by biological hardware are repeated
  23849. widely throughout nature. The behavioral and mathematical study
  23850. of ubiquitous decision processes like patch leaving and social
  23851. information seeking thus provides a powerful new approach to
  23852. uncovering the fundamental design structure of nervous systems.",
  23853. journal = "Curr. Opin. Neurobiol.",
  23854. volume = 22,
  23855. number = 6,
  23856. pages = "982--989",
  23857. month = dec,
  23858. year = 2012,
  23859. language = "en"
  23860. }
  23861. @ARTICLE{Drugowitsch2012-rd,
  23862. title = "Probabilistic vs. non-probabilistic approaches to the
  23863. neurobiology of perceptual decision-making",
  23864. author = "Drugowitsch, Jan and Pouget, Alexandre",
  23865. abstract = "Optimal binary perceptual decision making requires accumulation
  23866. of evidence in the form of a probability distribution that
  23867. specifies the probability of the choices being correct given the
  23868. evidence so far. Reward rates can then be maximized by stopping
  23869. the accumulation when the confidence about either option reaches
  23870. a threshold. Behavioral and neuronal evidence suggests that
  23871. humans and animals follow such a probabilitistic decision
  23872. strategy, although its neural implementation has yet to be fully
  23873. characterized. Here we show that that diffusion decision models
  23874. and attractor network models provide an approximation to the
  23875. optimal strategy only under certain circumstances. In particular,
  23876. neither model type is sufficiently flexible to encode the
  23877. reliability of both the momentary and the accumulated evidence,
  23878. which is a pre-requisite to accumulate evidence of time-varying
  23879. reliability. Probabilistic population codes, by contrast, can
  23880. encode these quantities and, as a consequence, have the potential
  23881. to implement the optimal strategy accurately.",
  23882. journal = "Curr. Opin. Neurobiol.",
  23883. volume = 22,
  23884. number = 6,
  23885. pages = "963--969",
  23886. month = dec,
  23887. year = 2012,
  23888. language = "en"
  23889. }
  23890. @ARTICLE{Dehaene2012-xq,
  23891. title = "From a single decision to a multi-step algorithm",
  23892. author = "Dehaene, Stanislas and Sigman, Mariano",
  23893. abstract = "Humans can perform sequential and recursive computations, as when
  23894. calculating 23$\times$74. However, this comes at a cost: flexible
  23895. computations are slow and effortful. We argue that this
  23896. competence involves serial chains of successive decisions, each
  23897. based on the accumulation of evidence up to a threshold and
  23898. forwarding the result to the subsequent step. Such serial
  23899. 'programs' require a specific neurobiological architecture,
  23900. approximating the operation of a slow serial Turing machine. We
  23901. review recent progress in understanding how the brain implements
  23902. such multi-step decisions and briefly examine how they might be
  23903. realized in models of primate cortex.",
  23904. journal = "Curr. Opin. Neurobiol.",
  23905. volume = 22,
  23906. number = 6,
  23907. pages = "937--945",
  23908. month = dec,
  23909. year = 2012,
  23910. language = "en"
  23911. }
  23912. @ARTICLE{Cisek2012-qy,
  23913. title = "Making decisions through a distributed consensus",
  23914. author = "Cisek, Paul",
  23915. abstract = "How does the brain decide between actions? Is it through
  23916. comparisons of abstract representations of outcomes or through a
  23917. competition in a sensorimotor map defining the actions
  23918. themselves? Here, I review strengths and limitations of both of
  23919. these proposals, and suggest that decisions emerge through a
  23920. distributed consensus across many levels of representation.",
  23921. journal = "Curr. Opin. Neurobiol.",
  23922. volume = 22,
  23923. number = 6,
  23924. pages = "927--936",
  23925. month = dec,
  23926. year = 2012,
  23927. language = "en"
  23928. }
  23929. @ARTICLE{Chapman2010-dv,
  23930. title = "Reaching for the unknown: multiple target encoding and real-time
  23931. decision-making in a rapid reach task",
  23932. author = "Chapman, Craig S and Gallivan, Jason P and Wood, Daniel K and
  23933. Milne, Jennifer L and Culham, Jody C and Goodale, Melvyn A",
  23934. abstract = "Decision-making is central to human cognition. Fundamental to
  23935. every decision is the ability to internally represent the
  23936. available choices and their relative costs and benefits. The most
  23937. basic and frequent decisions we make occur as our motor system
  23938. chooses and executes only those actions that achieve our current
  23939. goals. Although these interactions with the environment may
  23940. appear effortless, this belies what must be incredibly
  23941. sophisticated visuomotor decision-making processes. In order to
  23942. measure how visuomotor decisions unfold in real-time, we used a
  23943. unique reaching paradigm that forced participants to initiate
  23944. rapid hand movements toward multiple potential targets, with only
  23945. one being cued after reach onset. We show across three
  23946. experiments that, in cases of target uncertainty, trajectories
  23947. are spatially sensitive to the probabilistic distribution of
  23948. targets within the display. Specifically, when presented with two
  23949. or three target displays, subjects initiate their reaches toward
  23950. an intermediary or 'averaged' location before correcting their
  23951. trajectory in-flight to the cued target location. A control
  23952. experiment suggests that our effect depends on the targets acting
  23953. as potential reach locations and not as distractors. This study
  23954. is the first to show that the 'averaging' of target-directed
  23955. reaching movements depends not only on the spatial position of
  23956. the targets in the display but also the probability of acting at
  23957. each target location.",
  23958. journal = "Cognition",
  23959. volume = 116,
  23960. number = 2,
  23961. pages = "168--176",
  23962. month = aug,
  23963. year = 2010,
  23964. language = "en"
  23965. }
  23966. @ARTICLE{Thura2014-zb,
  23967. title = "Deliberation and commitment in the premotor and primary motor
  23968. cortex during dynamic decision making",
  23969. author = "Thura, David and Cisek, Paul",
  23970. abstract = "Neurophysiological studies of decision making have primarily
  23971. focused on decisions about information that is stable over time.
  23972. However, during natural behavior, animals make decisions in a
  23973. constantly changing environment. To investigate the neural
  23974. mechanisms of such dynamic choices, we recorded activity in
  23975. dorsal premotor (PMd) and primary motor cortex (M1) while monkeys
  23976. performed a two-choice reaching task in which sensory information
  23977. about the correct choice was changing within each trial and the
  23978. decision could be made at any time. During deliberation, activity
  23979. in both areas did not integrate sensory information but instead
  23980. tracked it and combined it with a growing urgency signal.
  23981. Approximately 280 ms before movement onset, PMd activity tuned to
  23982. the selected target reached a consistent peak while M1 activity
  23983. tuned to the unselected target was suppressed. We propose that
  23984. this reflects the resolution of a competition between the
  23985. potential responses and constitutes the volitional commitment to
  23986. an action choice.",
  23987. journal = "Neuron",
  23988. volume = 81,
  23989. number = 6,
  23990. pages = "1401--1416",
  23991. month = mar,
  23992. year = 2014,
  23993. language = "en"
  23994. }
  23995. @ARTICLE{Klaes2011-hi,
  23996. title = "Choosing goals, not rules: deciding among rule-based action plans",
  23997. author = "Klaes, Christian and Westendorff, Stephanie and Chakrabarti,
  23998. Shubhodeep and Gail, Alexander",
  23999. abstract = "In natural situations, movements are often directed toward
  24000. locations different from that of the evoking sensory stimulus.
  24001. Movement goals must then be inferred from the sensory cue based
  24002. on rules. When there is uncertainty about the rule that applies
  24003. for a given cue, planning a movement involves both choosing the
  24004. relevant rule and computing the movement goal based on that rule.
  24005. Under these conditions, it is not clear whether primates compute
  24006. multiple movement goals based on all possible rules before
  24007. choosing an action, or whether they first choose a rule and then
  24008. only represent the movement goal associated with that rule.
  24009. Supporting the former hypothesis, we show that neurons in the
  24010. frontoparietal reach areas of monkeys simultaneously represent
  24011. two different rule-based movement goals, which are biased by the
  24012. monkeys' choice preferences. Apparently, primates choose between
  24013. multiple behavioral options by weighing against each other the
  24014. movement goals associated with each option.",
  24015. journal = "Neuron",
  24016. volume = 70,
  24017. number = 3,
  24018. pages = "536--548",
  24019. month = may,
  24020. year = 2011,
  24021. language = "en"
  24022. }
  24023. @ARTICLE{Busemeyer1993-qq,
  24024. title = "Decision field theory: a dynamic-cognitive approach to decision
  24025. making in an uncertain environment",
  24026. author = "Busemeyer, J R and Townsend, J T",
  24027. abstract = "Decision field theory provides for a mathematical foundation
  24028. leading to a dynamic, stochastic theory of decision behavior in
  24029. an uncertain environment. This theory is used to explain (a)
  24030. violations of stochastic dominance, (b) violations of strong
  24031. stochastic transitivity, (c) violations of independence between
  24032. alternatives, (d) serial position effects on preference, (e)
  24033. speed-accuracy trade-off effects in decision making, (f) the
  24034. inverse relation between choice probability and decision time,
  24035. (g) changes in the direction of preference under time pressure,
  24036. (h) slower decision times for avoidance as compared with
  24037. approach conflicts, and (i) preference reversals between choice
  24038. and selling price measures of preference. The proposed theory is
  24039. compared with 4 other theories of decision making under
  24040. uncertainty.",
  24041. journal = "Psychol. Rev.",
  24042. publisher = "doi.apa.org",
  24043. volume = 100,
  24044. number = 3,
  24045. pages = "432--459",
  24046. month = jul,
  24047. year = 1993,
  24048. language = "en"
  24049. }
  24050. @ARTICLE{Basso1998-ng,
  24051. title = "Modulation of neuronal activity in superior colliculus by
  24052. changes in target probability",
  24053. author = "Basso, M A and Wurtz, R H",
  24054. abstract = "Complex visual scenes require that a target for an impending
  24055. saccadic eye movement be selected from a larger number of
  24056. possible targets. We investigated whether changing the
  24057. probability that a visual stimulus would be selected as the
  24058. target for a saccade altered activity of monkey superior
  24059. colliculus (SC) neurons in two experiments. First, we changed
  24060. the number of possible targets on each trial. Second, we kept
  24061. the visual display constant and presented a single saccade
  24062. target repeatedly so that target probability was established
  24063. over time. Buildup neurons in the SC, those with delay period
  24064. activity, showed a consistent reduction in activity as the
  24065. probability of the saccade decreased, independent of the visual
  24066. stimulus configuration. Other SC neurons, fixation and burst,
  24067. were largely unaffected by the changes in saccade target
  24068. probability. Because we had monkeys making saccades to many
  24069. locations within the visual field, we could examine activity
  24070. associated with saccades outside of the movement field of
  24071. neurons. We found the activity of buildup neurons to be similar
  24072. across the SC, before the target was identified, and reduced
  24073. when the number of possible targets increased. The results of
  24074. our experiments are consistent with a role for this activity in
  24075. establishing a motor set. We found, consistent with this
  24076. interpretation, that the activity of these neurons was
  24077. predictive of the latency of a saccadic eye movement and not
  24078. other saccade parameters such as end point or peak velocity.",
  24079. journal = "J. Neurosci.",
  24080. publisher = "Soc Neuroscience",
  24081. volume = 18,
  24082. number = 18,
  24083. pages = "7519--7534",
  24084. month = sep,
  24085. year = 1998,
  24086. language = "en"
  24087. }
  24088. @ARTICLE{Leon1998-oj,
  24089. title = "Exploring the neurophysiology of decisions",
  24090. author = "Leon, M I and Shadlen, M N",
  24091. journal = "Neuron",
  24092. volume = 21,
  24093. number = 4,
  24094. pages = "669--672",
  24095. month = oct,
  24096. year = 1998,
  24097. language = "en"
  24098. }
  24099. @ARTICLE{Schall2001-zu,
  24100. title = "Neural basis of deciding, choosing and acting",
  24101. author = "Schall, J D",
  24102. abstract = "The ability and opportunity to make decisions and carry out
  24103. effective actions in pursuit of goals is central to intelligent
  24104. life. Recent research has provided significant new insights into
  24105. how the brain arrives at decisions, makes choices, and produces
  24106. and evaluates the consequences of actions. In fact, by monitoring
  24107. or manipulating specific neurons, certain choices can now be
  24108. predicted or manipulated.",
  24109. journal = "Nat. Rev. Neurosci.",
  24110. volume = 2,
  24111. number = 1,
  24112. pages = "33--42",
  24113. month = jan,
  24114. year = 2001,
  24115. language = "en"
  24116. }
  24117. @ARTICLE{Evans2003-qh,
  24118. title = "In two minds: dual-process accounts of reasoning",
  24119. author = "Evans, Jonathan St B T",
  24120. abstract = "Researchers in thinking and reasoning have proposed recently that
  24121. there are two distinct cognitive systems underlying reasoning.
  24122. System 1 is old in evolutionary terms and shared with other
  24123. animals: it comprises a set of autonomous subsystems that include
  24124. both innate input modules and domain-specific knowledge acquired
  24125. by a domain-general learning mechanism. System 2 is
  24126. evolutionarily recent and distinctively human: it permits
  24127. abstract reasoning and hypothetical thinking, but is constrained
  24128. by working memory capacity and correlated with measures of
  24129. general intelligence. These theories essentially posit two minds
  24130. in one brain with a range of experimental psychological evidence
  24131. showing that the two systems compete for control of our
  24132. inferences and actions.",
  24133. journal = "Trends Cogn. Sci.",
  24134. volume = 7,
  24135. number = 10,
  24136. pages = "454--459",
  24137. month = oct,
  24138. year = 2003
  24139. }
  24140. @ARTICLE{Rangel2008-xe,
  24141. title = "A framework for studying the neurobiology of value-based decision
  24142. making",
  24143. author = "Rangel, Antonio and Camerer, Colin and Montague, P Read",
  24144. abstract = "Neuroeconomics is the study of the neurobiological and
  24145. computational basis of value-based decision making. Its goal is
  24146. to provide a biologically based account of human behaviour that
  24147. can be applied in both the natural and the social sciences. This
  24148. Review proposes a framework to investigate different aspects of
  24149. the neurobiology of decision making. The framework allows us to
  24150. bring together recent findings in the field, highlight some of
  24151. the most important outstanding problems, define a common lexicon
  24152. that bridges the different disciplines that inform
  24153. neuroeconomics, and point the way to future applications.",
  24154. journal = "Nat. Rev. Neurosci.",
  24155. volume = 9,
  24156. number = 7,
  24157. pages = "545--556",
  24158. month = jul,
  24159. year = 2008,
  24160. language = "en"
  24161. }
  24162. % The entry below contains non-ASCII chars that could not be converted
  24163. % to a LaTeX equivalent.
  24164. @MISC{Tamprateep2017-bd,
  24165. title = "Of Mice and Man",
  24166. author = "Tamprateep, V",
  24167. abstract = "… As mentioned, we considered two datasets, featuring
  24168. different maze environments, from Jeremy Freeman and Nicholas
  24169. Sofroniew at Janelia Research Campus. These two datasets
  24170. consisted of one maze each. To create a discrete task, these
  24171. mazes were discretized …",
  24172. publisher = "pdfs.semanticscholar.org",
  24173. year = 2017,
  24174. howpublished = "\url{https://pdfs.semanticscholar.org/e3a0/b91f0dbf28afcc608ba7190e7bf941693510.pdf}",
  24175. note = "Accessed: 2018-11-23"
  24176. }
  24177. @UNPUBLISHED{Steinmetz2018-zx,
  24178. title = "Distributed correlates of visually-guided behavior across the
  24179. mouse brain",
  24180. author = "Steinmetz, Nicholas and Zatka-Haas, Peter and Carandini, Matteo
  24181. and Harris, Kenneth",
  24182. abstract = "Behavior arises from neuronal activity, but it is not known how
  24183. the active neurons are distributed across brain regions and how
  24184. their activity unfolds in time. Here, we used high-density
  24185. Neuropixels probes to record from ~30,000 neurons in mice
  24186. performing a visual contrast discrimination task. The task
  24187. activated 60\% of the neurons, involving nearly all 42 recorded
  24188. brain regions, well beyond the regions activated by passive
  24189. visual stimulation. However, neurons selective for choice (left
  24190. vs. right) were rare, and found mostly in midbrain, striatum, and
  24191. frontal cortex. Those in midbrain were typically activated prior
  24192. to contralateral choices and suppressed prior to ipsilateral
  24193. choices, consistent with a competitive midbrain circuit for
  24194. adjudicating the subject9s choice. A brain-wide state shift
  24195. distinguished trials in which visual stimuli led to movement.
  24196. These results reveal concurrent representations of movement and
  24197. choice in neurons widely distributed across the brain.",
  24198. journal = "bioRxiv",
  24199. pages = "474437",
  24200. month = nov,
  24201. year = 2018,
  24202. language = "en"
  24203. }
  24204. @UNPUBLISHED{Hok2018-di,
  24205. title = "A spatial code in the dorsal lateral geniculate nucleus",
  24206. author = "Hok, Vincent and Jacob, Pierre-Yves and Bordiga, Pierrick and
  24207. Truchet, Bruno and Poucet, Bruno and Save, Etienne",
  24208. abstract = "Since their discovery in the early 970s, hippocampal place cells
  24209. have been studied in numerous animal and human spatial memory
  24210. paradigms. These pyramidal cells, along with other spatially
  24211. tuned types of neurons (e.g. grid cells, head direction cells),
  24212. are thought to provide the mammalian brain a unique spatial
  24213. signature characterizing a specific environment, and thereby a
  24214. memory trace of the subject9s place. While grid and head
  24215. direction cells are found in various brain regions, only few
  24216. hippocampal-related structures showing 9place cell9-like neurons
  24217. have been identified, thus reinforcing the central role of the
  24218. hippocampus in spatial memory. Concurrently, it is increasingly
  24219. suggested that visual areas play an important role in spatial
  24220. cognition as recent studies showed a clear spatial selectivity of
  24221. visual cortical (V1) neurons in freely moving rodents. We
  24222. therefore thought to investigate, in the rat, such spatial
  24223. correlates in a thalamic structure located one synapse upstream
  24224. of V1, the dorsal Lateral Geniculate Nucleus (dLGN), and
  24225. discovered that a substantial proportion (ca. 30\%) of neurons
  24226. exhibits spatio-selective activity. We found that dLGN place
  24227. cells maintain their spatial selectivity in the absence of visual
  24228. inputs, presumably relying on odor and locomotor inputs. We also
  24229. found that dLGN place cells maintain their place selectivity
  24230. across sessions in a familiar environment and that contextual
  24231. modifications yield separated representations. Our results show
  24232. that dLGN place cells are likely to participate in spatial
  24233. cognition processes, creating as early as the thalamic stage a
  24234. comprehensive representation of one given environment.",
  24235. journal = "bioRxiv",
  24236. pages = "473520",
  24237. month = nov,
  24238. year = 2018,
  24239. language = "en"
  24240. }
  24241. @UNPUBLISHED{Long2018-zl,
  24242. title = "A novel somatosensory spatial navigation system outside the
  24243. hippocampal formation",
  24244. author = "Long, Xiaoyang and Zhang, Sheng-Jia",
  24245. abstract = "The hippocampal-parahippocampal formation has long been regarded
  24246. as the only site for the brain9s navigational system.
  24247. Nevertheless, studies from patients with medial temporal lobe
  24248. (MTL) lesions suggest that the hippocampal formation is not
  24249. essential for space memory, indicating that spatial navigation
  24250. might be computed with another unknown representation system
  24251. outside the MTL. Such an extra-hippocampal navigational system
  24252. has never been identified, however. Here we report the existence,
  24253. in the rat somatosensory cortex only, of a novel navigational
  24254. system, which contains the full spectrum of all distinct spatial
  24255. cell types including place, head-direction, border/boundary,
  24256. conjunctive, speed and grid cells. All somatosensory spatial
  24257. cells show similar firing characteristics to those detected
  24258. previously in the hippocampal-parahippocampal structures. The
  24259. somatosensory navigational system extends the classical theory of
  24260. a cognitive map from two discrete hippocampal-entorhinal regions
  24261. to only one neocortical domain, providing possible alternative
  24262. and more sophisticated computational algorithms for spatial
  24263. memory and cognitive mapping.",
  24264. journal = "bioRxiv",
  24265. pages = "473090",
  24266. month = nov,
  24267. year = 2018,
  24268. language = "en"
  24269. }
  24270. @ARTICLE{Johnson2007-ri,
  24271. title = "Integrating hippocampus and striatum in decision-making",
  24272. author = "Johnson, Adam and van der Meer, Matthijs A A and Redish, A David",
  24273. abstract = "Learning and memory and navigation literatures emphasize
  24274. interactions between multiple memory systems: a flexible,
  24275. planning-based system and a rigid, cached-value system. This has
  24276. profound implications for decision-making. Recent
  24277. conceptualizations of flexible decision-making employ
  24278. prospection and projection arising from a network involving the
  24279. hippocampus. Recent recordings from rodent hippocampus in
  24280. decision-making situations have found transient forward-shifted
  24281. representations. Evaluation of that prediction and subsequent
  24282. action-selection probably occurs downstream (e.g. in
  24283. orbitofrontal cortex, in ventral and dorsomedial striatum).
  24284. Classically, striatum has been identified as a crucial component
  24285. of the less-flexible, incremental system. Current evidence,
  24286. however, suggests that striatum is involved in both flexible and
  24287. stimulus-response decision-making, with dorsolateral striatum
  24288. involved in stimulus-response strategies and ventral and
  24289. dorsomedial striatum involved in goal-directed strategies.",
  24290. journal = "Curr. Opin. Neurobiol.",
  24291. publisher = "Elsevier",
  24292. volume = 17,
  24293. number = 6,
  24294. pages = "692--697",
  24295. month = dec,
  24296. year = 2007,
  24297. language = "en"
  24298. }
  24299. @ARTICLE{Gupta2010-gx,
  24300. title = "Hippocampal replay is not a simple function of experience",
  24301. author = "Gupta, Anoopum S and van der Meer, Matthijs A A and Touretzky,
  24302. David S and Redish, A David",
  24303. abstract = "Replay of behavioral sequences in the hippocampus during sharp
  24304. wave ripple complexes (SWRs) provides a potential mechanism for
  24305. memory consolidation and the learning of knowledge structures.
  24306. Current hypotheses imply that replay should straightforwardly
  24307. reflect recent experience. However, we find these hypotheses to
  24308. be incompatible with the content of replay on a task with two
  24309. distinct behavioral sequences (A and B). We observed forward and
  24310. backward replay of B even when rats had been performing A for
  24311. >10 min. Furthermore, replay of nonlocal sequence B occurred
  24312. more often when B was infrequently experienced. Neither forward
  24313. nor backward sequences preferentially represented highly
  24314. experienced trajectories within a session. Additionally, we
  24315. observed the construction of never-experienced novel-path
  24316. sequences. These observations challenge the idea that sequence
  24317. activation during SWRs is a simple replay of recent experience.
  24318. Instead, replay reflected all physically available trajectories
  24319. within the environment, suggesting a potential role in active
  24320. learning and maintenance of the cognitive map.",
  24321. journal = "Neuron",
  24322. publisher = "Elsevier",
  24323. volume = 65,
  24324. number = 5,
  24325. pages = "695--705",
  24326. month = mar,
  24327. year = 2010,
  24328. language = "en"
  24329. }
  24330. @ARTICLE{Wu2014-ud,
  24331. title = "Hippocampal replay captures the unique topological structure of a
  24332. novel environment",
  24333. author = "Wu, Xiaojing and Foster, David J",
  24334. abstract = "Hippocampal place-cell replay has been proposed as a fundamental
  24335. mechanism of learning and memory, which might support
  24336. navigational learning and planning. An important hypothesis of
  24337. relevance to these proposed functions is that the information
  24338. encoded in replay should reflect the topological structure of
  24339. experienced environments; that is, which places in the
  24340. environment are connected with which others. Here we report
  24341. several attributes of replay observed in rats exploring a novel
  24342. forked environment that support the hypothesis. First, we
  24343. observed that overlapping replays depicting divergent
  24344. trajectories through the fork recruited the same population of
  24345. cells with the same firing rates to represent the common portion
  24346. of the trajectories. Second, replay tended to be directional and
  24347. to flip the represented direction at the fork. Third,
  24348. replay-associated sharp-wave-ripple events in the local field
  24349. potential exhibited substructure that mapped onto the maze
  24350. topology. Thus, the spatial complexity of our recording
  24351. environment was accurately captured by replay: the underlying
  24352. neuronal activities reflected the bifurcating shape, and both
  24353. directionality and associated ripple structure reflected the
  24354. segmentation of the maze. Finally, we observed that replays
  24355. occurred rapidly after small numbers of experiences. Our results
  24356. suggest that hippocampal replay captures learned information
  24357. about environmental topology to support a role in navigation.",
  24358. journal = "J. Neurosci.",
  24359. volume = 34,
  24360. number = 19,
  24361. pages = "6459--6469",
  24362. month = may,
  24363. year = 2014,
  24364. language = "en"
  24365. }
  24366. @ARTICLE{Dragoi2011-ty,
  24367. title = "Preplay of future place cell sequences by hippocampal cellular
  24368. assemblies",
  24369. author = "Dragoi, George and Tonegawa, Susumu",
  24370. abstract = "During spatial exploration, hippocampal neurons show a
  24371. sequential firing pattern in which individual neurons fire
  24372. specifically at particular locations along the animal's
  24373. trajectory (place cells). According to the dominant model of
  24374. hippocampal cell assembly activity, place cell firing order is
  24375. established for the first time during exploration, to encode the
  24376. spatial experience, and is subsequently replayed during rest or
  24377. slow-wave sleep for consolidation of the encoded experience.
  24378. Here we report that temporal sequences of firing of place cells
  24379. expressed during a novel spatial experience occurred on a
  24380. significant number of occasions during the resting or sleeping
  24381. period preceding the experience. This phenomenon, which is
  24382. called preplay, occurred in disjunction with sequences of replay
  24383. of a familiar experience. These results suggest that internal
  24384. neuronal dynamics during resting or sleep organize hippocampal
  24385. cellular assemblies into temporal sequences that contribute to
  24386. the encoding of a related novel experience occurring in the
  24387. future.",
  24388. journal = "Nature",
  24389. publisher = "nature.com",
  24390. volume = 469,
  24391. number = 7330,
  24392. pages = "397--401",
  24393. month = jan,
  24394. year = 2011,
  24395. language = "en"
  24396. }
  24397. @ARTICLE{Barron2013-ho,
  24398. title = "Online evaluation of novel choices by simultaneous
  24399. representation of multiple memories",
  24400. author = "Barron, Helen C and Dolan, Raymond J and Behrens, Timothy E J",
  24401. abstract = "Prior experience is critical for decision-making. It enables
  24402. explicit representation of potential outcomes and provides
  24403. training to valuation mechanisms. However, we can also make
  24404. choices in the absence of prior experience by merely imagining
  24405. the consequences of a new experience. Using functional magnetic
  24406. resonance imaging repetition suppression in humans, we examined
  24407. how neuronal representations of novel rewards can be constructed
  24408. and evaluated. A likely novel experience was constructed by
  24409. invoking multiple independent memories in hippocampus and medial
  24410. prefrontal cortex. This construction persisted for only a short
  24411. time period, during which new associations were observed between
  24412. the memories for component items. Together, these findings
  24413. suggest that, in the absence of direct experience, coactivation
  24414. of multiple relevant memories can provide a training signal to
  24415. the valuation system that allows the consequences of new
  24416. experiences to be imagined and acted on.",
  24417. journal = "Nat. Neurosci.",
  24418. publisher = "nature.com",
  24419. volume = 16,
  24420. number = 10,
  24421. pages = "1492--1498",
  24422. month = oct,
  24423. year = 2013,
  24424. language = "en"
  24425. }
  24426. @ARTICLE{Olafsdottir2015-dj,
  24427. title = "Hippocampal place cells construct reward related sequences
  24428. through unexplored space",
  24429. author = "{\'O}lafsd{\'o}ttir, H Freyja and Barry, Caswell and Saleem,
  24430. Aman B and Hassabis, Demis and Spiers, Hugo J",
  24431. abstract = "Dominant theories of hippocampal function propose that place
  24432. cell representations are formed during an animal's first
  24433. encounter with a novel environment and are subsequently replayed
  24434. during off-line states to support consolidation and future
  24435. behaviour. Here we report that viewing the delivery of food to
  24436. an unvisited portion of an environment leads to off-line
  24437. pre-activation of place cells sequences corresponding to that
  24438. space. Such 'preplay' was not observed for an unrewarded but
  24439. otherwise similar portion of the environment. These results
  24440. suggest that a hippocampal representation of a visible, yet
  24441. unexplored environment can be formed if the environment is of
  24442. motivational relevance to the animal. We hypothesise such
  24443. goal-biased preplay may support preparation for future
  24444. experiences in novel environments.",
  24445. journal = "Elife",
  24446. publisher = "cdn.elifesciences.org",
  24447. volume = 4,
  24448. pages = "e06063",
  24449. month = jun,
  24450. year = 2015,
  24451. keywords = "consolidation; hippocampus; neuroscience; place cells; preplay;
  24452. rat; replay; spatial memory",
  24453. language = "en"
  24454. }
  24455. @ARTICLE{Foster2017-ss,
  24456. title = "Replay Comes of Age",
  24457. author = "Foster, David J",
  24458. abstract = "Hippocampal place cells take part in sequenced patterns of
  24459. reactivation after behavioral experience, known as replay. Since
  24460. replay was first reported, nearly 20 years ago, many new results
  24461. have been found, necessitating revision of the original
  24462. interpretations. We review some of these results with a focus on
  24463. the phenomenology of replay.",
  24464. journal = "Annu. Rev. Neurosci.",
  24465. publisher = "annualreviews.org",
  24466. volume = 40,
  24467. pages = "581--602",
  24468. month = jul,
  24469. year = 2017,
  24470. keywords = "hippocampus; memory; place cell; replay",
  24471. language = "en"
  24472. }
  24473. @ARTICLE{Daw2005-hs,
  24474. title = "Uncertainty-based competition between prefrontal and
  24475. dorsolateral striatal systems for behavioral control",
  24476. author = "Daw, Nathaniel D and Niv, Yael and Dayan, Peter",
  24477. abstract = "A broad range of neural and behavioral data suggests that the
  24478. brain contains multiple systems for behavioral choice, including
  24479. one associated with prefrontal cortex and another with
  24480. dorsolateral striatum. However, such a surfeit of control raises
  24481. an additional choice problem: how to arbitrate between the
  24482. systems when they disagree. Here, we consider dual-action choice
  24483. systems from a normative perspective, using the computational
  24484. theory of reinforcement learning. We identify a key trade-off
  24485. pitting computational simplicity against the flexible and
  24486. statistically efficient use of experience. The trade-off is
  24487. realized in a competition between the dorsolateral striatal and
  24488. prefrontal systems. We suggest a Bayesian principle of
  24489. arbitration between them according to uncertainty, so each
  24490. controller is deployed when it should be most accurate. This
  24491. provides a unifying account of a wealth of experimental evidence
  24492. about the factors favoring dominance by either system.",
  24493. journal = "Nat. Neurosci.",
  24494. publisher = "nature.com",
  24495. volume = 8,
  24496. number = 12,
  24497. pages = "1704--1711",
  24498. month = dec,
  24499. year = 2005,
  24500. language = "en"
  24501. }
  24502. @ARTICLE{Redish1997-pq,
  24503. title = "Cognitive maps beyond the hippocampus",
  24504. author = "Redish, A D and Touretzky, D S",
  24505. abstract = "We present a conceptual framework for the role of the
  24506. hippocampus and its afferent and efferent structures in rodent
  24507. navigation. Our proposal is compatible with the behavioral,
  24508. neurophysiological, anatomical, and neuropharmacological
  24509. literature, and suggests a number of practical experiments that
  24510. could support or refute it. We begin with a review of place
  24511. cells and how the place code for an environment might be aligned
  24512. with sensory cues and updated by self-motion information. The
  24513. existence of place fields in the dark suggests that location
  24514. information is maintained by path integration, which requires an
  24515. internal representation of direction of motion. This leads to a
  24516. consideration of the organization of the rodent head direction
  24517. system, and thence into a discussion of the computational
  24518. structure and anatomical locus of the path integrator. If the
  24519. place code is used in navigation, there must be a mechanism for
  24520. selecting an action based on this information. We review
  24521. evidence that the nucleus accumbens subserves this function.
  24522. From there, we move to interactions between the hippocampal
  24523. system and the environment, emphasizing mechanisms for learning
  24524. novel environments and for aligning the various subsystems upon
  24525. re-entry into familiar environments. We conclude with a
  24526. discussion of the relationship between navigation and
  24527. declarative memory.",
  24528. journal = "Hippocampus",
  24529. publisher = "Wiley Online Library",
  24530. volume = 7,
  24531. number = 1,
  24532. pages = "15--35",
  24533. year = 1997,
  24534. language = "en"
  24535. }
  24536. % The entry below contains non-ASCII chars that could not be converted
  24537. % to a LaTeX equivalent.
  24538. @ARTICLE{Tolman1948-ym,
  24539. title = "Cognitive maps in rats and men",
  24540. author = "Tolman, E C",
  24541. abstract = "This paper is devoted to a description of experiments with rats,
  24542. mostly at the author's laboratory, and to indicating the
  24543. significance of these findings on rats for the clinical behavior
  24544. of men. While all students agree as to the facts reported, they
  24545. disagree on theory and …",
  24546. journal = "Psychol. Rev.",
  24547. publisher = "psycnet.apa.org",
  24548. volume = 55,
  24549. number = 4,
  24550. pages = "189--208",
  24551. month = jul,
  24552. year = 1948,
  24553. keywords = "CONDITIONING THERAPY",
  24554. language = "en"
  24555. }
  24556. @ARTICLE{Kepecs2008-ys,
  24557. title = "Neural correlates, computation and behavioural impact of decision
  24558. confidence",
  24559. author = "Kepecs, Adam and Uchida, Naoshige and Zariwala, Hatim A and
  24560. Mainen, Zachary F",
  24561. abstract = "Humans and other animals must often make decisions on the basis
  24562. of imperfect evidence. Statisticians use measures such as P
  24563. values to assign degrees of confidence to propositions, but
  24564. little is known about how the brain computes confidence estimates
  24565. about decisions. We explored this issue using behavioural
  24566. analysis and neural recordings in rats in combination with
  24567. computational modelling. Subjects were trained to perform an
  24568. odour categorization task that allowed decision confidence to be
  24569. manipulated by varying the distance of the test stimulus to the
  24570. category boundary. To understand how confidence could be computed
  24571. along with the choice itself, using standard models of
  24572. decision-making, we defined a simple measure that quantified the
  24573. quality of the evidence contributing to a particular decision.
  24574. Here we show that the firing rates of many single neurons in the
  24575. orbitofrontal cortex match closely to the predictions of
  24576. confidence models and cannot be readily explained by alternative
  24577. mechanisms, such as learning stimulus-outcome associations.
  24578. Moreover, when tested using a delayed reward version of the task,
  24579. we found that rats' willingness to wait for rewards increased
  24580. with confidence, as predicted by the theoretical model. These
  24581. results indicate that confidence estimates, previously suggested
  24582. to require 'metacognition' and conscious awareness are available
  24583. even in the rodent brain, can be computed with relatively simple
  24584. operations, and can drive adaptive behaviour. We suggest that
  24585. confidence estimation may be a fundamental and ubiquitous
  24586. component of decision-making.",
  24587. journal = "Nature",
  24588. volume = 455,
  24589. number = 7210,
  24590. pages = "227--231",
  24591. month = sep,
  24592. year = 2008,
  24593. language = "en"
  24594. }
  24595. @ARTICLE{Mattar2018-ro,
  24596. title = "Prioritized memory access explains planning and hippocampal
  24597. replay",
  24598. author = "Mattar, Marcelo G and Daw, Nathaniel D",
  24599. abstract = "To make decisions, animals must evaluate candidate choices by
  24600. accessing memories of relevant experiences. Yet little is known
  24601. about which experiences are considered or ignored during
  24602. deliberation, which ultimately governs choice. We propose a
  24603. normative theory predicting which memories should be accessed at
  24604. each moment to optimize future decisions. Using nonlocal 'replay'
  24605. of spatial locations in hippocampus as a window into memory
  24606. access, we simulate a spatial navigation task in which an agent
  24607. accesses memories of locations sequentially, ordered by utility:
  24608. how much extra reward would be earned due to better choices. This
  24609. prioritization balances two desiderata: the need to evaluate
  24610. imminent choices versus the gain from propagating newly
  24611. encountered information to preceding locations. Our theory offers
  24612. a simple explanation for numerous findings about place cells;
  24613. unifies seemingly disparate proposed functions of replay
  24614. including planning, learning, and consolidation; and posits a
  24615. mechanism whose dysfunction may underlie pathologies like
  24616. rumination and craving.",
  24617. journal = "Nat. Neurosci.",
  24618. volume = 21,
  24619. number = 11,
  24620. pages = "1609--1617",
  24621. month = nov,
  24622. year = 2018,
  24623. language = "en"
  24624. }
  24625. @ARTICLE{Daw2018-vt,
  24626. title = "Are we of two minds?",
  24627. author = "Daw, Nathaniel D",
  24628. journal = "Nat. Neurosci.",
  24629. volume = 21,
  24630. number = 11,
  24631. pages = "1497--1499",
  24632. month = nov,
  24633. year = 2018,
  24634. language = "en"
  24635. }
  24636. @ARTICLE{Vander_Weele2018-qt,
  24637. title = "Dopamine enhances signal-to-noise ratio in cortical-brainstem
  24638. encoding of aversive stimuli",
  24639. author = "Vander Weele, Caitlin M and Siciliano, Cody A and Matthews,
  24640. Gillian A and Namburi, Praneeth and Izadmehr, Ehsan M and
  24641. Espinel, Isabella C and Nieh, Edward H and Schut, Evelien H S and
  24642. Padilla-Coreano, Nancy and Burgos-Robles, Anthony and Chang,
  24643. Chia-Jung and Kimchi, Eyal Y and Beyeler, Anna and Wichmann, Romy
  24644. and Wildes, Craig P and Tye, Kay M",
  24645. abstract = "Dopamine modulates medial prefrontal cortex (mPFC) activity to
  24646. mediate diverse behavioural functions1,2; however, the precise
  24647. circuit computations remain unknown. One potentially unifying
  24648. model by which dopamine may underlie a diversity of functions is
  24649. by modulating the signal-to-noise ratio in subpopulations of mPFC
  24650. neurons3-6, where neural activity conveying sensory information
  24651. (signal) is amplified relative to spontaneous firing (noise).
  24652. Here we demonstrate that dopamine increases the signal-to-noise
  24653. ratio of responses to aversive stimuli in mPFC neurons projecting
  24654. to the dorsal periaqueductal grey (dPAG). Using an
  24655. electrochemical approach, we reveal the precise time course of
  24656. pinch-evoked dopamine release in the mPFC, and show that mPFC
  24657. dopamine biases behavioural responses to aversive stimuli.
  24658. Activation of mPFC-dPAG neurons is sufficient to drive place
  24659. avoidance and defensive behaviours. mPFC-dPAG neurons display
  24660. robust shock-induced excitations, as visualized by single-cell,
  24661. projection-defined microendoscopic calcium imaging. Finally,
  24662. photostimulation of dopamine terminals in the mPFC reveals an
  24663. increase in the signal-to-noise ratio in mPFC-dPAG responses to
  24664. aversive stimuli. Together, these data highlight how dopamine in
  24665. the mPFC can selectively route sensory information to specific
  24666. downstream circuits, representing a potential circuit mechanism
  24667. for valence processing.",
  24668. journal = "Nature",
  24669. volume = 563,
  24670. number = 7731,
  24671. pages = "397--401",
  24672. month = nov,
  24673. year = 2018,
  24674. language = "en"
  24675. }
  24676. @UNPUBLISHED{Bagur2018-lr,
  24677. title = "Dissociation of fear initiation and maintenance by
  24678. breathing-driven prefrontal oscillations",
  24679. author = "Bagur, Sophie and Lefort, Julie M and Lacroix, Marie M and de
  24680. Lavilleon, Gaetan and Herry, Cyril and Billand, Clara and
  24681. Geoffroy, Helene and Benchenane, Karim",
  24682. abstract = "Does the body play an active role in emotions? Since the original
  24683. James/Cannon controversy this debate has mainly been fueled by
  24684. introspective accounts of human experience. Here, we use the
  24685. animal model to demonstrate a physiological mechanism for bodily
  24686. feedback and its causal role in the stabilization of emotional
  24687. states. We report that during fear-related freezing mice breathe
  24688. at 4Hz and show, using probabilistic modelling, that optogenetic
  24689. perturbation of this feedback specifically reduces freezing
  24690. maintenance without impacting its initiation. This rhythm is
  24691. transmitted by the olfactory bulb to the prefrontal cortex where
  24692. it organizes neural firing and optogenetic probing of the circuit
  24693. demonstrates frequency-specific tuning that maximizes prefrontal
  24694. cortex responsivity at 4Hz, the breathing frequency during
  24695. freezing. These results point to a brain-body-brain loop in which
  24696. the initiation of emotional behavior engenders somatic changes
  24697. which then feedback to the cortex to directly participate in
  24698. sustaining emotional states.",
  24699. journal = "bioRxiv",
  24700. pages = "468264",
  24701. month = nov,
  24702. year = 2018,
  24703. language = "en"
  24704. }
  24705. @ARTICLE{Webb2001-kg,
  24706. title = "Can robots make good models of biological behaviour?",
  24707. author = "Webb, B",
  24708. abstract = "UNLABELLED: How should biological behaviour be modelled? A
  24709. relatively new approach is to investigate problems in
  24710. neuroethology by building physical robot models of biological
  24711. sensorimotor systems. The explication and justification of this
  24712. approach are here placed within a framework for describing and
  24713. comparing models in the behavioural and biological sciences.
  24714. First, simulation models--the representation of a hypothesis
  24715. about a target system--are distinguished from several other
  24716. relationships also termed ``modelling'' in discussions of
  24717. scientific explanation. Seven dimensions on which simulation
  24718. models can differ are defined and distinctions between them
  24719. discussed: 1. RELEVANCE: whether the model tests and generates
  24720. hypotheses applicable to biology. 2. Level: the elemental units
  24721. of the model in the hierarchy from atoms to societies. 3.
  24722. Generality: the range of biological systems the model can
  24723. represent. 4. Abstraction: the complexity, relative to the
  24724. target, or amount of detail included in the model. 5. Structural
  24725. accuracy: how well the model represents the actual mechanisms
  24726. underlying the behaviour. 6. Performance match: to what extent
  24727. the model behaviour matches the target behaviour. 7. Medium: the
  24728. physical basis by which the model is implemented. No specific
  24729. position in the space of models thus defined is the only correct
  24730. one, but a good modelling methodology should be explicit about
  24731. its position and the justification for that position. It is
  24732. argued that in building robot models biological relevance is more
  24733. effective than loose biological inspiration; multiple levels can
  24734. be integrated; that generality cannot be assumed but might emerge
  24735. from studying specific instances; abstraction is better done by
  24736. simplification than idealisation; accuracy can be approached
  24737. through iterations of complete systems; that the model should be
  24738. able to match and predict target behaviour; and that a physical
  24739. medium can have significant advantages. These arguments reflect
  24740. the view that biological behaviour needs to be studied and
  24741. modelled in context, that is, in terms of the real problems faced
  24742. by real animals in real environments.",
  24743. journal = "Behav. Brain Sci.",
  24744. volume = 24,
  24745. number = 6,
  24746. pages = "1033--50; discussion 1050--94",
  24747. month = dec,
  24748. year = 2001,
  24749. language = "en"
  24750. }
  24751. @ARTICLE{Webb2009-nq,
  24752. title = "Animals Versus Animats: Or Why Not Model the Real Iguana?",
  24753. author = "Webb, Barbara",
  24754. abstract = "The overlapping fields of adaptive behavior and artificial life
  24755. are often described as novel approaches to biology. They focus
  24756. attention on bottom-up explanations and how lifelike phenomena
  24757. can result from relatively simple systems interacting
  24758. dynamically with their environments. They are also characterized
  24759. by the use of synthetic methodologies, that is, building
  24760. artificial systems as a means of exploring these ideas. Two
  24761. differing approaches can be distinguished: building models of
  24762. specific animal systems and assessing them within complete
  24763. behavior?environment loops; and exploring the behavior of
  24764. invented artificial animals, often called animats, under similar
  24765. conditions. An obvious question about the latter approach is,
  24766. how can we learn about real biology from simulation of
  24767. non-existent animals? In this article I will argue, first, that
  24768. animat research, to the extent that it is relevant to biology,
  24769. should also be considered as model building. Animat simulations
  24770. do, implicitly, represent hypotheses about, and should be
  24771. evaluated by comparison to, animals. Casting this research in
  24772. terms of invented agents serves only to limit the ability to
  24773. draw useful conclusions from it by deflecting or deferring any
  24774. serious comparisons of the model mechanisms and results with
  24775. real biological systems. Claims that animat models are meant to
  24776. be existence proofs, idealizations, or represent general
  24777. problems in biology do not make these models qualitatively
  24778. different from more conventional models of specific animals, nor
  24779. undermine the ultimate requirement to justify this work by
  24780. making concrete comparisons with empirical data. It is thus
  24781. suggested that we will learn more by choosing real, and not
  24782. made-up, targets for our models.",
  24783. journal = "Adapt. Behav.",
  24784. publisher = "SAGE Publications Ltd STM",
  24785. volume = 17,
  24786. number = 4,
  24787. pages = "269--286",
  24788. month = aug,
  24789. year = 2009
  24790. }
  24791. @ARTICLE{Yan2017-hj,
  24792. title = "Network control principles predict neuron function in the
  24793. Caenorhabditis elegans connectome",
  24794. author = "Yan, Gang and V{\'e}rtes, Petra E and Towlson, Emma K and Chew,
  24795. Yee Lian and Walker, Denise S and Schafer, William R and
  24796. Barab{\'a}si, Albert-L{\'a}szl{\'o}",
  24797. abstract = "Recent studies on the controllability of complex systems offer a
  24798. powerful mathematical framework to systematically explore the
  24799. structure-function relationship in biological, social, and
  24800. technological networks. Despite theoretical advances, we lack
  24801. direct experimental proof of the validity of these widely used
  24802. control principles. Here we fill this gap by applying a control
  24803. framework to the connectome of the nematode Caenorhabditis
  24804. elegans, allowing us to predict the involvement of each C.
  24805. elegans neuron in locomotor behaviours. We predict that control
  24806. of the muscles or motor neurons requires 12 neuronal classes,
  24807. which include neuronal groups previously implicated in locomotion
  24808. by laser ablation, as well as one previously uncharacterized
  24809. neuron, PDB. We validate this prediction experimentally, finding
  24810. that the ablation of PDB leads to a significant loss of
  24811. dorsoventral polarity in large body bends. Importantly, control
  24812. principles also allow us to investigate the involvement of
  24813. individual neurons within each neuronal class. For example, we
  24814. predict that, within the class of DD motor neurons, only three
  24815. (DD04, DD05, or DD06) should affect locomotion when ablated
  24816. individually. This prediction is also confirmed; single cell
  24817. ablations of DD04 or DD05 specifically affect posterior body
  24818. movements, whereas ablations of DD02 or DD03 do not. Our
  24819. predictions are robust to deletions of weak connections, missing
  24820. connections, and rewired connections in the current connectome,
  24821. indicating the potential applicability of this analytical
  24822. framework to larger and less well-characterized connectomes.",
  24823. journal = "Nature",
  24824. volume = 550,
  24825. number = 7677,
  24826. pages = "519--523",
  24827. month = oct,
  24828. year = 2017,
  24829. language = "en"
  24830. }
  24831. @ARTICLE{Najafi2018-vk,
  24832. title = "Perceptual {Decision-Making}: A Field in the Midst of a
  24833. Transformation",
  24834. author = "Najafi, Farzaneh and Churchland, Anne K",
  24835. abstract = "Major changes are underway in the field of perceptual
  24836. decision-making. Single-neuron studies have given way to
  24837. population recordings with identified cell types, traditional
  24838. analyses have been extended to accommodate these large and
  24839. diverse collections of neurons, and novel methods of neural
  24840. disruption have provided insights about causal circuits.
  24841. Further, the field has expanded to include multiple new species:
  24842. rodents and invertebrates, for example, have been instrumental
  24843. in demonstrating the importance of internal state on neural
  24844. responses. Finally, a renewed interest in ethological stimuli
  24845. prompted development of new behaviors, frequently analyzed by
  24846. new, automated movement tracking methods. Taken together, these
  24847. advances constitute a seismic shift in both our approach and
  24848. understanding of how incoming sensory signals are used to guide
  24849. decisions.",
  24850. journal = "Neuron",
  24851. publisher = "Elsevier",
  24852. volume = 100,
  24853. number = 2,
  24854. pages = "453--462",
  24855. month = oct,
  24856. year = 2018,
  24857. keywords = "decision-making; cognition; mice; imaging; neural data analysis;
  24858. computational models",
  24859. language = "en"
  24860. }
  24861. @ARTICLE{Lee2017-fw,
  24862. title = "Flexibility to contingency changes distinguishes habitual and
  24863. goal-directed strategies in humans",
  24864. author = "Lee, Julie J and Keramati, Mehdi",
  24865. abstract = "Decision-making in the real world presents the challenge of
  24866. requiring flexible yet prompt behavior, a balance that has been
  24867. characterized in terms of a trade-off between a slower,
  24868. prospective goal-directed model-based (MB) strategy and a fast,
  24869. retrospective habitual model-free (MF) strategy. Theory predicts
  24870. that flexibility to changes in both reward values and transition
  24871. contingencies can determine the relative influence of the two
  24872. systems in reinforcement learning, but few studies have
  24873. manipulated the latter. Therefore, we developed a novel two-level
  24874. contingency change task in which transition contingencies between
  24875. states change every few trials; MB and MF control predict
  24876. different responses following these contingency changes, allowing
  24877. their relative influence to be inferred. Additionally, we
  24878. manipulated the rate of contingency changes in order to determine
  24879. whether contingency change volatility would play a role in
  24880. shifting subjects between a MB and MF strategy. We found that
  24881. human subjects employed a hybrid MB/MF strategy on the task,
  24882. corroborating the parallel contribution of MB and MF systems in
  24883. reinforcement learning. Further, subjects did not remain at one
  24884. level of MB/MF behaviour but rather displayed a shift towards
  24885. more MB behavior over the first two blocks that was not
  24886. attributable to the rate of contingency changes but rather to the
  24887. extent of training. We demonstrate that flexibility to
  24888. contingency changes can distinguish MB and MF strategies, with
  24889. human subjects utilizing a hybrid strategy that shifts towards
  24890. more MB behavior over blocks, consequently corresponding to a
  24891. higher payoff.",
  24892. journal = "PLoS Comput. Biol.",
  24893. volume = 13,
  24894. number = 9,
  24895. pages = "e1005753",
  24896. month = sep,
  24897. year = 2017,
  24898. language = "en"
  24899. }
  24900. @ARTICLE{Reiter2018-gu,
  24901. title = "Elucidating the control and development of skin patterning in
  24902. cuttlefish",
  24903. author = "Reiter, Sam and H{\"u}lsdunk, Philipp and Woo, Theodosia and
  24904. Lauterbach, Marcel A and Eberle, Jessica S and Akay, Leyla Anne
  24905. and Longo, Amber and Meier-Credo, Jakob and Kretschmer, Friedrich
  24906. and Langer, Julian D and Kaschube, Matthias and Laurent, Gilles",
  24907. abstract = "Few animals provide a readout that is as objective of their
  24908. perceptual state as camouflaging cephalopods. Their skin display
  24909. system includes an extensive array of pigment cells
  24910. (chromatophores), each expandable by radial muscles controlled by
  24911. motor neurons. If one could track the individual expansion states
  24912. of the chromatophores, one would obtain a quantitative
  24913. description-and potentially even a neural description by proxy-of
  24914. the perceptual state of the animal in real time. Here we present
  24915. the use of computational and analytical methods to achieve this
  24916. in behaving animals, quantifying the states of tens of thousands
  24917. of chromatophores at sixty frames per second, at single-cell
  24918. resolution, and over weeks. We infer a statistical hierarchy of
  24919. motor control, reveal an underlying low-dimensional structure to
  24920. pattern dynamics and uncover rules that govern the development of
  24921. skin patterns. This approach provides an objective description of
  24922. complex perceptual behaviour, and a powerful means to uncover the
  24923. organizational principles that underlie the function, dynamics
  24924. and morphogenesis of neural systems.",
  24925. journal = "Nature",
  24926. volume = 562,
  24927. number = 7727,
  24928. pages = "361--366",
  24929. month = oct,
  24930. year = 2018,
  24931. language = "en"
  24932. }
  24933. @MISC{noauthor_undated-bu,
  24934. institution = "bioRxiv"
  24935. }
  24936. @TECHREPORT{Kahneman1977-iq,
  24937. title = "Prospect Theory. An Analysis of Decision Making Under Risk",
  24938. author = "Kahneman, Daniel and Tversky, Amos",
  24939. publisher = "Defense Technical Information Center",
  24940. month = apr,
  24941. year = 1977
  24942. }
  24943. @MISC{noauthor_undated-zp,
  24944. institution = "bioRxiv"
  24945. }
  24946. @ARTICLE{Dhawale2017-yp,
  24947. title = "Automated long-term recording and analysis of neural activity in
  24948. behaving animals",
  24949. author = "Dhawale, Ashesh K and Poddar, Rajesh and Wolff, Steffen Be and
  24950. Normand, Valentin A and Kopelowitz, Evi and {\"O}lveczky, Bence P",
  24951. abstract = "Addressing how neural circuits underlie behavior is routinely
  24952. done by measuring electrical activity from single neurons in
  24953. experimental sessions. While such recordings yield snapshots of
  24954. neural dynamics during specified tasks, they are ill-suited for
  24955. tracking single-unit activity over longer timescales relevant for
  24956. most developmental and learning processes, or for capturing
  24957. neural dynamics across different behavioral states. Here we
  24958. describe an automated platform for continuous long-term
  24959. recordings of neural activity and behavior in freely moving
  24960. rodents. An unsupervised algorithm identifies and tracks the
  24961. activity of single units over weeks of recording, dramatically
  24962. simplifying the analysis of large datasets. Months-long
  24963. recordings from motor cortex and striatum made and analyzed with
  24964. our system revealed remarkable stability in basic neuronal
  24965. properties, such as firing rates and inter-spike interval
  24966. distributions. Interneuronal correlations and the representation
  24967. of different movements and behaviors were similarly stable. This
  24968. establishes the feasibility of high-throughput long-term
  24969. extracellular recordings in behaving animals.",
  24970. journal = "Elife",
  24971. volume = 6,
  24972. month = sep,
  24973. year = 2017,
  24974. keywords = "behavior; neural recordings; neuroscience; rat; systems
  24975. neuroscience",
  24976. language = "en"
  24977. }
  24978. @ARTICLE{Wolff2018-ir,
  24979. title = "The promise and perils of causal circuit manipulations",
  24980. author = "Wolff, Steffen Be and {\"O}lveczky, Bence P",
  24981. abstract = "The development of increasingly sophisticated methods for
  24982. recording and manipulating neural activity is revolutionizing
  24983. neuroscience. By probing how activity patterns in different types
  24984. of neurons and circuits contribute to behavior, these tools can
  24985. help inform mechanistic models of brain function and explain the
  24986. roles of distinct circuit elements. However, in systems where
  24987. functions are distributed over large networks, interpreting
  24988. causality experiments can be challenging. Here we review common
  24989. assumptions underlying circuit manipulations in behaving animals
  24990. and discuss the strengths and limitations of different
  24991. approaches.",
  24992. journal = "Curr. Opin. Neurobiol.",
  24993. volume = 49,
  24994. pages = "84--94",
  24995. month = apr,
  24996. year = 2018,
  24997. language = "en"
  24998. }
  24999. @UNPUBLISHED{Low2018-gr,
  25000. title = "Probing variability in a cognitive map using manifold inference
  25001. from neural dynamics",
  25002. author = "Low, Ryan J and Lewallen, Sam and Aronov, Dmitriy and Nevers,
  25003. Rhino and Tank, David W",
  25004. abstract = "Hippocampal neurons fire selectively in local behavioral contexts
  25005. such as the position in an environment or phase of a task, and
  25006. are thought to form a cognitive map of task-relevant variables.
  25007. However, their activity varies over repeated behavioral
  25008. conditions, such as different runs through the same position or
  25009. repeated trials. Although widely observed across the brain, such
  25010. variability is not well understood, and could reflect noise or
  25011. structure, such as the encoding of additional cognitive
  25012. information. Here, we introduce a conceptual model to explain
  25013. variability in terms of underlying, population-level structure in
  25014. single-trial neural activity. To test this model, we developed a
  25015. novel unsupervised learning algorithm incorporating temporal
  25016. dynamics, in order to characterize population activity as a
  25017. trajectory on a nonlinear manifold--a space of possible network
  25018. states. The manifold9s structure captures correlations between
  25019. neurons and temporal relationships between states, constraints
  25020. arising from underlying network architecture and inputs. Using
  25021. measurements of activity over time but no information about
  25022. exogenous behavioral variables, we recovered hippocampal activity
  25023. manifolds during spatial and non-spatial cognitive tasks in rats.
  25024. Manifolds were low dimensional and smoothly encoded task-related
  25025. variables, but contained an extra dimension reflecting
  25026. information beyond the measured behavioral variables. Consistent
  25027. with our model, neurons fired as a function of overall network
  25028. state, and fluctuations in their activity across trials
  25029. corresponded to variation in the underlying trajectory on the
  25030. manifold. In particular, the extra dimension allowed the system
  25031. to take different trajectories despite repeated behavioral
  25032. conditions. Furthermore, the trajectory could temporarily
  25033. decouple from current behavioral conditions and traverse
  25034. neighboring manifold points corresponding to past, future, or
  25035. nearby behavioral states. Our results suggest that trial-to-trial
  25036. variability in the hippocampus is structured, and may reflect the
  25037. operation of internal cognitive processes. The manifold structure
  25038. of population activity is well-suited for organizing information
  25039. to support memory, planning, and reinforcement learning. In
  25040. general, our approach could find broader use in probing the
  25041. organization and computational role of circuit dynamics in other
  25042. brain regions.",
  25043. journal = "bioRxiv",
  25044. pages = "418939",
  25045. month = sep,
  25046. year = 2018,
  25047. language = "en"
  25048. }
  25049. @ARTICLE{Pandarinath2018-mm,
  25050. title = "Inferring single-trial neural population dynamics using
  25051. sequential auto-encoders",
  25052. author = "Pandarinath, Chethan and O'Shea, Daniel J and Collins, Jasmine
  25053. and Jozefowicz, Rafal and Stavisky, Sergey D and Kao, Jonathan C
  25054. and Trautmann, Eric M and Kaufman, Matthew T and Ryu, Stephen I
  25055. and Hochberg, Leigh R and Henderson, Jaimie M and Shenoy, Krishna
  25056. V and Abbott, L F and Sussillo, David",
  25057. abstract = "Neuroscience is experiencing a revolution in which simultaneous
  25058. recording of thousands of neurons is revealing population
  25059. dynamics that are not apparent from single-neuron responses. This
  25060. structure is typically extracted from data averaged across many
  25061. trials, but deeper understanding requires studying phenomena
  25062. detected in single trials, which is challenging due to incomplete
  25063. sampling of the neural population, trial-to-trial variability,
  25064. and fluctuations in action potential timing. We introduce latent
  25065. factor analysis via dynamical systems, a deep learning method to
  25066. infer latent dynamics from single-trial neural spiking data. When
  25067. applied to a variety of macaque and human motor cortical
  25068. datasets, latent factor analysis via dynamical systems accurately
  25069. predicts observed behavioral variables, extracts precise firing
  25070. rate estimates of neural dynamics on single trials, infers
  25071. perturbations to those dynamics that correlate with behavioral
  25072. choices, and combines data from non-overlapping recording
  25073. sessions spanning months to improve inference of underlying
  25074. dynamics.",
  25075. journal = "Nat. Methods",
  25076. month = sep,
  25077. year = 2018,
  25078. language = "en"
  25079. }
  25080. @UNPUBLISHED{Ahilan2018-qy,
  25081. title = "Forgetful inference in a sophisticated world model",
  25082. author = "Ahilan, Sanjeevan and Solomon, Rebecca B and Breton,
  25083. Yannick-Andr{\'e} and Conover, Kent and Niyogi, Ritwik K and
  25084. Shizgal, Peter and Dayan, Peter",
  25085. abstract = "Humans and other animals are able to discover underlying
  25086. statistical structure in their environments and exploit it to
  25087. achieve efficient and effective performance. However, such
  25088. structure is often difficult to learn and use because it is
  25089. obscure, involving long-range temporal dependencies. Here, we
  25090. analysed behavioural data from an extended experiment with rats,
  25091. showing that the subjects learned the underlying statistical
  25092. structure, albeit suffering at times from immediate inferential
  25093. imperfections as to their current state within it. We accounted
  25094. for their behaviour using a Hidden Markov Model, in which recent
  25095. observations are integrated with the recollections of an
  25096. imperfect memory. We found that over the course of training,
  25097. subjects came to track their progress through the task more
  25098. accurately, a change that our model largely attributed to
  25099. decreased forgetting. This 9learning to remember9 decreased
  25100. reliance on recent observations, which may be misleading, in
  25101. favour of a longer-term memory.",
  25102. journal = "bioRxiv",
  25103. pages = "419317",
  25104. month = sep,
  25105. year = 2018,
  25106. language = "en"
  25107. }
  25108. @ARTICLE{Han2018-vj,
  25109. title = "A Neural Circuit for {Gut-Induced} Reward",
  25110. author = "Han, Wenfei and Tellez, Luis A and Perkins, Matthew H and Perez,
  25111. Isaac O and Qu, Taoran and Ferreira, Jozelia and Ferreira,
  25112. Tatiana L and Quinn, Daniele and Liu, Zhong-Wu and Gao, Xiao-Bing
  25113. and Kaelberer, Melanie M and Boh{\'o}rquez, Diego V and
  25114. Shammah-Lagnado, Sara J and de Lartigue, Guillaume and de Araujo,
  25115. Ivan E",
  25116. abstract = "The gut is now recognized as a major regulator of motivational
  25117. and emotional states. However, the relevant gut-brain neuronal
  25118. circuitry remains unknown. We show that optical activation of
  25119. gut-innervating vagal sensory neurons recapitulates the hallmark
  25120. effects of stimulating brain reward neurons. Specifically, right,
  25121. but not left, vagal sensory ganglion activation sustained
  25122. self-stimulation behavior, conditioned both flavor and place
  25123. preferences, and induced dopamine release from Substantia nigra.
  25124. Cell-specific transneuronal tracing revealed asymmetric ascending
  25125. pathways of vagal origin throughout the CNS. In particular,
  25126. transneuronal labeling identified the glutamatergic neurons of
  25127. the dorsolateral parabrachial region as the obligatory relay
  25128. linking the right vagal sensory ganglion to dopamine cells in
  25129. Substantia nigra. Consistently, optical activation of
  25130. parabrachio-nigral projections replicated the rewarding effects
  25131. of right vagus excitation. Our findings establish the vagal
  25132. gut-to-brain axis as an integral component of the neuronal reward
  25133. pathway. They also suggest novel vagal stimulation approaches to
  25134. affective disorders.",
  25135. journal = "Cell",
  25136. month = sep,
  25137. year = 2018,
  25138. keywords = "dopamine; gut-brain axis; reward; vagus nerve",
  25139. language = "en"
  25140. }
  25141. @UNPUBLISHED{Grundemann2018-yn,
  25142. title = "Amygdala neuronal ensembles dynamically encode behavioral states",
  25143. author = "Gr{\"u}ndemann, Jan and Bitterman, Yael and Lu, Tingjia and
  25144. Krabbe, Sabine and Grewe, Benjamin F and Schnitzer, Mark J and
  25145. L{\"u}thi, Andreas",
  25146. abstract = "Internal states, including affective or homeostatic states, are
  25147. important behavioral motivators. The amygdala is a key brain
  25148. region involved in the regulation of motivated behaviors, yet how
  25149. distinct internal states are represented in amygdala circuits is
  25150. not known. Here, by imaging somatic neural calcium dynamics in
  25151. freely moving mice, we demonstrate that changes in the relative
  25152. activity levels of two major, non-overlapping populations of
  25153. principal neurons in the basal nucleus of the amygdala (BA)
  25154. predict switches between exploratory and anxiety-like or
  25155. defensive behavioral states across different environments.
  25156. Moreover, we found that the amygdala widely broadcasts internal
  25157. state information via several output pathways to larger brain
  25158. networks, and that sensory responses in the BA were not
  25159. correlated with behavioral states. Our data indicate that the
  25160. brain processes external stimuli and internal states in an
  25161. orthogonal manner, which may facilitate rapid and flexible
  25162. selection of appropriate, state-dependent behavioral responses.",
  25163. journal = "bioRxiv",
  25164. pages = "425736",
  25165. month = sep,
  25166. year = 2018,
  25167. language = "en"
  25168. }
  25169. @ARTICLE{Buzsaki2018-je,
  25170. title = "Space and Time: The Hippocampus as a Sequence Generator",
  25171. author = "Buzs{\'a}ki, Gy{\"o}rgy and Tingley, David",
  25172. abstract = "Neural computations are often compared to instrument-measured
  25173. distance or duration, and such relationships are interpreted by a
  25174. human observer. However, neural circuits do not depend on
  25175. human-made instruments but perform computations relative to an
  25176. internally defined rate-of-change. While neuronal correlations
  25177. with external measures, such as distance or duration, can be
  25178. observed in spike rates or other measures of neuronal activity,
  25179. what matters for the brain is how such activity patterns are
  25180. utilized by downstream neural observers. We suggest that
  25181. hippocampal operations can be described by the sequential
  25182. activity of neuronal assemblies and their internally defined rate
  25183. of change without resorting to the concept of space or time.",
  25184. journal = "Trends Cogn. Sci.",
  25185. volume = 22,
  25186. number = 10,
  25187. pages = "853--869",
  25188. month = oct,
  25189. year = 2018,
  25190. keywords = "place cells; time cells; theta oscillation; phase coding; lateral
  25191. septum"
  25192. }
  25193. @ARTICLE{Moita2004-fi,
  25194. title = "Putting fear in its place: remapping of hippocampal place cells
  25195. during fear conditioning",
  25196. author = "Moita, Marta A P and Rosis, Svetlana and Zhou, Yu and LeDoux,
  25197. Joseph E and Blair, Hugh T",
  25198. abstract = "We recorded hippocampal place cells in two spatial environments:
  25199. a training environment in which rats underwent fear conditioning
  25200. and a neutral control environment. Fear conditioning caused many
  25201. place cells to alter (or remap) their preferred firing locations
  25202. in the training environment, whereas most cells remained stable
  25203. in the control environment. This finding indicates that aversive
  25204. reinforcement can induce place cell remapping even when the
  25205. environment itself remains unchanged. Furthermore, contextual
  25206. fear conditioning caused significantly more remapping of place
  25207. cells than auditory fear conditioning, suggesting that place
  25208. cell remapping was related to the rat's learned fear of the
  25209. environment. These results suggest that one possible function of
  25210. place cell remapping may be to generate new spatial
  25211. representations of a single environment, which could help the
  25212. animal to discriminate among different motivational contexts
  25213. within that environment.",
  25214. journal = "J. Neurosci.",
  25215. publisher = "Soc Neuroscience",
  25216. volume = 24,
  25217. number = 31,
  25218. pages = "7015--7023",
  25219. month = aug,
  25220. year = 2004,
  25221. language = "en"
  25222. }
  25223. @ARTICLE{Nusbaum2017-sy,
  25224. title = "Functional consequences of neuropeptide and small-molecule
  25225. co-transmission",
  25226. author = "Nusbaum, Michael P and Blitz, Dawn M and Marder, Eve",
  25227. abstract = "Colocalization of small-molecule and neuropeptide transmitters is
  25228. common throughout the nervous system of all animals. The
  25229. resulting co-transmission, which provides conjoint ionotropic
  25230. ('classical') and metabotropic ('modulatory') actions, includes
  25231. neuropeptide- specific aspects that are qualitatively different
  25232. from those that result from metabotropic actions of
  25233. small-molecule transmitter release. Here, we focus on the
  25234. flexibility afforded to microcircuits by such co-transmission,
  25235. using examples from various nervous systems. Insights from such
  25236. studies indicate that co-transmission mediated even by a single
  25237. neuron can configure microcircuit activity via an array of
  25238. contributing mechanisms, operating on multiple timescales, to
  25239. enhance both behavioural flexibility and robustness.",
  25240. journal = "Nat. Rev. Neurosci.",
  25241. volume = 18,
  25242. number = 7,
  25243. pages = "389--403",
  25244. month = jul,
  25245. year = 2017,
  25246. keywords = "synapses",
  25247. language = "en"
  25248. }
  25249. @ARTICLE{Ryan2009-qm,
  25250. title = "The origin and evolution of synapses",
  25251. author = "Ryan, Tom{\'a}s J and Grant, Seth G N",
  25252. abstract = "Understanding the evolutionary origins of behaviour is a central
  25253. aim in the study of biology and may lead to insights into human
  25254. disorders. Synaptic transmission is observed in a wide range of
  25255. invertebrate and vertebrate organisms and underlies their
  25256. behaviour. Proteomic studies of the molecular components of the
  25257. highly complex mammalian postsynaptic machinery point to an
  25258. ancestral molecular machinery in unicellular organisms--the
  25259. protosynapse--that existed before the evolution of metazoans and
  25260. neurons, and hence challenges existing views on the origins of
  25261. the brain. The phylogeny of the molecular components of the
  25262. synapse provides a new model for studying synapse diversity and
  25263. complexity, and their implications for brain evolution.",
  25264. journal = "Nat. Rev. Neurosci.",
  25265. volume = 10,
  25266. number = 10,
  25267. pages = "701--712",
  25268. month = oct,
  25269. year = 2009,
  25270. keywords = "synapses",
  25271. language = "en"
  25272. }
  25273. @ARTICLE{Masi2015-wb,
  25274. title = "Electrical spiking in bacterial biofilms",
  25275. author = "Masi, Elisa and Ciszak, Marzena and Santopolo, Luisa and
  25276. Frascella, Arcangela and Giovannetti, Luciana and Marchi,
  25277. Emmanuela and Viti, Carlo and Mancuso, Stefano",
  25278. abstract = "In nature, biofilms are the most common form of bacterial growth.
  25279. In biofilms, bacteria display coordinated behaviour to perform
  25280. specific functions. Here, we investigated electrical signalling
  25281. as a possible driver in biofilm sociobiology. Using a
  25282. multi-electrode array system that enables high spatio-temporal
  25283. resolution, we studied the electrical activity in two
  25284. biofilm-forming strains and one non-biofilm-forming strain. The
  25285. action potential rates monitored during biofilm-forming bacterial
  25286. growth exhibited a one-peak maximum with a long tail,
  25287. corresponding to the highest biofilm development. This peak was
  25288. not observed for the non-biofilm-forming strain, demonstrating
  25289. that the intensity of the electrical activity was not linearly
  25290. related to the bacterial density, but was instead correlated with
  25291. biofilm formation. Results obtained indicate that the analysis of
  25292. the spatio-temporal electrical activity of bacteria during
  25293. biofilm formation can open a new frontier in the study of the
  25294. emergence of collective microbial behaviour.",
  25295. journal = "J. R. Soc. Interface",
  25296. volume = 12,
  25297. number = 102,
  25298. pages = "20141036",
  25299. month = jan,
  25300. year = 2015,
  25301. keywords = "bacteria; biofilm; electrical spiking; multi-electrode array;
  25302. sociobiology;synapses",
  25303. language = "en"
  25304. }
  25305. @ARTICLE{Emes2012-ea,
  25306. title = "Evolution of synapse complexity and diversity",
  25307. author = "Emes, Richard D and Grant, Seth G N",
  25308. abstract = "Proteomic studies of the composition of mammalian synapses have
  25309. revealed a high degree of complexity. The postsynaptic and
  25310. presynaptic terminals are molecular systems with highly organized
  25311. protein networks producing emergent physiological and behavioral
  25312. properties. The major classes of synapse proteins and their
  25313. respective functions in intercellular communication and adaptive
  25314. responses evolved in prokaryotes and eukaryotes prior to the
  25315. origins of neurons in metazoa. In eukaryotes, the organization of
  25316. individual proteins into multiprotein complexes comprising
  25317. scaffold proteins, receptors, and signaling enzymes formed the
  25318. precursor to the core adaptive machinery of the metazoan
  25319. postsynaptic terminal. Multiplicative increases in the complexity
  25320. of this protosynapse machinery secondary to genome duplications
  25321. drove synaptic, neuronal, and behavioral novelty in vertebrates.
  25322. Natural selection has constrained diversification in mammalian
  25323. postsynaptic mechanisms and the repertoire of adaptive and innate
  25324. behaviors. The evolution and organization of synapse proteomes
  25325. underlie the origins and complexity of nervous systems and
  25326. behavior.",
  25327. journal = "Annu. Rev. Neurosci.",
  25328. volume = 35,
  25329. pages = "111--131",
  25330. year = 2012,
  25331. keywords = "synapses",
  25332. language = "en"
  25333. }
  25334. @ARTICLE{Burkhardt2017-rp,
  25335. title = "Evolutionary origin of synapses and neurons - Bridging the gap",
  25336. author = "Burkhardt, Pawel and Sprecher, Simon G",
  25337. abstract = "The evolutionary origin of synapses and neurons is an enigmatic
  25338. subject that inspires much debate. Non-bilaterian metazoans, both
  25339. with and without neurons and their closest relatives already
  25340. contain many components of the molecular toolkits for synapse
  25341. functions. The origin of these components and their assembly into
  25342. ancient synaptic signaling machineries are particularly important
  25343. in light of recent findings on the phylogeny of non-bilaterian
  25344. metazoans. The evolution of synapses and neurons are often
  25345. discussed only from a metazoan perspective leaving a considerable
  25346. gap in our understanding. By taking an integrative approach we
  25347. highlight the need to consider different, but extremely relevant
  25348. phyla and to include the closest unicellular relatives of
  25349. metazoans, the ichthyosporeans, filastereans and
  25350. choanoflagellates, to fully understand the evolutionary origin of
  25351. synapses and neurons. This approach allows for a detailed
  25352. understanding of when and how the first pre- and postsynaptic
  25353. signaling machineries evolved.",
  25354. journal = "Bioessays",
  25355. volume = 39,
  25356. number = 10,
  25357. month = oct,
  25358. year = 2017,
  25359. keywords = "evolution; neuron; origin; protein-protein interactions;
  25360. synapse;synapses",
  25361. language = "en"
  25362. }
  25363. @ARTICLE{Brunet2016-lp,
  25364. title = "From damage response to action potentials: early evolution of
  25365. neural and contractile modules in stem eukaryotes",
  25366. author = "Brunet, Thibaut and Arendt, Detlev",
  25367. abstract = "Eukaryotic cells convert external stimuli into membrane
  25368. depolarization, which in turn triggers effector responses such as
  25369. secretion and contraction. Here, we put forward an evolutionary
  25370. hypothesis for the origin of the
  25371. depolarization-contraction-secretion (DCS) coupling, the
  25372. functional core of animal neuromuscular circuits. We propose that
  25373. DCS coupling evolved in unicellular stem eukaryotes as part of an
  25374. 'emergency response' to calcium influx upon membrane rupture. We
  25375. detail how this initial response was subsequently modified into
  25376. an ancient mechanosensory-effector arc, present in the last
  25377. eukaryotic common ancestor, which enabled contractile amoeboid
  25378. movement that is widespread in extant eukaryotes. Elaborating on
  25379. calcium-triggered membrane depolarization, we reason that the
  25380. first action potentials evolved alongside the membrane of
  25381. sensory-motile cilia, with the first voltage-sensitive
  25382. sodium/calcium channels (Nav/Cav) enabling a fast and coordinated
  25383. response of the entire cilium to mechanosensory stimuli. From the
  25384. cilium, action potentials then spread across the entire cell,
  25385. enabling global cellular responses such as concerted contraction
  25386. in several independent eukaryote lineages. In animals, this
  25387. process led to the invention of mechanosensory contractile cells.
  25388. These gave rise to mechanosensory receptor cells, neurons and
  25389. muscle cells by division of labour and can be regarded as the
  25390. founder cell type of the nervous system.",
  25391. journal = "Philos. Trans. R. Soc. Lond. B Biol. Sci.",
  25392. volume = 371,
  25393. number = 1685,
  25394. pages = "20150043",
  25395. month = jan,
  25396. year = 2016,
  25397. keywords = "action potentials; electrophysiology; evo-devo; evolution;
  25398. musculature; nervous systems;synapses",
  25399. language = "en"
  25400. }
  25401. @ARTICLE{Brette_undated-lb,
  25402. title = "Theory of action potentials",
  25403. author = "Brette, Romain",
  25404. keywords = "synapses"
  25405. }
  25406. @MISC{noauthor_undated-oe,
  25407. title = "418228.full.pdf",
  25408. keywords = "synapses"
  25409. }
  25410. @ARTICLE{Cox2000-ax,
  25411. title = "Action potentials reliably invade axonal arbors of rat
  25412. neocortical neurons",
  25413. author = "Cox, C L and Denk, W and Tank, D W and Svoboda, K",
  25414. abstract = "Neocortical pyramidal neurons have extensive axonal arborizations
  25415. that make thousands of synapses. Action potentials can invade
  25416. these arbors and cause calcium influx that is required for
  25417. neurotransmitter release and excitation of postsynaptic targets.
  25418. Thus, the regulation of action potential invasion in axonal
  25419. branches might shape the spread of excitation in cortical neural
  25420. networks. To measure the reliability and extent of action
  25421. potential invasion into axonal arbors, we have used two-photon
  25422. excitation laser scanning microscopy to directly image
  25423. action-potential-mediated calcium influx in single varicosities
  25424. of layer 2/3 pyramidal neurons in acute brain slices. Our data
  25425. show that single action potentials or bursts of action potentials
  25426. reliably invade axonal arbors over a range of developmental ages
  25427. (postnatal 10-24 days) and temperatures (24 degrees C-30 degrees
  25428. C). Hyperpolarizing current steps preceding action potential
  25429. initiation, protocols that had previously been observed to
  25430. produce failures of action potential propagation in cultured
  25431. preparations, were ineffective in modulating the spread of action
  25432. potentials in acute slices. Our data show that action potentials
  25433. reliably invade the axonal arbors of neocortical pyramidal
  25434. neurons. Failures in synaptic transmission must therefore
  25435. originate downstream of action potential invasion. We also
  25436. explored the function of modulators that inhibit presynaptic
  25437. calcium influx. Consistent with previous studies, we find that
  25438. adenosine reduces action-potential-mediated calcium influx in
  25439. presynaptic terminals. This reduction was observed in all
  25440. terminals tested, suggesting that some modulatory systems are
  25441. expressed homogeneously in most terminals of the same neuron.",
  25442. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  25443. volume = 97,
  25444. number = 17,
  25445. pages = "9724--9728",
  25446. month = aug,
  25447. year = 2000,
  25448. keywords = "synapses",
  25449. language = "en"
  25450. }
  25451. @ARTICLE{Huguenard2000-qu,
  25452. title = "Reliability of axonal propagation: the spike doesn't stop here",
  25453. author = "Huguenard, J R",
  25454. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  25455. volume = 97,
  25456. number = 17,
  25457. pages = "9349--9350",
  25458. month = aug,
  25459. year = 2000,
  25460. language = "en"
  25461. }
  25462. @ARTICLE{Tao2014-bv,
  25463. title = "Flexible stop and double-cascaded stop to improve shock
  25464. reliability of {MEMS} accelerometer",
  25465. author = "Tao, Yong-Kang and Liu, Yun-Feng and Dong, Jing-Xin",
  25466. abstract = "Flexible stop could provide shock protection for MEMS
  25467. accelerometer. By modeling and simulation, the paper studied the
  25468. response of a closed-loop MEMS accelerometer with stop under
  25469. shock of different amplitudes and pulse width. Contact force
  25470. plays an important role and the shock response shows strong
  25471. nonlinearity due to the contact mechanism. A kind of
  25472. double-cascaded stop is proposed to mitigate high-frequency shock
  25473. failure. MEMS accelerometers with flexible stop and with
  25474. double-cascaded stop are both designed and fabricated based on
  25475. SOG (silicon on glass) technology. Compared with shock tests of
  25476. accelerometers with hard cylinder stop, flexible stop could
  25477. withstand more than 1e4g shock with about 100$\mu$s pulse width.
  25478. Double-cascaded stop is more robust to high frequency shock.",
  25479. journal = "Microelectron. Reliab.",
  25480. volume = 54,
  25481. number = 6,
  25482. pages = "1328--1337",
  25483. month = jun,
  25484. year = 2014
  25485. }
  25486. @ARTICLE{Pfeiffer2013-un,
  25487. title = "Hippocampal place-cell sequences depict future paths to
  25488. remembered goals",
  25489. author = "Pfeiffer, Brad E and Foster, David J",
  25490. abstract = "Effective navigation requires planning extended routes to
  25491. remembered goal locations. Hippocampal place cells have been
  25492. proposed to have a role in navigational planning, but direct
  25493. evidence has been lacking. Here we show that before goal-directed
  25494. navigation in an open arena, the rat hippocampus generates brief
  25495. sequences encoding spatial trajectories strongly biased to
  25496. progress from the subject's current location to a known goal
  25497. location. These sequences predict immediate future behaviour,
  25498. even in cases in which the specific combination of start and goal
  25499. locations is novel. These results indicate that hippocampal
  25500. sequence events characterized previously in linearly constrained
  25501. environments as 'replay' are also capable of supporting a
  25502. goal-directed, trajectory-finding mechanism, which identifies
  25503. important places and relevant behavioural paths, at specific
  25504. times when memory retrieval is required, and in a manner that
  25505. could be used to control subsequent navigational behaviour.",
  25506. journal = "Nature",
  25507. volume = 497,
  25508. number = 7447,
  25509. pages = "74--79",
  25510. month = may,
  25511. year = 2013,
  25512. language = "en"
  25513. }
  25514. @ARTICLE{Lin2007-yy,
  25515. title = "Neural encoding of the concept of nest in the mouse brain",
  25516. author = "Lin, Longnian and Chen, Guifen and Kuang, Hui and Wang, Dong and
  25517. Tsien, Joe Z",
  25518. abstract = "As important as memory is to our daily functions, the ability to
  25519. extract fundamental features and commonalities from various
  25520. episodic experiences and to then generalize them into abstract
  25521. concepts is even more crucial for both humans and animals to
  25522. adapt to novel and complex situations. Here, we report the neural
  25523. correlates of the abstract concept of nests or beds in mice.
  25524. Specifically, we find hippocampal neurons that selectively fire
  25525. or cease to fire when the mouse perceives nests or beds,
  25526. regardless of their locations and environments. Parametric
  25527. analyses show that responses of nest cells remain invariant over
  25528. changes in the nests' physical shape, style, color, odor, or
  25529. construction materials; rather, their responses are driven by
  25530. conscious awareness and physical determination of the categorical
  25531. features that would functionally define nests. Such
  25532. functionality-based abstraction and generalization of conceptual
  25533. knowledge, emerging from episodic experiences, suggests that the
  25534. hippocampus is an intrinsic part of the hierarchical structure
  25535. for generating concepts and knowledge in the brain.",
  25536. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  25537. volume = 104,
  25538. number = 14,
  25539. pages = "6066--6071",
  25540. month = apr,
  25541. year = 2007,
  25542. language = "en"
  25543. }
  25544. @ARTICLE{Yoo2018-fd,
  25545. title = "Economic Choice as an Untangling of Options into Actions",
  25546. author = "Yoo, Seng Bum Michael and Hayden, Benjamin Yost",
  25547. abstract = "We propose that economic choice can be understood as a gradual
  25548. transformation from a domain of options to one of the actions. We
  25549. draw an analogy with the idea of untangling information in the
  25550. form vision system and propose that form vision and economic
  25551. choice may be two aspects of a larger process that sculpts
  25552. actions based on sensory inputs. From this viewpoint, choice
  25553. results from the accumulated effect of repetitions of simple
  25554. computations. These may consist primarily of relative valuations
  25555. (evaluations relative to the value of rejection, perhaps in a
  25556. manner akin to divisive normalization) applied to individual
  25557. offers. With regard to economic choice, cortical brain regions
  25558. differ primarily in their position and in what information they
  25559. prioritize, and do not-with a few exceptions-have categorically
  25560. distinct roles. Each region's specific contribution is determined
  25561. largely by its inputs; thus, understanding connectivity is
  25562. crucial for understanding choice. This view suggests that there
  25563. is no single site of choice, that there is no meaningful
  25564. distinction between pre- and post-decisionality, and that there
  25565. is no explicit representation of value in the brain.",
  25566. journal = "Neuron",
  25567. volume = 99,
  25568. number = 3,
  25569. pages = "434--447",
  25570. month = aug,
  25571. year = 2018,
  25572. keywords = "Untangling; affordance competition; economic choice; functional
  25573. neuroanatomy; value",
  25574. language = "en"
  25575. }
  25576. @ARTICLE{Huk2018-ng,
  25577. title = "Beyond {Trial-Based} Paradigms: Continuous Behavior, Ongoing
  25578. Neural Activity, and Natural Stimuli",
  25579. author = "Huk, Alexander and Bonnen, Kathryn and He, Biyu J",
  25580. abstract = "The vast majority of experiments examining perception and
  25581. behavior are conducted using experimental paradigms that adhere
  25582. to a rigid trial structure: each trial consists of a brief and
  25583. discrete series of events and is regarded as independent from all
  25584. other trials. The assumptions underlying this structure ignore
  25585. the reality that natural behavior is rarely discrete, brain
  25586. activity follows multiple time courses that do not necessarily
  25587. conform to the trial structure, and the natural environment has
  25588. statistical structure and dynamics that exhibit long-range
  25589. temporal correlation. Modern advances in statistical modeling and
  25590. analysis offer tools that make it feasible for experiments to
  25591. move beyond rigid independent and identically distributed trial
  25592. structures. Here we review literature that serves as evidence for
  25593. the feasibility and advantages of moving beyond trial-based
  25594. paradigms to understand the neural basis of perception and
  25595. cognition. Furthermore, we propose a synthesis of these efforts,
  25596. integrating the characterization of natural stimulus properties
  25597. with measurements of continuous neural activity and behavioral
  25598. outputs within the framework of sensory-cognitive-motor loops.
  25599. Such a framework provides a basis for the study of natural
  25600. statistics, naturalistic tasks, and/or slow fluctuations in brain
  25601. activity, which should provide starting points for important
  25602. generalizations of analytical tools in neuroscience and
  25603. subsequent progress in understanding the neural basis of
  25604. perception and cognition.",
  25605. journal = "J. Neurosci.",
  25606. volume = 38,
  25607. number = 35,
  25608. pages = "7551--7558",
  25609. month = aug,
  25610. year = 2018,
  25611. language = "en"
  25612. }
  25613. @ARTICLE{Bicanski2018-jz,
  25614. title = "A neural-level model of spatial memory and imagery",
  25615. author = "Bicanski, Andrej and Burgess, Neil",
  25616. abstract = "We present a model of how neural representations of egocentric
  25617. spatial experiences in parietal cortex interface with
  25618. viewpoint-independent representations in medial temporal areas,
  25619. via retrosplenial cortex, to enable many key aspects of spatial
  25620. cognition. This account shows how previously reported neural
  25621. responses (place, head-direction and grid cells, allocentric
  25622. boundary- and object-vector cells, gain-field neurons) can map
  25623. onto higher cognitive function in a modular way, and predicts new
  25624. cell types (egocentric and head-direction-modulated boundary- and
  25625. object-vector cells). The model predicts how these neural
  25626. populations should interact across multiple brain regions to
  25627. support spatial memory, scene construction, novelty-detection,
  25628. 'trace cells', and mental navigation. Simulated behavior and
  25629. firing rate maps are compared to experimental data, for example
  25630. showing how object-vector cells allow items to be remembered
  25631. within a contextual representation based on environmental
  25632. boundaries, and how grid cells could update the viewpoint in
  25633. imagery during planning and short-cutting by driving sequential
  25634. place cell activity.",
  25635. journal = "Elife",
  25636. volume = 7,
  25637. month = sep,
  25638. year = 2018,
  25639. keywords = "computational model; episodic memory; neuroscience; none; scene
  25640. construction; spatial cognition; spatially selective cells; trace
  25641. cells",
  25642. language = "en"
  25643. }
  25644. @UNPUBLISHED{Zhang2018-tv,
  25645. title = "Modeling {Sensory-Motor} Decisions in Natural Behavior",
  25646. author = "Zhang, Ruohan and Zhang, Shun and Tong, Matthew H and Cui, Yuchen
  25647. and Rothkopf, Constantin A and Ballard, Dana H and Hayhoe, Mary M",
  25648. abstract = "Although a standard reinforcement learning model can capture many
  25649. aspects of reward-seeking behaviors, it may not be practical for
  25650. modeling human natural behaviors because of the richness of
  25651. dynamic environments and limitations in cognitive resources. We
  25652. propose a modular reinforcement learning model that addresses
  25653. these factors. Based on this model, a modular inverse
  25654. reinforcement learning algorithm is developed to estimate both
  25655. the rewards and discount factors from human behavioral data,
  25656. which allows predictions of human navigation behaviors in virtual
  25657. reality with high accuracy across different subjects and with
  25658. different tasks. Complex human navigation trajectories in novel
  25659. environments can be reproduced by an artificial agent that is
  25660. based on the modular model. This model provides a strategy for
  25661. estimating the subjective value of actions and how they influence
  25662. sensory-motor decisions in natural behavior.",
  25663. journal = "bioRxiv",
  25664. pages = "412155",
  25665. month = sep,
  25666. year = 2018,
  25667. language = "en"
  25668. }
  25669. % The entry below contains non-ASCII chars that could not be converted
  25670. % to a LaTeX equivalent.
  25671. @ARTICLE{Swanson2000-yg,
  25672. title = "Cerebral hemisphere regulation of motivated behavior",
  25673. author = "Swanson, Larry W",
  25674. abstract = "The goals of this article are to suggest a basic wiring diagram
  25675. for the motor neural network that controls motivated behavior,
  25676. and to provide a model for the organization of cerebral
  25677. hemisphere inputs to this network. Cerebral projections mediate
  25678. voluntary regulation of a …",
  25679. journal = "Brain Res.",
  25680. publisher = "Elsevier",
  25681. volume = 886,
  25682. number = "1-2",
  25683. pages = "113--164",
  25684. year = 2000
  25685. }
  25686. @ARTICLE{Garrett2018-np,
  25687. title = "Updating Beliefs Under Perceived Threat",
  25688. author = "Garrett, Neil and Gonz{\'a}lez-Garz{\'o}n, Ana Mar{\'\i}a and
  25689. Foulkes, Lucy and Levita, Liat and Sharot, Tali",
  25690. abstract = "Humans are better at integrating desirable information into
  25691. their beliefs than undesirable. This asymmetry poses an
  25692. evolutionary puzzle, as it can lead to an underestimation of
  25693. risk and thus failure to take precautionary action. Here, we
  25694. suggest a mechanism that can speak to this conundrum. In
  25695. particular, we show that the bias vanishes in response to
  25696. perceived threat in the environment. We report that an
  25697. improvement in participants' tendency to incorporate bad news
  25698. into their beliefs is associated with physiological arousal in
  25699. response to threat indexed by galvanic skin response and
  25700. self-reported anxiety. This pattern of results was observed in a
  25701. controlled laboratory setting (Experiment I), where perceived
  25702. threat was manipulated, and in firefighters on duty (Experiment
  25703. II), where it naturally varied. Such flexibility in how
  25704. individuals integrate information may enhance the likelihood of
  25705. responding to warnings with caution in environments rife with
  25706. threat, while maintaining a positivity bias otherwise, a
  25707. strategy that can increase well-being.SIGNIFICANCE STATEMENTThe
  25708. human tendency to be overly optimistic has mystified scholars
  25709. and lay people for decades: how could biased beliefs have been
  25710. selected for over unbiased beliefs? Scholars have suggested that
  25711. while the optimism bias can lead to negative outcomes, including
  25712. financial collapse and war, it can also facilitate health and
  25713. productivity. Here, we demonstrate that a mechanism generating
  25714. the optimism bias, namely asymmetric information integration,
  25715. evaporates under threat. Such flexibility could result in
  25716. enhanced caution in dangerous environments while supporting an
  25717. optimism bias otherwise, potentially increasing well-being.",
  25718. journal = "J. Neurosci.",
  25719. publisher = "papers.ssrn.com",
  25720. month = aug,
  25721. year = 2018,
  25722. language = "en"
  25723. }
  25724. @ARTICLE{Papale2012-xl,
  25725. title = "Interactions between deliberation and delay-discounting in rats",
  25726. author = "Papale, Andrew E and Stott, Jeffrey J and Powell, Nathaniel J
  25727. and Regier, Paul S and Redish, A David",
  25728. abstract = "When faced with decisions, rats sometimes pause and look back
  25729. and forth between possible alternatives, a phenomenon termed
  25730. vicarious trial and error (VTE). When it was first observed in
  25731. the 1930s, VTE was theorized to be a mechanism for exploration.
  25732. Later theories suggested that VTE aided the resolution of
  25733. sensory or neuroeconomic conflict. In contrast, recent
  25734. neurophysiological data suggest that VTE reflects a dynamic
  25735. search and evaluation process. These theories make unique
  25736. predictions about the timing of VTE on behavioral tasks. We
  25737. tested these theories of VTE on a T-maze with return rails,
  25738. where rats were given a choice between a smaller reward
  25739. available after one delay or a larger reward available after an
  25740. adjustable delay. Rats showed three clear phases of behavior on
  25741. this task: investigation, characterized by discovery of task
  25742. parameters; titration, characterized by iterative adjustment of
  25743. the delay to a preferred interval; and exploitation,
  25744. characterized by alternation to hold the delay at the preferred
  25745. interval. We found that VTE events occurred during adjustment
  25746. laps more often than during alternation laps. Results were
  25747. incompatible with theories of VTE as an exploratory behavior, as
  25748. reflecting sensory conflict, or as a simple neuroeconomic
  25749. valuation process. Instead, our results were most consistent
  25750. with VTE as reflecting a search process during deliberative
  25751. decision making. This pattern of VTE that we observed is
  25752. reminiscent of current navigational theories proposing a
  25753. transition from a deliberative to a habitual decision-making
  25754. mechanism.",
  25755. journal = "Cogn. Affect. Behav. Neurosci.",
  25756. publisher = "Springer",
  25757. volume = 12,
  25758. number = 3,
  25759. pages = "513--526",
  25760. month = sep,
  25761. year = 2012,
  25762. language = "en"
  25763. }
  25764. @ARTICLE{Gallivan2018-rt,
  25765. title = "Decision-making in sensorimotor control",
  25766. author = "Gallivan, Jason P and Chapman, Craig S and Wolpert, Daniel M and
  25767. Flanagan, J Randall",
  25768. abstract = "Skilled sensorimotor interactions with the world result from a
  25769. series of decision-making processes that determine, on the basis
  25770. of information extracted during the unfolding sequence of events,
  25771. which movements to make and when and how to make them. Despite
  25772. this inherent link between decision-making and sensorimotor
  25773. control, research into each of these two areas has largely
  25774. evolved in isolation, and it is only fairly recently that
  25775. researchers have begun investigating how they interact and,
  25776. together, influence behaviour. Here, we review recent
  25777. behavioural, neurophysiological and computational research that
  25778. highlights the role of decision-making processes in the
  25779. selection, planning and control of goal-directed movements in
  25780. humans and nonhuman primates.",
  25781. journal = "Nat. Rev. Neurosci.",
  25782. month = aug,
  25783. year = 2018,
  25784. language = "en"
  25785. }
  25786. @ARTICLE{Gilad2018-cq,
  25787. title = "Behavioral Strategy Determines Frontal or Posterior Location of
  25788. {Short-Term} Memory in Neocortex",
  25789. author = "Gilad, Ariel and Gallero-Salas, Yasir and Groos, Dominik and
  25790. Helmchen, Fritjof",
  25791. abstract = "The location of short-term memory in mammalian neocortex remains
  25792. elusive. Here we show that distinct neocortical areas maintain
  25793. short-term memory depending on behavioral strategy. Using
  25794. wide-field and single-cell calcium imaging, we measured layer 2/3
  25795. neuronal activity in mice performing a whisker-based texture
  25796. discrimination task with delayed response. Mice either deployed
  25797. an active strategy-engaging their body toward the approaching
  25798. texture-or passively awaited the touch. Independent of strategy,
  25799. whisker-related posterior areas encoded choice early after touch.
  25800. During the delay, in contrast, persistent cortical activity was
  25801. located medio-frontally in active trials but in a lateral
  25802. posterior area in passive trials. Perturbing these areas impaired
  25803. performance for the associated strategy and also provoked
  25804. strategy switches. Frontally maintained information related to
  25805. future action, whereas activity in the posterior cortex reflected
  25806. past stimulus identity. Thus, depending on behavioral strategy,
  25807. cortical activity is routed differentially to hold information
  25808. either frontally or posteriorly before converging to similar
  25809. action.",
  25810. journal = "Neuron",
  25811. month = jul,
  25812. year = 2018,
  25813. keywords = "barrel cortex; calcium imaging; fronto-posterior interactions;
  25814. motor cortex; optogenetics; posterolateral cortex; secondary
  25815. motor cortex; whisker; wide-field imaging; working memory",
  25816. language = "en"
  25817. }
  25818. @UNPUBLISHED{Karalis2018-wn,
  25819. title = "Breathing coordinates limbic network dynamics underlying memory
  25820. consolidation",
  25821. author = "Karalis, Nikolaos and Sirota, Anton",
  25822. abstract = "The coordinated activity between remote brain regions underlies
  25823. cognition and memory function. Although neuronal oscillations
  25824. have been proposed as a mechanistic substrate for the
  25825. coordination of information transfer and memory consolidation
  25826. during sleep, little is known about the mechanisms that support
  25827. the widespread synchronization of brain regions and the
  25828. relationship of neuronal dynamics with other bodily rhythms, such
  25829. as breathing. Here we address this question using large-scale
  25830. recordings from a number of structures, including the medial
  25831. prefrontal cortex, hippocampus, thalamus, amygdala and nucleus
  25832. accumbens in mice. We identify a dual mechanism of respiratory
  25833. entrainment, in the form of an intracerebral corollary discharge
  25834. that acts jointly with an olfactory reafference to coordinate
  25835. limbic network dynamics, such as hippocampal ripples and cortical
  25836. UP and DOWN states, involved in memory consolidation. These
  25837. results highlight breathing, a perennial rhythmic input to the
  25838. brain, as an oscillatory scaffold for the functional coordination
  25839. of the limbic circuit, enabling the segregation and integration
  25840. of information flow across neuronal networks.",
  25841. journal = "bioRxiv",
  25842. pages = "392530",
  25843. month = aug,
  25844. year = 2018,
  25845. language = "en"
  25846. }
  25847. @UNPUBLISHED{Carrillo-Reid2018-tp,
  25848. title = "Triggering visually-guided behavior by holographic activation of
  25849. pattern completion neurons in cortical ensembles",
  25850. author = "Carrillo-Reid, Luis and Han, Shuting and Yang, Weijian and
  25851. Akrouh, Alejandro and Yuste, Rafael",
  25852. abstract = "Neuronal ensembles are building blocks of cortical activity yet
  25853. it is unclear if they have any causal role in behavior. Here we
  25854. tested if the precise activation of neuronal ensembles with
  25855. two-photon holographic optogenetics in mouse primary visual
  25856. cortex alters behavioral performance in a visual task. Disruption
  25857. of behaviorally relevant cortical ensembles by activation of
  25858. non-selective neurons decreased behavioral performance whereas
  25859. optogenetic targeting of as few as two neurons with pattern
  25860. completion capability from behaviorally relevant ensembles
  25861. improved task performance by reliably recalling the whole
  25862. ensemble. Moreover, in some cases, activation of two pattern
  25863. completion neurons, in the absence of visual stimulus, triggered
  25864. correct behavioral responses. Our results demonstrate a causal
  25865. role of neuronal ensembles in a visually guided behavior and
  25866. suggest that ensembles could represent perceptual states.",
  25867. journal = "bioRxiv",
  25868. pages = "394999",
  25869. month = aug,
  25870. year = 2018,
  25871. language = "en"
  25872. }
  25873. @UNPUBLISHED{Kim2018-uq,
  25874. title = "Task complexity interacts with state-space uncertainty in the
  25875. arbitration process between model-based and model-free
  25876. reinforcement-learning at both behavioral and neural levels",
  25877. author = "Kim, Dongjae and Park, Geon Yeong and O'Doherty, John P and Lee,
  25878. Sang Wan",
  25879. abstract = "A major open question concerns how the brain governs the
  25880. allocation of control between two distinct strategies for
  25881. learning from reinforcement: model-based and model-free
  25882. reinforcement learning. While there is evidence to suggest that
  25883. the reliability of the predictions of the two systems is a key
  25884. variable responsible for the arbitration process, another key
  25885. variable has remained relatively unexplored: the role of task
  25886. complexity. By using a combination of novel task design,
  25887. computational modeling, and model-based fMRI analysis, we
  25888. examined the role of task complexity alongside state-space
  25889. uncertainty in the arbitration process between model-based and
  25890. model-free RL. We found evidence to suggest that task complexity
  25891. plays a role in influencing the arbitration process alongside
  25892. state-space uncertainty. Participants tended to increase
  25893. model-based RL control in response to increasing task complexity.
  25894. However, they resorted to model-free RL when both uncertainty and
  25895. task complexity were high, suggesting that these two variables
  25896. interact during the arbitration process. Computational fMRI
  25897. revealed that task complexity interacts with neural
  25898. representations of the reliability of the two systems in the
  25899. inferior prefrontal cortex bilaterally. These findings provide
  25900. insight into how the inferior prefrontal cortex negotiates the
  25901. trade-off between model-based and model-free RL in the presence
  25902. of uncertainty and complexity, and more generally, illustrates
  25903. how the brain resolves uncertainty and complexity in dynamically
  25904. changing environment.",
  25905. journal = "bioRxiv",
  25906. pages = "393983",
  25907. month = aug,
  25908. year = 2018,
  25909. language = "en"
  25910. }
  25911. % The entry below contains non-ASCII chars that could not be converted
  25912. % to a LaTeX equivalent.
  25913. @ARTICLE{Niv2006-yu,
  25914. title = "A normative perspective on motivation",
  25915. author = "Niv, Y and Joel, D and Dayan, P",
  25916. abstract = "Understanding the effects of motivation on instrumental action
  25917. selection, and specifically on its two main forms, goal-directed
  25918. and habitual control, is fundamental to the study of decision
  25919. making. Motivational states have been shown to
  25920. 'direct'goal-directed behavior rather straightforwardly towards
  25921. more valuable outcomes. However, how motivational states can
  25922. influence outcome-insensitive habitual behavior is more
  25923. mysterious. We adopt a normative perspective, assuming that
  25924. animals seek to maximize the utilities they achieve, and viewing
  25925. …",
  25926. journal = "Trends Cogn. Sci.",
  25927. publisher = "Elsevier",
  25928. year = 2006
  25929. }
  25930. % The entry below contains non-ASCII chars that could not be converted
  25931. % to a LaTeX equivalent.
  25932. @ARTICLE{Van_der_Meer2012-lh,
  25933. title = "Information processing in decision-making systems",
  25934. author = "van der Meer, M and Kurth-Nelson, Z and {others}",
  25935. abstract = "Decisions result from an interaction between multiple functional
  25936. systems acting in parallel to process information in very
  25937. different ways, each with strengths and weaknesses. In this
  25938. review, the authors address three action-selection components of
  25939. decision-making: The Pavlovian system releases an action from a
  25940. limited repertoire of potential actions, such as approaching
  25941. learned stimuli. Like the Pavlovian system, the habit system is
  25942. computationally fast but, unlike the Pavlovian system permits
  25943. arbitrary stimulus-action pairings. These …",
  25944. journal = "The",
  25945. publisher = "journals.sagepub.com",
  25946. year = 2012
  25947. }
  25948. @ARTICLE{Gardner2013-bg,
  25949. title = "A secondary working memory challenge preserves primary place
  25950. strategies despite overtraining",
  25951. author = "Gardner, Robert S and Uttaro, Michael R and Fleming, Samantha E
  25952. and Suarez, Daniel F and Ascoli, Giorgio A and Dumas, Theodore C",
  25953. abstract = "Learning by repetition engages distinct cognitive strategies
  25954. whose contributions are adjusted with experience. Early in
  25955. learning, performance relies upon flexible, attentive strategies.
  25956. With extended practice, inflexible, automatic strategies emerge.
  25957. This transition is thought fundamental to habit formation and
  25958. applies to human and animal cognition. In the context of spatial
  25959. navigation, place strategies are flexible, typically employed
  25960. early in training, and rely on the spatial arrangement of
  25961. landmarks to locate a goal. Response strategies are inflexible,
  25962. become dominant after overtraining, and utilize fixed motor
  25963. sequences. Although these strategies can operate independently,
  25964. they have also been shown to interact. However, since previous
  25965. work has focused on single-choice learning, if and how these
  25966. strategies interact across sequential choices remains unclear. To
  25967. test strategy interactions across sequential choices, we utilized
  25968. various two-choice spatial navigation tasks administered on the
  25969. Opposing Ts maze, an apparatus for rodents that permits
  25970. experimental control over strategy recruitment. We found that
  25971. when a second choice required spatial working memory, the
  25972. transition to response navigation on the first choice was
  25973. blocked. Control experiments specified this effect to the
  25974. cognitive aspects of the secondary task. In addition, response
  25975. navigation, once established on a single choice, was not reversed
  25976. by subsequent introduction of a secondary choice reliant on
  25977. spatial working memory. These results demonstrate that
  25978. performance strategies interact across choices, highlighting the
  25979. sensitivity of strategy use to the cognitive demands of
  25980. subsequent actions, an influence from which overtrained rigid
  25981. actions may be protected.",
  25982. journal = "Learn. Mem.",
  25983. volume = 20,
  25984. number = 11,
  25985. pages = "648--656",
  25986. month = oct,
  25987. year = 2013,
  25988. language = "en"
  25989. }
  25990. @ARTICLE{Hasz2018-hy,
  25991. title = "Deliberation and Procedural Automation on a {Two-Step} Task for
  25992. Rats",
  25993. author = "Hasz, Brendan M and Redish, A David",
  25994. abstract = "Current theories suggest that decision-making arises from
  25995. multiple systems. Human studies dissociate model-based and
  25996. model-free systems, while rodent studies dissociate deliberation
  25997. and habit. However, the relationship between these constructs
  25998. remains unresolved. We adapted for rats a two-step task which has
  25999. been used to dissociate model based from model-free decisions in
  26000. humans. We found that an uncertainty-based algorithm predicted
  26001. rats' choices on the maze better than an algorithm with a
  26002. constant weighting between systems, supporting theoretical work
  26003. suggesting decision-making systems are more likely to be used
  26004. when their valuation is less uncertain. We also found that path
  26005. stereotypy, a measure of behavioral consistency associated with
  26006. procedural learning, was correlated with model-free certainty,
  26007. while vicarious trial and error, a deliberative behavior, was
  26008. increased during model-free uncertainty.",
  26009. journal = "Front. Integr. Neurosci.",
  26010. volume = 12,
  26011. pages = "30",
  26012. year = 2018
  26013. }
  26014. @ARTICLE{McNaughton2018-bc,
  26015. title = "Survival circuits and risk assessment",
  26016. author = "McNaughton, Neil and Corr, Philip J",
  26017. abstract = "Risk assessment (RA) behaviour is unusual in the context of
  26018. survival circuits. An external object elicits eating, mating or
  26019. fleeing; but conflict between internal approach and withdrawal
  26020. tendencies elicits RA-specific behaviour that scans the
  26021. environment for new information to bring closure. Recently rodent
  26022. and human threat responses have been compared using `predators'
  26023. that can be real (e.g. a tarantula), robot, virtual, or symbolic
  26024. (with the last three rendered predatory by the use of shock).
  26025. `Quick and dirty' survival circuits in the periaqueductal grey,
  26026. hypothalamus, and amygdala control external RA behaviour. These
  26027. subcortical circuits activate, and are partially inhibited by,
  26028. higher-order internal RA processes (anxiety, memory scanning,
  26029. evaluation and sometimes---maladaptive rumination) in the ventral
  26030. hippocampus and medial prefrontal cortex.",
  26031. journal = "Current Opinion in Behavioral Sciences",
  26032. volume = 24,
  26033. pages = "14--20",
  26034. month = dec,
  26035. year = 2018
  26036. }
  26037. @UNPUBLISHED{Javer2018-js,
  26038. title = "Powerful and interpretable behavioural features for quantitative
  26039. phenotyping of C. elegans",
  26040. author = "Javer, Avelino and Ripoll-Sanchez, Lidia and Brown, Andr{\'e} E X",
  26041. abstract = "Behaviour is a sensitive and integrative readout of nervous
  26042. system function and therefore an attractive measure for assessing
  26043. the effects of mutation or drug treatment on animals. Video data
  26044. provides a rich but high-dimensional representation of behaviour
  26045. and so the first step of analysis is often some form of tracking
  26046. and feature extraction to reduce dimensionality while maintaining
  26047. relevant information. Modern machine learning methods are
  26048. powerful but notoriously difficult to interpret, while
  26049. handcrafted features are interpretable but do not always perform
  26050. as well. Here we report a new set of handcrafted features to
  26051. compactly quantify C. elegans behaviour. The features are
  26052. designed to be interpretable but to capture as much of the
  26053. phenotypic differences between worms as possible. We show that
  26054. the full feature set is more powerful than a previously defined
  26055. feature set in classifying mutant strains. We then use a
  26056. combination of automated and manual feature selection to define a
  26057. core set of interpretable features that still provides sufficient
  26058. power to detect behavioural differences between mutant strains
  26059. and the wild type. Finally, we apply the new features to detect
  26060. time- resolved behavioural differences in a series of optogenetic
  26061. experiments targeting different neural subsets.",
  26062. journal = "bioRxiv",
  26063. pages = "389023",
  26064. month = aug,
  26065. year = 2018,
  26066. language = "en"
  26067. }
  26068. @ARTICLE{Laubach2018-sh,
  26069. title = "What, if anything, is rodent prefrontal cortex?",
  26070. author = "Laubach, Mark and Amarante, Linda and Swanson, Kyra and White,
  26071. Samantha R",
  26072. publisher = "PsyArXiv",
  26073. year = 2018
  26074. }
  26075. @ARTICLE{Stern2015-kb,
  26076. title = "Analyzing animal behavior via classifying each video frame using
  26077. convolutional neural networks",
  26078. author = "Stern, Ulrich and He, Ruo and Yang, Chung-Hui",
  26079. abstract = "High-throughput analysis of animal behavior requires software to
  26080. analyze videos. Such software analyzes each frame individually,
  26081. detecting animals' body parts. But the image analysis rarely
  26082. attempts to recognize ``behavioral states''-e.g., actions or
  26083. facial expressions-directly from the image instead of using the
  26084. detected body parts. Here, we show that convolutional neural
  26085. networks (CNNs)-a machine learning approach that recently became
  26086. the leading technique for object recognition, human pose
  26087. estimation, and human action recognition-were able to recognize
  26088. directly from images whether Drosophila were ``on'' (standing or
  26089. walking) or ``off'' (not in physical contact with) egg-laying
  26090. substrates for each frame of our videos. We used multiple nets
  26091. and image transformations to optimize accuracy for our
  26092. classification task, achieving a surprisingly low error rate of
  26093. just 0.072\%. Classifying one of our 8 h videos took less than 3
  26094. h using a fast GPU. The approach enabled uncovering a novel
  26095. egg-laying-induced behavior modification in Drosophila.
  26096. Furthermore, it should be readily applicable to other behavior
  26097. analysis tasks.",
  26098. journal = "Sci. Rep.",
  26099. volume = 5,
  26100. pages = "14351",
  26101. month = sep,
  26102. year = 2015,
  26103. keywords = "Analysis/Modelling [Behaviour]",
  26104. language = "en"
  26105. }
  26106. @ARTICLE{Felsen2012-ng,
  26107. title = "Midbrain contributions to sensorimotor decision making",
  26108. author = "Felsen, Gidon and Mainen, Zachary F",
  26109. abstract = "Making decisions about future actions is a fundamental function
  26110. of the nervous system. Classical theories hold that separate sets
  26111. of brain regions are responsible for selecting and implementing
  26112. an action. Traditionally, action selection has been considered
  26113. the domain of high-level regions, such as the prefrontal cortex,
  26114. whereas action generation is thought to be carried out by
  26115. dedicated cortical and subcortical motor regions. However,
  26116. increasing evidence suggests that the activity of individual
  26117. neurons in cortical motor structures reflects abstract properties
  26118. of ``decision variables'' rather than conveying simple motor
  26119. commands. Less is known, though, about the role of subcortical
  26120. structures in decision making. In particular, the superior
  26121. colliculus (SC) is critical for planning and initiating visually
  26122. guided, gaze-displacing movements and selecting visual targets,
  26123. but whether and how it contributes more generally to sensorimotor
  26124. decisions are unclear. Here, we show that the SC is intimately
  26125. involved in orienting decisions based on odor cues, even though
  26126. the SC does not explicitly process olfactory stimuli. Neurons
  26127. were recorded from the intermediate and deep SC layers in rats
  26128. trained to perform a delayed-response, odor-cued spatial choice
  26129. task. SC neurons commonly fired well in advance of movement
  26130. initiation, predicting the chosen direction nearly 1 s before
  26131. movement. Moreover, under conditions of sensory uncertainty, SC
  26132. activity varied with task difficulty and reward outcome,
  26133. reflecting the influence of decision variables on the
  26134. intercollicular competition thought to underlie orienting
  26135. movements. These results indicate that the SC plays a more
  26136. general role in decisions than previously appreciated, extending
  26137. beyond visuomotor functions.",
  26138. journal = "J. Neurophysiol.",
  26139. volume = 108,
  26140. number = 1,
  26141. pages = "135--147",
  26142. month = jul,
  26143. year = 2012,
  26144. language = "en"
  26145. }
  26146. @ARTICLE{Anderson2014-hr,
  26147. title = "Toward a science of computational ethology",
  26148. author = "Anderson, David J and Perona, Pietro",
  26149. abstract = "The new field of ``Computational Ethology'' is made possible by
  26150. advances in technology, mathematics, and engineering that allow
  26151. scientists to automate the measurement and the analysis of animal
  26152. behavior. We explore the opportunities and long-term directions
  26153. of research in this area.",
  26154. journal = "Neuron",
  26155. volume = 84,
  26156. number = 1,
  26157. pages = "18--31",
  26158. month = oct,
  26159. year = 2014,
  26160. language = "en"
  26161. }
  26162. @ARTICLE{Kastner2013-gi,
  26163. title = "When can a Computer Simulation act as Substitute for an
  26164. Experiment? A {Case-Study} from Chemisty",
  26165. author = "K{\"a}stner, Johannes and Arnold, Eckhart",
  26166. year = 2013,
  26167. keywords = "Theoretical"
  26168. }
  26169. @INCOLLECTION{Peschard2013-ga,
  26170. title = "Modeling and experimenting",
  26171. booktitle = "Models, simulations, and representations",
  26172. author = "Peschard, Isabelle",
  26173. publisher = "Routledge",
  26174. pages = "60--79",
  26175. year = 2013,
  26176. keywords = "Theoretical"
  26177. }
  26178. @ARTICLE{Peschard2011-sz,
  26179. title = "Is Simulation an Epistemic Substitute for Experimentation?",
  26180. author = "Peschard, Isabelle",
  26181. abstract = "It is sometimes said that simulation can serve as epistemic
  26182. substitute for experimentation. Such a claim might be suggested
  26183. by the fast-spreading use of computer simulation to investigate
  26184. phenomena not accessible to experimentation (in astrophysics,
  26185. ecology, economics, climatology, etc.). But what does that mean?
  26186. The paper starts with a clarification of the terms of the issue
  26187. and then focuses on two powerful arguments for the view that
  26188. simulation and experimentation are `epistemically on a par'. One
  26189. is based on the claim that, in experimentation, no less than in
  26190. simulation, it is not the system under study that is manipulated
  26191. but a system that `stands-in' for it. The other one highlights
  26192. the pervasive use of models in experimentation. It will be argued
  26193. that these arguments, as compelling as they might seem, are each
  26194. based on a mistaken interpretation of experimentation and that,
  26195. far from simulation and experimentation being epistemically on a
  26196. par, they do not have the same epistemic function, do not produce
  26197. the same kind of epistemic results.",
  26198. month = aug,
  26199. year = 2011,
  26200. keywords = "simulation, experiment, experimentation, substitute, modeling,
  26201. target system, surrogate.;Theoretical",
  26202. language = "en"
  26203. }
  26204. @INCOLLECTION{Guala2002-ap,
  26205. title = "Models, Simulations, and Experiments",
  26206. booktitle = "{Model-Based} Reasoning: Science, Technology, Values",
  26207. author = "Guala, Francesco",
  26208. editor = "Magnani, Lorenzo and Nersessian, Nancy J",
  26209. abstract = "I discuss the difference between models, simulations, and
  26210. experiments from an epistemological and an ontological
  26211. perspective. I first distinguish between ``static'' models (like
  26212. a map) and ``dynamic'' models endowed with the capacity to
  26213. generate processes. Only the latter can be used to simulate. I
  26214. then criticize the view according to which the difference
  26215. between models/simulations and experiments is fundamentally
  26216. epistemic in character. Following Herbert Simon, I argue that
  26217. the difference is ontological. Simulations merely require the
  26218. existence of an abstract correspondence between the simulating
  26219. and the simulated system. In experiments, in contrast, the
  26220. causal relations governing the experimental and the target
  26221. systems are grounded in the same material. Simulations can
  26222. produce new knowledge just as experiments do, but the prior
  26223. knowledge needed to run a good simulation is not the same as
  26224. that needed to run a good experiment. I conclude by discussing
  26225. ``hybrid'' cases of ``experimental simulations'' or ``simulating
  26226. experiments''.",
  26227. publisher = "Springer US",
  26228. pages = "59--74",
  26229. year = 2002,
  26230. address = "Boston, MA",
  26231. keywords = "Theoretical"
  26232. }
  26233. @ARTICLE{Frigg2009-bp,
  26234. title = "The philosophy of simulation: hot new issues or same old stew?",
  26235. author = "Frigg, Roman and Reiss, Julian",
  26236. abstract = "Computer simulations are an exciting tool that plays important
  26237. roles in many scientific disciplines. This has attracted the
  26238. attention of a number of philosophers of science. The main tenor
  26239. in this literature is that computer simulations not only
  26240. constitute interesting and powerful new science, but that they
  26241. also raise a host of new philosophical issues. The protagonists
  26242. in this debate claim no less than that simulations call into
  26243. question our philosophical understanding of scientific ontology,
  26244. the epistemology and semantics of models and theories, and the
  26245. relation between experimentation and theorising, and submit that
  26246. simulations demand a fundamentally new philosophy of science in
  26247. many respects. The aim of this paper is to critically evaluate
  26248. these claims. Our conclusion will be sober. We argue that these
  26249. claims are overblown and that simulations, far from demanding a
  26250. new metaphysics, epistemology, semantics and methodology, raise
  26251. few if any new philosophical problems. The philosophical
  26252. problems that do come up in connection with simulations are not
  26253. specific to simulations and most of them are variants of
  26254. problems that have been discussed in other contexts before.",
  26255. journal = "Synthese",
  26256. publisher = "Springer Netherlands",
  26257. volume = 169,
  26258. number = 3,
  26259. pages = "593--613",
  26260. month = aug,
  26261. year = 2009,
  26262. keywords = "Theoretical",
  26263. language = "en"
  26264. }
  26265. @ARTICLE{Huxter2003-tz,
  26266. title = "Independent rate and temporal coding in hippocampal pyramidal
  26267. cells",
  26268. author = "Huxter, John and Burgess, Neil and O'Keefe, John",
  26269. abstract = "In the brain, hippocampal pyramidal cells use temporal as well as
  26270. rate coding to signal spatial aspects of the animal's environment
  26271. or behaviour. The temporal code takes the form of a phase
  26272. relationship to the concurrent cycle of the hippocampal
  26273. electroencephalogram theta rhythm. These two codes could each
  26274. represent a different variable. However, this requires the rate
  26275. and phase to vary independently, in contrast to recent
  26276. suggestions that they are tightly coupled, both reflecting the
  26277. amplitude of the cell's input. Here we show that the time of
  26278. firing and firing rate are dissociable, and can represent two
  26279. independent variables: respectively the animal's location within
  26280. the place field, and its speed of movement through the field.
  26281. Independent encoding of location together with actions and
  26282. stimuli occurring there may help to explain the dual roles of the
  26283. hippocampus in spatial and episodic memory, or may indicate a
  26284. more general role of the hippocampus in relational/declarative
  26285. memory.",
  26286. journal = "Nature",
  26287. volume = 425,
  26288. number = 6960,
  26289. pages = "828--832",
  26290. month = oct,
  26291. year = 2003,
  26292. language = "en"
  26293. }
  26294. @ARTICLE{Jazayeri2017-on,
  26295. title = "Navigating the Neural Space in Search of the Neural Code",
  26296. author = "Jazayeri, Mehrdad and Afraz, Arash",
  26297. abstract = "The advent of powerful perturbation tools, such as optogenetics,
  26298. has created new frontiers for probing causal dependencies in
  26299. neural and behavioral states. These approaches have significantly
  26300. enhanced the ability to characterize the contribution of
  26301. different cells and circuits to neural function in health and
  26302. disease. They have shifted the emphasis of research toward causal
  26303. interrogations and increased the demand for more precise and
  26304. powerful tools to control and manipulate neural activity. Here,
  26305. we clarify the conditions under which measurements and
  26306. perturbations support causal inferences. We note that the brain
  26307. functions at multiple scales and that causal dependencies may be
  26308. best inferred with perturbation tools that interface with the
  26309. system at the appropriate scale. Finally, we develop a geometric
  26310. framework to facilitate the interpretation of causal experiments
  26311. when brain perturbations do or do not respect the intrinsic
  26312. patterns of brain activity. We describe the challenges and
  26313. opportunities of applying perturbations in the presence of
  26314. dynamics, and we close with a general perspective on navigating
  26315. the activity space of neurons in the search for neural codes.",
  26316. journal = "Neuron",
  26317. volume = 93,
  26318. number = 5,
  26319. pages = "1003--1014",
  26320. month = mar,
  26321. year = 2017,
  26322. keywords = "behavior; causation; correlation; neural code; neural manifold;
  26323. perturbation;Theoretical",
  26324. language = "en"
  26325. }
  26326. @UNPUBLISHED{Slezak2018-ix,
  26327. title = "Astrocytes integrate local sensory and brain-wide neuromodulatory
  26328. signals",
  26329. author = "Slezak, Michal and Kandler, Steffen and Van Veldhoven, Paul P and
  26330. Bonin, Vincent and Holt, Matthew G",
  26331. abstract = "Astrocytes play multiple functions in the central nervous system,
  26332. from control of blood flow through to modulation of synaptic
  26333. activity. Transient increases in intracellular Ca2+ are thought
  26334. to control these activities. The prevailing concept is that these
  26335. Ca2+ transients are triggered by distinct pathways, with little
  26336. mechanistic and functional overlap. Here we demonstrate that
  26337. astrocytes in visual cortex of mice encode local visual signals
  26338. in conjunction with arousal state, functioning as multi-modal
  26339. integrators. Such activity adds an additional layer of complexity
  26340. to astrocyte function and may enable astrocytes to specifically
  26341. and subtly regulate local network activity and plasticity.",
  26342. journal = "bioRxiv",
  26343. pages = "381434",
  26344. month = jul,
  26345. year = 2018,
  26346. language = "en"
  26347. }
  26348. @ARTICLE{Kaplan2011-vm,
  26349. title = "Explanation and description in computational neuroscience",
  26350. author = "Kaplan, David Michael",
  26351. abstract = "The central aim of this paper is to shed light on the nature of
  26352. explanation in computational neuroscience. I argue that
  26353. computational models in this domain possess explanatory force to
  26354. the extent that they describe the mechanisms responsible for
  26355. producing a given phenomenon---paralleling how other mechanistic
  26356. models explain. Conceiving computational explanation as a
  26357. species of mechanistic explanation affords an important
  26358. distinction between computational models that play genuine
  26359. explanatory roles and those that merely provide accurate
  26360. descriptions or predictions of phenomena. It also serves to
  26361. clarify the pattern of model refinement and elaboration
  26362. undertaken by computational neuroscientists.",
  26363. journal = "Synthese",
  26364. publisher = "Springer Netherlands",
  26365. volume = 183,
  26366. number = 3,
  26367. pages = "339",
  26368. month = dec,
  26369. year = 2011,
  26370. keywords = "Theoretical",
  26371. language = "en"
  26372. }
  26373. @UNPUBLISHED{George2018-aw,
  26374. title = "Cortical Microcircuits from a Generative Vision Model",
  26375. author = "George, Dileep and Lavin, Alexander and Swaroop Guntupalli, J and
  26376. Mely, David and Hay, Nick and Lazaro-Gredilla, Miguel",
  26377. abstract = "Understanding the information processing roles of cortical
  26378. circuits is an outstanding problem in neuroscience and artificial
  26379. intelligence. The theoretical setting of Bayesian inference has
  26380. been suggested as a framework for understanding cortical
  26381. computation. Based on a recently published generative model for
  26382. visual inference (George et al., 2017), we derive a family of
  26383. anatomically instantiated and functional cortical circuit models.
  26384. In contrast to simplistic models of Bayesian inference, the
  26385. underlying generative model9s representational choices are
  26386. validated with real-world tasks that required efficient inference
  26387. and strong generalization. The cortical circuit model is derived
  26388. by systematically comparing the computational requirements of
  26389. this model with known anatomical constraints. The derived model
  26390. suggests precise functional roles for the feedforward, feedback
  26391. and lateral connections observed in different laminae and
  26392. columns, and assigns a computational role for the path through
  26393. the thalamus.",
  26394. journal = "bioRxiv",
  26395. pages = "379313",
  26396. month = aug,
  26397. year = 2018,
  26398. language = "en"
  26399. }
  26400. @UNPUBLISHED{Genewsky2018-bj,
  26401. title = "How much fear is in anxiety?",
  26402. author = "Genewsky, Andreas and Albrecht, Nina and Bura, Simona A and
  26403. Kaplick, Paul M and Heinz, Daniel E and Nu{\ss}baumer, Markus and
  26404. Engel, Mareen and Gr{\"u}necker, Barbara and Kaltwasser,
  26405. Sebastian F and Riebe, Caitlin J and Bedenk, Benedikt T and
  26406. Czisch, Michael and Wotjak, Carsten T",
  26407. abstract = "The selective breeding for extreme behavior on the elevated
  26408. plus-maze (EPM) resulted in two mouse lines namely high-anxiety
  26409. behaving (HAB) and low-anxiety behaving (LAB) mice. Using novel
  26410. behavioral tests we demonstrate that HAB animals additionally
  26411. exhibit maladaptive escape behavior and defensive vocalizations,
  26412. whereas LAB mice show profound deficits in escaping from
  26413. approaching threats which partially results from sensory
  26414. deficits. We could relate these behavioral distortions to tonic
  26415. changes in brain activity within the periaqueductal gray (PAG) in
  26416. HAB mice and the superior colliculus (SC) in LAB mice, using in
  26417. vivo manganese-enhanced MRI (MEMRI) followed by pharmacological
  26418. or chemogenetic interventions. Therefore, midbrain-tectal
  26419. structures govern the expression of both anxiety-like behavior
  26420. and defensive responses. Our results challenge the uncritical use
  26421. of the anthropomorphic terms anxiety or anxiety-like for the
  26422. description of mouse behavior, as they imply higher cognitive
  26423. processes, which are not necessarily in place.",
  26424. journal = "bioRxiv",
  26425. pages = "385823",
  26426. month = aug,
  26427. year = 2018,
  26428. keywords = "Threat response",
  26429. language = "en"
  26430. }
  26431. @ARTICLE{Juavinett2018-al,
  26432. title = "Decision-making behaviors: weighing ethology, complexity, and
  26433. sensorimotor compatibility",
  26434. author = "Juavinett, Ashley L and Erlich, Jeffrey C and Churchland, Anne K",
  26435. abstract = "Rodent decision-making research aims to uncover the neural
  26436. circuitry underlying the ability to evaluate alternatives and
  26437. select appropriate actions. Designing behavioral paradigms that
  26438. provide a solid foundation to ask questions about decision-making
  26439. computations and mechanisms is a difficult and often
  26440. underestimated challenge. Here, we propose three dimensions on
  26441. which we can consider rodent decision-making tasks: ethological
  26442. validity, task complexity, and stimulus-response compatibility.
  26443. We review recent research through this lens, and provide
  26444. practical guidance for researchers in the decision-making field.",
  26445. journal = "Curr. Opin. Neurobiol.",
  26446. volume = 49,
  26447. pages = "42--50",
  26448. month = apr,
  26449. year = 2018,
  26450. keywords = "Decision Making",
  26451. language = "en"
  26452. }
  26453. % The entry below contains non-ASCII chars that could not be converted
  26454. % to a LaTeX equivalent.
  26455. @ARTICLE{Meyer2018-ex,
  26456. title = "An ultralight head-mounted camera system integrates detailed
  26457. behavioral monitoring with multichannel electrophysiology in
  26458. freely moving mice",
  26459. author = "Meyer, A F and Poort, J and O'Keefe, J and Sahani, M and Linden,
  26460. J F",
  26461. abstract = "Breakthroughs in understanding the neural basis of natural
  26462. behavior require neural recording and intervention to be paired
  26463. with high-fidelity multimodal behavioral monitoring. An
  26464. extensive genetic toolkit for neural circuit dissection, and
  26465. well-developed neural recording technology, make the mouse a
  26466. powerful model organism for systems neuroscience. However,
  26467. methods for high-bandwidth acquisition of behavioral signals in
  26468. mice remain limited to fixed-position cameras and other
  26469. off-animal devices, complicating the …",
  26470. journal = "bioRxiv",
  26471. publisher = "biorxiv.org",
  26472. year = 2018,
  26473. keywords = "Analysis/Modelling [Behaviour]"
  26474. }
  26475. @ARTICLE{Sejnowski2014-rn,
  26476. title = "Putting big data to good use in neuroscience",
  26477. author = "Sejnowski, Terrence J and Churchland, Patricia S and Movshon, J
  26478. Anthony",
  26479. abstract = "Big data has transformed fields such as physics and genomics.
  26480. Neuroscience is set to collect its own big data sets, but to
  26481. exploit its full potential, there need to be ways to
  26482. standardize, integrate and synthesize diverse types of data from
  26483. different levels of analysis and across species. This will
  26484. require a cultural shift in sharing data across labs, as well as
  26485. to a central role for theorists in neuroscience research.",
  26486. journal = "Nat. Neurosci.",
  26487. publisher = "nature.com",
  26488. volume = 17,
  26489. number = 11,
  26490. pages = "1440--1441",
  26491. month = nov,
  26492. year = 2014,
  26493. keywords = "Theoretical",
  26494. language = "en"
  26495. }
  26496. @ARTICLE{Carandini2013-hh,
  26497. title = "Probing perceptual decisions in rodents",
  26498. author = "Carandini, Matteo and Churchland, Anne K",
  26499. abstract = "The study of perceptual decision-making offers insight into how
  26500. the brain uses complex, sometimes ambiguous information to guide
  26501. actions. Understanding the underlying processes and their neural
  26502. bases requires that one pair recordings and manipulations of
  26503. neural activity with rigorous psychophysics. Though this
  26504. research has been traditionally performed in primates, it seems
  26505. increasingly promising to pursue it at least partly in mice and
  26506. rats. However, rigorous psychophysical methods are not yet as
  26507. developed for these rodents as they are for primates. Here we
  26508. give a brief overview of the sensory capabilities of rodents and
  26509. of their cortical areas devoted to sensation and decision. We
  26510. then review methods of psychophysics, focusing on the technical
  26511. issues that arise in their implementation in rodents. These
  26512. methods represent a rich set of challenges and opportunities.",
  26513. journal = "Nat. Neurosci.",
  26514. publisher = "nature.com",
  26515. volume = 16,
  26516. number = 7,
  26517. pages = "824--831",
  26518. month = jul,
  26519. year = 2013,
  26520. keywords = "Decision Making",
  26521. language = "en"
  26522. }
  26523. @ARTICLE{Shadlen2013-hz,
  26524. title = "Decision making as a window on cognition",
  26525. author = "Shadlen, Michael N and Kiani, Roozbeh",
  26526. abstract = "A decision is a commitment to a proposition or plan of action
  26527. based on information and values associated with the possible
  26528. outcomes. The process operates in a flexible timeframe that is
  26529. free from the immediacy of evidence acquisition and the real
  26530. time demands of action itself. Thus, it involves deliberation,
  26531. planning, and strategizing. This Perspective focuses on
  26532. perceptual decision making in nonhuman primates and the
  26533. discovery of neural mechanisms that support accuracy, speed, and
  26534. confidence in a decision. We suggest that these mechanisms
  26535. expose principles of cognitive function in general, and we
  26536. speculate about the challenges and directions before the field.",
  26537. journal = "Neuron",
  26538. publisher = "Elsevier",
  26539. volume = 80,
  26540. number = 3,
  26541. pages = "791--806",
  26542. month = oct,
  26543. year = 2013,
  26544. keywords = "Decision Making",
  26545. language = "en"
  26546. }
  26547. @ARTICLE{DeCharms2000-ng,
  26548. title = "Neural representation and the cortical code",
  26549. author = "deCharms, R C and Zador, A",
  26550. abstract = "The principle function of the central nervous system is to
  26551. represent and transform information and thereby mediate
  26552. appropriate decisions and behaviors. The cerebral cortex is one
  26553. of the primary seats of the internal representations maintained
  26554. and used in perception, memory, decision making, motor control,
  26555. and subjective experience, but the basic coding scheme by which
  26556. this information is carried and transformed by neurons is not
  26557. yet fully understood. This article defines and reviews how
  26558. information is represented in the firing rates and temporal
  26559. patterns of populations of cortical neurons, with a particular
  26560. emphasis on how this information mediates behavior and
  26561. experience.",
  26562. journal = "Annu. Rev. Neurosci.",
  26563. publisher = "annualreviews.org",
  26564. volume = 23,
  26565. pages = "613--647",
  26566. year = 2000,
  26567. keywords = "Theoretical",
  26568. language = "en"
  26569. }
  26570. @ARTICLE{Bastos2012-fc,
  26571. title = "Canonical microcircuits for predictive coding",
  26572. author = "Bastos, Andre M and Usrey, W Martin and Adams, Rick A and Mangun,
  26573. George R and Fries, Pascal and Friston, Karl J",
  26574. abstract = "This Perspective considers the influential notion of a canonical
  26575. (cortical) microcircuit in light of recent theories about
  26576. neuronal processing. Specifically, we conciliate quantitative
  26577. studies of microcircuitry and the functional logic of neuronal
  26578. computations. We revisit the established idea that message
  26579. passing among hierarchical cortical areas implements a form of
  26580. Bayesian inference-paying careful attention to the implications
  26581. for intrinsic connections among neuronal populations. By deriving
  26582. canonical forms for these computations, one can associate
  26583. specific neuronal populations with specific computational roles.
  26584. This analysis discloses a remarkable correspondence between the
  26585. microcircuitry of the cortical column and the connectivity
  26586. implied by predictive coding. Furthermore, it provides some
  26587. intuitive insights into the functional asymmetries between
  26588. feedforward and feedback connections and the characteristic
  26589. frequencies over which they operate.",
  26590. journal = "Neuron",
  26591. volume = 76,
  26592. number = 4,
  26593. pages = "695--711",
  26594. month = nov,
  26595. year = 2012,
  26596. keywords = "Theoretical",
  26597. language = "en"
  26598. }
  26599. @ARTICLE{Spratling2017-te,
  26600. title = "A review of predictive coding algorithms",
  26601. author = "Spratling, M W",
  26602. abstract = "Predictive coding is a leading theory of how the brain performs
  26603. probabilistic inference. However, there are a number of distinct
  26604. algorithms which are described by the term ``predictive coding''.
  26605. This article provides a concise review of these different
  26606. predictive coding algorithms, highlighting their similarities and
  26607. differences. Five algorithms are covered: linear predictive
  26608. coding which has a long and influential history in the signal
  26609. processing literature; the first neuroscience-related application
  26610. of predictive coding to explaining the function of the retina;
  26611. and three versions of predictive coding that have been proposed
  26612. to model cortical function. While all these algorithms aim to fit
  26613. a generative model to sensory data, they differ in the type of
  26614. generative model they employ, in the process used to optimise the
  26615. fit between the model and sensory data, and in the way that they
  26616. are related to neurobiology.",
  26617. journal = "Brain Cogn.",
  26618. volume = 112,
  26619. pages = "92--97",
  26620. month = mar,
  26621. year = 2017,
  26622. keywords = "Cortex; Free energy; Neural networks; Predictive coding; Retina;
  26623. Signal processing;Theoretical",
  26624. language = "en"
  26625. }
  26626. @UNPUBLISHED{Gomez-Marin2014-qb,
  26627. title = "Big Behavioral Data: Psychology, Ethology and the Foundations of
  26628. Neuroscience",
  26629. author = "Gomez-Marin, Alex and Paton, Joseph J and Kampff, Adam R and
  26630. Costa, Rui M and Mainen, Zachary M",
  26631. abstract = "Behavior is a unifying organismal process where genes, neural
  26632. function, anatomy and environment converge and interrelate. Here
  26633. we review the current state and discuss the future impact of
  26634. accelerating advances in technology for behavioral studies,
  26635. focusing on rodents as an exemplar. We frame our perspective in
  26636. three dimensions: degree of experimental constraint,
  26637. dimensionality of data, and level of description. We argue that
  26638. ``big behavioral data'' presents challenges proportionate to its
  26639. promise and describe how these challenges might be met through
  26640. opportunities afforded by the two rival conceptual legacies of
  26641. 20th century behavioral science, ethology and psychology. We
  26642. conclude that although ``more is not necessarily better'',
  26643. copious, quantitative and open behavioral data has the potential
  26644. to transform and unify these two disciplines and to solidify the
  26645. foundations of others, including neuroscience, but only if the
  26646. development of novel theoretical frameworks and improved
  26647. experimental designs matches the technological progress.",
  26648. journal = "bioRxiv",
  26649. pages = "006809",
  26650. month = jul,
  26651. year = 2014,
  26652. keywords = "Analysis/Modelling [Behaviour]",
  26653. language = "en"
  26654. }
  26655. @ARTICLE{Dudman2016-gy,
  26656. title = "The basal ganglia: from motor commands to the control of vigor",
  26657. author = "Dudman, Joshua T and Krakauer, John W",
  26658. abstract = "Vertebrates are remarkable for their ability to select and
  26659. execute goal-directed actions: motor skills critical for thriving
  26660. in complex, competitive environments. A key aspect of a motor
  26661. skill is the ability to execute its component movements over a
  26662. range of speeds, amplitudes and frequencies (vigor). Recent work
  26663. has indicated that a subcortical circuit, the basal ganglia, is a
  26664. critical determinant of movement vigor in rodents and primates.
  26665. We propose that the basal ganglia evolved from a circuit that in
  26666. lower vertebrates and some mammals is sufficient to directly
  26667. command simple or stereotyped movements to one that indirectly
  26668. controls the vigor of goal-directed movements. The implications
  26669. of a dual role of the basal ganglia in the control of vigor and
  26670. response to reward are also discussed.",
  26671. journal = "Curr. Opin. Neurobiol.",
  26672. volume = 37,
  26673. pages = "158--166",
  26674. month = apr,
  26675. year = 2016,
  26676. language = "en"
  26677. }
  26678. @ARTICLE{Brody2016-ud,
  26679. title = "Neural underpinnings of the evidence accumulator",
  26680. author = "Brody, Carlos D and Hanks, Timothy D",
  26681. abstract = "Gradual accumulation of evidence favoring one or another choice
  26682. is considered a core component of many different types of
  26683. decisions, and has been the subject of many neurophysiological
  26684. studies in non-human primates. But its neural circuit mechanisms
  26685. remain mysterious. Investigating it in rodents has recently
  26686. become possible, facilitating perturbation experiments to
  26687. delineate the relevant causal circuit, as well as the application
  26688. of other tools more readily available in rodents. In addition,
  26689. advances in stimulus design and analysis have aided studying the
  26690. relevant neural encoding. In complement to ongoing non-human
  26691. primate studies, these newly available model systems and tools
  26692. place the field at an exciting time that suggests that the
  26693. dynamical circuit mechanisms underlying accumulation of evidence
  26694. could soon be revealed.",
  26695. journal = "Curr. Opin. Neurobiol.",
  26696. volume = 37,
  26697. pages = "149--157",
  26698. month = apr,
  26699. year = 2016,
  26700. language = "en"
  26701. }
  26702. @ARTICLE{Fetsch2016-zu,
  26703. title = "The importance of task design and behavioral control for
  26704. understanding the neural basis of cognitive functions",
  26705. author = "Fetsch, Christopher R",
  26706. abstract = "The success of systems neuroscience depends on the ability to
  26707. forge quantitative links between neural activity and behavior.
  26708. Traditionally, this process has benefited from the rigorous
  26709. development and testing of hypotheses using tools derived from
  26710. classical psychophysics and computational motor control. As our
  26711. capacity for measuring neural activity improves, accompanied by
  26712. powerful new analysis strategies, it seems prudent to remember
  26713. what these traditional approaches have to offer. Here I present a
  26714. perspective on the merits of principled task design and tight
  26715. behavioral control, along with some words of caution about
  26716. interpretation in unguided, large-scale neural recording studies.
  26717. I argue that a judicious combination of new and old approaches is
  26718. the best way to advance our understanding of higher brain
  26719. function in health and disease.",
  26720. journal = "Curr. Opin. Neurobiol.",
  26721. volume = 37,
  26722. pages = "16--22",
  26723. month = apr,
  26724. year = 2016,
  26725. keywords = "Theoretical",
  26726. language = "en"
  26727. }
  26728. @ARTICLE{Jovanic2016-ok,
  26729. title = "Competitive Disinhibition Mediates Behavioral Choice and
  26730. Sequences in Drosophila",
  26731. author = "Jovanic, Tihana and Schneider-Mizell, Casey Martin and Shao, Mei
  26732. and Masson, Jean-Baptiste and Denisov, Gennady and Fetter,
  26733. Richard Doty and Mensh, Brett Daren and Truman, James William and
  26734. Cardona, Albert and Zlatic, Marta",
  26735. abstract = "Even a simple sensory stimulus can elicit distinct innate
  26736. behaviors and sequences. During sensorimotor decisions,
  26737. competitive interactions among neurons that promote distinct
  26738. behaviors must ensure the selection and maintenance of one
  26739. behavior, while suppressing others. The circuit implementation of
  26740. these competitive interactions is still an open question. By
  26741. combining comprehensive electron microscopy reconstruction of
  26742. inhibitory interneuron networks, modeling, electrophysiology, and
  26743. behavioral studies, we determined the circuit mechanisms that
  26744. contribute to the Drosophila larval sensorimotor decision to
  26745. startle, explore, or perform a sequence of the two in response to
  26746. a mechanosensory stimulus. Together, these studies reveal that,
  26747. early in sensory processing, (1) reciprocally connected
  26748. feedforward inhibitory interneurons implement behavioral choice,
  26749. (2) local feedback disinhibition provides positive feedback that
  26750. consolidates and maintains the chosen behavior, and (3) lateral
  26751. disinhibition promotes sequence transitions. The combination of
  26752. these interconnected circuit motifs can implement both behavior
  26753. selection and the serial organization of behaviors into a
  26754. sequence.",
  26755. journal = "Cell",
  26756. volume = 167,
  26757. number = 3,
  26758. pages = "858--870.e19",
  26759. month = oct,
  26760. year = 2016,
  26761. keywords = "Drosophila; EM connectome; behavioral choice; behavioral
  26762. sequences; disinihibition; recurrent inhibition; sensory
  26763. processing",
  26764. language = "en"
  26765. }
  26766. @ARTICLE{Stephens2011-ce,
  26767. title = "Searching for simplicity in the analysis of neurons and behavior",
  26768. author = "Stephens, Greg J and Osborne, Leslie C and Bialek, William",
  26769. abstract = "What fascinates us about animal behavior is its richness and
  26770. complexity, but understanding behavior and its neural basis
  26771. requires a simpler description. Traditionally, simplification has
  26772. been imposed by training animals to engage in a limited set of
  26773. behaviors, by hand scoring behaviors into discrete classes, or by
  26774. limiting the sensory experience of the organism. An alternative
  26775. is to ask whether we can search through the dynamics of natural
  26776. behaviors to find explicit evidence that these behaviors are
  26777. simpler than they might have been. We review two mathematical
  26778. approaches to simplification, dimensionality reduction and the
  26779. maximum entropy method, and we draw on examples from different
  26780. levels of biological organization, from the crawling behavior of
  26781. Caenorhabditis elegans to the control of smooth pursuit eye
  26782. movements in primates, and from the coding of natural scenes by
  26783. networks of neurons in the retina to the rules of English
  26784. spelling. In each case, we argue that the explicit search for
  26785. simplicity uncovers new and unexpected features of the biological
  26786. system and that the evidence for simplification gives us a
  26787. language with which to phrase new questions for the next
  26788. generation of experiments. The fact that similar mathematical
  26789. structures succeed in taming the complexity of very different
  26790. biological systems hints that there is something more general to
  26791. be discovered.",
  26792. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  26793. volume = "108 Suppl 3",
  26794. pages = "15565--15571",
  26795. month = sep,
  26796. year = 2011,
  26797. keywords = "Analysis/Modelling [Behaviour];Theoretical",
  26798. language = "en"
  26799. }
  26800. @ARTICLE{Stephens2008-vt,
  26801. title = "Dimensionality and dynamics in the behavior of C. elegans",
  26802. author = "Stephens, Greg J and Johnson-Kerner, Bethany and Bialek, William
  26803. and Ryu, William S",
  26804. abstract = "A major challenge in analyzing animal behavior is to discover
  26805. some underlying simplicity in complex motor actions. Here, we
  26806. show that the space of shapes adopted by the nematode
  26807. Caenorhabditis elegans is low dimensional, with just four
  26808. dimensions accounting for 95\% of the shape variance. These
  26809. dimensions provide a quantitative description of worm behavior,
  26810. and we partially reconstruct ``equations of motion'' for the
  26811. dynamics in this space. These dynamics have multiple attractors,
  26812. and we find that the worm visits these in a rapid and almost
  26813. completely deterministic response to weak thermal stimuli.
  26814. Stimulus-dependent correlations among the different modes suggest
  26815. that one can generate more reliable behaviors by synchronizing
  26816. stimuli to the state of the worm in shape space. We confirm this
  26817. prediction, effectively ``steering'' the worm in real time.",
  26818. journal = "PLoS Comput. Biol.",
  26819. volume = 4,
  26820. number = 4,
  26821. pages = "e1000028",
  26822. month = apr,
  26823. year = 2008,
  26824. keywords = "Analysis/Modelling [Behaviour]",
  26825. language = "en"
  26826. }
  26827. @ARTICLE{Poehlmann2018-yj,
  26828. title = "A unifying model to predict multiple object orienting behaviors
  26829. in tethered flies",
  26830. author = "Poehlmann, A and Soselisa, S and Fenk, L and Straw, A",
  26831. journal = "BiorXiv",
  26832. year = 2018,
  26833. keywords = "Analysis/Modelling [Behaviour]"
  26834. }
  26835. @ARTICLE{Todd2017-xy,
  26836. title = "Systematic exploration of unsupervised methods for mapping
  26837. behavior",
  26838. author = "Todd, Jeremy G and Kain, Jamey S and de Bivort, Benjamin L",
  26839. abstract = "To fully understand the mechanisms giving rise to behavior, we
  26840. need to be able to precisely measure it. When coupled with large
  26841. behavioral data sets, unsupervised clustering methods offer the
  26842. potential of unbiased mapping of behavioral spaces. However,
  26843. unsupervised techniques to map behavioral spaces are in their
  26844. infancy, and there have been few systematic considerations of all
  26845. the methodological options. We compared the performance of seven
  26846. distinct mapping methods in clustering a wavelet-transformed data
  26847. set consisting of the x- and y-positions of the six legs of
  26848. individual flies. Legs were automatically tracked by small pieces
  26849. of fluorescent dye, while the fly was tethered and walking on an
  26850. air-suspended ball. We find that there is considerable variation
  26851. in the performance of these mapping methods, and that better
  26852. performance is attained when clustering is done in higher
  26853. dimensional spaces (which are otherwise less preferable because
  26854. they are hard to visualize). High dimensionality means that some
  26855. algorithms, including the non-parametric watershed cluster
  26856. assignment algorithm, cannot be used. We developed an alternative
  26857. watershed algorithm which can be used in high-dimensional spaces
  26858. when a probability density estimate can be computed directly.
  26859. With these tools in hand, we examined the behavioral space of fly
  26860. leg postural dynamics and locomotion. We find a striking division
  26861. of behavior into modes involving the fore legs and modes
  26862. involving the hind legs, with few direct transitions between
  26863. them. By computing behavioral clusters using the data from all
  26864. flies simultaneously, we show that this division appears to be
  26865. common to all flies. We also identify individual-to-individual
  26866. differences in behavior and behavioral transitions. Lastly, we
  26867. suggest a computational pipeline that can achieve satisfactory
  26868. levels of performance without the taxing computational demands of
  26869. a systematic combinatorial approach.",
  26870. journal = "Phys. Biol.",
  26871. volume = 14,
  26872. number = 1,
  26873. pages = "015002",
  26874. month = feb,
  26875. year = 2017,
  26876. keywords = "Analysis/Modelling [Behaviour]",
  26877. language = "en"
  26878. }
  26879. % The entry below contains non-ASCII chars that could not be converted
  26880. % to a LaTeX equivalent.
  26881. @ARTICLE{Miller2018-vn,
  26882. title = "Retrosplenial cortical representations of space and future goal
  26883. locations develop with learning",
  26884. author = "Miller, A M P and Mau, W and Smith, D M",
  26885. abstract = "The retrosplenial cortex (RSC) is important for long-term
  26886. contextual memory and spatial navigation, but little is known
  26887. about how RSC neural representations develop with experience. We
  26888. recorded neuronal activity in the RSC of rats as they learned a
  26889. continuous spatial alternation task and found that the RSC
  26890. slowly developed a population-level representation of the rat9s
  26891. spatial location and current trajectory to the goal. After the
  26892. rats reached peak performance, RSC firing patterns became
  26893. predictive of navigation accuracy …",
  26894. journal = "bioRxiv",
  26895. publisher = "biorxiv.org",
  26896. year = 2018,
  26897. keywords = "Spatial Navigation"
  26898. }
  26899. @ARTICLE{Krakauer2017-va,
  26900. title = "Neuroscience Needs Behavior: Correcting a Reductionist Bias",
  26901. author = "Krakauer, John W and Ghazanfar, Asif A and Gomez-Marin, Alex and
  26902. MacIver, Malcolm A and Poeppel, David",
  26903. abstract = "There are ever more compelling tools available for neuroscience
  26904. research, ranging from selective genetic targeting to optogenetic
  26905. circuit control to mapping whole connectomes. These approaches
  26906. are coupled with a deep-seated, often tacit, belief in the
  26907. reductionist program for understanding the link between the brain
  26908. and behavior. The aim of this program is causal explanation
  26909. through neural manipulations that allow testing of necessity and
  26910. sufficiency claims. We argue, however, that another equally
  26911. important approach seeks an alternative form of understanding
  26912. through careful theoretical and experimental decomposition of
  26913. behavior. Specifically, the detailed analysis of tasks and of the
  26914. behavior they elicit is best suited for discovering component
  26915. processes and their underlying algorithms. In most cases, we
  26916. argue that study of the neural implementation of behavior is best
  26917. investigated after such behavioral work. Thus, we advocate a more
  26918. pluralistic notion of neuroscience when it comes to the
  26919. brain-behavior relationship: behavioral work provides
  26920. understanding, whereas neural interventions test causality.",
  26921. journal = "Neuron",
  26922. volume = 93,
  26923. number = 3,
  26924. pages = "480--490",
  26925. month = feb,
  26926. year = 2017,
  26927. keywords = "Theoretical",
  26928. language = "en"
  26929. }
  26930. @ARTICLE{Egnor2016-ix,
  26931. title = "Computational Analysis of Behavior",
  26932. author = "Egnor, S E Roian and Branson, Kristin",
  26933. abstract = "In this review, we discuss the emerging field of computational
  26934. behavioral analysis-the use of modern methods from computer
  26935. science and engineering to quantitatively measure animal
  26936. behavior. We discuss aspects of experiment design important to
  26937. both obtaining biologically relevant behavioral data and enabling
  26938. the use of machine vision and learning techniques for automation.
  26939. These two goals are often in conflict. Restraining or restricting
  26940. the environment of the animal can simplify automatic behavior
  26941. quantification, but it can also degrade the quality or alter
  26942. important aspects of behavior. To enable biologists to design
  26943. experiments to obtain better behavioral measurements, and
  26944. computer scientists to pinpoint fruitful directions for algorithm
  26945. improvement, we review known effects of artificial manipulation
  26946. of the animal on behavior. We also review machine vision and
  26947. learning techniques for tracking, feature extraction, automated
  26948. behavior classification, and automated behavior discovery, the
  26949. assumptions they make, and the types of data they work best with.",
  26950. journal = "Annu. Rev. Neurosci.",
  26951. volume = 39,
  26952. pages = "217--236",
  26953. month = jul,
  26954. year = 2016,
  26955. keywords = "animal behavior; automated behavioral analysis; computer vision;
  26956. machine learning; tracking;Analysis/Modelling [Behaviour]",
  26957. language = "en"
  26958. }
  26959. @UNPUBLISHED{Goode2018-px,
  26960. title = "Bed nucleus of the stria terminalis mediates fear to ambiguous
  26961. threat signals",
  26962. author = "Goode, Travis D and Ressler, Reed L and Acca, Gillian M and
  26963. Maren, Stephen",
  26964. abstract = "The bed nucleus of the stria terminalis (BNST) has been
  26965. implicated in fear and anxiety, but the specific factors that
  26966. engage the BNST in defensive behavior are unclear. Here we
  26967. explore the possibility that ambiguous threats recruit the BNST
  26968. during Pavlovian fear conditioning in rats. We arranged a
  26969. conditioned stimulus (CS) to either precede or follow an aversive
  26970. unconditioned stimulus (US), a procedure that established
  26971. reliable (forward) or ambiguous (backward) signals for US onset.
  26972. After conditioning, reversible inactivation of the BNST
  26973. selectively reduced freezing to the backward CS; BNST
  26974. inactivation did not affect freezing to the forward CS even when
  26975. that CS predicted a variable magnitude US. Backward CSs increased
  26976. Fos in the ventral BNST and in BNST-projecting neurons in the
  26977. infralimbic cortex, but not the hippocampus or amygdala. These
  26978. data reveal that BNST circuits process ambiguous threat signals
  26979. central to the etiology and expression of anxiety.",
  26980. journal = "bioRxiv",
  26981. pages = "376228",
  26982. month = jul,
  26983. year = 2018,
  26984. keywords = "Threat response",
  26985. language = "en"
  26986. }
  26987. @ARTICLE{Atiya_undated-kw,
  26988. title = "Neural Circuit Mechanism of Decision Uncertainty and
  26989. {Change-of-Mind}",
  26990. author = "Atiya, N and Ra{\~n}{\'o}, I and Prasad, G and Wong-Lin, K",
  26991. journal = "BiorXiv",
  26992. keywords = "Decision Making;Theoretical"
  26993. }
  26994. @ARTICLE{Clark_undated-jz,
  26995. title = "Identifying the cognitive processes underpinning
  26996. hippocampal-dependent tasks",
  26997. author = "Clark, I and Hotchin, V and Monk, A and Pizzamiglio, G and
  26998. Liefgreen, A and Maguire, E",
  26999. keywords = "Spatial Navigation"
  27000. }
  27001. @ARTICLE{Tingley2018-yq,
  27002. title = "Transformation of a Spatial Map across the {Hippocampal-Lateral}
  27003. Septal Circuit",
  27004. author = "Tingley, David and Buzs{\'a}ki, Gy{\"o}rgy",
  27005. abstract = "The hippocampus constructs a map of the environment. How this
  27006. ``cognitive map'' is utilized by other brain regions to guide
  27007. behavior remains unexplored. To examine how neuronal firing
  27008. patterns in the hippocampus are transmitted and transformed, we
  27009. recorded neurons in its principal subcortical target, the lateral
  27010. septum (LS). We observed that LS neurons carry reliable spatial
  27011. information in the phase of action potentials, relative to
  27012. hippocampal theta oscillations, while the firing rates of LS
  27013. neurons remained uninformative. Furthermore, this spatial phase
  27014. code had an anatomical microstructure within the LS and was bound
  27015. to the hippocampal spatial code by synchronous gamma frequency
  27016. cell assemblies. Using a data-driven model, we show that
  27017. rate-independent spatial tuning arises through the dynamic
  27018. weighting of CA1 and CA3 cell assemblies. Our findings
  27019. demonstrate that transformation of the hippocampal spatial map
  27020. depends on higher-order theta-dependent neuronal sequences. VIDEO
  27021. ABSTRACT.",
  27022. journal = "Neuron",
  27023. volume = 98,
  27024. number = 6,
  27025. pages = "1229--1242.e5",
  27026. month = jun,
  27027. year = 2018,
  27028. keywords = "cell assemblies; dynamic weighting; hippocampus; information
  27029. transfer; lateral septum; phase coding; rate coding; theta
  27030. sequences; transformation;Spatial Navigation",
  27031. language = "en"
  27032. }
  27033. @ARTICLE{Akam2015-tt,
  27034. title = "Simple Plans or Sophisticated Habits? State, Transition and
  27035. Learning Interactions in the {Two-Step} Task",
  27036. author = "Akam, Thomas and Costa, Rui and Dayan, Peter",
  27037. abstract = "The recently developed 'two-step' behavioural task promises to
  27038. differentiate model-based from model-free reinforcement learning,
  27039. while generating neurophysiologically-friendly decision datasets
  27040. with parametric variation of decision variables. These desirable
  27041. features have prompted its widespread adoption. Here, we analyse
  27042. the interactions between a range of different strategies and the
  27043. structure of transitions and outcomes in order to examine
  27044. constraints on what can be learned from behavioural performance.
  27045. The task involves a trade-off between the need for stochasticity,
  27046. to allow strategies to be discriminated, and a need for
  27047. determinism, so that it is worth subjects' investment of effort
  27048. to exploit the contingencies optimally. We show through
  27049. simulation that under certain conditions model-free strategies
  27050. can masquerade as being model-based. We first show that seemingly
  27051. innocuous modifications to the task structure can induce
  27052. correlations between action values at the start of the trial and
  27053. the subsequent trial events in such a way that analysis based on
  27054. comparing successive trials can lead to erroneous conclusions. We
  27055. confirm the power of a suggested correction to the analysis that
  27056. can alleviate this problem. We then consider model-free
  27057. reinforcement learning strategies that exploit correlations
  27058. between where rewards are obtained and which actions have high
  27059. expected value. These generate behaviour that appears model-based
  27060. under these, and also more sophisticated, analyses. Exploiting
  27061. the full potential of the two-step task as a tool for behavioural
  27062. neuroscience requires an understanding of these issues.",
  27063. journal = "PLoS Comput. Biol.",
  27064. volume = 11,
  27065. number = 12,
  27066. pages = "e1004648",
  27067. month = dec,
  27068. year = 2015,
  27069. language = "en"
  27070. }
  27071. @ARTICLE{Markowitz2018-fk,
  27072. title = "The Striatum Organizes {3D} Behavior via {Moment-to-Moment}
  27073. Action Selection",
  27074. author = "Markowitz, Jeffrey E and Gillis, Winthrop F and Beron, Celia C
  27075. and Neufeld, Shay Q and Robertson, Keiramarie and Bhagat, Neha D
  27076. and Peterson, Ralph E and Peterson, Emalee and Hyun, Minsuk and
  27077. Linderman, Scott W and Sabatini, Bernardo L and Datta, Sandeep
  27078. Robert",
  27079. abstract = "Many naturalistic behaviors are built from modular components
  27080. that are expressed sequentially. Although striatal circuits have
  27081. been implicated in action selection and implementation, the
  27082. neural mechanisms that compose behavior in unrestrained animals
  27083. are not well understood. Here, we record bulk and cellular neural
  27084. activity in the direct and indirect pathways of dorsolateral
  27085. striatum (DLS) as mice spontaneously express action sequences.
  27086. These experiments reveal that DLS neurons systematically encode
  27087. information about the identity and ordering of sub-second 3D
  27088. behavioral motifs; this encoding is facilitated by fast-timescale
  27089. decorrelations between the direct and indirect pathways.
  27090. Furthermore, lesioning the DLS prevents appropriate sequence
  27091. assembly during exploratory or odor-evoked behaviors. By
  27092. characterizing naturalistic behavior at neural timescales, these
  27093. experiments identify a code for elemental 3D pose dynamics built
  27094. from complementary pathway dynamics, support a role for DLS in
  27095. constructing meaningful behavioral sequences, and suggest models
  27096. for how actions are sculpted over time.",
  27097. journal = "Cell",
  27098. volume = 174,
  27099. number = 1,
  27100. pages = "44--58.e17",
  27101. month = jun,
  27102. year = 2018,
  27103. keywords = "basal ganglia; behavior; coding; direct pathway; ethology;
  27104. indirect pathway; machine learning; mouse; photometry; striatum",
  27105. language = "en"
  27106. }
  27107. @ARTICLE{Wiltschko2015-wd,
  27108. title = "Mapping {Sub-Second} Structure in Mouse Behavior",
  27109. author = "Wiltschko, Alexander B and Johnson, Matthew J and Iurilli,
  27110. Giuliano and Peterson, Ralph E and Katon, Jesse M and
  27111. Pashkovski, Stan L and Abraira, Victoria E and Adams, Ryan P and
  27112. Datta, Sandeep Robert",
  27113. abstract = "Complex animal behaviors are likely built from simpler modules,
  27114. but their systematic identification in mammals remains a
  27115. significant challenge. Here we use depth imaging to show that 3D
  27116. mouse pose dynamics are structured at the sub-second timescale.
  27117. Computational modeling of these fast dynamics effectively
  27118. describes mouse behavior as a series of reused and stereotyped
  27119. modules with defined transition probabilities. We demonstrate
  27120. this combined 3D imaging and machine learning method can be used
  27121. to unmask potential strategies employed by the brain to adapt to
  27122. the environment, to capture both predicted and previously hidden
  27123. phenotypes caused by genetic or neural manipulations, and to
  27124. systematically expose the global structure of behavior within an
  27125. experiment. This work reveals that mouse body language is built
  27126. from identifiable components and is organized in a predictable
  27127. fashion; deciphering this language establishes an objective
  27128. framework for characterizing the influence of environmental
  27129. cues, genes and neural activity on behavior.",
  27130. journal = "Neuron",
  27131. publisher = "Elsevier",
  27132. volume = 88,
  27133. number = 6,
  27134. pages = "1121--1135",
  27135. month = dec,
  27136. year = 2015,
  27137. language = "en"
  27138. }
  27139. @ARTICLE{Friston2018-lp,
  27140. title = "Does predictive coding have a future?",
  27141. author = "Friston, Karl",
  27142. journal = "Nat. Neurosci.",
  27143. month = jul,
  27144. year = 2018,
  27145. keywords = "Threat response",
  27146. language = "en"
  27147. }
  27148. @ARTICLE{Campbell2018-tx,
  27149. title = "Principles governing the integration of landmark and self-motion
  27150. cues in entorhinal cortical codes for navigation",
  27151. author = "Campbell, Malcolm G and Ocko, Samuel A and Mallory, Caitlin S and
  27152. Low, Isabel I C and Ganguli, Surya and Giocomo, Lisa M",
  27153. abstract = "To guide navigation, the nervous system integrates multisensory
  27154. self-motion and landmark information. We dissected how these
  27155. inputs generate spatial representations by recording entorhinal
  27156. grid, border and speed cells in mice navigating virtual
  27157. environments. Manipulating the gain between the animal's
  27158. locomotion and the visual scene revealed that border cells
  27159. responded to landmark cues while grid and speed cells responded
  27160. to combinations of locomotion, optic flow and landmark cues in a
  27161. context-dependent manner, with optic flow becoming more
  27162. influential when it was faster than expected. A network model
  27163. explained these results by revealing a phase transition between
  27164. two regimes in which grid cells remain coherent with or break
  27165. away from the landmark reference frame. Moreover, during
  27166. path-integration-based navigation, mice estimated their position
  27167. following principles predicted by our recordings. Together, these
  27168. results provide a theoretical framework for understanding how
  27169. landmark and self-motion cues combine during navigation to
  27170. generate spatial representations and guide behavior.",
  27171. journal = "Nat. Neurosci.",
  27172. month = jul,
  27173. year = 2018,
  27174. keywords = "Spatial Navigation",
  27175. language = "en"
  27176. }
  27177. @ARTICLE{Branco2009-gn,
  27178. title = "The probability of neurotransmitter release: variability and
  27179. feedback control at single synapses",
  27180. author = "Branco, Tiago and Staras, Kevin",
  27181. abstract = "Information transfer at chemical synapses occurs when vesicles
  27182. fuse with the plasma membrane and release neurotransmitter. This
  27183. process is stochastic and its likelihood of occurrence is a
  27184. crucial factor in the regulation of signal propagation in
  27185. neuronal networks. The reliability of neurotransmitter release
  27186. can be highly variable: experimental data from
  27187. electrophysiological, molecular and imaging studies have
  27188. demonstrated that synaptic terminals can individually set their
  27189. neurotransmitter release probability dynamically through local
  27190. feedback regulation. This local tuning of transmission has
  27191. important implications for current models of single-neuron
  27192. computation.",
  27193. journal = "Nat. Rev. Neurosci.",
  27194. publisher = "nature.com",
  27195. volume = 10,
  27196. number = 5,
  27197. pages = "373--383",
  27198. month = may,
  27199. year = 2009,
  27200. language = "en"
  27201. }
  27202. @ARTICLE{Wood2018-fu,
  27203. title = "The honeycomb maze provides a novel test to study
  27204. hippocampal-dependent spatial navigation",
  27205. author = "Wood, Ruth A and Bauza, Marius and Krupic, Julija and Burton,
  27206. Stephen and Delekate, Andrea and Chan, Dennis and O'Keefe, John",
  27207. abstract = "Here we describe the honeycomb maze, a behavioural paradigm for
  27208. the study of spatial navigation in rats. The maze consists of 37
  27209. platforms that can be raised or lowered independently. Place
  27210. navigation requires an animal to go to a goal platform from any
  27211. of several start platforms via a series of sequential choices.
  27212. For each, the animal is confined to a raised platform and
  27213. allowed to choose between two of the six adjacent platforms, the
  27214. correct one being the platform with the smallest angle to the
  27215. goal-heading direction. Rats learn rapidly and their choices are
  27216. influenced by three factors: the angle between the two choice
  27217. platforms, the distance from the goal, and the angle between the
  27218. correct platform and the direction of the goal. Rats with
  27219. hippocampal damage are impaired in learning and their
  27220. performance is affected by all three factors. The honeycomb maze
  27221. represents a marked improvement over current spatial navigation
  27222. tests, such as the Morris water maze, because it controls the
  27223. choices of the animal at each point in the maze, provides the
  27224. ability to assess knowledge of the goal direction from any
  27225. location, enables the identification of factors influencing task
  27226. performance and provides the possibility for concomitant
  27227. single-cell recording.",
  27228. journal = "Nature",
  27229. publisher = "nature.com",
  27230. volume = 554,
  27231. number = 7690,
  27232. pages = "102--105",
  27233. month = feb,
  27234. year = 2018,
  27235. keywords = "Spatial Navigation",
  27236. language = "en"
  27237. }
  27238. % The entry below contains non-ASCII chars that could not be converted
  27239. % to a LaTeX equivalent.
  27240. @ARTICLE{Tolman1939-vg,
  27241. title = "Prediction of vicarious trial and error by means of the
  27242. schematic sowbug",
  27243. author = "Tolman, Edward Chace",
  27244. abstract = "An experiment is described in white-black and white-gray
  27245. discrimination by rats, in which considerable``
  27246. looking-back-and-forth'' or vicarious trial -and- error
  27247. (Muenzinger's terminology), hereafter called VTE behavior,
  27248. occurred. 3 groups of 10 rats each were used …",
  27249. journal = "Psychol. Rev.",
  27250. publisher = "American Psychological Association",
  27251. volume = 46,
  27252. number = 4,
  27253. pages = "318",
  27254. year = 1939,
  27255. keywords = "Decision Making"
  27256. }
  27257. @ARTICLE{Schmidt2013-bd,
  27258. title = "Conflict between place and response navigation strategies:
  27259. effects on vicarious trial and error ({VTE}) behaviors",
  27260. author = "Schmidt, Brandy and Papale, Andrew and Redish, A David and
  27261. Markus, Etan J",
  27262. abstract = "Navigation can be accomplished through multiple decision-making
  27263. strategies, using different information-processing computations.
  27264. A well-studied dichotomy in these decision-making strategies
  27265. compares hippocampal-dependent ``place'' and dorsal-lateral
  27266. striatal-dependent ``response'' strategies. A place strategy
  27267. depends on the ability to flexibly respond to environmental
  27268. cues, while a response strategy depends on the ability to
  27269. quickly recognize and react to situations with well-learned
  27270. action-outcome relationships. When rats reach decision points,
  27271. they sometimes pause and orient toward the potential routes of
  27272. travel, a process termed vicarious trial and error (VTE). VTE
  27273. co-occurs with neurophysiological information processing,
  27274. including sweeps of representation ahead of the animal in the
  27275. hippocampus and transient representations of reward in the
  27276. ventral striatum and orbitofrontal cortex. To examine the
  27277. relationship between VTE and the place/response strategy
  27278. dichotomy, we analyzed data in which rats were cued to switch
  27279. between place and response strategies on a plus maze. The
  27280. configuration of the maze allowed for place and response
  27281. strategies to work competitively or cooperatively. Animals
  27282. showed increased VTE on trials entailing competition between
  27283. navigational systems, linking VTE with deliberative
  27284. decision-making. Even in a well-learned task, VTE was
  27285. preferentially exhibited when a spatial selection was required,
  27286. further linking VTE behavior with decision-making associated
  27287. with hippocampal processing.",
  27288. journal = "Learn. Mem.",
  27289. publisher = "learnmem.cshlp.org",
  27290. volume = 20,
  27291. number = 3,
  27292. pages = "130--138",
  27293. month = feb,
  27294. year = 2013,
  27295. keywords = "Decision Making",
  27296. language = "en"
  27297. }
  27298. % The entry below contains non-ASCII chars that could not be converted
  27299. % to a LaTeX equivalent.
  27300. @ARTICLE{Goss1956-dh,
  27301. title = "Vicarious trial and error and related behavior",
  27302. author = "Goss, A E and Wischner, G J",
  27303. abstract = "Empirical material relating to`` vicarious trial -and- error
  27304. ''(VTE) is summarized and evaluated critically in terms of
  27305. criteria for VTE, of antecedents to and response correlates of
  27306. VTE, and of VTE and learning efficiency. It is proposed that the
  27307. criterion for scoring VTE behavior should …",
  27308. journal = "Psychol. Bull.",
  27309. publisher = "psycnet.apa.org",
  27310. volume = 53,
  27311. number = 1,
  27312. pages = "35--54",
  27313. month = jan,
  27314. year = 1956,
  27315. keywords = "LEARNING;Decision Making",
  27316. language = "en"
  27317. }
  27318. @ARTICLE{Redish2016-id,
  27319. title = "Vicarious trial and error",
  27320. author = "Redish, A David",
  27321. abstract = "When rats come to a decision point, they sometimes pause and
  27322. look back and forth as if deliberating over the choice; at other
  27323. times, they proceed as if they have already made their decision.
  27324. In the 1930s, this pause-and-look behaviour was termed
  27325. 'vicarious trial and error' (VTE), with the implication that the
  27326. rat was 'thinking about the future'. The discovery in 2007 that
  27327. the firing of hippocampal place cells gives rise to alternating
  27328. representations of each of the potential path options in a
  27329. serial manner during VTE suggested a possible neural mechanism
  27330. that could underlie the representations of future outcomes.
  27331. More-recent experiments examining VTE in rats suggest that there
  27332. are direct parallels to human processes of deliberative decision
  27333. making, working memory and mental time travel.",
  27334. journal = "Nat. Rev. Neurosci.",
  27335. publisher = "nature.com",
  27336. volume = 17,
  27337. number = 3,
  27338. pages = "147--159",
  27339. month = mar,
  27340. year = 2016,
  27341. keywords = "Decision Making",
  27342. language = "en"
  27343. }
  27344. @ARTICLE{Amsel1993-qk,
  27345. title = "Hippocampal function in the rat: cognitive mapping or vicarious
  27346. trial and error?",
  27347. author = "Amsel, A",
  27348. abstract = "The most prominent hypothesis of hippocampal function likens the
  27349. hippocampus to a ``cognitive map,'' a term used by a famous
  27350. learning theorist, E. C. Tolman, to explain maze learning. The
  27351. usual application of this concept of cognitive map, as it
  27352. applies to the hippocampus, is to what is called spatial
  27353. learning, mainly in the radial-arm maze of Olton and the Morris
  27354. water maze. In a recent Hippocampus Forum, evidence for the
  27355. cognitive map hypothesis was reviewed in a lead article by
  27356. Nadel, followed by a series of commentaries by leading
  27357. investigators of hippocampal function. This speculative
  27358. commentary offers an alternative not represented in the
  27359. forum--that the function of the hippocampus in spatial learning
  27360. is not as a cognitive map, but that it subserves another
  27361. function proposed by Tolman in his work on simple discrimination
  27362. learning, vicarious trial and error, based on incipient,
  27363. conflicting dispositions to approach and avoid.",
  27364. journal = "Hippocampus",
  27365. publisher = "Wiley Online Library",
  27366. volume = 3,
  27367. number = 3,
  27368. pages = "251--256",
  27369. month = jul,
  27370. year = 1993,
  27371. keywords = "Decision Making",
  27372. language = "en"
  27373. }
  27374. @ARTICLE{Hu1995-ip,
  27375. title = "A simple test of the vicarious trial-and-error hypothesis of
  27376. hippocampal function",
  27377. author = "Hu, D and Amsel, A",
  27378. abstract = "Vicarious trial-and-error (VTE) is a term that Muenzinger and
  27379. Tolman used to describe the rat's conflict-like behavior before
  27380. responding to choice. Recently, VTE was proposed as a mechanism
  27381. alternative to the concept of ``cognitive map'' in accounts of
  27382. hippocampal function. That is, many phenomena of impaired
  27383. learning and memory related to hippocampal interventions may be
  27384. explained by behavioral first principles: reduced conflicting,
  27385. incipient, pre-choice tendencies to approach and avoid. The
  27386. nonspatial black-white discrimination learning and VTE behavior
  27387. of the rat were investigated. Hippocampal-lesioned and
  27388. sham-lesioned animals were trained for 25 days (20 trials per
  27389. day) starting at 60 days of age. Each movement of the head from
  27390. one discriminative stimulus to the other was counted as a VTE
  27391. instance. Lesioned rats had fewer VTEs than sham controls, and
  27392. the former learned much more slowly or never learned. After
  27393. learning, VTE frequency declined. Male and female rats showed no
  27394. significant differences in VTE behavior or discrimination
  27395. learning.",
  27396. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  27397. publisher = "National Acad Sciences",
  27398. volume = 92,
  27399. number = 12,
  27400. pages = "5506--5509",
  27401. month = jun,
  27402. year = 1995,
  27403. keywords = "Decision Making",
  27404. language = "en"
  27405. }
  27406. @ARTICLE{Muenzinger1938-ex,
  27407. title = "Vicarious Trial and Error at a Point of Choice: I. A General
  27408. Survey of its Relation to Learning Efficiency",
  27409. author = "Muenzinger, Karl F",
  27410. abstract = "* Received in the Editorial Office on December 7, 1937.",
  27411. journal = "The Pedagogical Seminary and Journal of Genetic Psychology",
  27412. publisher = "Routledge",
  27413. volume = 53,
  27414. number = 1,
  27415. pages = "75--86",
  27416. month = sep,
  27417. year = 1938,
  27418. keywords = "Decision Making"
  27419. }
  27420. @ARTICLE{De_Franceschi2016-vo,
  27421. title = "Vision Guides Selection of Freeze or Flight Defense Strategies
  27422. in Mice",
  27423. author = "De Franceschi, Gioia and Vivattanasarn, Tipok and Saleem, Aman B
  27424. and Solomon, Samuel G",
  27425. abstract = "In prey species such as mice, avoidance of predators is key to
  27426. survival and drives instinctual behaviors like freeze or flight
  27427. [1, 2]. Sensory signals guide the selection of appropriate
  27428. behavior [3], and for aerial predators only vision provides
  27429. useful information. Surprisingly, there is no evidence that
  27430. vision can guide the selection of escape strategies. Fleeing
  27431. behavior can be readily triggered by a rapidly looming overhead
  27432. stimulus [4]. Freezing behavior, however, has previously been
  27433. induced by real predators or their odors [5]. Here, we discover
  27434. that a small moving disk, simulating the sweep of a predator
  27435. cruising overhead, is sufficient to induce freezing response in
  27436. mice. Looming and sweeping therefore provide visual triggers for
  27437. opposing flight and freeze behaviors and provide evidence that
  27438. mice innately make behavioral choices based on vision alone.
  27439. VIDEO ABSTRACT.",
  27440. journal = "Curr. Biol.",
  27441. publisher = "Elsevier",
  27442. volume = 26,
  27443. number = 16,
  27444. pages = "2150--2154",
  27445. month = aug,
  27446. year = 2016,
  27447. keywords = "innate behavior; mouse; predator and prey; visual pathways",
  27448. language = "en"
  27449. }
  27450. @ARTICLE{Vale2017-cv,
  27451. title = "Rapid Spatial Learning Controls Instinctive Defensive Behavior
  27452. in Mice",
  27453. author = "Vale, Ruben and Evans, Dominic A and Branco, Tiago",
  27454. abstract = "Instinctive defensive behaviors are essential for animal
  27455. survival. Across the animal kingdom, there are sensory stimuli
  27456. that innately represent threat and trigger stereotyped behaviors
  27457. such as escape or freezing [1-4]. While innate behaviors are
  27458. considered to be hard-wired stimulus-responses [5], they act
  27459. within dynamic environments, and factors such as the properties
  27460. of the threat [6-9] and its perceived intensity [1, 10, 11],
  27461. access to food sources [12-14], and expectations from past
  27462. experience [15, 16] have been shown to influence defensive
  27463. behaviors, suggesting that their expression can be modulated.
  27464. However, despite recent work [2, 4, 17-21], little is known
  27465. about how flexible mouse innate defensive behaviors are and how
  27466. quickly they can be modified by experience. To address this, we
  27467. have investigated the dependence of escape behavior on learned
  27468. knowledge about the spatial environment and how the behavior is
  27469. updated when the environment changes acutely. Using behavioral
  27470. assays with innately threatening visual and auditory stimuli, we
  27471. show that the primary goal of escape in mice is to reach a
  27472. previously memorized shelter location. Memory of the escape
  27473. target can be formed in a single shelter visit lasting less than
  27474. 20 s, and changes in the spatial environment lead to a rapid
  27475. update of the defensive action, including changing the defensive
  27476. strategy from escape to freezing. Our results show that although
  27477. there are innate links between specific sensory features and
  27478. defensive behavior, instinctive defensive actions are
  27479. surprisingly flexible and can be rapidly updated by experience
  27480. to adapt to changing spatial environments.",
  27481. journal = "Curr. Biol.",
  27482. publisher = "Elsevier",
  27483. volume = 27,
  27484. number = 9,
  27485. pages = "1342--1349",
  27486. month = may,
  27487. year = 2017,
  27488. keywords = "defensive behavior; escape; freezing; innate behavior; mouse;
  27489. shelter; spatial learning; spatial memory",
  27490. language = "en"
  27491. }
  27492. @ARTICLE{Evans2018-rs,
  27493. title = "A synaptic threshold mechanism for computing escape decisions",
  27494. author = "Evans, Dominic A and Stempel, A Vanessa and Vale, Ruben and
  27495. Ruehle, Sabine and Lefler, Yaara and Branco, Tiago",
  27496. abstract = "Escaping from imminent danger is an instinctive behaviour that
  27497. is fundamental for survival, and requires the classification of
  27498. sensory stimuli as harmless or threatening. The absence of
  27499. threat enables animals to forage for essential resources, but as
  27500. the level of threat and potential for harm increases, they have
  27501. to decide whether or not to seek safety 1 . Despite previous
  27502. work on instinctive defensive behaviours in rodents2-11, little
  27503. is known about how the brain computes the threat level for
  27504. initiating escape. Here we show that the probability and vigour
  27505. of escape in mice scale with the saliency of innate threats, and
  27506. are well described by a model that computes the distance between
  27507. the threat level and an escape threshold. Calcium imaging and
  27508. optogenetics in the midbrain of freely behaving mice show that
  27509. the activity of excitatory neurons in the deep layers of the
  27510. medial superior colliculus (mSC) represents the saliency of the
  27511. threat stimulus and is predictive of escape, whereas
  27512. glutamatergic neurons of the dorsal periaqueductal grey (dPAG)
  27513. encode exclusively the choice to escape and control escape
  27514. vigour. We demonstrate a feed-forward monosynaptic excitatory
  27515. connection from mSC to dPAG neurons, which is weak and
  27516. unreliable-yet required for escape behaviour-and provides a
  27517. synaptic threshold for dPAG activation and the initiation of
  27518. escape. This threshold can be overcome by high mSC network
  27519. activity because of short-term synaptic facilitation and
  27520. recurrent excitation within the mSC, which amplifies and
  27521. sustains synaptic drive to the dPAG. Therefore, dPAG
  27522. glutamatergic neurons compute escape decisions and escape vigour
  27523. using a synaptic mechanism to threshold threat information
  27524. received from the mSC, and provide a biophysical model of how
  27525. the brain performs a critical behavioural computation.",
  27526. journal = "Nature",
  27527. publisher = "nature.com",
  27528. volume = 558,
  27529. number = 7711,
  27530. pages = "590--594",
  27531. month = jun,
  27532. year = 2018,
  27533. language = "en"
  27534. }
  27535. @ARTICLE{Branco2010-cb,
  27536. title = "Dendritic discrimination of temporal input sequences in cortical
  27537. neurons",
  27538. author = "Branco, Tiago and Clark, Beverley A and H{\"a}usser, Michael",
  27539. abstract = "The detection and discrimination of temporal sequences is
  27540. fundamental to brain function and underlies perception,
  27541. cognition, and motor output. By applying patterned, two-photon
  27542. glutamate uncaging, we found that single dendrites of cortical
  27543. pyramidal neurons exhibit sensitivity to the sequence of
  27544. synaptic activation. This sensitivity is encoded by both local
  27545. dendritic calcium signals and somatic depolarization, leading to
  27546. sequence-selective spike output. The mechanism involves
  27547. dendritic impedance gradients and nonlinear synaptic
  27548. N-methyl-D-aspartate receptor activation and is generalizable to
  27549. dendrites in different neuronal types. This enables
  27550. discrimination of patterns delivered to a single dendrite, as
  27551. well as patterns distributed randomly across the dendritic tree.
  27552. Pyramidal cell dendrites can thus act as processing compartments
  27553. for the detection of synaptic sequences, thereby implementing a
  27554. fundamental cortical computation.",
  27555. journal = "Science",
  27556. publisher = "science.sciencemag.org",
  27557. volume = 329,
  27558. number = 5999,
  27559. pages = "1671--1675",
  27560. month = sep,
  27561. year = 2010,
  27562. language = "en"
  27563. }
  27564. @UNPUBLISHED{Brown2018-bm,
  27565. title = "Ethology as a physical science",
  27566. author = "Brown, Andre E X and de Bivort, Benjamin",
  27567. abstract = "Behaviour is the ultimate output of an animal9s nervous system
  27568. and choosing the right action at the right time can be critical
  27569. for survival. The study of the organisation of behaviour in its
  27570. natural context, ethology, has historically been a primarily
  27571. qualitative science. A quantitative theory of behaviour would
  27572. advance research in neuroscience as well as ecology and
  27573. evolution. However, animal posture typically has many degrees of
  27574. freedom and behavioural dynamics vary on timescales ranging from
  27575. milliseconds to years, presenting both technical and conceptual
  27576. challenges. Here we review 1) advances in imaging and computer
  27577. vision that are making it possible to capture increasingly
  27578. complete records of animal motion and 2) new approaches to
  27579. understanding the resulting behavioural data sets. With the right
  27580. analytical approaches, these data are allowing researchers to
  27581. revisit longstanding questions about the structure and
  27582. organisation of animal behaviour and to put unifying principles
  27583. on a quantitative footing. Contributions from both
  27584. experimentalists and theorists are leading to the emergence of a
  27585. physics of behaviour and the prospect of discovering laws and
  27586. developing theories with broad applicability. We believe that
  27587. there now exists an opportunity to develop theories of behaviour
  27588. which can be tested using these data sets leading to a deeper
  27589. understanding of how and why animals behave.",
  27590. journal = "bioRxiv",
  27591. pages = "220855",
  27592. month = feb,
  27593. year = 2018,
  27594. keywords = "Analysis/Modelling [Behaviour]",
  27595. language = "en"
  27596. }
  27597. @ARTICLE{Mathis2018-uu,
  27598. title = "Markerless tracking of user-defined features with deep
  27599. learning",
  27600. author = "Mathis, Alexander and Mamidanna, Pranav and Abe, Taiga and
  27601. Cury, Kevin M and Murthy, Venkatesh N and Mathis, Mackenzie
  27602. W and Bethge, Matthias",
  27603. abstract = "Quantifying behavior is crucial for many applications in
  27604. neuroscience. Videography provides easy methods for the
  27605. observation and recording of animal behavior in diverse
  27606. settings, yet extracting particular aspects of a behavior
  27607. for further analysis can be highly time consuming. In motor
  27608. control studies, humans or other animals are often marked
  27609. with reflective markers to assist with computer-based
  27610. tracking, yet markers are intrusive (especially for smaller
  27611. animals), and the number and location of the markers must be
  27612. determined a priori. Here, we present a highly efficient
  27613. method for markerless tracking based on transfer learning
  27614. with deep neural networks that achieves excellent results
  27615. with minimal training data. We demonstrate the versatility
  27616. of this framework by tracking various body parts in a broad
  27617. collection of experimental settings: mice odor
  27618. trail-tracking, egg-laying behavior in drosophila, and mouse
  27619. hand articulation in a skilled forelimb task. For example,
  27620. during the skilled reaching behavior, individual joints can
  27621. be automatically tracked (and a confidence score is
  27622. reported). Remarkably, even when a small number of frames
  27623. are labeled ($\approx 200$), the algorithm achieves
  27624. excellent tracking performance on test frames that is
  27625. comparable to human accuracy.",
  27626. month = apr,
  27627. year = 2018,
  27628. keywords = "Analysis/Modelling [Behaviour]",
  27629. archivePrefix = "arXiv",
  27630. primaryClass = "cs.CV",
  27631. eprint = "1804.03142"
  27632. }
  27633. @ARTICLE{Ainge2007-ue,
  27634. title = "Hippocampal {CA1} place cells encode intended destination on a
  27635. maze with multiple choice points",
  27636. author = "Ainge, James A and Tamosiunaite, Minija and Woergoetter,
  27637. Florentin and Dudchenko, Paul A",
  27638. abstract = "The hippocampus encodes both spatial and nonspatial aspects of a
  27639. rat's ongoing behavior at the single-cell level. In this study,
  27640. we examined the encoding of intended destination by hippocampal
  27641. (CA1) place cells during performance of a serial reversal task on
  27642. a double Y-maze. On the maze, rats had to make two choices to
  27643. access one of four possible goal locations, two of which
  27644. contained reward. Reward locations were kept constant within
  27645. blocks of 10 trials but changed between blocks, and the session
  27646. of each day comprised three or more trial blocks. A
  27647. disproportionate number of place fields were observed in the
  27648. start box and beginning stem of the maze, relative to other
  27649. locations on the maze. Forty-six percent of these place fields
  27650. had different firing rates on journeys to different goal boxes.
  27651. Another group of cells had place fields before the second choice
  27652. point, and, of these, 44\% differentiated between journeys to
  27653. specific goal boxes. In a second experiment, we observed that
  27654. rats with hippocampal damage made significantly more errors than
  27655. control rats on the Y-maze when reward locations were reversed.
  27656. Together, these results suggest that, at the start of the maze,
  27657. the hippocampus encodes both current location and the intended
  27658. destination of the rat, and this encoding is necessary for the
  27659. flexible response to changes in reinforcement contingencies.",
  27660. journal = "J. Neurosci.",
  27661. volume = 27,
  27662. number = 36,
  27663. pages = "9769--9779",
  27664. month = sep,
  27665. year = 2007,
  27666. keywords = "Spatial Navigation;Decision Making",
  27667. language = "en"
  27668. }
  27669. @ARTICLE{Grieves2017-ay,
  27670. title = "The representation of space in the brain",
  27671. author = "Grieves, Roddy M and Jeffery, Kate J",
  27672. journal = "Behav. Processes",
  27673. publisher = "Elsevier",
  27674. volume = 135,
  27675. pages = "113--131",
  27676. year = 2017,
  27677. keywords = "Spatial Navigation"
  27678. }
  27679. @ARTICLE{Gauthier2018-db,
  27680. title = "A Dedicated Population for Reward Coding in the Hippocampus",
  27681. author = "Gauthier, Jeffrey L and Tank, David W",
  27682. abstract = "The hippocampus plays a critical role in goal-directed
  27683. navigation. Across different environments, however, hippocampal
  27684. maps are randomized, making it unclear how goal locations could
  27685. be encoded consistently. To address this question, we developed a
  27686. virtual reality task with shifting reward contingencies to
  27687. distinguish place versus reward encoding. In mice performing the
  27688. task, large-scale recordings in CA1 and subiculum revealed a
  27689. small, specialized cell population that was only active near
  27690. reward yet whose activity could not be explained by sensory cues
  27691. or stereotyped reward anticipation behavior. Across different
  27692. virtual environments, most cells remapped randomly, but reward
  27693. encoding consistently arose from a single pool of cells,
  27694. suggesting that they formed a dedicated channel for reward. These
  27695. observations represent a significant departure from the current
  27696. understanding of CA1 as a relatively homogeneous ensemble without
  27697. fixed coding properties and provide a new candidate for the
  27698. cellular basis of goal memory in the hippocampus.",
  27699. journal = "Neuron",
  27700. volume = 99,
  27701. number = 1,
  27702. pages = "179--193.e7",
  27703. month = jul,
  27704. year = 2018,
  27705. keywords = "CA1; hippocampus; navigation; place cells; place fields; reward;
  27706. subiculum; virtual reality;Spatial Navigation",
  27707. language = "en"
  27708. }
  27709. % The entry below contains non-ASCII chars that could not be converted
  27710. % to a LaTeX equivalent.
  27711. @ARTICLE{Horndli2018-ja,
  27712. title = "Machine Learning Reveals Modules of Economic Behavior from
  27713. Foraging Mice",
  27714. author = "Horndli, C N S and Wong, E and Ferris, E and Rhodes, A N and
  27715. {others}",
  27716. abstract = "The mechanisms shaping most ethological behavior patterns are
  27717. elusive because we do not understand how complex patterns are
  27718. constructed. Here, we develop a behavioral paradigm and data
  27719. analysis methods to dissect foraging patterns in mice. We
  27720. uncover discrete behavioral modules linked to round trip
  27721. excursions from the home. Using machine learning, 59 modules are
  27722. revealed across different genetic backgrounds and ages.
  27723. Different modules develop at different ages and are linked to
  27724. different aspects of economic behavior, including …",
  27725. journal = "bioRxiv",
  27726. publisher = "biorxiv.org",
  27727. year = 2018,
  27728. keywords = "Analysis/Modelling [Behaviour]"
  27729. }
  27730. @ARTICLE{Luo2018-cf,
  27731. title = "A dopaminergic switch for fear to safety transitions",
  27732. author = "Luo, Ray and Uematsu, Akira and Weitemier, Adam and Aquili, Luca
  27733. and Koivumaa, Jenny and McHugh, Thomas J and Johansen, Joshua P",
  27734. abstract = "Overcoming aversive emotional memories requires neural systems
  27735. that detect when fear responses are no longer appropriate so that
  27736. they can be extinguished. The midbrain ventral tegmental area
  27737. (VTA) dopamine system has been implicated in reward and more
  27738. broadly in signaling when a better-than-expected outcome has
  27739. occurred. This suggests that it may be important in guiding fear
  27740. to safety transitions. We report that when an expected aversive
  27741. outcome does not occur, activity in midbrain dopamine neurons is
  27742. necessary to extinguish behavioral fear responses and engage
  27743. molecular signaling events in extinction learning circuits.
  27744. Furthermore, a specific dopamine projection to the nucleus
  27745. accumbens medial shell is partially responsible for this effect.
  27746. In contrast, a separate dopamine projection to the medial
  27747. prefrontal cortex opposes extinction learning. This demonstrates
  27748. a novel function for the canonical VTA-dopamine reward system and
  27749. reveals opposing behavioral roles for different dopamine neuron
  27750. projections in fear extinction learning.",
  27751. journal = "Nat. Commun.",
  27752. volume = 9,
  27753. number = 1,
  27754. pages = "2483",
  27755. month = jun,
  27756. year = 2018,
  27757. keywords = "Threat response",
  27758. language = "en"
  27759. }
  27760. @UNPUBLISHED{Torok2018-fl,
  27761. title = "A novel virtual plus-maze for studying electrophysiological
  27762. correlates of spatial reorientation",
  27763. author = "Torok, Agoston and Kobor, Andrea and Honbolygo, Ferenc and Baker,
  27764. Travis",
  27765. abstract = "Quick reorientation is an essential part of successful
  27766. navigation. Despite growing attention to this ability, little is
  27767. known about how reorientation happens in humans. To this aim, we
  27768. recorded EEG from 34 participants. Participants were navigating a
  27769. simple virtual reality plus-maze where at the beginning of each
  27770. trial they were randomly teleported to either the North or the
  27771. South alley. Results show that the teleportation event caused a
  27772. quick reorientation effect over occipito-parietal areas as early
  27773. as 100 msec; meaning that despite the known stochastic nature of
  27774. the teleportation, participants built up expectations for their
  27775. place of arrival. This result has important consequences for the
  27776. optimal design of virtual reality locomotion.",
  27777. journal = "bioRxiv",
  27778. pages = "369207",
  27779. month = jul,
  27780. year = 2018,
  27781. keywords = "Spatial Navigation",
  27782. language = "en"
  27783. }
  27784. @ARTICLE{OConnell2018-st,
  27785. title = "Bridging Neural and Computational Viewpoints on Perceptual
  27786. {Decision-Making}",
  27787. author = "O'Connell, Redmond G and Shadlen, Michael N and Wong-Lin,
  27788. Kongfatt and Kelly, Simon P",
  27789. abstract = "Sequential sampling models have provided a dominant theoretical
  27790. framework guiding computational and neurophysiological
  27791. investigations of perceptual decision-making. While these models
  27792. share the basic principle that decisions are formed by
  27793. accumulating sensory evidence to a bound, they come in many forms
  27794. that can make similar predictions of choice behaviour despite
  27795. invoking fundamentally different mechanisms. The identification
  27796. of neural signals that reflect some of the core computations
  27797. underpinning decision formation offers new avenues for
  27798. empirically testing and refining key model assumptions. Here, we
  27799. highlight recent efforts to explore these avenues and, in so
  27800. doing, consider the conceptual and methodological challenges that
  27801. arise when seeking to infer decision computations from complex
  27802. neural data.",
  27803. journal = "Trends Neurosci.",
  27804. month = jul,
  27805. year = 2018,
  27806. keywords = "computational modelling; lateral intraparietal area (LIP);
  27807. perceptual decision-making; sequential sampling;Decision Making",
  27808. language = "en"
  27809. }
  27810. @ARTICLE{Stringer2048-xi,
  27811. title = "High-dimensional geometry of population responses in visual
  27812. cortex",
  27813. author = "Stringer, C and Pachitariu, M and Steinmetz, N and Carandini, M
  27814. and Harris, K",
  27815. abstract = "A neuronal population encodes information most efficiently when
  27816. its activity is uncorrelated and high-dimensional, but cor-
  27817. related lower-dimensional codes provide robustness against noise.
  27818. Here, we analyzed the correlation structure of natural image
  27819. coding, in large visual cortical populations recorded from awake
  27820. mice. Evoked population activity was high dimen- sional, with
  27821. correlations obeying an unexpected power-law: the nth principal
  27822. component variance scaled as 1/n. This was not inherited from the
  27823. 1/f spectrum of natural images, because it persisted after
  27824. stimulus whitening. We proved mathemat- ically that the variance
  27825. spectrum must decay at least this fast if a population code is
  27826. smooth, i.e. if small changes in input cannot dominate population
  27827. activity. The theory also predicted larger power-law exponents
  27828. for lower-dimensional stimu- lus ensembles, which we validated
  27829. experimentally. These results suggest that coding smoothness
  27830. represents a fundamental constraint governing correlations in
  27831. neural population codes.",
  27832. journal = "bioRxiv",
  27833. pages = "374090",
  27834. month = jul,
  27835. year = 2048,
  27836. keywords = "Authors;Authors/Harris/Carandini",
  27837. language = "en"
  27838. }
  27839. @ARTICLE{Sweis2018-sg,
  27840. title = "Sensitivity to ``sunk costs'' in mice, rats, and humans",
  27841. author = "Sweis, Brian M and Abram, Samantha V and Schmidt, Brandy J and
  27842. Seeland, Kelsey D and MacDonald, Angus W and Thomas, Mark J and
  27843. David Redish, A",
  27844. abstract = "Sunk costs are irrecoverable investments that should not
  27845. influence decisions, because decisions should be made on the
  27846. basis of expected future consequences. Both human and nonhuman
  27847. animals can show sensitivity to sunk costs, but reports from
  27848. across species are inconsistent. In a temporal context, a
  27849. sensitivity to sunk costs arises when an individual resists
  27850. ending an activity, even if it seems unproductive, because of
  27851. the time already invested. In two parallel foraging tasks that
  27852. we designed, we found that mice, rats, and humans show similar
  27853. sensitivities to sunk costs in their decision-making.
  27854. Unexpectedly, sensitivity to time invested accrued only after an
  27855. initial decision had been made. These findings suggest that
  27856. sensitivity to temporal sunk costs lies in a vulnerability
  27857. distinct from deliberation processes and that this distinction
  27858. is present across species.",
  27859. journal = "Science",
  27860. publisher = "American Association for the Advancement of Science",
  27861. volume = 361,
  27862. number = 6398,
  27863. pages = "178--181",
  27864. month = jul,
  27865. year = 2018,
  27866. keywords = "Decision Making",
  27867. language = "en"
  27868. }
  27869. @ARTICLE{Djurdjevic2018-yn,
  27870. title = "Accuracy of Rats in Discriminating Visual Objects Is Explained by
  27871. the Complexity of Their Perceptual Strategy",
  27872. author = "Djurdjevic, Vladimir and Ansuini, Alessio and Bertolini, Daniele
  27873. and Macke, Jakob H and Zoccolan, Davide",
  27874. abstract = "Despite their growing popularity as models of visual functions,
  27875. it remains unclear whether rodents are capable of deploying
  27876. advanced shape-processing strategies when engaged in visual
  27877. object recognition. In rats, for instance, pattern vision has
  27878. been reported to range from mere detection of overall object
  27879. luminance to view-invariant processing of discriminative shape
  27880. features. Here we sought to clarify how refined object vision is
  27881. in rodents, and how variable the complexity of their visual
  27882. processing strategy is across individuals. To this aim, we
  27883. measured how well rats could discriminate a reference object from
  27884. 11 distractors, which spanned a spectrum of image-level
  27885. similarity to the reference. We also presented the animals with
  27886. random variations of the reference, and processed their responses
  27887. to these stimuli to derive subject-specific models of rat
  27888. perceptual choices. Our models successfully captured the highly
  27889. variable discrimination performance observed across subjects and
  27890. object conditions. In particular, they revealed that the animals
  27891. that succeeded with the most challenging distractors were those
  27892. that integrated the wider variety of discriminative features into
  27893. their perceptual strategies. Critically, these strategies were
  27894. largely preserved when the rats were required to discriminate
  27895. outlined and scaled versions of the stimuli, thus showing that
  27896. rat object vision can be characterized as a
  27897. transformation-tolerant, feature-based filtering process.
  27898. Overall, these findings indicate that rats are capable of
  27899. advanced processing of shape information, and point to the
  27900. rodents as powerful models for investigating the neuronal
  27901. underpinnings of visual object recognition and other high-level
  27902. visual functions.",
  27903. journal = "Curr. Biol.",
  27904. volume = 28,
  27905. number = 7,
  27906. pages = "1005--1015.e5",
  27907. month = apr,
  27908. year = 2018,
  27909. keywords = "classification; filtering; image; object; perception; processing;
  27910. recognition; rodent; shape; vision",
  27911. language = "en"
  27912. }
  27913. @ARTICLE{Berman2016-it,
  27914. title = "Predictability and hierarchy in Drosophila behavior",
  27915. author = "Berman, Gordon J and Bialek, William and Shaevitz, Joshua W",
  27916. abstract = "Even the simplest of animals exhibit behavioral sequences with
  27917. complex temporal dynamics. Prominent among the proposed
  27918. organizing principles for these dynamics has been the idea of a
  27919. hierarchy, wherein the movements an animal makes can be
  27920. understood as a set of nested subclusters. Although this type of
  27921. organization holds potential advantages in terms of motion
  27922. control and neural circuitry, measurements demonstrating this for
  27923. an animal's entire behavioral repertoire have been limited in
  27924. scope and temporal complexity. Here, we use a recently developed
  27925. unsupervised technique to discover and track the occurrence of
  27926. all stereotyped behaviors performed by fruit flies moving in a
  27927. shallow arena. Calculating the optimally predictive
  27928. representation of the fly's future behaviors, we show that fly
  27929. behavior exhibits multiple time scales and is organized into a
  27930. hierarchical structure that is indicative of its underlying
  27931. behavioral programs and its changing internal states.",
  27932. journal = "Proc. Natl. Acad. Sci. U. S. A.",
  27933. volume = 113,
  27934. number = 42,
  27935. pages = "11943--11948",
  27936. month = oct,
  27937. year = 2016,
  27938. keywords = "Drosophila; behavior; hierarchy; information
  27939. bottleneck;Authors/Berman",
  27940. language = "en"
  27941. }
  27942. @ARTICLE{Berman2014-pw,
  27943. title = "Mapping the stereotyped behaviour of freely moving fruit flies",
  27944. author = "Berman, Gordon J and Choi, Daniel M and Bialek, William and
  27945. Shaevitz, Joshua W",
  27946. abstract = "A frequent assumption in behavioural science is that most of an
  27947. animal's activities can be described in terms of a small set of
  27948. stereotyped motifs. Here, we introduce a method for mapping an
  27949. animal's actions, relying only upon the underlying structure of
  27950. postural movement data to organize and classify behaviours.
  27951. Applying this method to the ground-based behaviour of the fruit
  27952. fly, Drosophila melanogaster, we find that flies perform
  27953. stereotyped actions roughly 50\% of the time, discovering over
  27954. 100 distinguishable, stereotyped behavioural states. These
  27955. include multiple modes of locomotion and grooming. We use the
  27956. resulting measurements as the basis for identifying subtle
  27957. sex-specific behavioural differences and revealing the
  27958. low-dimensional nature of animal motions.",
  27959. journal = "J. R. Soc. Interface",
  27960. volume = 11,
  27961. number = 99,
  27962. month = oct,
  27963. year = 2014,
  27964. keywords = "Drosophila; behaviour; phase reconstruction; stereotypy;
  27965. unsupervised learning;Authors/Berman;Analysis/Modelling
  27966. [Behaviour]",
  27967. language = "en"
  27968. }
  27969. @ARTICLE{Klibaite2017-kd,
  27970. title = "An unsupervised method for quantifying the behavior of paired
  27971. animals",
  27972. author = "Klibaite, Ugne and Berman, Gordon J and Cande, Jessica and Stern,
  27973. David L and Shaevitz, Joshua W",
  27974. abstract = "Behaviors involving the interaction of multiple individuals are
  27975. complex and frequently crucial for an animal's survival. These
  27976. interactions, ranging across sensory modalities, length scales,
  27977. and time scales, are often subtle and difficult to characterize.
  27978. Contextual effects on the frequency of behaviors become even more
  27979. difficult to quantify when physical interaction between animals
  27980. interferes with conventional data analysis, e.g. due to visual
  27981. occlusion. We introduce a method for quantifying behavior in
  27982. fruit fly interaction that combines high-throughput video
  27983. acquisition and tracking of individuals with recent unsupervised
  27984. methods for capturing an animal's entire behavioral repertoire.
  27985. We find behavioral differences between solitary flies and those
  27986. paired with an individual of the opposite sex, identifying
  27987. specific behaviors that are affected by social and spatial
  27988. context. Our pipeline allows for a comprehensive description of
  27989. the interaction between two individuals using unsupervised
  27990. machine learning methods, and will be used to answer questions
  27991. about the depth of complexity and variance in fruit fly
  27992. courtship.",
  27993. journal = "Phys. Biol.",
  27994. volume = 14,
  27995. number = 1,
  27996. pages = "015006",
  27997. month = feb,
  27998. year = 2017,
  27999. keywords = "Authors/Berman;Analysis/Modelling [Behaviour]",
  28000. language = "en"
  28001. }
  28002. @ARTICLE{Bassett2018-ni,
  28003. title = "On the nature and use of models in network neuroscience",
  28004. author = "Bassett, Danielle S and Zurn, Perry and Gold, Joshua I",
  28005. abstract = "Network theory provides an intuitively appealing framework for
  28006. studying relationships among interconnected brain mechanisms and
  28007. their relevance to behaviour. As the space of its applications
  28008. grows, so does the diversity of meanings of the term network
  28009. model. This diversity can cause confusion, complicate efforts to
  28010. assess model validity and efficacy, and hamper interdisciplinary
  28011. collaboration. In this Review, we examine the field of network
  28012. neuroscience, focusing on organizing principles that can help
  28013. overcome these challenges. First, we describe the fundamental
  28014. goals in constructing network models. Second, we review the most
  28015. common forms of network models, which can be described
  28016. parsimoniously along the following three primary dimensions: from
  28017. data representations to first-principles theory; from biophysical
  28018. realism to functional phenomenology; and from elementary
  28019. descriptions to coarse-grained approximations. Third, we draw on
  28020. biology, philosophy and other disciplines to establish validation
  28021. principles for these models. We close with a discussion of
  28022. opportunities to bridge model types and point to exciting
  28023. frontiers for future pursuits.",
  28024. journal = "Nat. Rev. Neurosci.",
  28025. month = jul,
  28026. year = 2018,
  28027. language = "en"
  28028. }
  28029. @ARTICLE{Tovote2016-of,
  28030. title = "Midbrain circuits for defensive behaviour",
  28031. author = "Tovote, Philip and Esposito, Maria Soledad and Botta, Paolo and
  28032. Chaudun, Fabrice and Fadok, Jonathan P and Markovic, Milica and
  28033. Wolff, Steffen B E and Ramakrishnan, Charu and Fenno, Lief and
  28034. Deisseroth, Karl and Herry, Cyril and Arber, Silvia and
  28035. L{\"u}thi, Andreas",
  28036. abstract = "Survival in threatening situations depends on the selection and
  28037. rapid execution of an appropriate active or passive defensive
  28038. response, yet the underlying brain circuitry is not understood.
  28039. Here we use circuit-based optogenetic, in vivo and in vitro
  28040. electrophysiological, and neuroanatomical tracing methods to
  28041. define midbrain periaqueductal grey circuits for specific
  28042. defensive behaviours. We identify an inhibitory pathway from the
  28043. central nucleus of the amygdala to the ventrolateral
  28044. periaqueductal grey that produces freezing by disinhibition of
  28045. ventrolateral periaqueductal grey excitatory outputs to pre-motor
  28046. targets in the magnocellular nucleus of the medulla. In addition,
  28047. we provide evidence for anatomical and functional interaction of
  28048. this freezing pathway with long-range and local circuits
  28049. mediating flight. Our data define the neuronal circuitry
  28050. underlying the execution of freezing, an evolutionarily conserved
  28051. defensive behaviour, which is expressed by many species including
  28052. fish, rodents and primates. In humans, dysregulation of this
  28053. 'survival circuit' has been implicated in anxiety-related
  28054. disorders.",
  28055. journal = "Nature",
  28056. volume = 534,
  28057. number = 7606,
  28058. pages = "206--212",
  28059. month = jun,
  28060. year = 2016,
  28061. keywords = "Threat response",
  28062. language = "en"
  28063. }
  28064. @ARTICLE{Tovote2015-ds,
  28065. title = "Neuronal circuits for fear and anxiety",
  28066. author = "Tovote, Philip and Fadok, Jonathan Paul and L{\"u}thi, Andreas",
  28067. abstract = "Decades of research has identified the brain areas that are
  28068. involved in fear, fear extinction, anxiety and related defensive
  28069. behaviours. Newly developed genetic and viral tools, optogenetics
  28070. and advanced in vivo imaging techniques have now made it possible
  28071. to characterize the activity, connectivity and function of
  28072. specific cell types within complex neuronal circuits. Recent
  28073. findings that have been made using these tools and techniques
  28074. have provided mechanistic insights into the exquisite
  28075. organization of the circuitry underlying internal defensive
  28076. states. This Review focuses on studies that have used
  28077. circuit-based approaches to gain a more detailed, and also more
  28078. comprehensive and integrated, view on how the brain governs fear
  28079. and anxiety and how it orchestrates adaptive defensive
  28080. behaviours.",
  28081. journal = "Nat. Rev. Neurosci.",
  28082. volume = 16,
  28083. number = 6,
  28084. pages = "317--331",
  28085. month = jun,
  28086. year = 2015,
  28087. keywords = "Threat response",
  28088. language = "en"
  28089. }
  28090. @ARTICLE{Alexander2017-ho,
  28091. title = "Spatially Periodic Activation Patterns of Retrosplenial Cortex
  28092. Encode Route Sub-spaces and Distance Traveled",
  28093. author = "Alexander, Andrew S and Nitz, Douglas A",
  28094. abstract = "Traversal of a complicated route is often facilitated by
  28095. considering it as a set of related sub-spaces. Such
  28096. compartmentalization processes could occur within retrosplenial
  28097. cortex, a structure whose neurons simultaneously encode position
  28098. within routes and other spatial coordinate systems. Here,
  28099. retrosplenial cortex neurons were recorded as rats traversed a
  28100. track having recurrent structure at multiple scales. Consistent
  28101. with a major role in compartmentalization of complex routes,
  28102. individual retrosplenial cortex (RSC) neurons exhibited periodic
  28103. activation patterns that repeated across route segments having
  28104. the same shape. Concurrently, a larger population of RSC neurons
  28105. exhibited single-cycle periodicity over the full route,
  28106. effectively defining a framework for encoding of sub-route
  28107. positions relative to the whole. The same population
  28108. simultaneously provides a novel metric for distance from each
  28109. route position to all others. Together, the findings implicate
  28110. retrosplenial cortex in the extraction of path sub-spaces, the
  28111. encoding of their spatial relationships to each other, and path
  28112. integration.",
  28113. journal = "Curr. Biol.",
  28114. volume = 27,
  28115. number = 11,
  28116. pages = "1551--1560.e4",
  28117. month = jun,
  28118. year = 2017,
  28119. keywords = "distance; fragmentation; hippocampus; path integration;
  28120. periodicity; retrosplenial cortex; spatial navigation; spatial
  28121. representation; sub-route; sub-space;Spatial
  28122. Navigation;Authors/Nitz",
  28123. language = "en"
  28124. }
  28125. @ARTICLE{Behrens_undated-kg,
  28126. title = "What is a cognitive map? Organising knowledge for flexible
  28127. behaviour",
  28128. author = "Behrens, Timothy E J and Muller, Timothy H and Whittington, James
  28129. C R and Mark, Shirley and Baram, Alon B and Stachenfeld,
  28130. Kimberley L and Kurth-Nelson, Zeb",
  28131. abstract = "It is proposed that a cognitive map encoding the relationships
  28132. between entities in the world supports flexible behaviour, but
  28133. the majority of the neural evidence for such a system comes from
  28134. studies of spatial navigation. Recent work describing neuronal
  28135. parallels between spatial and non-spatial behaviours has
  28136. rekindled the notion of a systematic organisation of knowledge
  28137. across multiple domains. We review experimental evidence and
  28138. theoretical frameworks that point to principles unifying these
  28139. apparently disparate functions. These principles describe how to
  28140. learn and use abstract, generalisable knowledge and suggest
  28141. map-like representations observed in a spatial context may be an
  28142. instance of general coding mechanisms capable of organising
  28143. knowledge of all kinds. We highlight how artificial agents
  28144. endowed with such principles exhibit flexible behaviour and learn
  28145. map-like representations observed in the brain. Finally, we
  28146. speculate on how these principles may offer insight into the
  28147. extreme generalisations, abstractions and inferences that
  28148. characterise human cognition.",
  28149. journal = "Biorxiv",
  28150. keywords = "Spatial Navigation"
  28151. }
  28152. % The entry below contains non-ASCII chars that could not be converted
  28153. % to a LaTeX equivalent.
  28154. @ARTICLE{Patai2018-io,
  28155. title = "Neural signatures of detours, shortcuts and back-tracking during
  28156. navigation",
  28157. author = "Patai, E Z and Javadi, A H and Margois, A and Tan, H R and
  28158. Kumaran, D and {others}",
  28159. abstract = "Central to the concept of the 9cognitive map9 is that it confers
  28160. flexibility in behaviour allowing animals to take efficient
  28161. detours, exploit shortcuts and realise when they need to
  28162. back-track rather than continue on a poorly chosen route.
  28163. Currently the neural underpinnings of such behaviour remains
  28164. unclear. During fMRI we tested human subjects on their ability
  28165. to navigate to a set of goal locations in a virtual desert
  28166. island riven by lava, which occasionally shifted to block
  28167. selected paths (necessitating detours) or receded to …",
  28168. journal = "bioRxiv",
  28169. publisher = "biorxiv.org",
  28170. year = 2018,
  28171. keywords = "Spatial Navigation"
  28172. }
  28173. @ARTICLE{Berman2018-xu,
  28174. title = "Measuring behavior across scales",
  28175. author = "Berman, Gordon J",
  28176. abstract = "The need for high-throughput, precise, and meaningful methods for
  28177. measuring behavior has been amplified by our recent successes in
  28178. measuring and manipulating neural circuitry. The largest
  28179. challenges associated with moving in this direction, however, are
  28180. not technical but are instead conceptual: what numbers should one
  28181. put on the movements an animal is performing (or not performing)?
  28182. In this review, I will describe how theoretical and data
  28183. analytical ideas are interfacing with recently-developed
  28184. computational and experimental methodologies to answer these
  28185. questions across a variety of contexts, length scales, and time
  28186. scales. I will attempt to highlight commonalities between
  28187. approaches and areas where further advances are necessary to
  28188. place behavior on the same quantitative footing as other
  28189. scientific fields.",
  28190. journal = "BMC Biol.",
  28191. volume = 16,
  28192. number = 1,
  28193. pages = "23",
  28194. month = feb,
  28195. year = 2018,
  28196. keywords = "Authors/Berman;Analysis/Modelling [Behaviour]",
  28197. language = "en"
  28198. }
  28199. @ARTICLE{Severi2014-ox,
  28200. title = "Neural control and modulation of swimming speed in the larval
  28201. zebrafish",
  28202. author = "Severi, Kristen E and Portugues, Ruben and Marques, Jo{\~a}o C
  28203. and O'Malley, Donald M and Orger, Michael B and Engert, Florian",
  28204. abstract = "Vertebrate locomotion at different speeds is driven by descending
  28205. excitatory connections to central pattern generators in the
  28206. spinal cord. To investigate how these inputs determine locomotor
  28207. kinematics, we used whole-field visual motion to drive zebrafish
  28208. to swim at different speeds. Larvae match the stimulus speed by
  28209. utilizing more locomotor events, or modifying kinematic
  28210. parameters such as the duration and speed of swimming bouts, the
  28211. tail-beat frequency, and the choice of gait. We used laser
  28212. ablations, electrical stimulation, and activity recordings in
  28213. descending neurons of the nucleus of the medial longitudinal
  28214. fasciculus (nMLF) to dissect their contribution to controlling
  28215. forward movement. We found that the activity of single identified
  28216. neurons within the nMLF is correlated with locomotor kinematics,
  28217. and modulates both the duration and oscillation frequency of tail
  28218. movements. By identifying the contribution of individual
  28219. supraspinal circuit elements to locomotion kinematics, we build a
  28220. better understanding of how the brain controls movement.",
  28221. journal = "Neuron",
  28222. volume = 83,
  28223. number = 3,
  28224. pages = "692--707",
  28225. month = aug,
  28226. year = 2014,
  28227. language = "en"
  28228. }
  28229. @ARTICLE{Marques2018-kf,
  28230. title = "Structure of the Zebrafish Locomotor Repertoire Revealed with
  28231. Unsupervised Behavioral Clustering",
  28232. author = "Marques, Jo{\~a}o C and Lackner, Simone and F{\'e}lix, Rita and
  28233. Orger, Michael B",
  28234. abstract = "An important concept in ethology is that complex behaviors can be
  28235. constructed from a set of basic motor patterns. Identifying the
  28236. set of patterns available to an animal is key to making
  28237. quantitative descriptions of behavior that reflect the underlying
  28238. motor system organization. We addressed these questions in
  28239. zebrafish larvae, which swim in bouts that are naturally
  28240. segmented in time. We developed a robust and general purpose
  28241. clustering method (clusterdv) to ensure accurate identification
  28242. of movement clusters and applied it to a dataset consisting of
  28243. millions of swim bouts, captured at high temporal resolution from
  28244. a comprehensive set of behavioral contexts. We identified a set
  28245. of thirteen basic swimming patterns that are used flexibly in
  28246. various combinations across different behavioral contexts and
  28247. show that this classification can be used to dissect the
  28248. sensorimotor transformations underlying larval social behavior
  28249. and hunting. Furthermore, using the same approach at different
  28250. levels in the behavioral hierarchy, we show that the set of swim
  28251. bouts are themselves constructed from a basic set of tail
  28252. movements and that bouts are executed in sequences specific to
  28253. different behaviors.",
  28254. journal = "Curr. Biol.",
  28255. volume = 28,
  28256. number = 2,
  28257. pages = "181--195.e5",
  28258. month = jan,
  28259. year = 2018,
  28260. keywords = "behavior; behavioral motifs; cluster analysis; clusterdv;
  28261. locomotion; motor control; sequences; unsupervised machine
  28262. learning; visual behavior; zebrafish",
  28263. language = "en"
  28264. }
  28265. % The entry below contains non-ASCII chars that could not be converted
  28266. % to a LaTeX equivalent.
  28267. @ARTICLE{Mimica2018-ls,
  28268. title = "Efficient cortical coding of {3D} posture in freely behaving
  28269. rats",
  28270. author = "Mimica, B and Dunn, B A and Tombaz, T and Bojja, V S and
  28271. Whitlock, J R",
  28272. abstract = "In order to meet physical and behavioural demands of their
  28273. environments animals constantly update their body posture, but
  28274. little is known about the neural signals on which this ability
  28275. depends. To better understand the role of cortex in coordinating
  28276. natural pose and movement, we tracked the heads and backs of
  28277. freely foraging rats in 3D while recording simultaneously from
  28278. posterior parietal cortex (PPC) and frontal motor cortex (M2),
  28279. areas critical for spatial movement planning and navigation.
  28280. Single units in both regions were …",
  28281. journal = "bioRxiv",
  28282. publisher = "biorxiv.org",
  28283. year = 2018,
  28284. keywords = "Spatial Navigation;Analysis/Modelling [Behaviour]"
  28285. }
  28286. @ARTICLE{Dunn2016-zt,
  28287. title = "Neural Circuits Underlying Visually Evoked Escapes in Larval
  28288. Zebrafish",
  28289. author = "Dunn, Timothy W and Gebhardt, Christoph and Naumann, Eva A and
  28290. Riegler, Clemens and Ahrens, Misha B and Engert, Florian and Del
  28291. Bene, Filippo",
  28292. abstract = "Escape behaviors deliver organisms away from imminent
  28293. catastrophe. Here, we characterize behavioral responses of freely
  28294. swimming larval zebrafish to looming visual stimuli simulating
  28295. predators. We report that the visual system alone can recruit
  28296. lateralized, rapid escape motor programs, similar to those
  28297. elicited by mechanosensory modalities. Two-photon calcium imaging
  28298. of retino-recipient midbrain regions isolated the optic tectum as
  28299. an important center processing looming stimuli, with ensemble
  28300. activity encoding the critical image size determining escape
  28301. latency. Furthermore, we describe activity in retinal ganglion
  28302. cell terminals and superficial inhibitory interneurons in the
  28303. tectum during looming and propose a model for how temporal
  28304. dynamics in tectal periventricular neurons might arise from
  28305. computations between these two fundamental constituents. Finally,
  28306. laser ablations of hindbrain circuitry confirmed that visual and
  28307. mechanosensory modalities share the same premotor output network.
  28308. We establish a circuit for the processing of aversive stimuli in
  28309. the context of an innate visual behavior.",
  28310. journal = "Neuron",
  28311. volume = 89,
  28312. number = 3,
  28313. pages = "613--628",
  28314. month = feb,
  28315. year = 2016,
  28316. language = "en"
  28317. }
  28318. @ARTICLE{Floyd2000-rj,
  28319. title = "Orbitomedial prefrontal cortical projections to distinct
  28320. longitudinal columns of the periaqueductal gray in the rat",
  28321. author = "Floyd, N S and Price, J L and Ferry, A T and Keay, K A and
  28322. Bandler, R",
  28323. abstract = "We utilised retrograde and anterograde tracing procedures to
  28324. study the origin and termination of prefrontal cortical (PFC)
  28325. projections to the periaqueductal gray (PAG) in the rat. A
  28326. previous study, in the primate, had demonstrated that distinct
  28327. subgroups of PFC areas project to specific PAG columns.
  28328. Retrograde tracing experiments revealed that projections to
  28329. dorsolateral (dlPAG) and ventrolateral (vlPAG) periaqueductal
  28330. gray columns arose from medial PFC, specifically prelimbic,
  28331. infralimbic, and anterior cingulate cortices. Injections made in
  28332. the vlPAG also labeled cells in medial, ventral, and dorsolateral
  28333. orbital cortex and dorsal and posterior agranular insular cortex.
  28334. Other orbital and insular regions, including lateral and
  28335. ventrolateral orbital, ventral agranular insular, and dysgranular
  28336. and granular insular cortex did not give rise to appreciable
  28337. projections to the PAG. Anterograde tracing experiments revealed
  28338. that the projections to different PAG columns arose from specific
  28339. PFC areas. Projections from the caudodorsal medial PFC (caudal
  28340. prelimbic and anterior cingulate cortices) terminated
  28341. predominantly in dlPAG, whereas projections from the
  28342. rostroventral medial PFC (rostral prelimbic cortex) innervated
  28343. predominantly the vlPAG. As well, consistent with the retrograde
  28344. data, projections arising from select orbital and agranular
  28345. insular cortical areas terminated selectively in the vlPAG. The
  28346. results indicate: (1) that rat orbital and medial PFC possesses
  28347. an organisation broadly similar to that of the primate; and (2)
  28348. that subdivisions within the rat orbital and medial PFC can be
  28349. recognised on the basis of projections to distinct PAG columns.",
  28350. journal = "J. Comp. Neurol.",
  28351. volume = 422,
  28352. number = 4,
  28353. pages = "556--578",
  28354. month = jul,
  28355. year = 2000,
  28356. keywords = "Threat response",
  28357. language = "en"
  28358. }
  28359. @ARTICLE{Arber2018-kx,
  28360. title = "Connecting neuronal circuits for movement",
  28361. author = "Arber, Silvia and Costa, Rui M",
  28362. journal = "Science",
  28363. volume = 360,
  28364. number = 6396,
  28365. pages = "1403--1404",
  28366. month = jun,
  28367. year = 2018,
  28368. language = "en"
  28369. }
  28370. @ARTICLE{Gris2017-ow,
  28371. title = "Supervised and Unsupervised Learning Technology in the Study of
  28372. Rodent Behavior",
  28373. author = "Gris, Katsiaryna V and Coutu, Jean-Philippe and Gris, Denis",
  28374. abstract = "Quantifying behavior is a challenge for scientists studying
  28375. neuroscience, ethology, psychology, pathology, etc. Until now,
  28376. behavior was mostly considered as qualitative descriptions of
  28377. postures or labor intensive counting of bouts of individual
  28378. movements. Many prominent behavioral scientists conducted studies
  28379. describing postures of mice and rats, depicting step by step
  28380. eating, grooming, courting, and other behaviors. Automated video
  28381. assessment technologies permit scientists to quantify daily
  28382. behavioral patterns/routines, social interactions, and postural
  28383. changes in an unbiased manner. Here, we extensively reviewed
  28384. published research on the topic of the structural blocks of
  28385. behavior and proposed a structure of behavior based on the latest
  28386. publications. We discuss the importance of defining a clear
  28387. structure of behavior to allow professionals to write viable
  28388. algorithms. We presented a discussion of technologies that are
  28389. used in automated video assessment of behavior in mice and rats.
  28390. We considered advantages and limitations of supervised and
  28391. unsupervised learning. We presented the latest scientific
  28392. discoveries that were made using automated video assessment. In
  28393. conclusion, we proposed that the automated quantitative approach
  28394. to evaluating animal behavior is the future of understanding the
  28395. effect of brain signaling, pathologies, genetic content, and
  28396. environment on behavior.",
  28397. journal = "Front. Behav. Neurosci.",
  28398. volume = 11,
  28399. pages = "141",
  28400. month = jul,
  28401. year = 2017,
  28402. keywords = "animal behavior; automatic analysis; computer learning;
  28403. supervised; unsupervised;Analysis/Modelling [Behaviour]",
  28404. language = "en"
  28405. }
  28406. @ARTICLE{Robie2017-fu,
  28407. title = "Mapping the Neural Substrates of Behavior",
  28408. author = "Robie, Alice A and Hirokawa, Jonathan and Edwards, Austin W and
  28409. Umayam, Lowell A and Lee, Allen and Phillips, Mary L and Card,
  28410. Gwyneth M and Korff, Wyatt and Rubin, Gerald M and Simpson, Julie
  28411. H and Reiser, Michael B and Branson, Kristin",
  28412. abstract = "Assigning behavioral functions to neural structures has long been
  28413. a central goal in neuroscience and is a necessary first step
  28414. toward a circuit-level understanding of how the brain generates
  28415. behavior. Here, we map the neural substrates of locomotion and
  28416. social behaviors for Drosophila melanogaster using automated
  28417. machine-vision and machine-learning techniques. From videos of
  28418. 400,000 flies, we quantified the behavioral effects of activating
  28419. 2,204 genetically targeted populations of neurons. We combined a
  28420. novel quantification of anatomy with our behavioral analysis to
  28421. create brain-behavior correlation maps, which are shared as
  28422. browsable web pages and interactive software. Based on these
  28423. maps, we generated hypotheses of regions of the brain causally
  28424. related to sensory processing, locomotor control, courtship,
  28425. aggression, and sleep. Our maps directly specify genetic tools to
  28426. target these regions, which we used to identify a small
  28427. population of neurons with a role in the control of walking.",
  28428. journal = "Cell",
  28429. volume = 170,
  28430. number = 2,
  28431. pages = "393--406.e28",
  28432. month = jul,
  28433. year = 2017,
  28434. keywords = "Drosophila; behavior; computer vision; machine learning; neural
  28435. activation; neural anatomy; neural substrates; neuroscience;
  28436. whole-brain mapping;Analysis/Modelling [Behaviour]",
  28437. language = "en"
  28438. }
  28439. @ARTICLE{Grundemann2015-zj,
  28440. title = "Ensemble coding in amygdala circuits for associative learning",
  28441. author = "Gr{\"u}ndemann, Jan and L{\"u}thi, Andreas",
  28442. abstract = "Associative fear learning in the basolateral amygdala (BLA) is
  28443. crucial for an animal's survival upon environmental threats. BLA
  28444. neurons are defined on the basis of their projection target,
  28445. genetic markers, and associated function. BLA principal neuron
  28446. responses to threat signaling stimuli are potentiated upon
  28447. associative fear learning, which is tightly controlled by defined
  28448. interneuron subpopulations. In addition, BLA population activity
  28449. correlates with behavioral states and threat or safety signals.
  28450. BLA neuronal ensembles activated by different behavioral signals
  28451. can be identified using immediate early gene markers. The next
  28452. challenge will be to determine the activity patterns and coding
  28453. properties of defined BLA ensembles in relation to the whole
  28454. neuronal population.",
  28455. journal = "Curr. Opin. Neurobiol.",
  28456. volume = 35,
  28457. pages = "200--206",
  28458. month = dec,
  28459. year = 2015,
  28460. keywords = "Threat response;Authors/Luthi",
  28461. language = "en"
  28462. }
  28463. @ARTICLE{Colgin2016-ny,
  28464. title = "Rhythms of the hippocampal network",
  28465. author = "Colgin, Laura Lee",
  28466. abstract = "The hippocampal local field potential (LFP) shows three major
  28467. types of rhythms: theta, sharp wave-ripples and gamma. These
  28468. rhythms are defined by their frequencies, they have behavioural
  28469. correlates in several species including rats and humans, and they
  28470. have been proposed to carry out distinct functions in hippocampal
  28471. memory processing. However, recent findings have challenged
  28472. traditional views on these behavioural functions. In this Review,
  28473. I discuss our current understanding of the origins and the
  28474. mnemonic functions of hippocampal theta, sharp wave-ripples and
  28475. gamma rhythms on the basis of findings from rodent studies. In
  28476. addition, I present an updated synthesis of their roles and
  28477. interactions within the hippocampal network.",
  28478. journal = "Nat. Rev. Neurosci.",
  28479. volume = 17,
  28480. number = 4,
  28481. pages = "239--249",
  28482. month = apr,
  28483. year = 2016,
  28484. language = "en"
  28485. }
  28486. @ARTICLE{Bianco2015-sb,
  28487. title = "Visuomotor transformations underlying hunting behavior in
  28488. zebrafish",
  28489. author = "Bianco, Isaac H and Engert, Florian",
  28490. abstract = "Visuomotor circuits filter visual information and determine
  28491. whether or not to engage downstream motor modules to produce
  28492. behavioral outputs. However, the circuit mechanisms that mediate
  28493. and link perception of salient stimuli to execution of an
  28494. adaptive response are poorly understood. We combined a virtual
  28495. hunting assay for tethered larval zebrafish with two-photon
  28496. functional calcium imaging to simultaneously monitor neuronal
  28497. activity in the optic tectum during naturalistic behavior.
  28498. Hunting responses showed mixed selectivity for combinations of
  28499. visual features, specifically stimulus size, speed, and contrast
  28500. polarity. We identified a subset of tectal neurons with similar
  28501. highly selective tuning, which show non-linear mixed selectivity
  28502. for visual features and are likely to mediate the perceptual
  28503. recognition of prey. By comparing neural dynamics in the optic
  28504. tectum during response versus non-response trials, we discovered
  28505. premotor population activity that specifically preceded
  28506. initiation of hunting behavior and exhibited anatomical
  28507. localization that correlated with motor variables. In summary,
  28508. the optic tectum contains non-linear mixed selectivity neurons
  28509. that are likely to mediate reliable detection of ethologically
  28510. relevant sensory stimuli. Recruitment of small tectal assemblies
  28511. appears to link perception to action by providing the premotor
  28512. commands that release hunting responses. These findings allow us
  28513. to propose a model circuit for the visuomotor transformations
  28514. underlying a natural behavior.",
  28515. journal = "Curr. Biol.",
  28516. volume = 25,
  28517. number = 7,
  28518. pages = "831--846",
  28519. month = mar,
  28520. year = 2015,
  28521. language = "en"
  28522. }
  28523. @ARTICLE{Romero-Ferrero2018-dh,
  28524. title = "idtracker.ai: Tracking all individuals in large collectives of
  28525. unmarked animals",
  28526. author = "Romero-Ferrero, F and Bergomi, M and Hinz, Robert and Heras, F
  28527. and de Polavieja, G",
  28528. journal = "bioRxiv",
  28529. month = mar,
  28530. year = 2018,
  28531. keywords = "Analysis/Modelling [Behaviour]"
  28532. }
  28533. @ARTICLE{Skaggs1995-gd,
  28534. title = "A model of the neural basis of the rat's sense of direction",
  28535. author = "Skaggs, W E and Knierim, J J and Kudrimoti, H S and McNaughton, B
  28536. L",
  28537. abstract = "In the last decade the outlines of the neural structures
  28538. subserving the sense of direction have begun to emerge. Several
  28539. investigations have shed light on the effects of vestibular input
  28540. and visual input on the head direction representation. In this
  28541. paper, a model is formulated of the neural mechanisms underlying
  28542. the head direction system. The model is built out of simple
  28543. ingredients, depending on nothing more complicated than
  28544. connectional specificity, attractor dynamics, Hebbian learning,
  28545. and sigmoidal nonlinearities, but it behaves in a sophisticated
  28546. way and is consistent with most of the observed properties of
  28547. real head direction cells. In addition it makes a number of
  28548. predictions that ought to be testable by reasonably
  28549. straightforward experiments.",
  28550. journal = "Adv. Neural Inf. Process. Syst.",
  28551. volume = 7,
  28552. pages = "173--180",
  28553. year = 1995,
  28554. keywords = "Spatial Navigation",
  28555. language = "en"
  28556. }
  28557. @INCOLLECTION{Nitz2014-ca,
  28558. title = "The Posterior Parietal Cortex: Interface Between Maps of
  28559. External Spaces and the Generation of Action Sequences",
  28560. booktitle = "{Space,Time} and Memory in the Hippocampal Formation",
  28561. author = "Nitz, Douglas A",
  28562. editor = "Derdikman, Dori and Knierim, James J",
  28563. abstract = "In primates as well as rodents, the posterior parietal cortex
  28564. maps spatial relationships having both egocentric and external
  28565. frames of reference. In this chapter, the form in which rat
  28566. posterior parietal cortex neuronal activity maps position within
  28567. trajectories through the environment is considered in detail and
  28568. compared to the forms of spatial mapping observed for neurons of
  28569. the hippocampus and entorhinal cortex. Evidence is presented to
  28570. indicate that posterior parietal neurons simultaneously map
  28571. positions both within and across segments of paths through an
  28572. environment. It is suggested that the specific nature of
  28573. posterior parietal cortex mapping of space serves, in part, to
  28574. transition knowledge of position in the environment, given by
  28575. hippocampus and entorhinal cortex, into efficient path-running
  28576. behavior via projections to primary and secondary sensory and
  28577. motor cortices. Posterior parietal cortex activity is also
  28578. hypothesized to play a role both in driving trajectory
  28579. dependence of hippocampal place cells and in anchoring spatially
  28580. specific hippocampal and entorhinal cortical activity to the
  28581. boundaries of the observable environment.",
  28582. publisher = "Springer Vienna",
  28583. pages = "27--54",
  28584. year = 2014,
  28585. address = "Vienna",
  28586. keywords = "Spatial Navigation;Authors/Nitz"
  28587. }
  28588. @ARTICLE{Nitz2009-vz,
  28589. title = "Parietal cortex, navigation, and the construction of arbitrary
  28590. reference frames for spatial information",
  28591. author = "Nitz, Douglas",
  28592. abstract = "The registration of spatial information by neurons of the
  28593. parietal cortex takes on many forms. In most experiments,
  28594. spatially modulated parietal activity patterns are found to take
  28595. as their frame of reference some part of the body such as the
  28596. retina. However, recent findings obtained in single neuron
  28597. recordings from both rat and monkey parietal cortex suggest that
  28598. the frame of reference utilized by parietal cortex may also be
  28599. abstract or arbitrary in nature. Evidence in rats comes from work
  28600. indicating that parietal activity in freely behaving rodents is
  28601. organized according to the space defined by routes taken through
  28602. an environment. In monkeys, evidence for an object-centered frame
  28603. of reference has recently been presented. The present work
  28604. reviews single neuron recording experiments in parietal cortex of
  28605. freely behaving rats and considers the potential contribution of
  28606. parietal cortex in solving navigational tasks. It is proposed
  28607. that parietal cortex, in interaction with the hippocampus, plays
  28608. a critical role in the selection of the most appropriate route
  28609. between two points and, in addition, produces a route-based
  28610. positional signal capable of guiding sensorimotor transitions.",
  28611. journal = "Neurobiol. Learn. Mem.",
  28612. volume = 91,
  28613. number = 2,
  28614. pages = "179--185",
  28615. month = feb,
  28616. year = 2009,
  28617. keywords = "navigation;Spatial Navigation;Authors/Nitz",
  28618. language = "en"
  28619. }
  28620. @ARTICLE{Nitz2006-mn,
  28621. title = "Tracking route progression in the posterior parietal cortex",
  28622. author = "Nitz, Douglas A",
  28623. abstract = "Quick and efficient traversal of learned routes is critical to
  28624. the survival of many animals. Routes can be defined by both the
  28625. ordering of navigational epochs, such as continued forward motion
  28626. or execution of a turn, and the distances separating them. The
  28627. neural substrates conferring the ability to fluidly traverse
  28628. complex routes are not well understood, but likely entail
  28629. interactions between frontal, parietal, and rhinal cortices and
  28630. the hippocampus. This paper demonstrates that posterior parietal
  28631. cortical neurons map both individual and multiple navigational
  28632. epochs with respect to their order in a route. In direct contrast
  28633. to spatial firing patterns of hippocampal neurons, parietal
  28634. neurons discharged in a place- and direction-independent fashion.
  28635. Parietal route maps were scalable and versatile in that they were
  28636. independent of the size and spatial configuration of navigational
  28637. epochs. The results provide a framework in which to consider
  28638. parietal function in spatial cognition.",
  28639. journal = "Neuron",
  28640. volume = 49,
  28641. number = 5,
  28642. pages = "747--756",
  28643. month = mar,
  28644. year = 2006,
  28645. keywords = "navigation;Spatial Navigation;Authors/Nitz",
  28646. language = "en"
  28647. }