example_library.bib 160 KB

1234567891011121314151617181920212223242526272829303132333435363738394041424344454647484950515253545556575859606162636465666768697071727374757677787980818283848586878889909192939495969798991001011021031041051061071081091101111121131141151161171181191201211221231241251261271281291301311321331341351361371381391401411421431441451461471481491501511521531541551561571581591601611621631641651661671681691701711721731741751761771781791801811821831841851861871881891901911921931941951961971981992002012022032042052062072082092102112122132142152162172182192202212222232242252262272282292302312322332342352362372382392402412422432442452462472482492502512522532542552562572582592602612622632642652662672682692702712722732742752762772782792802812822832842852862872882892902912922932942952962972982993003013023033043053063073083093103113123133143153163173183193203213223233243253263273283293303313323333343353363373383393403413423433443453463473483493503513523533543553563573583593603613623633643653663673683693703713723733743753763773783793803813823833843853863873883893903913923933943953963973983994004014024034044054064074084094104114124134144154164174184194204214224234244254264274284294304314324334344354364374384394404414424434444454464474484494504514524534544554564574584594604614624634644654664674684694704714724734744754764774784794804814824834844854864874884894904914924934944954964974984995005015025035045055065075085095105115125135145155165175185195205215225235245255265275285295305315325335345355365375385395405415425435445455465475485495505515525535545555565575585595605615625635645655665675685695705715725735745755765775785795805815825835845855865875885895905915925935945955965975985996006016026036046056066076086096106116126136146156166176186196206216226236246256266276286296306316326336346356366376386396406416426436446456466476486496506516526536546556566576586596606616626636646656666676686696706716726736746756766776786796806816826836846856866876886896906916926936946956966976986997007017027037047057067077087097107117127137147157167177187197207217227237247257267277287297307317327337347357367377387397407417427437447457467477487497507517527537547557567577587597607617627637647657667677687697707717727737747757767777787797807817827837847857867877887897907917927937947957967977987998008018028038048058068078088098108118128138148158168178188198208218228238248258268278288298308318328338348358368378388398408418428438448458468478488498508518528538548558568578588598608618628638648658668678688698708718728738748758768778788798808818828838848858868878888898908918928938948958968978988999009019029039049059069079089099109119129139149159169179189199209219229239249259269279289299309319329339349359369379389399409419429439449459469479489499509519529539549559569579589599609619629639649659669679689699709719729739749759769779789799809819829839849859869879889899909919929939949959969979989991000100110021003100410051006100710081009101010111012101310141015101610171018101910201021102210231024102510261027102810291030103110321033103410351036103710381039104010411042104310441045104610471048104910501051105210531054105510561057105810591060106110621063106410651066106710681069107010711072107310741075107610771078107910801081108210831084108510861087108810891090109110921093109410951096109710981099110011011102110311041105110611071108110911101111111211131114111511161117111811191120112111221123112411251126112711281129113011311132113311341135113611371138113911401141114211431144114511461147114811491150115111521153115411551156115711581159116011611162116311641165116611671168116911701171117211731174117511761177117811791180118111821183118411851186118711881189119011911192119311941195119611971198119912001201120212031204120512061207120812091210121112121213121412151216121712181219122012211222122312241225122612271228122912301231123212331234123512361237123812391240124112421243124412451246124712481249125012511252125312541255125612571258125912601261126212631264126512661267126812691270127112721273127412751276127712781279128012811282128312841285128612871288128912901291129212931294129512961297129812991300130113021303130413051306130713081309131013111312131313141315131613171318131913201321132213231324132513261327132813291330133113321333133413351336133713381339134013411342134313441345134613471348134913501351135213531354135513561357135813591360136113621363136413651366136713681369137013711372137313741375137613771378137913801381138213831384138513861387138813891390139113921393139413951396139713981399140014011402140314041405140614071408140914101411141214131414141514161417141814191420142114221423142414251426142714281429143014311432143314341435143614371438143914401441144214431444144514461447144814491450145114521453145414551456145714581459146014611462146314641465146614671468146914701471147214731474147514761477147814791480148114821483148414851486148714881489149014911492149314941495149614971498149915001501150215031504150515061507150815091510151115121513151415151516151715181519152015211522152315241525152615271528152915301531153215331534153515361537153815391540154115421543154415451546154715481549155015511552155315541555155615571558155915601561156215631564156515661567156815691570157115721573157415751576157715781579158015811582158315841585158615871588158915901591159215931594159515961597159815991600160116021603160416051606160716081609161016111612161316141615161616171618161916201621162216231624162516261627162816291630163116321633163416351636163716381639164016411642164316441645164616471648164916501651165216531654165516561657165816591660166116621663166416651666166716681669167016711672167316741675167616771678167916801681168216831684168516861687168816891690169116921693169416951696169716981699170017011702170317041705170617071708170917101711171217131714171517161717171817191720172117221723172417251726172717281729173017311732173317341735173617371738173917401741174217431744174517461747174817491750175117521753175417551756175717581759176017611762176317641765176617671768176917701771177217731774177517761777177817791780178117821783178417851786178717881789179017911792179317941795179617971798179918001801180218031804180518061807180818091810181118121813181418151816181718181819182018211822182318241825182618271828182918301831183218331834183518361837183818391840184118421843184418451846184718481849185018511852185318541855185618571858185918601861186218631864186518661867186818691870187118721873187418751876187718781879188018811882188318841885188618871888188918901891189218931894189518961897189818991900190119021903190419051906190719081909191019111912191319141915191619171918191919201921192219231924192519261927192819291930193119321933193419351936193719381939194019411942194319441945194619471948194919501951195219531954195519561957195819591960196119621963196419651966196719681969197019711972197319741975197619771978197919801981198219831984198519861987198819891990199119921993199419951996199719981999200020012002200320042005200620072008200920102011201220132014201520162017201820192020202120222023202420252026202720282029203020312032203320342035203620372038203920402041204220432044204520462047204820492050205120522053205420552056205720582059206020612062206320642065206620672068206920702071207220732074207520762077207820792080208120822083208420852086208720882089209020912092209320942095209620972098209921002101210221032104210521062107210821092110211121122113211421152116211721182119212021212122212321242125212621272128212921302131213221332134213521362137213821392140214121422143214421452146214721482149215021512152215321542155215621572158215921602161216221632164216521662167216821692170217121722173217421752176217721782179218021812182218321842185218621872188218921902191219221932194219521962197219821992200220122022203220422052206220722082209221022112212221322142215221622172218221922202221222222232224222522262227222822292230223122322233223422352236223722382239224022412242224322442245224622472248224922502251225222532254225522562257225822592260226122622263226422652266226722682269227022712272227322742275227622772278227922802281228222832284228522862287228822892290229122922293229422952296229722982299230023012302230323042305230623072308230923102311231223132314231523162317231823192320232123222323232423252326232723282329233023312332233323342335233623372338233923402341234223432344234523462347234823492350235123522353235423552356235723582359236023612362236323642365236623672368236923702371237223732374237523762377237823792380238123822383238423852386238723882389239023912392239323942395239623972398239924002401240224032404240524062407240824092410241124122413241424152416241724182419242024212422242324242425242624272428242924302431243224332434243524362437243824392440244124422443244424452446244724482449245024512452245324542455245624572458245924602461246224632464246524662467246824692470247124722473247424752476247724782479248024812482248324842485248624872488248924902491249224932494249524962497249824992500250125022503250425052506250725082509251025112512251325142515251625172518251925202521252225232524252525262527252825292530253125322533253425352536253725382539254025412542254325442545254625472548254925502551255225532554255525562557255825592560256125622563256425652566256725682569257025712572257325742575257625772578257925802581258225832584258525862587258825892590259125922593259425952596259725982599260026012602260326042605260626072608260926102611261226132614261526162617261826192620262126222623262426252626262726282629263026312632263326342635263626372638263926402641264226432644264526462647264826492650265126522653265426552656265726582659266026612662266326642665266626672668266926702671267226732674267526762677267826792680268126822683268426852686268726882689269026912692269326942695269626972698269927002701270227032704270527062707270827092710271127122713271427152716271727182719272027212722272327242725272627272728272927302731273227332734273527362737273827392740274127422743274427452746274727482749275027512752275327542755275627572758275927602761276227632764276527662767276827692770277127722773277427752776277727782779278027812782278327842785278627872788278927902791279227932794279527962797279827992800280128022803280428052806280728082809281028112812281328142815281628172818281928202821282228232824282528262827282828292830283128322833283428352836283728382839284028412842284328442845284628472848284928502851285228532854285528562857285828592860286128622863286428652866286728682869287028712872287328742875287628772878287928802881288228832884288528862887288828892890289128922893289428952896289728982899290029012902290329042905290629072908290929102911291229132914291529162917291829192920292129222923292429252926292729282929293029312932293329342935293629372938293929402941294229432944294529462947294829492950295129522953295429552956295729582959296029612962296329642965296629672968296929702971297229732974
  1. % Generated by Paperpile. Check out https://paperpile.com for more information.
  2. % BibTeX export options can be customized via Settings -> BibTeX.
  3. % The entry below contains non-ASCII chars that could not be converted
  4. % to a LaTeX equivalent.
  5. @ARTICLE{Mizumori2005-mq,
  6. title = "Head direction codes in hippocampal afferent and efferent
  7. systems: what functions do they serve",
  8. author = "Mizumori, Sheri J Y and Puryear, Corey B and Gill, Kathryn M and
  9. Guazzelli, Alex",
  10. abstract = "The discovery of head direction cells in many limbic and
  11. limbic-afferent structures could lead one to suggest that the
  12. limbic system is specialized for spatial analysis, and in
  13. particular the processing of directional orientation.
  14. Considering this hypothesis, it becomes important to know
  15. whether head direction codes are unique to the limbic system.
  16. Previous chapters provide convincing evidence that the mechanism
  17. for the generation of head direction signals involves sequential
  18. processing through the tegmentum, mammillary nucleus, anterior
  19. dorsal …",
  20. journal = "Head direction cells and the neural mechanisms of spatial
  21. orientation",
  22. publisher = "books.google.com",
  23. pages = "203--220",
  24. year = 2005
  25. }
  26. @ARTICLE{Melzer2017-ek,
  27. title = "Distinct Corticostriatal {GABAergic} Neurons Modulate Striatal
  28. Output Neurons and Motor Activity",
  29. author = "Melzer, Sarah and Gil, Mariana and Koser, David E and Michael,
  30. Magdalena and Huang, Kee Wui and Monyer, Hannah",
  31. abstract = "The motor cortico-basal ganglion loop is critical for motor
  32. planning, execution, and learning. Balanced excitation and
  33. inhibition in this loop is crucial for proper motor output.
  34. Excitatory neurons have been thought to be the only source of
  35. motor cortical input to the striatum. Here, we identify
  36. long-range projecting GABAergic neurons in the primary (M1) and
  37. secondary (M2) motor cortex that target the dorsal striatum.
  38. This population of projecting GABAergic neurons comprises both
  39. somatostatin-positive (SOM+) and parvalbumin-positive (PV+)
  40. neurons that target direct and indirect pathway striatal output
  41. neurons as well as cholinergic interneurons differentially.
  42. Notably, optogenetic stimulation of M1 PV+ and M2 SOM+
  43. projecting neurons reduced locomotion, whereas stimulation of M1
  44. SOM+ projecting neurons enhanced locomotion. Thus,
  45. corticostriatal GABAergic projections modulate striatal output
  46. and motor activity.",
  47. journal = "Cell Rep.",
  48. publisher = "Elsevier",
  49. volume = 19,
  50. number = 5,
  51. pages = "1045--1055",
  52. month = may,
  53. year = 2017,
  54. keywords = "GABA; locomotion; long-range; motor cortex; optogenetics;
  55. parvalbumin; somatostatin; striatum;Locomotion",
  56. language = "en",
  57. issn = "2211-1247",
  58. pmid = "28467898",
  59. doi = "10.1016/j.celrep.2017.04.024",
  60. pmc = "PMC5437725"
  61. }
  62. @ARTICLE{Murakami2014-lh,
  63. title = "Neural antecedents of self-initiated actions in secondary motor
  64. cortex",
  65. author = "Murakami, Masayoshi and Vicente, M In{\^e}s and Costa, Gil M and
  66. Mainen, Zachary F",
  67. abstract = "The neural origins of spontaneous or self-initiated actions are
  68. not well understood and their interpretation is controversial.
  69. To address these issues, we used a task in which rats decide
  70. when to abort waiting for a delayed tone. We recorded neurons in
  71. the secondary motor cortex (M2) and interpreted our findings in
  72. light of an integration-to-bound decision model. A first
  73. population of M2 neurons ramped to a constant threshold at rates
  74. proportional to waiting time, strongly resembling integrator
  75. output. A second population, which we propose provide input to
  76. the integrator, fired in sequences and showed trial-to-trial
  77. rate fluctuations correlated with waiting times. An integration
  78. model fit to these data also quantitatively predicted the
  79. observed inter-neuronal correlations. Together, these results
  80. reinforce the generality of the integration-to-bound model of
  81. decision-making. These models identify the initial intention to
  82. act as the moment of threshold crossing while explaining how
  83. antecedent subthreshold neural activity can influence an action
  84. without implying a decision.",
  85. journal = "Nat. Neurosci.",
  86. publisher = "nature.com",
  87. volume = 17,
  88. number = 11,
  89. pages = "1574--1582",
  90. month = nov,
  91. year = 2014,
  92. language = "en",
  93. issn = "1097-6256, 1546-1726",
  94. pmid = "25262496",
  95. doi = "10.1038/nn.3826"
  96. }
  97. @ARTICLE{Gremel2013-im,
  98. title = "Premotor cortex is critical for goal-directed actions",
  99. author = "Gremel, Christina M and Costa, Rui M",
  100. abstract = "Shifting between motor plans is often necessary for adaptive
  101. behavior. When faced with changing consequences of one's
  102. actions, it is often imperative to switch from automatic actions
  103. to deliberative and controlled actions. The pre-supplementary
  104. motor area (pre-SMA) in primates, akin to the premotor cortex
  105. (M2) in mice, has been implicated in motor learning and
  106. planning, and action switching. We hypothesized that M2 would be
  107. differentially involved in goal-directed actions, which are
  108. controlled by their consequences vs. habits, which are more
  109. dependent on their past reinforcement history and less on their
  110. consequences. To investigate this, we performed M2 lesions in
  111. mice and then concurrently trained them to press the same lever
  112. for the same food reward using two different schedules of
  113. reinforcement that differentially bias towards the use of
  114. goal-directed versus habitual action strategies. We then probed
  115. whether actions were dependent on their expected consequence
  116. through outcome revaluation testing. We uncovered that M2
  117. lesions did not affect the acquisition of lever-pressing.
  118. However, in mice with M2 lesions, lever-pressing was insensitive
  119. to changes in expected outcome value following goal-directed
  120. training. However, habitual actions were intact. We confirmed a
  121. role for M2 in goal-directed but not habitual actions in
  122. separate groups of mice trained on the individual schedules
  123. biasing towards goal-directed versus habitual actions. These
  124. data indicate that M2 is critical for actions to be updated
  125. based on their consequences, and suggest that habitual action
  126. strategies may not require processing by M2 and the updating of
  127. motor plans.",
  128. journal = "Front. Comput. Neurosci.",
  129. publisher = "frontiersin.org",
  130. volume = 7,
  131. pages = "110",
  132. month = aug,
  133. year = 2013,
  134. keywords = "action selection; goal-directed actions; habitual actions;
  135. premotor cortex; value-based decision making",
  136. language = "en",
  137. issn = "1662-5188",
  138. pmid = "23964233",
  139. doi = "10.3389/fncom.2013.00110",
  140. pmc = "PMC3740264"
  141. }
  142. @ARTICLE{Erlich2011-rn,
  143. title = "A cortical substrate for memory-guided orienting in the rat",
  144. author = "Erlich, Jeffrey C and Bialek, Max and Brody, Carlos D",
  145. abstract = "Anatomical, stimulation, and lesion data have suggested a
  146. homology between the rat frontal orienting fields (FOF)
  147. (centered at +2 AP, $\pm$1.3 ML mm from Bregma) and primate
  148. frontal cortices such as the frontal or supplementary eye
  149. fields. We investigated the functional role of the FOF using
  150. rats trained to perform a memory-guided orienting task, in which
  151. there was a delay period between the end of a sensory stimulus
  152. instructing orienting direction and the time of the allowed
  153. motor response. Unilateral inactivation of the FOF resulted in
  154. impaired contralateral responses. Extracellular recordings of
  155. single units revealed that 37\% of FOF neurons had delay period
  156. firing rates that predicted the direction of the rats' later
  157. orienting motion. Our data provide the first
  158. electrophysiological and pharmacological evidence supporting the
  159. existence in the rat, as in the primate, of a frontal cortical
  160. area involved in the preparation and/or planning of orienting
  161. responses.",
  162. journal = "Neuron",
  163. publisher = "Elsevier",
  164. volume = 72,
  165. number = 2,
  166. pages = "330--343",
  167. month = oct,
  168. year = 2011,
  169. keywords = "To Read;navigation",
  170. language = "en",
  171. issn = "0896-6273, 1097-4199",
  172. pmid = "22017991",
  173. doi = "10.1016/j.neuron.2011.07.010",
  174. pmc = "PMC3212026"
  175. }
  176. @ARTICLE{Jeong2016-dq,
  177. title = "Comparative three-dimensional connectome map of motor cortical
  178. projections in the mouse brain",
  179. author = "Jeong, Minju and Kim, Yongsoo and Kim, Jeongjin and Ferrante,
  180. Daniel D and Mitra, Partha P and Osten, Pavel and Kim, Daesoo",
  181. abstract = "The motor cortex orchestrates simple to complex motor behaviors
  182. through its output projections to target areas. The primary (MOp)
  183. and secondary (MOs) motor cortices are known to produce specific
  184. output projections that are targeted to both similar and
  185. different target areas. These projections are further divided
  186. into layer 5 and 6 neuronal outputs, thereby producing four
  187. cortical outputs that may target other areas in a combinatorial
  188. manner. However, the precise network structure that integrates
  189. these four projections remains poorly understood. Here, we
  190. constructed a whole-brain, three-dimensional (3D) map showing the
  191. tract pathways and targeting locations of these four motor
  192. cortical outputs in mice. Remarkably, these motor cortical
  193. projections showed unique and separate tract pathways despite
  194. targeting similar areas. Within target areas, various
  195. combinations of these four projections were defined based on
  196. specific 3D spatial patterns, reflecting anterior-posterior,
  197. dorsal-ventral, and core-capsular relationships. This 3D
  198. topographic map ultimately provides evidence for the relevance of
  199. comparative connectomics: motor cortical projections known to be
  200. convergent are actually segregated in many target areas with
  201. unique targeting patterns, a finding that has anatomical value
  202. for revealing functional subdomains that have not been classified
  203. by conventional methods.",
  204. journal = "Sci. Rep.",
  205. volume = 6,
  206. pages = "20072",
  207. month = feb,
  208. year = 2016,
  209. keywords = "To Read",
  210. language = "en",
  211. issn = "2045-2322",
  212. pmid = "26830143",
  213. doi = "10.1038/srep20072",
  214. pmc = "PMC4735720"
  215. }
  216. % The entry below contains non-ASCII chars that could not be converted
  217. % to a LaTeX equivalent.
  218. @ARTICLE{Gradinaru2007-qv,
  219. title = "Targeting and readout strategies for fast optical neural control
  220. in vitro and in vivo",
  221. author = "Gradinaru, Viviana and Thompson, Kimberly R and Zhang, Feng and
  222. Mogri, Murtaza and Kay, Kenneth and Schneider, M Bret and
  223. Deisseroth, Karl",
  224. abstract = "Major obstacles faced by neuroscientists in attempting to
  225. unravel the complexity of brain function include both the
  226. heterogeneity of brain tissue (with a multitude of cell types
  227. present in vivo) and the high speed of brain information
  228. processing (with behaviorally relevant millisecondscale
  229. electrical activity patterns). To address different aspects of
  230. these technical constraints, genetically targetable neural
  231. modulation tools have been developed by a number of groups
  232. (Zemelman et al., 2002; Banghart et al., 2004; Karpova et al.,
  233. 2005; Lima …",
  234. journal = "J. Neurosci.",
  235. publisher = "Soc Neuroscience",
  236. volume = 27,
  237. number = 52,
  238. pages = "14231--14238",
  239. month = dec,
  240. year = 2007,
  241. keywords = "Locomotion",
  242. language = "en",
  243. issn = "0270-6474, 1529-2401",
  244. pmid = "18160630",
  245. doi = "10.1523/JNEUROSCI.3578-07.2007",
  246. pmc = "PMC6673457"
  247. }
  248. @ARTICLE{Magno2019-qz,
  249. title = "Optogenetic Stimulation of the {M2} Cortex Reverts Motor
  250. Dysfunction in a Mouse Model of Parkinson's Disease",
  251. author = "Magno, Luiz Alexandre Viana and Tenza-Ferrer, Helia and
  252. Collodetti, M{\'e}lcar and Aguiar, Matheus Felipe Guimar{\~a}es
  253. and Rodrigues, Ana Paula Carneiro and da Silva, Rodrigo Souza and
  254. Silva, Joice do Prado and Nicolau, Nycolle Ferreira and Rosa,
  255. Daniela Valad{\~a}o Freitas and Birbrair, Alexander and Miranda,
  256. D{\'e}bora Marques and Romano-Silva, Marco Aur{\'e}lio",
  257. abstract = "Neuromodulation of deep brain structures (deep brain stimulation)
  258. is the current surgical procedure for treatment of Parkinson's
  259. disease (PD). Less studied is the stimulation of cortical motor
  260. areas to treat PD symptoms, although also known to alleviate
  261. motor disturbances in PD. We were able to show that optogenetic
  262. activation of secondary (M2) motor cortex improves motor
  263. functions in dopamine-depleted male mice. The stimulated M2
  264. cortex harbors glutamatergic pyramidal neurons that project to
  265. subcortical structures, critically involved in motor control, and
  266. makes synaptic contacts with dopaminergic neurons. Strikingly,
  267. optogenetic activation of M2 neurons or axons into the
  268. dorsomedial striatum increases striatal levels of dopamine and
  269. evokes locomotor activity. We found that dopamine
  270. neurotransmission sensitizes the locomotor behavior elicited by
  271. activation of M2 neurons. Furthermore, combination of intranigral
  272. infusion of glutamatergic antagonists and circuit specific
  273. optogenetic stimulation revealed that behavioral response
  274. depended on the activity of M2 neurons projecting to SNc.
  275. Interestingly, repeated M2 stimulation combined with l-DOPA
  276. treatment produced an unanticipated improvement in working memory
  277. performance, which was absent in control mice under l-DOPA
  278. treatment only. Therefore, the M2-basal ganglia circuit is
  279. critical for the assembly of the motor and cognitive function,
  280. and this study demonstrates a therapeutic mechanism for cortical
  281. stimulation in PD that involves recruitment of long-range
  282. glutamatergic projection neurons.SIGNIFICANCE STATEMENT Some
  283. patients with Parkinson's disease are offered treatment through
  284. surgery, which consists of delivering electrical current to
  285. regions deep within the brain. This study shows that stimulation
  286. of an area located on the brain surface, known as the secondary
  287. motor cortex, can also reverse movement disorders in mice.
  288. Authors have used a brain stimulation technique called
  289. optogenetics, which allowed targeting a specific type of surface
  290. neuron that communicates with the deep part of the brain involved
  291. in movement control. The study also shows that a combination of
  292. this stimulation with drug treatment might be useful to treat
  293. memory impairment, a kind of cognitive problem in Parkinson's
  294. disease.",
  295. journal = "J. Neurosci.",
  296. volume = 39,
  297. number = 17,
  298. pages = "3234--3248",
  299. month = apr,
  300. year = 2019,
  301. keywords = "Parkinson's disorder; brain stimulation; cognition; movement;
  302. optogenetics; prefrontal cortex;Locomotion",
  303. language = "en",
  304. issn = "0270-6474, 1529-2401",
  305. pmid = "30782975",
  306. doi = "10.1523/JNEUROSCI.2277-18.2019",
  307. pmc = "PMC6788829"
  308. }
  309. @ARTICLE{Schiemann2015-th,
  310. title = "Cellular mechanisms underlying behavioral state-dependent
  311. bidirectional modulation of motor cortex output",
  312. author = "Schiemann, Julia and Puggioni, Paolo and Dacre, Joshua and
  313. Pelko, Miha and Domanski, Aleksander and van Rossum, Mark C W
  314. and Duguid, Ian",
  315. abstract = "Neuronal activity in primary motor cortex (M1) correlates with
  316. behavioral state, but the cellular mechanisms underpinning
  317. behavioral state-dependent modulation of M1 output remain
  318. largely unresolved. Here, we performed in vivo patch-clamp
  319. recordings from layer 5B (L5B) pyramidal neurons in awake mice
  320. during quiet wakefulness and self-paced, voluntary movement. We
  321. show that L5B output neurons display bidirectional (i.e.,
  322. enhanced or suppressed) firing rate changes during movement,
  323. mediated via two opposing subthreshold mechanisms: (1) a global
  324. decrease in membrane potential variability that reduced L5B
  325. firing rates (L5Bsuppressed neurons), and (2) a coincident
  326. noradrenaline-mediated increase in excitatory drive to a
  327. subpopulation of L5B neurons (L5Benhanced neurons) that elevated
  328. firing rates. Blocking noradrenergic receptors in forelimb M1
  329. abolished the bidirectional modulation of M1 output during
  330. movement and selectively impaired contralateral forelimb motor
  331. coordination. Together, our results provide a mechanism for how
  332. noradrenergic neuromodulation and network-driven input changes
  333. bidirectionally modulate M1 output during motor behavior.",
  334. journal = "Cell Rep.",
  335. publisher = "Elsevier",
  336. volume = 11,
  337. number = 8,
  338. pages = "1319--1330",
  339. month = may,
  340. year = 2015,
  341. keywords = "To Read",
  342. language = "en",
  343. issn = "2211-1247",
  344. pmid = "25981037",
  345. doi = "10.1016/j.celrep.2015.04.042",
  346. pmc = "PMC4451462"
  347. }
  348. @ARTICLE{Ebbesen2017-cm,
  349. title = "Motor cortex - to act or not to act?",
  350. author = "Ebbesen, Christian Laut and Brecht, Michael",
  351. abstract = "The motor cortex is a large frontal structure in the cerebral
  352. cortex of eutherian mammals. A vast array of evidence implicates
  353. the motor cortex in the volitional control of motor output, but
  354. how does the motor cortex exert this 'control'? Historically,
  355. ideas regarding motor cortex function have been shaped by the
  356. discovery of cortical 'motor maps' - that is, ordered
  357. representations of stimulation-evoked movements in anaesthetized
  358. animals. Volitional control, however, entails the initiation of
  359. movements and the ability to suppress undesired movements. In
  360. this article, we highlight classic and recent findings that
  361. emphasize that motor cortex neurons have a role in both
  362. processes.",
  363. journal = "Nat. Rev. Neurosci.",
  364. volume = 18,
  365. number = 11,
  366. pages = "694--705",
  367. month = oct,
  368. year = 2017,
  369. language = "en",
  370. issn = "1471-003X, 1471-0048",
  371. pmid = "29042690",
  372. doi = "10.1038/nrn.2017.119"
  373. }
  374. @ARTICLE{Calton2009-hj,
  375. title = "Where am {I} and how will {I} get there from here? A role for
  376. posterior parietal cortex in the integration of spatial
  377. information and route planning",
  378. author = "Calton, Jeffrey L and Taube, Jeffrey S",
  379. abstract = "The ability of an organism to accurately navigate from one place
  380. to another requires integration of multiple spatial constructs,
  381. including the determination of one's position and direction in
  382. space relative to allocentric landmarks, movement velocity, and
  383. the perceived location of the goal of the movement. In this
  384. review, we propose that while limbic areas are important for the
  385. sense of spatial orientation, the posterior parietal cortex is
  386. responsible for relating this sense with the location of a
  387. navigational goal and in formulating a plan to attain it. Hence,
  388. the posterior parietal cortex is important for the computation of
  389. the correct trajectory or route to be followed while navigating.
  390. Prefrontal and motor areas are subsequently responsible for
  391. executing the planned movement. Using this theory, we are able to
  392. bridge the gap between the rodent and primate literatures by
  393. suggesting that the allocentric role of the rodent PPC is largely
  394. analogous to the egocentric role typically emphasized in
  395. primates, that is, the integration of spatial orientation with
  396. potential goals in the planning of goal-directed movements.",
  397. journal = "Neurobiol. Learn. Mem.",
  398. volume = 91,
  399. number = 2,
  400. pages = "186--196",
  401. month = feb,
  402. year = 2009,
  403. keywords = "navigation;To Read",
  404. language = "en",
  405. issn = "1074-7427, 1095-9564",
  406. pmid = "18929674",
  407. doi = "10.1016/j.nlm.2008.09.015",
  408. pmc = "PMC2666283"
  409. }
  410. @ARTICLE{Cho2001-nw,
  411. title = "Head direction, place, and movement correlates for cells in the
  412. rat retrosplenial cortex",
  413. author = "Cho, J and Sharp, P E",
  414. abstract = "The retrosplenial cortex is strongly connected with brain
  415. regions involved in spatial signaling. To test whether it also
  416. codes space, single cells were recorded while rats navigated in
  417. an open field. As in earlier work (L. L. Chen, L. H. Lin, C. A.
  418. Barnes, \& B. L. McNaughton, 1994; L. L. Chen, L. H. Lin, E. J.
  419. Green, C. A. Barnes, \& B. L. McNaughton, 1994), the authors
  420. found head direction cells with properties similar to those in
  421. other areas. These cells were slightly anticipatory. Another
  422. cell type fired to particular combinations of location,
  423. direction, and movement, which suggested that they may fire
  424. whenever the rat approaches a particular location, using a
  425. particular locomotor behavior. The remaining cells could not be
  426. clearly categorized but also showed a significant correlation
  427. with one or more of the spatial-movement variables examined. The
  428. fact that the retrosplenial cortex contains spatial and
  429. movement-related signals and is connected with the motor cortex
  430. suggests that it may play a role in path integration or
  431. navigational motor planning.",
  432. journal = "Behav. Neurosci.",
  433. publisher = "psycnet.apa.org",
  434. volume = 115,
  435. number = 1,
  436. pages = "3--25",
  437. month = feb,
  438. year = 2001,
  439. keywords = "navigation;To Read",
  440. language = "en",
  441. issn = "0735-7044",
  442. pmid = "11256450",
  443. doi = "10.1037/0735-7044.115.1.3"
  444. }
  445. @ARTICLE{Li2015-tj,
  446. title = "A motor cortex circuit for motor planning and movement",
  447. author = "Li, Nuo and Chen, Tsai-Wen and Guo, Zengcai V and Gerfen, Charles
  448. R and Svoboda, Karel",
  449. abstract = "Activity in motor cortex predicts specific movements seconds
  450. before they occur, but how this preparatory activity relates to
  451. upcoming movements is obscure. We dissected the conversion of
  452. preparatory activity to movement within a structured motor cortex
  453. circuit. An anterior lateral region of the mouse cortex (a
  454. possible homologue of premotor cortex in primates) contains equal
  455. proportions of intermingled neurons predicting ipsi- or
  456. contralateral movements, yet unilateral inactivation of this
  457. cortical region during movement planning disrupts contralateral
  458. movements. Using cell-type-specific electrophysiology, cellular
  459. imaging and optogenetic perturbation, we show that layer 5
  460. neurons projecting within the cortex have unbiased laterality.
  461. Activity with a contralateral population bias arises specifically
  462. in layer 5 neurons projecting to the brainstem, and only late
  463. during movement planning. These results reveal the transformation
  464. of distributed preparatory activity into movement commands within
  465. hierarchically organized cortical circuits.",
  466. journal = "Nature",
  467. volume = 519,
  468. number = 7541,
  469. pages = "51--56",
  470. month = mar,
  471. year = 2015,
  472. keywords = "To Read",
  473. language = "en",
  474. issn = "0028-0836, 1476-4687",
  475. pmid = "25731172",
  476. doi = "10.1038/nature14178"
  477. }
  478. @ARTICLE{McNaughton1994-hv,
  479. title = "Cortical representation of motion during unrestrained spatial
  480. navigation in the rat",
  481. author = "McNaughton, B L and Mizumori, S J and Barnes, C A and Leonard, B
  482. J and Marquis, M and Green, E J",
  483. abstract = "Neural activity related to unrestrained movement through space
  484. was studied in rat sensorimotor and posterior parietal cortices
  485. during performance of an eight-arm, radial maze task. Nearly half
  486. of the cells exhibited movement-related activity that
  487. discriminated among three basic modes of locomotion: left turns,
  488. right turns, and forward motion. Correlates ranged from strong
  489. excitation (relative to the still condition) to strong
  490. inhibition, and were distributed among the movement modes in a
  491. variety of different ways. For example, cells that discriminated
  492. between clockwise and counterclockwise turns did so with either
  493. antagonistic responses or simple excitation or inhibition. Others
  494. showed either excitation or inhibition relative to both turning
  495. and the still condition, and hence were selective for forward
  496. motion. Many cells exhibited somatosensory responsiveness;
  497. however, in agreement with findings of others, motion correlates
  498. could rarely be sensibly explained by the somatosensory response.
  499. Moreover, movement correlates sometimes varied considerably with
  500. spatial context. Some cells exhibited more complex motion
  501. correlates, such as an apparent dependence on the nature of the
  502. preceding movement. Irrespective of the specific sensory or motor
  503. determinants of cell activity, which varied considerably among
  504. cells, the posterior neocortex of the rat appears to generate a
  505. robust and redundant internal representation of body motion
  506. through space. Such a representation could be useful in
  507. constructing ``cognitive maps'' of the environment.",
  508. journal = "Cereb. Cortex",
  509. volume = 4,
  510. number = 1,
  511. pages = "27--39",
  512. month = jan,
  513. year = 1994,
  514. keywords = "navigation",
  515. language = "en",
  516. issn = "1047-3211",
  517. pmid = "8180489",
  518. doi = "10.1093/cercor/4.1.27"
  519. }
  520. @ARTICLE{Kiehn2006-wi,
  521. title = "Locomotor circuits in the mammalian spinal cord",
  522. author = "Kiehn, Ole",
  523. abstract = "Intrinsic spinal networks, known as central pattern generators
  524. (CPGs), control the timing and pattern of the muscle activity
  525. underlying locomotion in mammals. This review discusses new
  526. advances in understanding the mammalian CPGs with a focus on
  527. experiments that address the overall network structure as well
  528. as the identification of CPG neurons. I address the
  529. identification of excitatory CPG neurons and their role in
  530. rhythm generation, the organization of flexor-extensor networks,
  531. and the diverse role of commissural interneurons in coordinating
  532. left-right movements. Molecular and genetic approaches that have
  533. the potential to elucidate the function of populations of CPG
  534. interneurons are also discussed.",
  535. journal = "Annu. Rev. Neurosci.",
  536. publisher = "annualreviews.org",
  537. volume = 29,
  538. pages = "279--306",
  539. year = 2006,
  540. keywords = "Locomotion",
  541. language = "en",
  542. issn = "0147-006X",
  543. pmid = "16776587",
  544. doi = "10.1146/annurev.neuro.29.051605.112910"
  545. }
  546. @ARTICLE{Fukuoka2015-ks,
  547. title = "A simple rule for quadrupedal gait generation determined by leg
  548. loading feedback: a modeling study",
  549. author = "Fukuoka, Yasuhiro and Habu, Yasushi and Fukui, Takahiro",
  550. abstract = "We discovered a specific rule for generating typical quadrupedal
  551. gaits (the order of the movement of four legs) through a
  552. simulated quadrupedal locomotion, in which unprogrammed gaits
  553. (diagonal/lateral sequence walks, left/right-lead canters, and
  554. left/right-lead transverse gallops) spontaneously emerged because
  555. of leg loading feedbacks to the CPGs hard-wired to produce a
  556. default trot. Additionally, all gaits transitioned according to
  557. speed, as seen in animals. We have therefore hypothesized that
  558. various gaits derive from a trot because of posture control
  559. through leg loading feedback. The body tilt on the two support
  560. legs of each diagonal pair during trotting was classified into
  561. three types (level, tilted up, or tilted down) according to
  562. speed. The load difference between the two legs led to the phase
  563. difference between their CPGs via the loading feedbacks,
  564. resulting in nine gaits (3(2): three tilts to the power of two
  565. diagonal pairs) including the aforementioned.",
  566. journal = "Sci. Rep.",
  567. volume = 5,
  568. pages = "8169",
  569. month = feb,
  570. year = 2015,
  571. keywords = "Locomotion",
  572. language = "en",
  573. issn = "2045-2322",
  574. pmid = "25639661",
  575. doi = "10.1038/srep08169",
  576. pmc = "PMC4313093"
  577. }
  578. @UNPUBLISHED{Roseberry2019-iz,
  579. title = "Locomotor suppression by a monosynaptic amygdala to brainstem
  580. circuit",
  581. author = "Roseberry, Thomas K and Lalive, Arnaud L and Margolin, Benjamin D
  582. and Kreitzer, Anatol C",
  583. abstract = "Abstract The control of locomotion is fundamental to vertebrate
  584. animal survival. Defensive situations require an animal to
  585. rapidly decide whether to run away or suppress locomotor activity
  586. to avoid detection. While much of the neural circuitry involved
  587. in defensive action selection has been elucidated, top-down
  588. modulation of brainstem locomotor circuitry remains unclear. Here
  589. we provide evidence for the existence and functionality of a
  590. monosynaptic connection from the central amygdala (CeA) to the
  591. mesencephalic locomotor region (MLR) that inhibits locomotion in
  592. unconditioned and conditioned defensive behavior in mice. We show
  593. that locomotion stimulated by airpuff coincides with increased
  594. activity of MLR glutamatergic neurons. Using retrograde tracing
  595. and ex vivo electrophysiology, we find that the CeA makes a
  596. monosynaptic connection with the MLR. In the open field, in vivo
  597. stimulation of this projection suppressed spontaneous locomotion,
  598. whereas inhibition of this projection had no effect. However,
  599. inhibiting CeA terminals within the MLR increased both neural
  600. activity and locomotor responses to airpuff. Finally, using a
  601. conditioned avoidance paradigm known to activate CeA neurons, we
  602. find that inhibition of the CeA projection increased successful
  603. escape, whereas activating the projection reduced escape.
  604. Together these results provide evidence for a new circuit
  605. substrate influencing locomotion and defensive behaviors.",
  606. journal = "Cold Spring Harbor Laboratory",
  607. pages = "724252",
  608. month = aug,
  609. year = 2019,
  610. keywords = "Locomotion",
  611. language = "en",
  612. doi = "10.1101/724252"
  613. }
  614. @ARTICLE{Carvalho2020-pw,
  615. title = "A Brainstem Locomotor Circuit Drives the Activity of Speed Cells
  616. in the Medial Entorhinal Cortex",
  617. author = "Carvalho, Miguel M and Tanke, Nouk and Kropff, Emilio and Witter,
  618. Menno P and Moser, May-Britt and Moser, Edvard I",
  619. abstract = "Locomotion activates an array of sensory inputs that may help
  620. build the self-position map of the medial entorhinal cortex
  621. (MEC). In this map, speed-coding neurons are thought to
  622. dynamically update representations of the animal's position. A
  623. possible origin for the entorhinal speed signal is the
  624. mesencephalic locomotor region (MLR), which is critically
  625. involved in the activation of locomotor programs. Here, we
  626. describe, in rats, a circuit connecting the pedunculopontine
  627. tegmental nucleus (PPN) of the MLR to the MEC via the horizontal
  628. limb of the diagonal band of Broca (HDB). At each level of this
  629. pathway, locomotion speed is linearly encoded in neuronal firing
  630. rates. Optogenetic activation of PPN cells drives locomotion and
  631. modulates activity of speed-modulated neurons in HDB and MEC. Our
  632. results provide evidence for a pathway by which brainstem speed
  633. signals can reach cortical structures implicated in navigation
  634. and higher-order dynamic representations of space.",
  635. journal = "Cell Rep.",
  636. volume = 32,
  637. number = 10,
  638. pages = "108123",
  639. month = sep,
  640. year = 2020,
  641. keywords = "diagonal band of Broca; medial entorhinal cortex; mesencephalic
  642. locomotor region; pedunculopontine tegmental nucleus; speed
  643. cells;Locomotion",
  644. language = "en",
  645. issn = "2211-1247",
  646. pmid = "32905779",
  647. doi = "10.1016/j.celrep.2020.108123",
  648. pmc = "PMC7487772"
  649. }
  650. @UNPUBLISHED{Dautan2020-lv,
  651. title = "Modulation of motor behavior by the mesencephalic locomotor
  652. region",
  653. author = "Dautan, Daniel and Kov{\'a}cs, Adrienn and Bayasgalan,
  654. Tsogbadrakh and Diaz-Acevedo, Miguel A and Pal, Balazs and
  655. Mena-Segovia, Juan",
  656. abstract = "The mesencephalic locomotor region (MLR) serves as an interface
  657. between higher-order motor systems and lower motor neurons. The
  658. excitatory module of the MLR is composed of the pedunculopontine
  659. nucleus (PPN) and the cuneiform nucleus (CnF), and their
  660. activation has been proposed to elicit different modalities of
  661. movement, but how the differences in connectivity and
  662. physiological properties explain their contributions to motor
  663. activity is not known. Here we report that CnF glutamatergic
  664. neurons are electrophysiologically homogeneous and have
  665. short-range axonal projections, whereas PPN glutamatergic neurons
  666. are heterogeneous and maintain long-range connections, most
  667. notably with the basal ganglia. Optogenetic activation of CnF
  668. neurons produced fast-onset, involuntary motor activity mediated
  669. by short-lasting muscle activation. In contrast, activation of
  670. PPN neurons produced long-lasting increases in muscle tone that
  671. reduced motor activity and disrupted gait. Our results thus
  672. reveal a differential contribution to motor behavior by the
  673. structures that compose the MLR. \#\#\# Competing Interest
  674. Statement The authors have declared no competing interest.",
  675. journal = "Cold Spring Harbor Laboratory",
  676. pages = "2020.06.25.172296",
  677. month = jun,
  678. year = 2020,
  679. keywords = "Locomotion",
  680. language = "en",
  681. doi = "10.1101/2020.06.25.172296"
  682. }
  683. @ARTICLE{Ruder2016-iv,
  684. title = "{Long-Distance} Descending Spinal Neurons Ensure Quadrupedal
  685. Locomotor Stability",
  686. author = "Ruder, Ludwig and Takeoka, Aya and Arber, Silvia",
  687. abstract = "Locomotion is an essential animal behavior used for
  688. translocation. The spinal cord acts as key executing center, but
  689. how it coordinates many body parts located across distance
  690. remains poorly understood. Here we employed mouse genetic and
  691. viral approaches to reveal organizational principles of
  692. long-projecting spinal circuits and their role in quadrupedal
  693. locomotion. Using neurotransmitter identity, developmental
  694. origin, and projection patterns as criteria, we uncover that
  695. spinal segments controlling forelimbs and hindlimbs are
  696. bidirectionally connected by symmetrically organized direct
  697. synaptic pathways that encompass multiple genetically tractable
  698. neuronal subpopulations. We demonstrate that selective ablation
  699. of descending spinal neurons linking cervical to lumbar segments
  700. impairs coherent locomotion, by reducing postural stability and
  701. speed during exploratory locomotion, as well as perturbing
  702. interlimb coordination during reinforced high-speed stepping.
  703. Together, our results implicate a highly organized long-distance
  704. projection system of spinal origin in the control of postural
  705. body stabilization and reliability during quadrupedal
  706. locomotion.",
  707. journal = "Neuron",
  708. publisher = "Elsevier",
  709. volume = 92,
  710. number = 5,
  711. pages = "1063--1078",
  712. month = dec,
  713. year = 2016,
  714. keywords = "genetic identity; interlimb coordination; locomotion; motor
  715. control; posture; spinal cord;Locomotion",
  716. language = "en",
  717. issn = "0896-6273, 1097-4199",
  718. pmid = "27866798",
  719. doi = "10.1016/j.neuron.2016.10.032"
  720. }
  721. @ARTICLE{Drew2015-sv,
  722. title = "Taking the next step: cortical contributions to the control of
  723. locomotion",
  724. author = "Drew, Trevor and Marigold, Daniel S",
  725. abstract = "The planning and execution of both discrete voluntary movements
  726. and visually guided locomotion depends on the contribution of
  727. multiple cortical areas. In this review, we discuss recent
  728. experiments that address the contribution of the posterior
  729. parietal cortex (PPC) and the motor cortex to the control of
  730. locomotion. The results from these experiments show that the PPC
  731. contributes to the planning of locomotion by providing an
  732. estimate of the position of an animal with respect to objects in
  733. its path. In contrast, the motor cortex contributes primarily to
  734. the execution of gait modifications by modulating the activity
  735. of groups of synergistic muscles active at different times
  736. during the gait cycle.",
  737. journal = "Curr. Opin. Neurobiol.",
  738. publisher = "Elsevier",
  739. volume = 33,
  740. pages = "25--33",
  741. month = aug,
  742. year = 2015,
  743. keywords = "Locomotion",
  744. language = "en",
  745. issn = "0959-4388, 1873-6882",
  746. pmid = "25643847",
  747. doi = "10.1016/j.conb.2015.01.011"
  748. }
  749. @ARTICLE{Ueno2011-tt,
  750. title = "Kinematic analyses reveal impaired locomotion following injury
  751. of the motor cortex in mice",
  752. author = "Ueno, Masaki and Yamashita, Toshihide",
  753. abstract = "Brain injury in the motor cortex can result in deleterious
  754. functional deficits of skilled and fine motor functions.
  755. However, in contrast to humans, the destruction of cortex and
  756. its descending fibers has been thought not to cause remarkable
  757. deficits in simple locomotion in quadropedal animals. In the
  758. present study, we aimed to investigate in detail how lesion of
  759. the sensorimotor cortex affected locomotion ability in mice
  760. using the KinemaTracer system, a novel video-based kinematic
  761. analyzer. We found that traumatic injury to the left
  762. sensorimotor cortex induced several apparent deficits in the
  763. movement of contralesional right limbs during treadmill
  764. locomotion. The step length of right limbs decreased, and the
  765. speed in the forward direction was abrogated in the swing phase.
  766. The coordinates and angle of each joint were also changed after
  767. the injury. Some of the abnormal values in these parameters
  768. gradually recovered near the control level. The number of
  769. cFos-expressing neurons following locomotion significantly
  770. decreased in the right side of the spinal cord in injured mice,
  771. suggesting a role for cortex and descending fibers in
  772. locomotion. In contrast, interlimb coordination did not change
  773. remarkably even after the injury, supporting the notion that the
  774. basic locomotor pattern was determined by intraspinal neural
  775. circuits. These results indicate that the motor cortex and its
  776. descending fibers regulate several aspects of fine limb movement
  777. during locomotion. Our findings provide practical parameters to
  778. assess motor deficits and recovery following cortical injury in
  779. mice.",
  780. journal = "Exp. Neurol.",
  781. publisher = "Elsevier",
  782. volume = 230,
  783. number = 2,
  784. pages = "280--290",
  785. month = aug,
  786. year = 2011,
  787. keywords = "Locomotion;To Read",
  788. language = "en",
  789. issn = "0014-4886, 1090-2430",
  790. pmid = "21619878",
  791. doi = "10.1016/j.expneurol.2011.05.006"
  792. }
  793. @ARTICLE{Holmes2006-fn,
  794. title = "The Dynamics of Legged Locomotion: Models, Analyses, and
  795. Challenges",
  796. author = "Holmes, Philip and Full, Robert J and Koditschek, Dan and
  797. Guckenheimer, John",
  798. abstract = "Cheetahs and beetles run, dolphins and salmon swim, and bees and
  799. birds fly with grace and economy surpassing our technology.
  800. Evolution has shaped the breathtaking abilities of animals,
  801. leaving us the challenge of reconstructing their targets of
  802. control and mechanisms of dexterity. In this review we explore a
  803. corner of this fascinating world. We describe mathematical
  804. models for legged animal locomotion, focusing on rapidly running
  805. insects and highlighting past achievements and challenges that
  806. remain. Newtonian body--limb dynamics are most naturally
  807. formulated as piecewise-holonomic rigid body mechanical systems,
  808. whose constraints change as legs touch down or lift off. Central
  809. pattern generators and proprioceptive sensing require models of
  810. spiking neurons and simplified phase oscillator descriptions of
  811. ensembles of them. A full neuromechanical model of a running
  812. animal requires integration of these elements, along with
  813. proprioceptive feedback and models of goal-oriented sensing,
  814. planning, and learning. We outline relevant background material
  815. from biomechanics and neurobiology, explain key properties of
  816. the hybrid dynamical systems that underlie legged locomotion
  817. models, and provide numerous examples of such models, from the
  818. simplest, completely soluble ``peg-leg walker'' to complex
  819. neuromuscular subsystems that are yet to be assembled into
  820. models of behaving animals. This final integration in a
  821. tractable and illuminating model is an outstanding challenge.",
  822. journal = "SIAM Rev.",
  823. publisher = "Society for Industrial and Applied Mathematics",
  824. volume = 48,
  825. number = 2,
  826. pages = "207--304",
  827. month = jan,
  828. year = 2006,
  829. keywords = "Locomotion;To Read",
  830. issn = "0036-1445",
  831. doi = "10.1137/S0036144504445133"
  832. }
  833. @ARTICLE{Schwenkgrub2020-yl,
  834. title = "Deep imaging in the brainstem reveals functional heterogeneity
  835. in V2a neurons controlling locomotion",
  836. author = "Schwenkgrub, Joanna and Harrell, Evan R and Bathellier, Brice
  837. and Bouvier, Julien",
  838. abstract = "V2a neurons are a genetically defined cell class that forms a
  839. major excitatory descending pathway from the brainstem reticular
  840. formation to the spinal cord. Their activation has been linked
  841. to the termination of locomotor activity based on broad
  842. optogenetic manipulations. However, because of the difficulties
  843. involved in accessing brainstem structures for in vivo cell
  844. type-specific recordings, V2a neuron function has never been
  845. directly observed during natural behaviors. Here, we imaged the
  846. activity of V2a neurons using micro-endoscopy in freely moving
  847. mice. We find that as many as half of the V2a neurons are
  848. excited at locomotion arrest and with low reliability. Other V2a
  849. neurons are inhibited at locomotor arrests and/or activated
  850. during other behaviors such as locomotion initiation or
  851. stationary grooming. Our results establish that V2a neurons not
  852. only drive stops as suggested by bulk optogenetics but also are
  853. stratified into subpopulations that likely contribute to diverse
  854. motor patterns.",
  855. journal = "Sci Adv",
  856. publisher = "advances.sciencemag.org",
  857. volume = 6,
  858. number = 49,
  859. month = dec,
  860. year = 2020,
  861. keywords = "Locomotion",
  862. language = "en",
  863. issn = "2375-2548",
  864. pmid = "33277252",
  865. doi = "10.1126/sciadv.abc6309"
  866. }
  867. @ARTICLE{Lemieux2019-yc,
  868. title = "Glutamatergic neurons of the gigantocellular reticular nucleus
  869. shape locomotor pattern and rhythm in the freely behaving mouse",
  870. author = "Lemieux, Maxime and Bretzner, Frederic",
  871. abstract = "Because of their intermediate position between supraspinal
  872. locomotor centers and spinal circuits, gigantocellular reticular
  873. nucleus (GRN) neurons play a key role in motor command. However,
  874. the functional contribution of glutamatergic GRN neurons in
  875. initiating, maintaining, and stopping locomotion is still
  876. unclear. Combining electromyographic recordings with optogenetic
  877. manipulations in freely behaving mice, we investigate the
  878. functional contribution of glutamatergic brainstem neurons of
  879. the GRN to motor and locomotor activity. Short-pulse
  880. photostimulation of one side of the glutamatergic GRN did not
  881. elicit locomotion but evoked distinct motor responses in flexor
  882. and extensor muscles at rest and during locomotion.
  883. Glutamatergic GRN outputs to the spinal cord appear to be gated
  884. according to the spinal locomotor network state. Increasing the
  885. duration of photostimulation increased motor and postural tone
  886. at rest and reset locomotor rhythm during ongoing locomotion. In
  887. contrast, photoinhibition impaired locomotor pattern and rhythm.
  888. We conclude that unilateral activation of glutamatergic GRN
  889. neurons triggered motor activity and modified ongoing locomotor
  890. pattern and rhythm.",
  891. journal = "PLoS Biol.",
  892. publisher = "journals.plos.org",
  893. volume = 17,
  894. number = 4,
  895. pages = "e2003880",
  896. month = apr,
  897. year = 2019,
  898. keywords = "Locomotion",
  899. language = "en",
  900. issn = "1544-9173, 1545-7885",
  901. pmid = "31017885",
  902. doi = "10.1371/journal.pbio.2003880",
  903. pmc = "PMC6502437"
  904. }
  905. @ARTICLE{Karadimas2020-ub,
  906. title = "Sensory cortical control of movement",
  907. author = "Karadimas, Spyridon K and Satkunendrarajah, Kajana and
  908. Laliberte, Alex M and Ringuette, Dene and Weisspapir, Iliya and
  909. Li, Lijun and Gosgnach, Simon and Fehlings, Michael G",
  910. abstract = "Walking in our complex environment requires continual higher
  911. order integrated spatiotemporal information. This information is
  912. processed in the somatosensory cortex, and it has long been
  913. presumed that it influences movement via descending tracts
  914. originating from the motor cortex. Here we show that neuronal
  915. activity in the primary somatosensory cortex tightly correlates
  916. with the onset and speed of locomotion in freely moving mice.
  917. Using optogenetics and pharmacogenetics in combination with in
  918. vivo and in vitro electrophysiology, we provide evidence for a
  919. direct corticospinal pathway from the primary somatosensory
  920. cortex that synapses with cervical excitatory neurons and
  921. modulates the lumbar locomotor network independently of the
  922. motor cortex and other supraspinal locomotor centers.
  923. Stimulation of this pathway enhances speed of locomotion, while
  924. inhibition decreases locomotor speed and ultimately terminates
  925. stepping. Our findings reveal a novel pathway for neural control
  926. of movement whereby the somatosensory cortex directly influences
  927. motor behavior, possibly in response to environmental cues.",
  928. journal = "Nat. Neurosci.",
  929. publisher = "nature.com",
  930. volume = 23,
  931. number = 1,
  932. pages = "75--84",
  933. month = jan,
  934. year = 2020,
  935. keywords = "Locomotion",
  936. language = "en",
  937. issn = "1097-6256, 1546-1726",
  938. pmid = "31740813",
  939. doi = "10.1038/s41593-019-0536-7"
  940. }
  941. @ARTICLE{Bouvier2015-nm,
  942. title = "Descending Command Neurons in the Brainstem that Halt Locomotion",
  943. author = "Bouvier, Julien and Caggiano, Vittorio and Leiras, Roberto and
  944. Caldeira, Vanessa and Bellardita, Carmelo and Balueva, Kira and
  945. Fuchs, Andrea and Kiehn, Ole",
  946. abstract = "The episodic nature of locomotion is thought to be controlled by
  947. descending inputs from the brainstem. Most studies have largely
  948. attributed this control to initiating excitatory signals, but
  949. little is known about putative commands that may specifically
  950. determine locomotor offset. To link identifiable brainstem
  951. populations to a potential locomotor stop signal, we used
  952. developmental genetics and considered a discrete neuronal
  953. population in the reticular formation: the V2a neurons. We find
  954. that those neurons constitute a major excitatory pathway to
  955. locomotor areas of the ventral spinal cord. Selective activation
  956. of V2a neurons of the rostral medulla stops ongoing locomotor
  957. activity, owing to an inhibition of premotor locomotor networks
  958. in the spinal cord. Moreover, inactivation of such neurons
  959. decreases spontaneous stopping in vivo. Therefore, the V2a ``stop
  960. neurons'' represent a glutamatergic descending pathway that
  961. favors immobility and may thus help control the episodic nature
  962. of locomotion.",
  963. journal = "Cell",
  964. volume = 163,
  965. number = 5,
  966. pages = "1191--1203",
  967. month = nov,
  968. year = 2015,
  969. keywords = "Locomotion",
  970. language = "en",
  971. issn = "0092-8674, 1097-4172",
  972. pmid = "26590422",
  973. doi = "10.1016/j.cell.2015.10.074",
  974. pmc = "PMC4899047"
  975. }
  976. @ARTICLE{Caggiano2018-td,
  977. title = "Midbrain circuits that set locomotor speed and gait selection",
  978. author = "Caggiano, V and Leiras, R and Go{\~n}i-Erro, H and Masini, D and
  979. Bellardita, C and Bouvier, J and Caldeira, V and Fisone, G and
  980. Kiehn, O",
  981. abstract = "Locomotion is a fundamental motor function common to the animal
  982. kingdom. It is implemented episodically and adapted to
  983. behavioural needs, including exploration, which requires slow
  984. locomotion, and escape behaviour, which necessitates faster
  985. speeds. The control of these functions originates in brainstem
  986. structures, although the neuronal substrate(s) that support them
  987. have not yet been elucidated. Here we show in mice that speed and
  988. gait selection are controlled by glutamatergic excitatory neurons
  989. (GlutNs) segregated in two distinct midbrain nuclei: the
  990. cuneiform nucleus (CnF) and the pedunculopontine nucleus (PPN).
  991. GlutNs in both of these regions contribute to the control of
  992. slower, alternating-gait locomotion, whereas only GlutNs in the
  993. CnF are able to elicit high-speed, synchronous-gait locomotion.
  994. Additionally, both the activation dynamics and the input and
  995. output connectivity matrices of GlutNs in the PPN and the CnF
  996. support explorative and escape locomotion, respectively. Our
  997. results identify two regions in the midbrain that act in
  998. conjunction to select context-dependent locomotor behaviours.",
  999. journal = "Nature",
  1000. volume = 553,
  1001. number = 7689,
  1002. pages = "455--460",
  1003. month = jan,
  1004. year = 2018,
  1005. keywords = "Locomotion",
  1006. language = "en",
  1007. issn = "0028-0836, 1476-4687",
  1008. pmid = "29342142",
  1009. doi = "10.1038/nature25448",
  1010. pmc = "PMC5937258"
  1011. }
  1012. @ARTICLE{Usseglio2020-nl,
  1013. title = "Control of Orienting Movements and Locomotion by
  1014. {Projection-Defined} Subsets of Brainstem V2a Neurons",
  1015. author = "Usseglio, Giovanni and Gatier, Edwin and Heuz{\'e}, Aur{\'e}lie
  1016. and H{\'e}rent, Coralie and Bouvier, Julien",
  1017. abstract = "Spatial orientation requires the execution of lateralized
  1018. movements and a change in the animal's heading in response to
  1019. multiple sensory modalities. While much research has focused on
  1020. the circuits for sensory integration, chiefly to the midbrain
  1021. superior colliculus (SC), the downstream cells and circuits that
  1022. engage adequate motor actions have remained elusive. Furthermore,
  1023. the mechanisms supporting trajectory changes are still
  1024. speculative. Here, using transneuronal viral tracings in mice, we
  1025. show that brainstem V2a neurons, a genetically defined subtype of
  1026. glutamatergic neurons of the reticular formation, receive
  1027. putative synaptic inputs from the contralateral SC. This makes
  1028. them a candidate relay of lateralized orienting commands. We next
  1029. show that unilateral optogenetic activations of brainstem V2a
  1030. neurons in vivo evoked ipsilateral orienting-like responses of
  1031. the head and the nose tip on stationary mice. When animals are
  1032. walking, similar stimulations impose a transient locomotor arrest
  1033. followed by a change of trajectory. Third, we reveal that these
  1034. distinct motor actions are controlled by dedicated V2a subsets
  1035. each projecting to a specific spinal cord segment, with at least
  1036. (1) a lumbar-projecting subset whose unilateral activation
  1037. specifically controls locomotor speed but neither impacts
  1038. trajectory nor evokes orienting movements, and (2) a
  1039. cervical-projecting subset dedicated to head orientation, but not
  1040. to locomotor speed. Activating the latter subset suffices to
  1041. steer the animals' directional heading, placing the head
  1042. orientation as the prime driver of locomotor trajectory. V2a
  1043. neurons and their modular organization may therefore underlie the
  1044. orchestration of multiple motor actions during multi-faceted
  1045. orienting behaviors.",
  1046. journal = "Curr. Biol.",
  1047. volume = 30,
  1048. number = 23,
  1049. pages = "4665--4681.e6",
  1050. month = dec,
  1051. year = 2020,
  1052. keywords = "V2a neurons; brainstem; circuit tracings; locomotion; motor
  1053. control; mouse; optogenetics; orientation; reticulospinal
  1054. neurons; spinal cord;Locomotion",
  1055. language = "en",
  1056. issn = "0960-9822, 1879-0445",
  1057. pmid = "33007251",
  1058. doi = "10.1016/j.cub.2020.09.014"
  1059. }
  1060. @INCOLLECTION{Matsuyama2004-fv,
  1061. title = "Locomotor role of the corticoreticular--reticulospinal--spinal
  1062. interneuronal system",
  1063. booktitle = "Progress in Brain Research",
  1064. author = "Matsuyama, Kiyoji and Mori, Futoshi and Nakajima, Katsumi and
  1065. Drew, Trevor and Aoki, Mamoru and Mori, Shigemi",
  1066. abstract = "In vertebrates, the descending reticulospinal pathway is the
  1067. primary means of conveying locomotor command signals from higher
  1068. motor centers to spinal interneuronal circuits, the latter
  1069. including the central pattern generators for locomotion. The
  1070. pathway is morphologically heterogeneous, being composed of
  1071. various types of in-parallel-descending axons, which terminate
  1072. with different arborization patterns in the spinal cord. Such
  1073. morphology suggests that this pathway and its target spinal
  1074. interneurons comprise varying types of functional subunits,
  1075. which have a wide variety of functional roles, as dictated by
  1076. command signals from the higher motor centers. Corticoreticular
  1077. fibers are one of the major output pathways from the motor
  1078. cortex to the brainstem. They project widely and diffusely
  1079. within the pontomedullary reticular formation. Such a diffuse
  1080. projection pattern seems well suited to combining and
  1081. integrating the function of the various types of reticulospinal
  1082. neurons, which are widely scattered throughout the
  1083. pontomedullary reticular formation. The
  1084. corticoreticular--reticulospinal--spinal interneuronal
  1085. connections appear to operate as a cohesive, yet flexible,
  1086. control system for the elaboration of a wide variety of
  1087. movements, including those that combine goal-directed locomotion
  1088. with other motor actions.",
  1089. publisher = "Elsevier",
  1090. volume = 143,
  1091. pages = "239--249",
  1092. month = jan,
  1093. year = 2004,
  1094. keywords = "Locomotion",
  1095. doi = "10.1016/S0079-6123(03)43024-0"
  1096. }
  1097. @ARTICLE{Maheswaranathan2020-fy,
  1098. title = "How recurrent networks implement contextual processing in
  1099. sentiment analysis",
  1100. author = "Maheswaranathan, Niru and Sussillo, David",
  1101. abstract = "Neural networks have a remarkable capacity for contextual
  1102. processing--using recent or nearby inputs to modify
  1103. processing of current input. For example, in natural
  1104. language, contextual processing is necessary to correctly
  1105. interpret negation (e.g. phrases such as ``not bad'').
  1106. However, our ability to understand how networks process
  1107. context is limited. Here, we propose general methods for
  1108. reverse engineering recurrent neural networks (RNNs) to
  1109. identify and elucidate contextual processing. We apply these
  1110. methods to understand RNNs trained on sentiment
  1111. classification. This analysis reveals inputs that induce
  1112. contextual effects, quantifies the strength and timescale of
  1113. these effects, and identifies sets of these inputs with
  1114. similar properties. Additionally, we analyze contextual
  1115. effects related to differential processing of the beginning
  1116. and end of documents. Using the insights learned from the
  1117. RNNs we improve baseline Bag-of-Words models with simple
  1118. extensions that incorporate contextual modification,
  1119. recovering greater than 90\% of the RNN's performance
  1120. increase over the baseline. This work yields a new
  1121. understanding of how RNNs process contextual information,
  1122. and provides tools that should provide similar insight more
  1123. broadly.",
  1124. month = apr,
  1125. year = 2020,
  1126. keywords = "RNN;RNN To read",
  1127. archivePrefix = "arXiv",
  1128. eprint = "2004.08013",
  1129. primaryClass = "cs.CL",
  1130. arxivid = "2004.08013"
  1131. }
  1132. @ARTICLE{Madhav2020-qs,
  1133. title = "The Synergy Between Neuroscience and Control Theory: The Nervous
  1134. System as Inspiration for Hard Control Challenges",
  1135. author = "Madhav, Manu S and Cowan, Noah J",
  1136. abstract = "Here, we review the role of control theory in modeling neural
  1137. control systems through a top-down analysis approach.
  1138. Specifically, we examine the role of the brain and central
  1139. nervous system as the controller in the organism, connected to
  1140. but isolated from the rest of the animal through insulated
  1141. interfaces. Though biological and engineering control systems
  1142. operate on similar principles, they differ in several critical
  1143. features, which makes drawing inspiration from biology for
  1144. engineering controllers challenging but worthwhile. We also
  1145. outline a procedure that the control theorist can use to draw
  1146. inspiration from the biological controller: starting from the
  1147. intact, behaving animal; designing experiments to deconstruct
  1148. and model hierarchies of feedback; modifying feedback
  1149. topologies; perturbing inputs and plant dynamics; using the
  1150. resultant outputs to perform system identification; and tuning
  1151. and validating the resultant control-theoretic model using
  1152. specially engineered robophysical models.",
  1153. journal = "Annu. Rev. Control Robot. Auton. Syst.",
  1154. publisher = "Annual Reviews",
  1155. volume = 3,
  1156. number = 1,
  1157. pages = "243--267",
  1158. month = may,
  1159. year = 2020,
  1160. issn = "2573-5144",
  1161. doi = "10.1146/annurev-control-060117-104856"
  1162. }
  1163. @ARTICLE{Fieseler2020-ne,
  1164. title = "Unsupervised learning of control signals and their encodings
  1165. in $\textit{C. elegans}$ whole-brain recordings",
  1166. author = "Fieseler, Charles and Zimmer, Manuel and Nathan Kutz, J",
  1167. abstract = "Recent whole brain imaging experiments on $\textit\{C.
  1168. elegans\}$ has revealed that the neural population dynamics
  1169. encode motor commands and stereotyped transitions between
  1170. behaviors on low dimensional manifolds. Efforts to
  1171. characterize the dynamics on this manifold have used
  1172. piecewise linear models to describe the entire state space,
  1173. but it is unknown how a single, global dynamical model can
  1174. generate the observed dynamics. Here, we propose a control
  1175. framework to achieve such a global model of the dynamics,
  1176. whereby underlying linear dynamics is actuated by sparse
  1177. control signals. This method learns the control signals in
  1178. an unsupervised way from data, then uses $\textit\{ Dynamic
  1179. Mode Decomposition with control\}$ (DMDc) to create the
  1180. first global, linear dynamical system that can reconstruct
  1181. whole-brain imaging data. These control signals are shown to
  1182. be implicated in transitions between behaviors. In addition,
  1183. we analyze the time-delay encoding of these control signals,
  1184. showing that these transitions can be predicted from neurons
  1185. previously implicated in behavioral transitions, but also
  1186. additional neurons previously unidentified. Moreover, our
  1187. decomposition method allows one to understand the observed
  1188. nonlinear global dynamics instead as linear dynamics with
  1189. control. The proposed mathematical framework is generic and
  1190. can be generalized to other neurosensory systems,
  1191. potentially revealing transitions and their encodings in a
  1192. completely unsupervised way.",
  1193. month = jan,
  1194. year = 2020,
  1195. archivePrefix = "arXiv",
  1196. eprint = "2001.08346",
  1197. primaryClass = "q-bio.QM",
  1198. arxivid = "2001.08346"
  1199. }
  1200. @UNPUBLISHED{Schaeffer2020-qv,
  1201. title = "Reverse-engineering Recurrent Neural Network solutions to a
  1202. hierarchical inference task for mice",
  1203. author = "Schaeffer, Rylan and Khona, Mikail and Meshulam, Leenoy and
  1204. {International Brain Laboratory} and Fiete, Ila Rani",
  1205. abstract = "We study how recurrent neural networks (RNNs) solve a
  1206. hierarchical inference task involving two latent variables and
  1207. disparate timescales separated by 1-2 orders of magnitude. The
  1208. task is of interest to the International Brain Laboratory, a
  1209. global collaboration of experimental and theoretical
  1210. neuroscientists studying how the mammalian brain generates
  1211. behavior. We make four discoveries. First, RNNs learn behavior
  1212. that is quantitatively similar to ideal Bayesian baselines.
  1213. Second, RNNs perform inference by learning a two-dimensional
  1214. subspace defining beliefs about the latent variables. Third, the
  1215. geometry of RNN dynamics reflects an induced coupling between the
  1216. two separate inference processes necessary to solve the task.
  1217. Fourth, we perform model compression through a novel form of
  1218. knowledge distillation on hidden representations --
  1219. Representations and Dynamics Distillation (RADD)-- to reduce the
  1220. RNN dynamics to a low-dimensional, highly interpretable model.
  1221. This technique promises a useful tool for interpretability of
  1222. high dimensional nonlinear dynamical systems. Altogether, this
  1223. work yields predictions to guide exploration and analysis of
  1224. mouse neural data and circuity. \#\#\# Competing Interest
  1225. Statement The authors have declared no competing interest.",
  1226. journal = "Cold Spring Harbor Laboratory",
  1227. pages = "2020.06.09.142745",
  1228. month = jun,
  1229. year = 2020,
  1230. keywords = "RNN",
  1231. language = "en",
  1232. doi = "10.1101/2020.06.09.142745"
  1233. }
  1234. % The entry below contains non-ASCII chars that could not be converted
  1235. % to a LaTeX equivalent.
  1236. @UNPUBLISHED{Van_der_Zouwen2020-zn,
  1237. title = "Freely behaving mice can brake and turn during optogenetic
  1238. stimulation of the Mesencephalic Locomotor Region",
  1239. author = "van der Zouwen, Cornelis Immanuel and Boutin, Jo{\"e}l and
  1240. Foug{\`e}re, Maxime and Flaive, Aur{\'e}lie and Vivancos,
  1241. M{\'e}lanie and Santuz, Alessandro and Akay, Turgay and Sarret,
  1242. Philippe and Ryczko, Dimitri",
  1243. abstract = "Background Stimulation of the Mesencephalic Locomotor Region (
  1244. MLR ) is increasingly considered as a target to improve locomotor
  1245. function in Parkinson's disease, spinal cord injury and stroke. A
  1246. key function of the MLR is to control the speed of forward
  1247. symmetrical locomotor movements. However, the ability of freely
  1248. moving mammals to integrate environmental cues to brake and turn
  1249. during MLR stimulation is poorly documented. Objective/hypothesis
  1250. We investigated whether freely behaving mice could brake or turn
  1251. based on environmental cues during MLR stimulation. Methods We
  1252. stimulated the cuneiform nucleus in mice expressing
  1253. channelrhodopsin in Vglut2-positive neurons in a Cre-dependent
  1254. manner (Vglut2-ChR2-EYFP) using optogenetics. We detected
  1255. locomotor movements using deep learning. We used patch-clamp
  1256. recordings to validate the functional expression of
  1257. channelrhodopsin and neuroanatomy to visualize the stimulation
  1258. sites. Results Optogenetic stimulation of the MLR evoked
  1259. locomotion and increasing laser power increased locomotor speed.
  1260. Gait diagram and limb kinematics were similar during spontaneous
  1261. and optogenetic-evoked locomotion. Mice could brake and make
  1262. sharp turns (∼90⁰) when approaching a corner during MLR
  1263. stimulation in an open-field arena. The speed during the turn was
  1264. scaled with the speed before the turn, and with the turn angle.
  1265. In a reporter mouse, many Vglut2-ZsGreen neurons were
  1266. immunopositive for glutamate in the MLR. Patch-clamp recordings
  1267. in Vglut2-ChR2-EYFP mice show that blue light evoked short
  1268. latency spiking in MLR neurons. Conclusion MLR glutamatergic
  1269. neurons are a relevant target to improve locomotor activity
  1270. without impeding the ability to brake and turn when approaching
  1271. an obstacle, thus ensuring smooth and adaptable navigation.
  1272. Highlights \#\#\# Competing Interest Statement The authors have
  1273. declared no competing interest.",
  1274. journal = "Cold Spring Harbor Laboratory",
  1275. pages = "2020.11.30.404525",
  1276. month = dec,
  1277. year = 2020,
  1278. keywords = "Locomotion",
  1279. language = "en",
  1280. doi = "10.1101/2020.11.30.404525"
  1281. }
  1282. @UNPUBLISHED{Michaels2020-ut,
  1283. title = "A modular neural network model of grasp movement generation",
  1284. author = "Michaels, Jonathan A and Schaffelhofer, Stefan and Agudelo-Toro,
  1285. Andres and Scherberger, Hansj{\"o}rg",
  1286. abstract = "Summary One of the primary ways we interact with the world is
  1287. using our hands. In macaques, the circuit spanning the anterior
  1288. intraparietal area, the hand area of the ventral premotor cortex,
  1289. and the primary motor cortex is necessary for transforming visual
  1290. information into grasping movements. We hypothesized that a
  1291. recurrent neural network mimicking the multi-area structure of
  1292. the anatomical circuit and using visual features to generate the
  1293. required muscle dynamics to grasp objects would explain the
  1294. neural and computational basis of the grasping circuit. Modular
  1295. networks with object feature input and sparse inter-module
  1296. connectivity outperformed other models at explaining neural data
  1297. and the inter-area relationships present in the biological
  1298. circuit, despite the absence of neural data during network
  1299. training. Network dynamics were governed by simple rules, and
  1300. targeted lesioning of modules produced deficits similar to those
  1301. observed in lesion studies, providing a potential explanation for
  1302. how grasping movements are generated.",
  1303. journal = "Cold Spring Harbor Laboratory",
  1304. pages = "742189",
  1305. month = feb,
  1306. year = 2020,
  1307. keywords = "RNN;RNN To read;To Read",
  1308. language = "en",
  1309. doi = "10.1101/742189"
  1310. }
  1311. @ARTICLE{Sussillo2015-xp,
  1312. title = "A neural network that finds a naturalistic solution for the
  1313. production of muscle activity",
  1314. author = "Sussillo, David and Churchland, Mark M and Kaufman, Matthew T and
  1315. Shenoy, Krishna V",
  1316. abstract = "It remains an open question how neural responses in motor cortex
  1317. relate to movement. We explored the hypothesis that motor cortex
  1318. reflects dynamics appropriate for generating temporally patterned
  1319. outgoing commands. To formalize this hypothesis, we trained
  1320. recurrent neural networks to reproduce the muscle activity of
  1321. reaching monkeys. Models had to infer dynamics that could
  1322. transform simple inputs into temporally and spatially complex
  1323. patterns of muscle activity. Analysis of trained models revealed
  1324. that the natural dynamical solution was a low-dimensional
  1325. oscillator that generated the necessary multiphasic commands.
  1326. This solution closely resembled, at both the single-neuron and
  1327. population levels, what was observed in neural recordings from
  1328. the same monkeys. Notably, data and simulations agreed only when
  1329. models were optimized to find simple solutions. An appealing
  1330. interpretation is that the empirically observed dynamics of motor
  1331. cortex may reflect a simple solution to the problem of generating
  1332. temporally patterned descending commands.",
  1333. journal = "Nat. Neurosci.",
  1334. volume = 18,
  1335. number = 7,
  1336. pages = "1025--1033",
  1337. month = jul,
  1338. year = 2015,
  1339. keywords = "RNN;RNN To read",
  1340. language = "en",
  1341. issn = "1097-6256, 1546-1726",
  1342. pmid = "26075643",
  1343. doi = "10.1038/nn.4042",
  1344. pmc = "PMC5113297"
  1345. }
  1346. @ARTICLE{Carlsson2020-lg,
  1347. title = "Topological methods for data modelling",
  1348. author = "Carlsson, Gunnar",
  1349. abstract = "The analysis of large and complex data sets is one of the most
  1350. important problems facing the scientific community, and physics
  1351. in particular. One response to this challenge has been the
  1352. development of topological data analysis (TDA), which models
  1353. data by graphs or networks rather than by linear algebraic
  1354. (matrix) methods or cluster analysis. TDA represents the shape
  1355. of the data (suitably defined) in a combinatorial fashion.
  1356. Methods for measuring shape have been developed within
  1357. mathematics, providing a toolkit referred to as homology. In
  1358. working with data, one can use this kind of modelling to obtain
  1359. an understanding of the overall structure of the data set. There
  1360. is a suite of methods for constructing vector representations of
  1361. various kinds of unstructured data. In this Review, we sketch
  1362. the basics of TDA and provide examples where this kind of
  1363. analysis has been carried out. The rapidly developing field of
  1364. topological data analysis represents data via graphs rather than
  1365. as solutions to equations or as decompositions into clusters.
  1366. This Review discusses the methods and provides examples from
  1367. physics and other sciences.",
  1368. journal = "Nature Reviews Physics",
  1369. publisher = "Nature Publishing Group",
  1370. pages = "1--12",
  1371. month = nov,
  1372. year = 2020,
  1373. keywords = "RNN",
  1374. language = "en",
  1375. issn = "2522-5820, 2522-5820",
  1376. doi = "10.1038/s42254-020-00249-3"
  1377. }
  1378. @UNPUBLISHED{Kalidindi2020-wd,
  1379. title = "Rotational dynamics in motor cortex are consistent with a
  1380. feedback controller",
  1381. author = "Kalidindi, Hari Teja and Cross, Kevin P and Lillicrap, Timothy P
  1382. and Omrani, Mohsen and Falotico, Egidio and Sabes, Philip N and
  1383. Scott, Stephen H",
  1384. abstract = "Recent studies hypothesize that motor cortical (MC) dynamics are
  1385. generated largely through its recurrent connections based on
  1386. observations that MC activity exhibits rotational structure.
  1387. However, behavioural and neurophysiological studies suggest that
  1388. MC behaves like a feedback controller where continuous sensory
  1389. feedback and interactions with other brain areas contribute
  1390. substantially to MC processing. We investigated these apparently
  1391. conflicting theories by building recurrent neural networks that
  1392. controlled a model arm and received sensory feedback about the
  1393. limb. Networks were trained to counteract perturbations to the
  1394. limb and to reach towards spatial targets. Network activities and
  1395. sensory feedback signals to the network exhibited rotational
  1396. structure even when the recurrent connections were removed.
  1397. Furthermore, neural recordings in monkeys performing similar
  1398. tasks also exhibited rotational structure not only in MC but also
  1399. in somatosensory cortex. Our results argue that rotational
  1400. structure may reflect dynamics throughout voluntary motor
  1401. circuits involved in online control of motor actions. \#\#\#
  1402. Competing Interest Statement SHS is co-founder and CSO of Kinarm
  1403. which commercializes the robotic technology used in the present
  1404. study.",
  1405. journal = "Cold Spring Harbor Laboratory",
  1406. pages = "2020.11.17.387043",
  1407. month = nov,
  1408. year = 2020,
  1409. keywords = "RNN",
  1410. language = "en",
  1411. doi = "10.1101/2020.11.17.387043"
  1412. }
  1413. @ARTICLE{Machado2015-ig,
  1414. title = "A quantitative framework for whole-body coordination reveals
  1415. specific deficits in freely walking ataxic mice",
  1416. author = "Machado, Ana S and Darmohray, Dana M and Fayad, Jo{\~a}o and
  1417. Marques, Hugo G and Carey, Megan R",
  1418. abstract = "The coordination of movement across the body is a fundamental,
  1419. yet poorly understood aspect of motor control. Mutant mice with
  1420. cerebellar circuit defects exhibit characteristic impairments in
  1421. locomotor coordination; however, the fundamental features of this
  1422. gait ataxia have not been effectively isolated. Here we describe
  1423. a novel system (LocoMouse) for analyzing limb, head, and tail
  1424. kinematics of freely walking mice. Analysis of visibly ataxic
  1425. Purkinje cell degeneration (pcd) mice reveals that while
  1426. differences in the forward motion of individual paws are fully
  1427. accounted for by changes in walking speed and body size, more
  1428. complex 3D trajectories and, especially, inter-limb and
  1429. whole-body coordination are specifically impaired. Moreover, the
  1430. coordination deficits in pcd are consistent with a failure to
  1431. predict and compensate for the consequences of movement across
  1432. the body. These results isolate specific impairments in
  1433. whole-body coordination in mice and provide a quantitative
  1434. framework for understanding cerebellar contributions to
  1435. coordinated locomotion.",
  1436. journal = "Elife",
  1437. volume = 4,
  1438. month = oct,
  1439. year = 2015,
  1440. keywords = "Purkinje cell; ataxia; cerebellum; locomotion; mouse;
  1441. neuroscience;Locomotion",
  1442. language = "en",
  1443. issn = "2050-084X",
  1444. pmid = "26433022",
  1445. doi = "10.7554/eLife.07892",
  1446. pmc = "PMC4630674"
  1447. }
  1448. @ARTICLE{Russo2018-me,
  1449. title = "Motor Cortex Embeds Muscle-like Commands in an Untangled
  1450. Population Response",
  1451. author = "Russo, Abigail A and Bittner, Sean R and Perkins, Sean M and
  1452. Seely, Jeffrey S and London, Brian M and Lara, Antonio H and
  1453. Miri, Andrew and Marshall, Najja J and Kohn, Adam and Jessell,
  1454. Thomas M and Abbott, Laurence F and Cunningham, John P and
  1455. Churchland, Mark M",
  1456. abstract = "Primate motor cortex projects to spinal interneurons and
  1457. motoneurons, suggesting that motor cortex activity may be
  1458. dominated by muscle-like commands. Observations during reaching
  1459. lend support to this view, but evidence remains ambiguous and
  1460. much debated. To provide a different perspective, we employed a
  1461. novel behavioral paradigm that facilitates comparison between
  1462. time-evolving neural and muscle activity. We found that single
  1463. motor cortex neurons displayed many muscle-like properties, but
  1464. the structure of population activity was not muscle-like. Unlike
  1465. muscle activity, neural activity was structured to avoid
  1466. ``tangling'': moments where similar activity patterns led to
  1467. dissimilar future patterns. Avoidance of tangling was present
  1468. across tasks and species. Network models revealed a potential
  1469. reason for this consistent feature: low tangling confers noise
  1470. robustness. Finally, we were able to predict motor cortex
  1471. activity from muscle activity by leveraging the hypothesis that
  1472. muscle-like commands are embedded in additional structure that
  1473. yields low tangling.",
  1474. journal = "Neuron",
  1475. volume = 97,
  1476. number = 4,
  1477. pages = "953--966.e8",
  1478. month = feb,
  1479. year = 2018,
  1480. keywords = "motor control; motor cortex; movement generation; neural
  1481. dynamics; neural network; pattern generation; rhythmic
  1482. movement;RNN;RNN To read",
  1483. language = "en",
  1484. issn = "0896-6273, 1097-4199",
  1485. pmid = "29398358",
  1486. doi = "10.1016/j.neuron.2018.01.004",
  1487. pmc = "PMC5823788"
  1488. }
  1489. @UNPUBLISHED{Russo2019-sw,
  1490. title = "Neural trajectories in the supplementary motor area and primary
  1491. motor cortex exhibit distinct geometries, compatible with
  1492. different classes of computation",
  1493. author = "Russo, Abigail A and Khajeh, Ramin and Bittner, Sean R and
  1494. Perkins, Sean M and Cunningham, John P and Abbott, Laurence F and
  1495. Churchland, Mark M",
  1496. abstract = "Abstract The supplementary motor area (SMA) is believed to
  1497. contribute to higher-order aspects of motor control. To examine
  1498. this contribution, we employed a novel cycling task and leveraged
  1499. an emerging strategy: testing whether population trajectories
  1500. possess properties necessary for a hypothesized class of
  1501. computations. We found that, at the single-neuron level, SMA
  1502. exhibited multiple response features absent in M1. We
  1503. hypothesized that these diverse features might contribute, at the
  1504. population level, to avoidance of `population trajectory
  1505. divergence' -- ensuring that two trajectories never followed the
  1506. same path before separating. Trajectory divergence was indeed
  1507. avoided in SMA but not in M1. Network simulations confirmed that
  1508. low trajectory divergence is necessary when guidance of future
  1509. action depends upon internally tracking contextual factors.
  1510. Furthermore, the empirical trajectory geometry -- helical in SMA
  1511. versus elliptical in M1 -- was naturally reproduced by networks
  1512. that did, versus did not, internally track context.",
  1513. journal = "Cold Spring Harbor Laboratory",
  1514. pages = "650002",
  1515. month = may,
  1516. year = 2019,
  1517. keywords = "RNN;RNN To read",
  1518. language = "en",
  1519. doi = "10.1101/650002"
  1520. }
  1521. @ARTICLE{Rivkind2017-pf,
  1522. title = "Local Dynamics in Trained Recurrent Neural Networks",
  1523. author = "Rivkind, Alexander and Barak, Omri",
  1524. abstract = "Learning a task induces connectivity changes in neural circuits,
  1525. thereby changing their dynamics. To elucidate task-related neural
  1526. dynamics, we study trained recurrent neural networks. We develop
  1527. a mean field theory for reservoir computing networks trained to
  1528. have multiple fixed point attractors. Our main result is that the
  1529. dynamics of the network's output in the vicinity of attractors is
  1530. governed by a low-order linear ordinary differential equation.
  1531. The stability of the resulting equation can be assessed,
  1532. predicting training success or failure. As a consequence,
  1533. networks of rectified linear units and of sigmoidal
  1534. nonlinearities are shown to have diametrically different
  1535. properties when it comes to learning attractors. Furthermore, a
  1536. characteristic time constant, which remains finite at the edge of
  1537. chaos, offers an explanation of the network's output robustness
  1538. in the presence of variability of the internal neural dynamics.
  1539. Finally, the proposed theory predicts state-dependent frequency
  1540. selectivity in the network response.",
  1541. journal = "Phys. Rev. Lett.",
  1542. volume = 118,
  1543. number = 25,
  1544. pages = "258101",
  1545. month = jun,
  1546. year = 2017,
  1547. keywords = "RNN",
  1548. language = "en",
  1549. issn = "0031-9007, 1079-7114",
  1550. pmid = "28696758",
  1551. doi = "10.1103/PhysRevLett.118.258101"
  1552. }
  1553. @ARTICLE{Sussillo2016-zn,
  1554. title = "{LFADS} - Latent Factor Analysis via Dynamical Systems",
  1555. author = "Sussillo, David and Jozefowicz, Rafal and Abbott, L F and
  1556. Pandarinath, Chethan",
  1557. abstract = "Neuroscience is experiencing a data revolution in which many
  1558. hundreds or thousands of neurons are recorded
  1559. simultaneously. Currently, there is little consensus on how
  1560. such data should be analyzed. Here we introduce LFADS
  1561. (Latent Factor Analysis via Dynamical Systems), a method to
  1562. infer latent dynamics from simultaneously recorded,
  1563. single-trial, high-dimensional neural spiking data. LFADS is
  1564. a sequential model based on a variational auto-encoder. By
  1565. making a dynamical systems hypothesis regarding the
  1566. generation of the observed data, LFADS reduces observed
  1567. spiking to a set of low-dimensional temporal factors,
  1568. per-trial initial conditions, and inferred inputs. We
  1569. compare LFADS to existing methods on synthetic data and show
  1570. that it significantly out-performs them in inferring neural
  1571. firing rates and latent dynamics.",
  1572. month = aug,
  1573. year = 2016,
  1574. keywords = "RNN To read;RNN",
  1575. archivePrefix = "arXiv",
  1576. eprint = "1608.06315",
  1577. primaryClass = "cs.LG",
  1578. arxivid = "1608.06315"
  1579. }
  1580. @ARTICLE{Maheswaranathan2019-ue,
  1581. title = "Reverse engineering recurrent networks for sentiment
  1582. classification reveals line attractor dynamics",
  1583. author = "Maheswaranathan, Niru and Williams, Alex H and Golub, Matthew D
  1584. and Ganguli, Surya and Sussillo, David",
  1585. abstract = "Recurrent neural networks (RNNs) are a widely used tool for
  1586. modeling sequential data, yet they are often treated as
  1587. inscrutable black boxes. Given a trained recurrent network, we
  1588. would like to reverse engineer it-to obtain a quantitative,
  1589. interpretable description of how it solves a particular task.
  1590. Even for simple tasks, a detailed understanding of how recurrent
  1591. networks work, or a prescription for how to develop such an
  1592. understanding, remains elusive. In this work, we use tools from
  1593. dynamical systems analysis to reverse engineer recurrent
  1594. networks trained to perform sentiment classification, a
  1595. foundational natural language processing task. Given a trained
  1596. network, we find fixed points of the recurrent dynamics and
  1597. linearize the nonlinear system around these fixed points.
  1598. Despite their theoretical capacity to implement complex,
  1599. high-dimensional computations, we find that trained networks
  1600. converge to highly interpretable, low-dimensional
  1601. representations. In particular, the topological structure of the
  1602. fixed points and corresponding linearized dynamics reveal an
  1603. approximate line attractor within the RNN, which we can use to
  1604. quantitatively understand how the RNN solves the sentiment
  1605. analysis task. Finally, we find this mechanism present across
  1606. RNN architectures (including LSTMs, GRUs, and vanilla RNNs)
  1607. trained on multiple datasets, suggesting that our findings are
  1608. not unique to a particular architecture or dataset. Overall,
  1609. these results demonstrate that surprisingly universal and human
  1610. interpretable computations can arise across a range of recurrent
  1611. networks.",
  1612. journal = "Adv. Neural Inf. Process. Syst.",
  1613. publisher = "papers.nips.cc",
  1614. volume = 32,
  1615. pages = "15696--15705",
  1616. month = dec,
  1617. year = 2019,
  1618. keywords = "RNN To read;RNN",
  1619. language = "en",
  1620. issn = "1049-5258",
  1621. pmid = "32782423",
  1622. pmc = "PMC7416638"
  1623. }
  1624. @ARTICLE{Maheswaranathan2019-ux,
  1625. title = "Universality and individuality in neural dynamics across large
  1626. populations of recurrent networks",
  1627. author = "Maheswaranathan, Niru and Williams, Alex H and Golub, Matthew D
  1628. and Ganguli, Surya and Sussillo, David",
  1629. abstract = "Task-based modeling with recurrent neural networks (RNNs) has
  1630. emerged as a popular way to infer the computational function of
  1631. different brain regions. These models are quantitatively assessed
  1632. by comparing the low-dimensional neural representations of the
  1633. model with the brain, for example using canonical correlation
  1634. analysis (CCA). However, the nature of the detailed
  1635. neurobiological inferences one can draw from such efforts remains
  1636. elusive. For example, to what extent does training neural
  1637. networks to solve common tasks uniquely determine the network
  1638. dynamics, independent of modeling architectural choices? Or
  1639. alternatively, are the learned dynamics highly sensitive to
  1640. different model choices? Knowing the answer to these questions
  1641. has strong implications for whether and how we should use
  1642. task-based RNN modeling to understand brain dynamics. To address
  1643. these foundational questions, we study populations of thousands
  1644. of networks, with commonly used RNN architectures, trained to
  1645. solve neuroscientifically motivated tasks and characterize their
  1646. nonlinear dynamics. We find the geometry of the RNN
  1647. representations can be highly sensitive to different network
  1648. architectures, yielding a cautionary tale for measures of
  1649. similarity that rely on representational geometry, such as CCA.
  1650. Moreover, we find that while the geometry of neural dynamics can
  1651. vary greatly across architectures, the underlying computational
  1652. scaffold-the topological structure of fixed points, transitions
  1653. between them, limit cycles, and linearized dynamics-often appears
  1654. universal across all architectures.",
  1655. journal = "Adv. Neural Inf. Process. Syst.",
  1656. volume = 2019,
  1657. pages = "15629--15641",
  1658. month = dec,
  1659. year = 2019,
  1660. keywords = "RNN",
  1661. language = "en",
  1662. issn = "1049-5258",
  1663. pmid = "32782422",
  1664. pmc = "PMC7416639"
  1665. }
  1666. @ARTICLE{Vyas2020-pw,
  1667. title = "Computation Through Neural Population Dynamics",
  1668. author = "Vyas, Saurabh and Golub, Matthew D and Sussillo, David and
  1669. Shenoy, Krishna V",
  1670. abstract = "Significant experimental, computational, and theoretical work has
  1671. identified rich structure within the coordinated activity of
  1672. interconnected neural populations. An emerging challenge now is
  1673. to uncover the nature of the associated computations, how they
  1674. are implemented, and what role they play in driving behavior. We
  1675. term this computation through neural population dynamics. If
  1676. successful, this framework will reveal general motifs of neural
  1677. population activity and quantitatively describe how neural
  1678. population dynamics implement computations necessary for driving
  1679. goal-directed behavior. Here, we start with a mathematical primer
  1680. on dynamical systems theory and analytical tools necessary to
  1681. apply this perspective to experimental data. Next, we highlight
  1682. some recent discoveries resulting from successful application of
  1683. dynamical systems. We focus on studies spanning motor control,
  1684. timing, decision-making, and working memory. Finally, we briefly
  1685. discuss promising recent lines of investigation and future
  1686. directions for the computation through neural population dynamics
  1687. framework.",
  1688. journal = "Annu. Rev. Neurosci.",
  1689. volume = 43,
  1690. pages = "249--275",
  1691. month = jul,
  1692. year = 2020,
  1693. keywords = "dynamical systems; neural computation; neural population
  1694. dynamics; state spaces;RNN",
  1695. language = "en",
  1696. issn = "0147-006X, 1545-4126",
  1697. pmid = "32640928",
  1698. doi = "10.1146/annurev-neuro-092619-094115",
  1699. pmc = "PMC7402639"
  1700. }
  1701. @ARTICLE{Bellardita2015-ut,
  1702. title = "Phenotypic characterization of speed-associated gait changes in
  1703. mice reveals modular organization of locomotor networks",
  1704. author = "Bellardita, Carmelo and Kiehn, Ole",
  1705. abstract = "Studies of locomotion in mice suggest that circuits controlling
  1706. the alternating between left and right limbs may have a modular
  1707. organization with distinct locomotor circuits being recruited at
  1708. different speeds. It is not clear, however, whether such a
  1709. modular organization reflects specific behavioral outcomes
  1710. expressed at different speeds of locomotion. Here, we use
  1711. detailed kinematic analyses to search for signatures of a
  1712. modular organization of locomotor circuits in intact and
  1713. genetically modified mice moving at different speeds of
  1714. locomotion. We show that wild-type mice display three distinct
  1715. gaits: two alternating, walk and trot, and one synchronous,
  1716. bound. Each gait is expressed in distinct ranges of speed with
  1717. phenotypic inter-limb and intra-limb coordination. A fourth
  1718. gait, gallop, closely resembled bound in most of the locomotor
  1719. parameters but expressed diverse inter-limb coordination.
  1720. Genetic ablation of commissural V0V neurons completely removed
  1721. the expression of one alternating gait, trot, but left intact
  1722. walk, gallop, and bound. Ablation of commissural V0V and V0D
  1723. neurons led to a loss of walk, trot, and gallop, leaving bound
  1724. as the default gait. Our study provides a benchmark for studies
  1725. of the neuronal control of locomotion in the full range of
  1726. speeds. It provides evidence that gait expression depends upon
  1727. selection of different modules of neuronal ensembles.",
  1728. journal = "Curr. Biol.",
  1729. publisher = "Elsevier",
  1730. volume = 25,
  1731. number = 11,
  1732. pages = "1426--1436",
  1733. month = jun,
  1734. year = 2015,
  1735. keywords = "Locomotion",
  1736. language = "en",
  1737. issn = "0960-9822, 1879-0445",
  1738. pmid = "25959968",
  1739. doi = "10.1016/j.cub.2015.04.005",
  1740. pmc = "PMC4469368"
  1741. }
  1742. % The entry below contains non-ASCII chars that could not be converted
  1743. % to a LaTeX equivalent.
  1744. @ARTICLE{Walker2018-qp,
  1745. title = "A comparison of two types of running wheel in terms of mouse
  1746. preference, health, and welfare",
  1747. author = "Walker, Michael and Mason, Georgia",
  1748. abstract = "Voluntary wheel running occurs in mice of all strains, sexes, and
  1749. ages. Mice find voluntary wheel running rewarding, and it leads
  1750. to numerous health benefits. For this reason wheels are used both
  1751. to enhance welfare and to create models of exercise. However,
  1752. many designs of running wheel are used. This makes between-study
  1753. comparisons difficult, as this variability could potentially
  1754. affect the amount, pattern, and/or intensity of running
  1755. behaviour, and thence the wheels' effects on welfare and
  1756. exercise-related changes in anatomy and physiology. This study
  1757. therefore evaluated two commercially available models, chosen
  1758. because safe for group-housed mice: Bio Serv\textregistered{}'s
  1759. ``fast-trac'' wheel combo and Ware Manufacturing Inc.'s stainless
  1760. steel mesh 5″ upright wheel. Working with a total of three
  1761. hundred and fifty one female C57BL/6, DBA/2 and BALB/c mice, we
  1762. assessed these wheels' relative utilization by mice when access
  1763. was free; the strength of motivation for each wheel-type when
  1764. access required crossing an electrified grid; and the impact each
  1765. wheel had on mouse well-being (inferred from acoustic startle
  1766. responses and neophobia) and exercise-related anatomical changes
  1767. (BMI; heart and hind limb masses). Mice ran more on the
  1768. ``fast-trac'' wheel regardless of whether both wheel-types were
  1769. available at once, or only if one was present. In terms of
  1770. motivation, subjects required to work to access a single wheel
  1771. worked equally hard for both wheel-types (even if locked and thus
  1772. not useable for running), but if provided with one working wheel
  1773. for free and the other type of wheel (again unlocked) accessible
  1774. via crossing the electrified grid, the ``fast-trac'' wheel
  1775. emerged as more motivating, as the Maximum Price Paid for the
  1776. Ware metal wheel was lower than that paid for the ``fast-trac''
  1777. plastic wheel, at least for C57BL/6s and DBA/2s. No deleterious
  1778. consequences were noted with either wheel in terms of health and
  1779. welfare, but only mice with plastic wheels developed
  1780. significantly larger hearts and hind limbs than control animals
  1781. with locked wheels. Thus, where differences emerged, Bio
  1782. Serv\textregistered{}'s ``fast-trac'' wheel combos appeared to
  1783. better meet the aims of exercise provision than Ware
  1784. Manufacturing's steel upright wheels.",
  1785. journal = "Physiol. Behav.",
  1786. volume = 191,
  1787. pages = "82--90",
  1788. month = jul,
  1789. year = 2018,
  1790. keywords = "Health; Motivation; Mouse; Preference; Welfare; Wheel running",
  1791. language = "en",
  1792. issn = "0031-9384, 1873-507X",
  1793. pmid = "29653112",
  1794. doi = "10.1016/j.physbeh.2018.04.006"
  1795. }
  1796. @ARTICLE{Lemieux2016-fx,
  1797. title = "{Speed-Dependent} Modulation of the Locomotor Behavior in Adult
  1798. Mice Reveals Attractor and Transitional Gaits",
  1799. author = "Lemieux, Maxime and Josset, Nicolas and Roussel, Marie and
  1800. Couraud, S{\'e}bastien and Bretzner, Fr{\'e}d{\'e}ric",
  1801. abstract = "Locomotion results from an interplay between biomechanical
  1802. constraints of the muscles attached to the skeleton and the
  1803. neuronal circuits controlling and coordinating muscle activities.
  1804. Quadrupeds exhibit a wide range of locomotor gaits. Given our
  1805. advances in the genetic identification of spinal and supraspinal
  1806. circuits important to locomotion in the mouse, it is now
  1807. important to get a better understanding of the full repertoire of
  1808. gaits in the freely walking mouse. To assess this range, young
  1809. adult C57BL/6J mice were trained to walk and run on a treadmill
  1810. at different locomotor speeds. Instead of using the classical
  1811. paradigm defining gaits according to their footfall pattern, we
  1812. combined the inter-limb coupling and the duty cycle of the stance
  1813. phase, thus identifying several types of gaits: lateral walk,
  1814. trot, out-of-phase walk, rotary gallop, transverse gallop, hop,
  1815. half-bound, and full-bound. Out-of-phase walk, trot, and
  1816. full-bound were robust and appeared to function as attractor
  1817. gaits (i.e., a state to which the network flows and stabilizes)
  1818. at low, intermediate, and high speeds respectively. In contrast,
  1819. lateral walk, hop, transverse gallop, rotary gallop, and
  1820. half-bound were more transient and therefore considered
  1821. transitional gaits (i.e., a labile state of the network from
  1822. which it flows to the attractor state). Surprisingly, lateral
  1823. walk was less frequently observed. Using graph analysis, we
  1824. demonstrated that transitions between gaits were predictable, not
  1825. random. In summary, the wild-type mouse exhibits a wider
  1826. repertoire of locomotor gaits than expected. Future locomotor
  1827. studies should benefit from this paradigm in assessing transgenic
  1828. mice or wild-type mice with neurotraumatic injury or
  1829. neurodegenerative disease affecting gait.",
  1830. journal = "Front. Neurosci.",
  1831. volume = 10,
  1832. pages = "42",
  1833. month = feb,
  1834. year = 2016,
  1835. keywords = "graph analysis; kinematic; locomotor gaits; mouse; speed;
  1836. steady-state;Locomotion",
  1837. language = "en",
  1838. issn = "1662-4548, 1662-453X",
  1839. pmid = "26941592",
  1840. doi = "10.3389/fnins.2016.00042",
  1841. pmc = "PMC4763020"
  1842. }
  1843. @ARTICLE{Herbin2006-mc,
  1844. title = "How does a mouse increase its velocity? A model for investigation
  1845. in the control of locomotion",
  1846. author = "Herbin, Marc and Gasc, Jean-Pierre and Renous, Sabine",
  1847. abstract = "We analysed treadmill locomotion of the adult SWISS-OF1 mice over
  1848. a large range of velocities. The use of a high-speed video camera
  1849. combined with cinefluoroscopic equipment allowed us to quantify
  1850. in detail the various space and time parameters of limb
  1851. kinematics. We find that velocity adjustments depend upon whether
  1852. animal used a symmetrical or non-symmetrical gait. In symmetrical
  1853. gaits, the increase of velocity generally results equally from an
  1854. increase in the stride frequency and the stride length. On the
  1855. other hand, in non-symmetrical gaits, the increase in velocity is
  1856. achieved differently according to the level of velocity used. As
  1857. speed increases, velocity increases first as a consequence of
  1858. increased stride frequency, then as in symmetrical gaits, by an
  1859. equal increase in both variables, and finally at high speed,
  1860. velocity increases through increased stride length. In both
  1861. symmetrical and non-symmetrical gaits, stance and swing-time
  1862. shortening contributed to the increase of the stride frequency,
  1863. with stance time decrease being the major contributor. The
  1864. pattern of locomotion obtained in the present study may be used
  1865. as a model mouse system for studying locomotor deficits resulting
  1866. from specific mutations in the nervous system. To cite this
  1867. article: M. Herbin et al., C. R. Palevol 5 (2006). R{\'e}sum{\'e}
  1868. Comment la souris augmente-elle sa vitesse ? Un mod{\`e}le pour
  1869. la recherche sur le contr{\^o}le moteur de la locomotion. La
  1870. locomotion sur tapis roulant de la souche de souris SWISS-OF1 a
  1871. {\'e}t{\'e} analys{\'e}e {\`a} travers une large gamme de
  1872. vitesses. L'utilisation de la vid{\'e}oradiographie {\`a} grande
  1873. vitesse a permis de quantifier de fa{\c c}on tr{\`e}s
  1874. d{\'e}taill{\'e}e tous les param{\`e}tres de la cin{\'e}matique
  1875. du membre de r{\'e}f{\'e}rence. Les r{\'e}sultats ainsi obtenus
  1876. montrent que la fr{\'e}quence et l'enjamb{\'e}e n'interviennent
  1877. pas de la m{\^e}me fa{\c c}on dans l'augmentation de la vitesse,
  1878. selon l'allure utilis{\'e}e. Lorsque l'animal est en allure
  1879. sym{\'e}trique, l'augmentation de la vitesse est
  1880. g{\'e}n{\'e}ralement obtenue par une {\'e}gale augmentation de la
  1881. fr{\'e}quence et de l'enjamb{\'e}e. En revanche, si la souris
  1882. utilise une allure non sym{\'e}trique, l'augmentation de la
  1883. vitesse est obtenue diff{\'e}remment selon la valeur de cette
  1884. derni{\`e}re. L'augmentation de la vitesse est d'abord surtout
  1885. assur{\'e}e par une augmentation de la fr{\'e}quence, puis par
  1886. l'augmentation {\'e}gale des deux variables et enfin surtout par
  1887. l'augmentation de l'enjamb{\'e}e. L'augmentation de la
  1888. fr{\'e}quence est, en revanche, surtout assur{\'e}e par une
  1889. diminution de la dur{\'e}e du pos{\'e} et cela, quelle que soit
  1890. l'allure utilis{\'e}e. Cette mod{\'e}lisation de la locomotion
  1891. normale de la souris pourra {\^e}tre utilis{\'e}e comme
  1892. r{\'e}f{\'e}rentiel pour les {\'e}tudes portant sur les
  1893. d{\'e}ficits moteurs de certaines souches de souris mutantes ou
  1894. transg{\'e}niques. Pour citer cet article : M. Herbin et al., C.
  1895. R. Palevol 5 (2006).",
  1896. journal = "C. R. Palevol",
  1897. volume = 5,
  1898. number = 3,
  1899. pages = "531--540",
  1900. month = mar,
  1901. year = 2006,
  1902. keywords = "Stride frequency; Stride length; Treadmill; Locomotion;
  1903. SWISS-OF1; Fr{\'e}quence; Enjamb{\'e}e; Tapis roulant;
  1904. Locomotion; SWISS-OF1;Locomotion",
  1905. issn = "1631-0683",
  1906. doi = "10.1016/j.crpv.2005.12.012"
  1907. }
  1908. @ARTICLE{Walter2003-pb,
  1909. title = "Kinematics of 90 degrees running turns in wild mice",
  1910. author = "Walter, Rebecca M",
  1911. abstract = "Turning is a requirement for locomotion on the variable terrain
  1912. that most terrestrial animals inhabit and is a deciding factor in
  1913. many predator-prey interactions. Despite this, the kinematics and
  1914. mechanics of quadrupedal turns are not well understood. To gain
  1915. insight to the turning kinematics of small quadrupedal mammals,
  1916. six adult wild mice were videotaped at 250 Hz from below as they
  1917. performed 90 degrees running turns. Four markers placed along the
  1918. sagittal axis were digitized to allow observation of lateral
  1919. bending and body rotation throughout the turn. Ground contact
  1920. periods of the fore- and hindlimbs were also noted for each
  1921. frame. During turning, mice increased their ground contact time,
  1922. but did not change their stride frequency relative to straight
  1923. running at maximum speed. Postcranial body rotation preceded
  1924. deflection in heading, and did not occur in one continuous
  1925. motion, but rather in bouts of 15-53 degrees. These bouts were
  1926. synchronized with the stride cycle, such that the majority of
  1927. rotation occurred during the second half of forelimb support and
  1928. the first half of hindlimb support. In this phase of the stride
  1929. cycle, the trunk was sagittally flexed and rotational inertia was
  1930. 65\% of that during maximal extension. By synchronizing body
  1931. rotation with this portion of the stride cycle, mice can achieve
  1932. a given angular acceleration with much lower applied torque.
  1933. Compared with humans running along curved trajectories, mice
  1934. maintained relatively higher speeds at proportionately smaller
  1935. radii. A possible explanation for this difference lies in the
  1936. more crouched limb posture of mice, which increases the
  1937. mechanical advantage for horizontal ground force production. The
  1938. occurrence of body rotation prior to deflection in heading may
  1939. facilitate acceleration in the new direction by making use of the
  1940. relatively greater force production inherent in the parasagittal
  1941. limb posture of mice.",
  1942. journal = "J. Exp. Biol.",
  1943. volume = 206,
  1944. number = "Pt 10",
  1945. pages = "1739--1749",
  1946. month = may,
  1947. year = 2003,
  1948. language = "en",
  1949. issn = "0022-0949",
  1950. pmid = "12682105",
  1951. doi = "10.1242/jeb.00349"
  1952. }
  1953. @ARTICLE{Herbin2004-ma,
  1954. title = "Symmetrical and asymmetrical gaits in the mouse: patterns to
  1955. increase velocity",
  1956. author = "Herbin, Marc and Gasc, Jean-Pierre and Renous, Sabine",
  1957. abstract = "The gaits of the adult SWISS mice during treadmill locomotion at
  1958. velocities ranging from 15 to 85 cm s(-1) have been analysed
  1959. using a high-speed video camera combined with cinefluoroscopic
  1960. equipment. The sequences of locomotion were analysed to determine
  1961. the various space and time parameters of limb kinematics. We
  1962. found that velocity adjustments are accounted for differently by
  1963. the stride frequency and the stride length if the animal showed a
  1964. symmetrical or an asymmetrical gait. In symmetrical gaits, the
  1965. increase of velocity is provided by an equal increase in the
  1966. stride length and the stride frequency. In asymmetrical gaits,
  1967. the increase in velocity is mainly assured by an increase in the
  1968. stride frequency in velocities ranging from 15 to 29 cm s(-1).
  1969. Above 68 cm s(-1), velocity increase is achieved by stride length
  1970. increase. In velocities ranging from 29 to 68 cm s(-1), the
  1971. contribution of both variables is equal as in symmetrical gaits.
  1972. Both stance time and swing time shortening contributed to the
  1973. increase of the stride frequency in both gaits, though with a
  1974. major contribution from stance time decrease. The pattern of
  1975. locomotion obtained in a normal mouse should be used as a
  1976. template for studying locomotor control deficits after lesions or
  1977. in different mutations affecting the nervous system.",
  1978. journal = "J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol.",
  1979. volume = 190,
  1980. number = 11,
  1981. pages = "895--906",
  1982. month = nov,
  1983. year = 2004,
  1984. keywords = "Locomotion",
  1985. language = "en",
  1986. issn = "0340-7594",
  1987. pmid = "15449091",
  1988. doi = "10.1007/s00359-004-0545-0"
  1989. }
  1990. @ARTICLE{Josset2018-js,
  1991. title = "Distinct Contributions of Mesencephalic Locomotor Region Nuclei
  1992. to Locomotor Control in the Freely Behaving Mouse",
  1993. author = "Josset, Nicolas and Roussel, Marie and Lemieux, Maxime and
  1994. Lafrance-Zoubga, David and Rastqar, Ali and Bretzner, Frederic",
  1995. abstract = "The mesencephalic locomotor region (MLR) has been initially
  1996. identified as a supraspinal center capable of initiating and
  1997. modulating locomotion. Whereas its functional contribution to
  1998. locomotion has been widely documented throughout the phylogeny
  1999. from the lamprey to humans, there is still debate about its exact
  2000. organization. Combining kinematic and electrophysiological
  2001. recordings in mouse genetics, our study reveals that
  2002. glutamatergic neurons of the cuneiform nucleus initiate
  2003. locomotion and induce running gaits, whereas glutamatergic and
  2004. cholinergic neurons of the pedunculopontine nucleus modulate
  2005. locomotor pattern and rhythm, contributing to slow-walking gaits.
  2006. By initiating, modulating, and accelerating locomotion, our study
  2007. identifies and characterizes distinct neuronal populations of
  2008. this functional region important to locomotor command.",
  2009. journal = "Curr. Biol.",
  2010. volume = 28,
  2011. number = 6,
  2012. pages = "884--901.e3",
  2013. month = mar,
  2014. year = 2018,
  2015. keywords = "cuneiform nucleus; electrophysiology; glutamatergic and
  2016. cholinergic neurons; kinematic analysis; locomotor command;
  2017. locomotor pattern rhythm and gait; mesencephalic locomotor
  2018. region; optogenetic tools; pedunculopontine nucleus;Locomotion",
  2019. language = "en",
  2020. issn = "0960-9822, 1879-0445",
  2021. pmid = "29526593",
  2022. doi = "10.1016/j.cub.2018.02.007"
  2023. }
  2024. @ARTICLE{Herbin2007-us,
  2025. title = "Gait parameters of treadmill versus overground locomotion in
  2026. mouse",
  2027. author = "Herbin, Marc and Hackert, R{\'e}mi and Gasc, Jean-Pierre and
  2028. Renous, Sabine",
  2029. abstract = "Many studies of interest in motor behaviour and motor impairment
  2030. in mice use equally treadmill or track as a routine test.
  2031. However, the literature in mammals shows a wide difference of
  2032. results between the kinematics of treadmill and overground
  2033. locomotion. To study these discrepancies, we analyzed the
  2034. locomotion of adult SWISS-OF1 mice over a large range of
  2035. velocities using treadmill and overground track. The use of a
  2036. high-speed video camera combined with cinefluoroscopic equipment
  2037. allowed us to quantify in detail the various space and time
  2038. parameters of limb kinematics. The results show that mice
  2039. maintain the same gait pattern in both conditions. However, they
  2040. also demonstrate that during treadmill exercise mice always
  2041. exhibit higher stride frequency and consequently lower stride
  2042. length. The relationship of the stance time and the swing time
  2043. against the stride frequency are still the same in both
  2044. conditions. We conclude that the conflict related to the
  2045. discrepancy between the proprioceptive, vestibular, and visual
  2046. inputs contribute to an increase in the stride frequency during
  2047. the treadmill locomotion.",
  2048. journal = "Behav. Brain Res.",
  2049. volume = 181,
  2050. number = 2,
  2051. pages = "173--179",
  2052. month = aug,
  2053. year = 2007,
  2054. keywords = "Locomotion",
  2055. language = "en",
  2056. issn = "0166-4328",
  2057. pmid = "17521749",
  2058. doi = "10.1016/j.bbr.2007.04.001"
  2059. }
  2060. % The entry below contains non-ASCII chars that could not be converted
  2061. % to a LaTeX equivalent.
  2062. @ARTICLE{Fiker2020-kd,
  2063. title = "Visual Gait Lab: A user-friendly approach to gait analysis",
  2064. author = "Fiker, Robert and Kim, Linda H and Molina, Leonardo A and
  2065. Chomiak, Taylor and Whelan, Patrick J",
  2066. abstract = "BACKGROUND: Gait analysis forms a critical part of many lab
  2067. workflows, ranging from those interested in preclinical
  2068. neurological models to others who use locomotion as part of a
  2069. standard battery of tests. Unfortunately, while paw detection can
  2070. be semi-automated, it becomes generally a time-consuming process
  2071. with error corrections. Improvement in paw tracking would aid in
  2072. better gait analysis performance and experience. NEW METHOD: Here
  2073. we show the use of Visual Gait Lab (VGL), a high-level software
  2074. with an intuitive, easy to use interface, that is built on
  2075. DeepLabCut™. VGL is optimized to generate gait metrics and allows
  2076. for quick manual error corrections. VGL comes with a single
  2077. executable, streamlining setup on Windows systems. We demonstrate
  2078. the use of VGL to analyze gait. RESULTS: Training and evaluation
  2079. of VGL were conducted using 200 frames (80/20 train-test split)
  2080. of video from mice walking on a treadmill. The trained network
  2081. was then used to visually track paw placements to compute gait
  2082. metrics. These are processed and presented on the screen where
  2083. the user can rapidly identify and correct errors. COMPARISON WITH
  2084. EXISTING METHODS: Gait analysis remains cumbersome, even with
  2085. commercial software due to paw detection errors. DeepLabCut™ is
  2086. an alternative that can improve visual tracking but is not
  2087. optimized for gait analysis functionality. CONCLUSIONS: VGL
  2088. allows for gait analysis to be performed in a rapid, unbiased
  2089. manner, with a set-up that can be easily implemented and executed
  2090. by those without a background in computer programming.",
  2091. journal = "J. Neurosci. Methods",
  2092. volume = 341,
  2093. pages = "108775",
  2094. month = may,
  2095. year = 2020,
  2096. keywords = "DeepLabCut™; Gait analysis; Gait tracking system; Motor control;
  2097. Mouse locomotion",
  2098. language = "en",
  2099. issn = "0165-0270, 1872-678X",
  2100. pmid = "32428621",
  2101. doi = "10.1016/j.jneumeth.2020.108775"
  2102. }
  2103. @MISC{Valente2007-qx,
  2104. title = "Analysis of the Trajectory of Drosophila melanogaster in a
  2105. Circular Open Field Arena",
  2106. author = "Valente, Dan and Golani, Ilan and Mitra, Partha P",
  2107. editor = "Scalas, Enrico",
  2108. abstract = "BACKGROUND Obtaining a complete phenotypic characterization of a
  2109. freely moving organism is a difficult task, yet such a
  2110. description is desired in many neuroethological studies. Many
  2111. metrics currently used in the literature to describe locomotor
  2112. and exploratory behavior are typically based on average
  2113. quantities or subjectively chosen spatial and temporal
  2114. thresholds. All of these measures are relatively coarse-grained
  2115. in the time domain. It is advantageous, however, to employ
  2116. metrics based on the entire trajectory that an organism takes
  2117. while exploring its environment. METHODOLOGY/PRINCIPAL FINDINGS
  2118. To characterize the locomotor behavior of Drosophila
  2119. melanogaster, we used a video tracking system to record the
  2120. trajectory of a single fly walking in a circular open field
  2121. arena. The fly was tracked for two hours. Here, we present
  2122. techniques with which to analyze the motion of the fly in this
  2123. paradigm, and we discuss the methods of calculation. The measures
  2124. we introduce are based on spatial and temporal probability
  2125. distributions and utilize the entire time-series trajectory of
  2126. the fly, thus emphasizing the dynamic nature of locomotor
  2127. behavior. Marginal and joint probability distributions of speed,
  2128. position, segment duration, path curvature, and reorientation
  2129. angle are examined and related to the observed behavior.
  2130. CONCLUSIONS/SIGNIFICANCE The measures discussed in this paper
  2131. provide a detailed profile of the behavior of a single fly and
  2132. highlight the interaction of the fly with the environment. Such
  2133. measures may serve as useful tools in any behavioral study in
  2134. which the movement of a fly is an important variable and can be
  2135. incorporated easily into many setups, facilitating
  2136. high-throughput phenotypic characterization.",
  2137. month = oct,
  2138. year = 2007,
  2139. doi = "10.1371/journal.pone.0001083"
  2140. }
  2141. % The entry below contains non-ASCII chars that could not be converted
  2142. % to a LaTeX equivalent.
  2143. @ARTICLE{Barthas2017-zs,
  2144. title = "Secondary motor cortex: where `sensory'meets `motor'in the
  2145. rodent frontal cortex",
  2146. author = "Barthas, Florent and Kwan, Alex C",
  2147. abstract = "In rodents, the medial aspect of the secondary motor cortex (M2)
  2148. is known by other names, including medial agranular cortex
  2149. (AGm), medial precentral cortex (PrCm), and frontal orienting
  2150. field (FOF). As a subdivision of the medial prefrontal cortex
  2151. (mPFC), M2 can be defined by a distinct set of afferent and
  2152. efferent connections, microstimulation responses, and lesion
  2153. outcomes. However, the behavioral role of M2 remains mysterious.
  2154. Here, we focus on evidence from rodent studies, highlighting
  2155. recent findings of early and context …",
  2156. journal = "Trends Neurosci.",
  2157. publisher = "Elsevier",
  2158. volume = 40,
  2159. number = 3,
  2160. pages = "181--193",
  2161. year = 2017,
  2162. keywords = "To Read",
  2163. issn = "0166-2236"
  2164. }
  2165. @ARTICLE{Olson2020-lm,
  2166. title = "Secondary Motor Cortex Transforms Spatial Information into
  2167. Planned Action during Navigation",
  2168. author = "Olson, Jacob M and Li, Jamie K and Montgomery, Sarah E and Nitz,
  2169. Douglas A",
  2170. abstract = "Fluid navigation requires constant updating of planned movements
  2171. to adapt to evolving obstacles and goals. For that reason, a
  2172. neural substrate for navigation demands spatial and
  2173. environmental information and the ability to effect actions
  2174. through efferents. The secondary motor cortex (M2) is a prime
  2175. candidate for this role given its interconnectivity with
  2176. association cortices that encode spatial relationships and its
  2177. projection to the primary motor cortex. Here, we report that M2
  2178. neurons robustly encode both planned and current left/right
  2179. turning actions across multiple turn locations in a multi-route
  2180. navigational task. Comparisons within a common statistical
  2181. framework reveal that M2 neurons differentiate contextual
  2182. factors, including environmental position, route, action
  2183. sequence, orientation, and choice availability. Despite
  2184. significant modulation by environmental factors, action
  2185. planning, and execution are the dominant output signals of M2
  2186. neurons. These results identify the M2 as a structure
  2187. integrating spatial information toward the updating of planned
  2188. movements.",
  2189. journal = "Curr. Biol.",
  2190. publisher = "Elsevier",
  2191. volume = 30,
  2192. number = 10,
  2193. pages = "1845--1854.e4",
  2194. month = may,
  2195. year = 2020,
  2196. keywords = "M2; action; allocentric; cortical circuits; decision making;
  2197. egocentric; in vivo electrophysiology; navigation; parietal
  2198. cortex; retrosplenial cortex; systems neuroscience;Locomotion",
  2199. language = "en",
  2200. issn = "0960-9822, 1879-0445",
  2201. pmid = "32302586",
  2202. doi = "10.1016/j.cub.2020.03.016"
  2203. }
  2204. @ARTICLE{Jordan2008-dx,
  2205. title = "Descending command systems for the initiation of locomotion in
  2206. mammals",
  2207. author = "Jordan, Larry M and Liu, Jun and Hedlund, Peter B and Akay,
  2208. Turgay and Pearson, Keir G",
  2209. abstract = "Neurons in the brainstem implicated in the initiation of
  2210. locomotion include glutamatergic, noradrenergic (NA),
  2211. dopaminergic (DA), and serotonergic (5-HT) neurons giving rise
  2212. to descending tracts. Glutamate antagonists block mesencephalic
  2213. locomotor region-induced and spontaneous locomotion, and
  2214. glutamatergic agonists induce locomotion in spinal animals. NA
  2215. and 5-HT inputs to the spinal cord originate in the brainstem,
  2216. while the descending dopaminergic pathway originates in the
  2217. hypothalamus. Agonists acting at NA, DA or 5-HT receptors
  2218. facilitate or induce locomotion in spinal animals. 5-HT neurons
  2219. located in the parapyramidal region (PPR) produce locomotion
  2220. when stimulated in the isolated neonatal rat brainstem-spinal
  2221. cord preparation, and they constitute the first anatomically
  2222. discrete group of spinally-projecting neurons demonstrated to be
  2223. involved in the initiation of locomotion in mammals. Neurons in
  2224. the PPR are activated during treadmill locomotion in adult rats.
  2225. Locomotion evoked from the PPR is mediated by 5-HT(7) and
  2226. 5-HT(2A) receptors, and 5-HT(7) antagonists block locomotion in
  2227. cat, rat and mouse preparations, but have little effect in mice
  2228. lacking 5-HT(7) receptors. 5-HT induced activity in 5-HT(7)
  2229. knockout mice is rhythmic, but coordination among flexor and
  2230. extensor motor nuclei and left and right sides of the spinal
  2231. cord is disrupted. In the adult wild-type mouse, 5-HT(7)
  2232. receptor antagonists impair locomotion, producing patterns of
  2233. activity resembling those induced by 5-HT in 5-HT(7) knockout
  2234. mice. 5-HT(7) receptor antagonists have a reduced effect on
  2235. locomotion in adult 5-HT(7) receptor knockout mice. We conclude
  2236. that the PPR is the source of a descending 5-HT command pathway
  2237. that activates the CPG via 5-HT(7) and 5-HT(2A) receptors.
  2238. Further experiments are necessary to define the putative
  2239. glutamatergic, DA, and NA command pathways.",
  2240. journal = "Brain Res. Rev.",
  2241. publisher = "Elsevier",
  2242. volume = 57,
  2243. number = 1,
  2244. pages = "183--191",
  2245. month = jan,
  2246. year = 2008,
  2247. language = "en",
  2248. issn = "0165-0173",
  2249. pmid = "17928060",
  2250. doi = "10.1016/j.brainresrev.2007.07.019"
  2251. }
  2252. @ARTICLE{Ryczko2013-pw,
  2253. title = "The multifunctional mesencephalic locomotor region",
  2254. author = "Ryczko, Dimitri and Dubuc, R{\'e}jean",
  2255. abstract = "In 1966, Shik, Severin and Orlovskii discovered that electrical
  2256. stimulation of a region at the junction between the midbrain and
  2257. hindbrain elicited controlled walking and running in the cat.
  2258. The region was named Mesencephalic Locomotor Region (MLR). Since
  2259. then, this locomotor center was shown to control locomotion in
  2260. various vertebrate species, including the lamprey, salamander,
  2261. stingray, rat, guinea-pig, rabbit or monkey. In human subjects
  2262. asked to imagine they are walking, there is an increased
  2263. activity in brainstem nuclei corresponding to the MLR (i.e.
  2264. pedunculopontine, cuneiform and subcuneiform nuclei). Clinicians
  2265. are now stimulating (deep brain stimulation) structures
  2266. considered to be part of the MLR to alleviate locomotor symptoms
  2267. of patients with Parkinson's disease. However, the anatomical
  2268. constituents of the MLR still remain a matter of debate,
  2269. especially relative to the pedunculopontine, cuneiform and
  2270. subcuneiform nuclei. Furthermore, recent studies in lampreys
  2271. have revealed that the MLR is more complex than a simple relay
  2272. in a serial descending pathway activating the spinal locomotor
  2273. circuits. It has multiple functions. Our goal is to review the
  2274. current knowledge relative to the anatomical constituents of the
  2275. MLR, and its physiological role, from lamprey to man. We will
  2276. discuss these results in the context of the recent clinical
  2277. studies involving stimulation of the MLR in patients with
  2278. Parkinson's disease.",
  2279. journal = "Curr. Pharm. Des.",
  2280. publisher = "ingentaconnect.com",
  2281. volume = 19,
  2282. number = 24,
  2283. pages = "4448--4470",
  2284. year = 2013,
  2285. keywords = "Locomotion",
  2286. language = "en",
  2287. issn = "1381-6128, 1873-4286",
  2288. pmid = "23360276",
  2289. doi = "10.2174/1381612811319240011"
  2290. }
  2291. @ARTICLE{Grillner2008-ev,
  2292. title = "Neural bases of goal-directed locomotion in vertebrates---An
  2293. overview",
  2294. author = "Grillner, Sten and Wall{\'e}n, Peter and Saitoh, Kazuya and
  2295. Kozlov, Alexander and Robertson, Brita",
  2296. abstract = "The different neural control systems involved in goal-directed
  2297. vertebrate locomotion are reviewed. They include not only the
  2298. central pattern generator networks in the spinal cord that
  2299. generate the basic locomotor synergy and the brainstem command
  2300. systems for locomotion but also the control systems for steering
  2301. and control of body orientation (posture) and finally the neural
  2302. structures responsible for determining which motor programs
  2303. should be turned on in a given instant. The role of the basal
  2304. ganglia is considered in this context. The review summarizes the
  2305. available information from a general vertebrate perspective, but
  2306. specific examples are often derived from the lamprey, which
  2307. provides the most detailed information when considering cellular
  2308. and network perspectives.",
  2309. journal = "Brain Res. Rev.",
  2310. publisher = "Elsevier",
  2311. volume = 57,
  2312. number = 1,
  2313. pages = "2--12",
  2314. month = jan,
  2315. year = 2008,
  2316. keywords = "Basal ganglia; Lamprey; Central pattern generator; Tectum; Brain
  2317. stem--spinal cord; Modeling",
  2318. issn = "0165-0173",
  2319. doi = "10.1016/j.brainresrev.2007.06.027"
  2320. }
  2321. @ARTICLE{Drew2004-kl,
  2322. title = "Cortical and brainstem control of locomotion",
  2323. author = "Drew, Trevor and Prentice, Stephen and Schepens,
  2324. B{\'e}n{\'e}dicte",
  2325. abstract = "While a basic locomotor rhythm is centrally generated by spinal
  2326. circuits, descending pathways are critical for ensuring
  2327. appropriate anticipatory modifications of gait to accommodate
  2328. uneven terrain. Neurons in the motor cortex command the changes
  2329. in muscle activity required to modify limb trajectory when
  2330. stepping over obstacles. Simultaneously, neurons in the
  2331. brainstem reticular formation ensure that these modifications
  2332. are superimposed on an appropriate base of postural support.
  2333. Recent experiments suggest that the same neurons in the same
  2334. structures also provide similar information during reaching
  2335. movements. It is suggested that, during both locomotion and
  2336. reaching movements, the final expression of descending signals
  2337. is influenced by the state and excitability of the spinal
  2338. circuits upon which they impinge.",
  2339. journal = "Prog. Brain Res.",
  2340. publisher = "Elsevier",
  2341. volume = 143,
  2342. pages = "251--261",
  2343. year = 2004,
  2344. keywords = "Locomotion",
  2345. language = "en",
  2346. issn = "0079-6123",
  2347. pmid = "14653170",
  2348. doi = "10.1016/S0079-6123(03)43025-2"
  2349. }
  2350. @UNPUBLISHED{Storchi2020-em,
  2351. title = "Beyond locomotion: in the mouse the mapping between sensations
  2352. and behaviours unfolds in a higher dimensional space",
  2353. author = "Storchi, Riccardo and Milosavljevic, Nina and Allen, Annette E
  2354. and Cootes, Timothy F and Lucas, Robert J",
  2355. abstract = "Abstract The ability of specific sensory stimuli to evoke
  2356. spontaneous behavioural responses in the mouse represents a
  2357. powerful approach to study how the mammalian brain processes
  2358. sensory information and selects appropriate motor actions. For
  2359. visually and auditory guided behaviours the relevant action has
  2360. been empirically identified as a change in locomotion state.
  2361. However, the extent to which locomotion alone captures the
  2362. diversity of those behaviours and their sensory specificity is
  2363. unknown.To tackle this problem we developed a method to obtain a
  2364. faithful 3D reconstruction of the mouse body that enabled us to
  2365. quantify a wide variety of movements and changes in postures.
  2366. This higher dimensional description of behaviour revealed that
  2367. responses to different sensory inputs is more stimulus-specific
  2368. than indicated by locomotion data alone. Thus, equivalent
  2369. locomotion patterns evoked by different stimuli (e.g. looming and
  2370. sound evoking locomotion arrest) could be well separated along
  2371. other dimensions. The enhanced stimulus-specificity was explained
  2372. by a surprising diversity of behavioural responses. A clustering
  2373. analysis revealed that distinct combinations of motor actions and
  2374. postures, giving rise to at least 7 different behaviours, were
  2375. required to account for stimulus-specificity. Moreover, each
  2376. stimulus evoked more than one behaviour revealing a robust
  2377. one-to-many mapping between sensations and behaviours that could
  2378. not be detected from locomotion data.Our results challenge the
  2379. current view of visually and auditory guided behaviours as purely
  2380. locomotion-based actions (e.g. freeze, escape) and indicate that
  2381. behavioural diversity and sensory specificity unfold in a higher
  2382. dimensional space spanning multiple motor actions.",
  2383. journal = "bioRxiv",
  2384. pages = "2020.02.24.961565",
  2385. month = mar,
  2386. year = 2020,
  2387. keywords = "Locomotion",
  2388. language = "en",
  2389. doi = "10.1101/2020.02.24.961565"
  2390. }
  2391. @ARTICLE{Cregg2020-ie,
  2392. title = "Brainstem neurons that command mammalian locomotor asymmetries",
  2393. author = "Cregg, Jared M and Leiras, Roberto and Montalant, Alexia and
  2394. Wanken, Paulina and Wickersham, Ian R and Kiehn, Ole",
  2395. abstract = "Descending command neurons instruct spinal networks to execute
  2396. basic locomotor functions, such as gait and speed. The command
  2397. functions for gait and speed are symmetric, implying that a
  2398. separate unknown system directs asymmetric movements, including
  2399. the ability to move left or right. In the present study, we
  2400. report that Chx10-lineage reticulospinal neurons act to control
  2401. the direction of locomotor movements in mammals. Chx10 neurons
  2402. exhibit mainly ipsilateral projection, and their selective
  2403. unilateral activation causes ipsilateral turning movements in
  2404. freely moving mice. Unilateral inhibition of Chx10 neurons causes
  2405. contralateral turning movements. Paired left--right motor
  2406. recordings identified distinct mechanisms for directional
  2407. movements mediated via limb and axial spinal circuits. Finally,
  2408. we identify sensorimotor brain regions that project on to Chx10
  2409. reticulospinal neurons, and demonstrate that their unilateral
  2410. activation can impart left--right directional commands. Together
  2411. these data identify the descending motor system that commands
  2412. left--right locomotor asymmetries in mammals.",
  2413. journal = "Nat. Neurosci.",
  2414. month = may,
  2415. year = 2020,
  2416. issn = "1097-6256, 1546-1726",
  2417. doi = "10.1038/s41593-020-0633-7"
  2418. }
  2419. @UNPUBLISHED{Rayshubskiy2020-ad,
  2420. title = "Neural control of steering in walking Drosophila",
  2421. author = "Rayshubskiy, Aleksandr and Holtz, Stephen L and D'Alessandro,
  2422. Isabel and Li, Anna A and Vanderbeck, Quinn X and Haber, Isabel S
  2423. and Gibb, Peter W and Wilson, Rachel I",
  2424. abstract = "During navigation, the brain must continuously integrate external
  2425. guidance cues with internal spatial maps to update steering
  2426. commands. However, it has been difficult to link spatial maps
  2427. with motor control. Here we identify 9descending steering9
  2428. neurons in the Drosophila brain that lie two synapses downstream
  2429. from the brain9s heading direction map in the central complex.
  2430. These steering neurons predict behavioral turns caused by
  2431. microstimulation of the spatial map. Moreover, these neurons
  2432. receive 9direct9 sensory input that bypasses the central complex,
  2433. and they predict steering evoked by multimodal stimuli.
  2434. Unilateral activation of these neurons can promote turning, while
  2435. bilateral silencing interferes with body and leg movements. In
  2436. short, these neurons combine internal maps with external cues to
  2437. predict and influence steering. They represent a key link between
  2438. cognitive maps, which use an abstract coordinate frame, and motor
  2439. commands, which use a body-centric coordinate frame.",
  2440. journal = "bioRxiv",
  2441. pages = "2020.04.04.024703",
  2442. month = apr,
  2443. year = 2020,
  2444. language = "en",
  2445. doi = "10.1101/2020.04.04.024703"
  2446. }
  2447. @ARTICLE{Remington2018-tc,
  2448. title = "Flexible Sensorimotor Computations through Rapid Reconfiguration
  2449. of Cortical Dynamics",
  2450. author = "Remington, Evan D and Narain, Devika and Hosseini, Eghbal A and
  2451. Jazayeri, Mehrdad",
  2452. abstract = "Neural mechanisms that support flexible sensorimotor computations
  2453. are not well understood. In a dynamical system whose state is
  2454. determined by interactions among neurons, computations can be
  2455. rapidly reconfigured by controlling the system's inputs and
  2456. initial conditions. To investigate whether the brain employs such
  2457. control mechanisms, we recorded from the dorsomedial frontal
  2458. cortex of monkeys trained to measure and produce time intervals
  2459. in two sensorimotor contexts. The geometry of neural trajectories
  2460. during the production epoch was consistent with a mechanism
  2461. wherein the measured interval and sensorimotor context exerted
  2462. control over cortical dynamics by adjusting the system's initial
  2463. condition and input, respectively. These adjustments, in turn,
  2464. set the speed at which activity evolved in the production epoch,
  2465. allowing the animal to flexibly produce different time intervals.
  2466. These results provide evidence that the language of dynamical
  2467. systems can be used to parsimoniously link brain activity to
  2468. sensorimotor computations.",
  2469. journal = "Neuron",
  2470. volume = 98,
  2471. number = 5,
  2472. pages = "1005--1019.e5",
  2473. month = jun,
  2474. year = 2018,
  2475. keywords = "Dynamical Systems; cognitive flexibility; electrophysiology;
  2476. frontal cortex; motor planning; population coding; recurrent
  2477. neural networks; sensorimotor coordination; timing",
  2478. language = "en",
  2479. issn = "0896-6273, 1097-4199",
  2480. pmid = "29879384",
  2481. doi = "10.1016/j.neuron.2018.05.020",
  2482. pmc = "PMC6009852"
  2483. }
  2484. @ARTICLE{Grillner2002-si,
  2485. title = "Cellular bases of a vertebrate locomotor system-steering,
  2486. intersegmental and segmental co-ordination and sensory control",
  2487. author = "Grillner, Sten and Wall{\'e}n, Peter",
  2488. abstract = "The isolated brainstem-spinal cord of the lamprey is used as an
  2489. experimental model in the analysis of the cellular bases of
  2490. vertebrate locomotor behaviour. In this article we review the
  2491. neural mechanisms involved in the control of steering,
  2492. intersegmental co-ordination, as well as the segmental burst
  2493. generation and the sensory contribution to motor pattern
  2494. generation. Within these four components of the control system
  2495. for locomotion, we now have good knowledge of not only the
  2496. neurones that take part and their synaptic interactions, but also
  2497. the membrane properties of these neurones, including ion channel
  2498. subtypes, and their contribution to motor pattern generation.",
  2499. journal = "Brain Res. Brain Res. Rev.",
  2500. volume = 40,
  2501. number = "1-3",
  2502. pages = "92--106",
  2503. month = oct,
  2504. year = 2002,
  2505. language = "en",
  2506. pmid = "12589909",
  2507. doi = "10.1016/s0165-0173(02)00193-5"
  2508. }
  2509. @ARTICLE{Saitoh2007-yr,
  2510. title = "Tectal control of locomotion, steering, and eye movements in
  2511. lamprey",
  2512. author = "Saitoh, Kazuya and M{\'e}nard, Ariane and Grillner, Sten",
  2513. abstract = "The intrinsic function of the brain stem-spinal cord networks
  2514. eliciting the locomotor synergy is well described in the
  2515. lamprey-a vertebrate model system. This study addresses the role
  2516. of tectum in integrating eye, body orientation, and locomotor
  2517. movements as in steering and goal-directed behavior. Electrical
  2518. stimuli were applied to different areas within the optic tectum
  2519. in head-restrained semi-intact lampreys (n = 40). Motions of the
  2520. eyes and body were recorded simultaneously (videotaped). Brief
  2521. pulse trains (0.5 s) lateral bending movements of the body
  2522. (orientation movements) were added, and with even longer stimuli
  2523. locomotor movements were initiated. Depending on the tectal area
  2524. stimulated, four characteristic response patterns were observed.
  2525. In a lateral area conjugate horizontal eye movements combined
  2526. with lateral bending movements of the body and locomotor
  2527. movements were elicited, depending on stimulus duration. The
  2528. amplitude of the eye movement and bending movements was site
  2529. specific within this region. In a rostromedial area, bilateral
  2530. downward vertical eye movements occurred. In a caudomedial tectal
  2531. area, large-amplitude undulatory body movements akin to
  2532. struggling behavior were elicited, combined with large-amplitude
  2533. eye movements that were antiphasic to the body movements. The
  2534. alternating eye movements were not dependent on vestibuloocular
  2535. reflexes. Finally, in a caudolateral area locomotor movements
  2536. without eye or bending movements could be elicited. These results
  2537. show that tectum can provide integrated motor responses of eye,
  2538. body orientation, and locomotion of the type that would be
  2539. required in goal-directed locomotion.",
  2540. journal = "J. Neurophysiol.",
  2541. volume = 97,
  2542. number = 4,
  2543. pages = "3093--3108",
  2544. month = apr,
  2545. year = 2007,
  2546. language = "en",
  2547. issn = "0022-3077",
  2548. pmid = "17303814",
  2549. doi = "10.1152/jn.00639.2006"
  2550. }
  2551. @ARTICLE{Sussillo2013-ey,
  2552. title = "Opening the black box: low-dimensional dynamics in
  2553. high-dimensional recurrent neural networks",
  2554. author = "Sussillo, David and Barak, Omri",
  2555. abstract = "Recurrent neural networks (RNNs) are useful tools for learning
  2556. nonlinear relationships between time-varying inputs and outputs
  2557. with complex temporal dependencies. Recently developed
  2558. algorithms have been successful at training RNNs to perform a
  2559. wide variety of tasks, but the resulting networks have been
  2560. treated as black boxes: their mechanism of operation remains
  2561. unknown. Here we explore the hypothesis that fixed points, both
  2562. stable and unstable, and the linearized dynamics around them,
  2563. can reveal crucial aspects of how RNNs implement their
  2564. computations. Further, we explore the utility of linearization
  2565. in areas of phase space that are not true fixed points but
  2566. merely points of very slow movement. We present a simple
  2567. optimization technique that is applied to trained RNNs to find
  2568. the fixed and slow points of their dynamics. Linearization
  2569. around these slow regions can be used to explore, or
  2570. reverse-engineer, the behavior of the RNN. We describe the
  2571. technique, illustrate it using simple examples, and finally
  2572. showcase it on three high-dimensional RNN examples: a 3-bit
  2573. flip-flop device, an input-dependent sine wave generator, and a
  2574. two-point moving average. In all cases, the mechanisms of
  2575. trained networks could be inferred from the sets of fixed and
  2576. slow points and the linearized dynamics around them.",
  2577. journal = "Neural Comput.",
  2578. publisher = "MIT Press",
  2579. volume = 25,
  2580. number = 3,
  2581. pages = "626--649",
  2582. month = mar,
  2583. year = 2013,
  2584. keywords = "RNN",
  2585. language = "en",
  2586. issn = "0899-7667, 1530-888X",
  2587. pmid = "23272922",
  2588. doi = "10.1162/NECO\_a\_00409"
  2589. }
  2590. @ARTICLE{Sussillo2014-mo,
  2591. title = "Neural circuits as computational dynamical systems",
  2592. author = "Sussillo, David",
  2593. abstract = "Many recent studies of neurons recorded from cortex reveal
  2594. complex temporal dynamics. How such dynamics embody the
  2595. computations that ultimately lead to behavior remains a mystery.
  2596. Approaching this issue requires developing plausible hypotheses
  2597. couched in terms of neural dynamics. A tool ideally suited to aid
  2598. in this question is the recurrent neural network (RNN). RNNs
  2599. straddle the fields of nonlinear dynamical systems and machine
  2600. learning and have recently seen great advances in both theory and
  2601. application. I summarize recent theoretical and technological
  2602. advances and highlight an example of how RNNs helped to explain
  2603. perplexing high-dimensional neurophysiological data in the
  2604. prefrontal cortex.",
  2605. journal = "Curr. Opin. Neurobiol.",
  2606. volume = 25,
  2607. pages = "156--163",
  2608. month = apr,
  2609. year = 2014,
  2610. keywords = "RNN To read;RNN",
  2611. language = "en",
  2612. issn = "0959-4388, 1873-6882",
  2613. pmid = "24509098",
  2614. doi = "10.1016/j.conb.2014.01.008"
  2615. }
  2616. @ARTICLE{Grillner2020-uq,
  2617. title = "Current Principles of Motor Control, with Special Reference to
  2618. Vertebrate Locomotion",
  2619. author = "Grillner, Sten and El Manira, Abdeljabbar",
  2620. abstract = "The vertebrate control of locomotion involves all levels of the
  2621. nervous system from cortex to the spinal cord. Here, we aim to
  2622. cover all main aspects of this complex behavior, from the
  2623. operation of the microcircuits in the spinal cord to the systems
  2624. and behavioral levels and extend from mammalian locomotion to
  2625. the basic undulatory movements of lamprey and fish. The cellular
  2626. basis of propulsion represents the core of the control system,
  2627. and it involves the spinal central pattern generator networks
  2628. (CPGs) controlling the timing of different muscles, the sensory
  2629. compensation for perturbations, and the brain stem command
  2630. systems controlling the level of activity of the CPGs and the
  2631. speed of locomotion. The forebrain and in particular the basal
  2632. ganglia are involved in determining which motor programs should
  2633. be recruited at a given point of time and can both initiate and
  2634. stop locomotor activity. The propulsive control system needs to
  2635. be integrated with the postural control system to maintain body
  2636. orientation. Moreover, the locomotor movements need to be
  2637. steered so that the subject approaches the goal of the locomotor
  2638. episode, or avoids colliding with elements in the environment or
  2639. simply escapes at high speed. These different aspects will all
  2640. be covered in the review.",
  2641. journal = "Physiol. Rev.",
  2642. publisher = "physiology.org",
  2643. volume = 100,
  2644. number = 1,
  2645. pages = "271--320",
  2646. month = jan,
  2647. year = 2020,
  2648. keywords = "basal ganglia; central pattern generators; cerebellum; spinal
  2649. cord; vestibular; visuomotor",
  2650. language = "en",
  2651. issn = "0031-9333, 1522-1210",
  2652. pmid = "31512990",
  2653. doi = "10.1152/physrev.00015.2019"
  2654. }
  2655. @ARTICLE{Woon2019-lj,
  2656. title = "Involvement of the rodent prelimbic and medial orbitofrontal
  2657. cortices in goal-directed action: A brief review",
  2658. author = "Woon, Ellen P and Sequeira, Michelle K and Barbee, Britton R and
  2659. Gourley, Shannon L",
  2660. abstract = "Goal-directed action refers to selecting behaviors based on the
  2661. expectation that they will be reinforced with desirable outcomes.
  2662. It is typically conceptualized as opposing habit-based behaviors,
  2663. which are instead supported by stimulus-response associations and
  2664. insensitive to consequences. The prelimbic prefrontal cortex (PL)
  2665. is positioned along the medial wall of the rodent prefrontal
  2666. cortex. It is indispensable for action-outcome-driven
  2667. (goal-directed) behavior, consolidating action-outcome
  2668. relationships and linking contextual information with
  2669. instrumental behavior. In this brief review, we will discuss the
  2670. growing list of molecular factors involved in PL function.
  2671. Ventral to the PL is the medial orbitofrontal cortex (mOFC). We
  2672. will also summarize emerging evidence from rodents (complementing
  2673. existing literature describing humans) that it too is involved in
  2674. action-outcome conditioning. We describe experiments using
  2675. procedures that quantify responding based on reward value, the
  2676. likelihood of reinforcement, or effort requirements, touching
  2677. also on experiments assessing food consumption more generally. We
  2678. synthesize these findings with the argument that the mOFC is
  2679. essential to goal-directed action when outcome value information
  2680. is not immediately observable and must be recalled and inferred.",
  2681. journal = "J. Neurosci. Res.",
  2682. month = dec,
  2683. year = 2019,
  2684. keywords = "action-outcome; contingency degradation; devaluation; habit;
  2685. mouse; rat; response-outcome; review; reward",
  2686. language = "en",
  2687. issn = "0360-4012, 1097-4547",
  2688. pmid = "31820488",
  2689. doi = "10.1002/jnr.24567"
  2690. }
  2691. @ARTICLE{Kao2019-hv,
  2692. title = "Considerations in using recurrent neural networks to probe neural
  2693. dynamics",
  2694. author = "Kao, Jonathan C",
  2695. abstract = "Recurrent neural networks (RNNs) are increasingly being used to
  2696. model complex cognitive and motor tasks performed by behaving
  2697. animals. RNNs are trained to reproduce animal behavior while also
  2698. capturing key statistics of empirically recorded neural activity.
  2699. In this manner, the RNN can be viewed as an in silico circuit
  2700. whose computational elements share similar motifs with the
  2701. cortical area it is modeling. Furthermore, because the RNN's
  2702. governing equations and parameters are fully known, they can be
  2703. analyzed to propose hypotheses for how neural populations
  2704. compute. In this context, we present important considerations
  2705. when using RNNs to model motor behavior in a delayed reach task.
  2706. First, by varying the network's nonlinear activation and rate
  2707. regularization, we show that RNNs reproducing single-neuron
  2708. firing rate motifs may not adequately capture important
  2709. population motifs. Second, we find that even when RNNs reproduce
  2710. key neurophysiological features on both the single neuron and
  2711. population levels, they can do so through distinctly different
  2712. dynamical mechanisms. To distinguish between these mechanisms, we
  2713. show that an RNN consistent with a previously proposed dynamical
  2714. mechanism is more robust to input noise. Finally, we show that
  2715. these dynamics are sufficient for the RNN to generalize to tasks
  2716. it was not trained on. Together, these results emphasize
  2717. important considerations when using RNN models to probe neural
  2718. dynamics.NEW \& NOTEWORTHY Artificial neurons in a recurrent
  2719. neural network (RNN) may resemble empirical single-unit activity
  2720. but not adequately capture important features on the neural
  2721. population level. Dynamics of RNNs can be visualized in
  2722. low-dimensional projections to provide insight into the RNN's
  2723. dynamical mechanism. RNNs trained in different ways may reproduce
  2724. neurophysiological motifs but do so with distinctly different
  2725. mechanisms. RNNs trained to only perform a delayed reach task can
  2726. generalize to perform tasks where the target is switched or the
  2727. target location is changed.",
  2728. journal = "J. Neurophysiol.",
  2729. volume = 122,
  2730. number = 6,
  2731. pages = "2504--2521",
  2732. month = dec,
  2733. year = 2019,
  2734. keywords = "artificial neural network; motor cortex; neural computation;
  2735. neural dynamics; recurrent neural network;RNN To read;RNN",
  2736. language = "en",
  2737. issn = "0022-3077, 1522-1598",
  2738. pmid = "31619125",
  2739. doi = "10.1152/jn.00467.2018"
  2740. }
  2741. @ARTICLE{Kao2019-wk,
  2742. title = "Neuroscience out of control: control-theoretic perspectives on
  2743. neural circuit dynamics",
  2744. author = "Kao, Ta-Chu and Hennequin, Guillaume",
  2745. abstract = "A major challenge in systems neuroscience is to understand how
  2746. the dynamics of neural circuits give rise to behaviour. Analysis
  2747. of complex dynamical systems is also at the heart of control
  2748. engineering, where it is central to the design of robust control
  2749. strategies. Although a rich engineering literature has grown over
  2750. decades to facilitate the analysis of such systems, little of it
  2751. has percolated into neuroscience so far. Here, we give a brief
  2752. introduction to a number of core control-theoretic concepts that
  2753. provide useful perspectives on neural circuit dynamics. We
  2754. introduce important mathematical tools related to these concepts,
  2755. and establish connections to neural circuit analysis, focusing on
  2756. a number of themes that have arisen from the modern 'state-space'
  2757. view on neural population dynamics.",
  2758. journal = "Curr. Opin. Neurobiol.",
  2759. volume = 58,
  2760. pages = "122--129",
  2761. month = oct,
  2762. year = 2019,
  2763. language = "en",
  2764. issn = "0959-4388, 1873-6882",
  2765. pmid = "31563084",
  2766. doi = "10.1016/j.conb.2019.09.001"
  2767. }
  2768. @INCOLLECTION{Nitz2014-ca,
  2769. title = "The Posterior Parietal Cortex: Interface Between Maps of
  2770. External Spaces and the Generation of Action Sequences",
  2771. booktitle = "{Space,Time} and Memory in the Hippocampal Formation",
  2772. author = "Nitz, Douglas A",
  2773. editor = "Derdikman, Dori and Knierim, James J",
  2774. abstract = "In primates as well as rodents, the posterior parietal cortex
  2775. maps spatial relationships having both egocentric and external
  2776. frames of reference. In this chapter, the form in which rat
  2777. posterior parietal cortex neuronal activity maps position within
  2778. trajectories through the environment is considered in detail and
  2779. compared to the forms of spatial mapping observed for neurons of
  2780. the hippocampus and entorhinal cortex. Evidence is presented to
  2781. indicate that posterior parietal neurons simultaneously map
  2782. positions both within and across segments of paths through an
  2783. environment. It is suggested that the specific nature of
  2784. posterior parietal cortex mapping of space serves, in part, to
  2785. transition knowledge of position in the environment, given by
  2786. hippocampus and entorhinal cortex, into efficient path-running
  2787. behavior via projections to primary and secondary sensory and
  2788. motor cortices. Posterior parietal cortex activity is also
  2789. hypothesized to play a role both in driving trajectory
  2790. dependence of hippocampal place cells and in anchoring spatially
  2791. specific hippocampal and entorhinal cortical activity to the
  2792. boundaries of the observable environment.",
  2793. publisher = "Springer Vienna",
  2794. pages = "27--54",
  2795. year = 2014,
  2796. address = "Vienna",
  2797. keywords = "Spatial Navigation;Authors/Nitz",
  2798. isbn = "9783709112922",
  2799. doi = "10.1007/978-3-7091-1292-2\_2"
  2800. }
  2801. @ARTICLE{Nitz2009-vz,
  2802. title = "Parietal cortex, navigation, and the construction of arbitrary
  2803. reference frames for spatial information",
  2804. author = "Nitz, Douglas",
  2805. abstract = "The registration of spatial information by neurons of the
  2806. parietal cortex takes on many forms. In most experiments,
  2807. spatially modulated parietal activity patterns are found to take
  2808. as their frame of reference some part of the body such as the
  2809. retina. However, recent findings obtained in single neuron
  2810. recordings from both rat and monkey parietal cortex suggest that
  2811. the frame of reference utilized by parietal cortex may also be
  2812. abstract or arbitrary in nature. Evidence in rats comes from work
  2813. indicating that parietal activity in freely behaving rodents is
  2814. organized according to the space defined by routes taken through
  2815. an environment. In monkeys, evidence for an object-centered frame
  2816. of reference has recently been presented. The present work
  2817. reviews single neuron recording experiments in parietal cortex of
  2818. freely behaving rats and considers the potential contribution of
  2819. parietal cortex in solving navigational tasks. It is proposed
  2820. that parietal cortex, in interaction with the hippocampus, plays
  2821. a critical role in the selection of the most appropriate route
  2822. between two points and, in addition, produces a route-based
  2823. positional signal capable of guiding sensorimotor transitions.",
  2824. journal = "Neurobiol. Learn. Mem.",
  2825. volume = 91,
  2826. number = 2,
  2827. pages = "179--185",
  2828. month = feb,
  2829. year = 2009,
  2830. keywords = "navigation;Spatial Navigation;Authors/Nitz",
  2831. language = "en",
  2832. issn = "1074-7427, 1095-9564",
  2833. pmid = "18804545",
  2834. doi = "10.1016/j.nlm.2008.08.007"
  2835. }
  2836. @ARTICLE{Nitz2006-mn,
  2837. title = "Tracking route progression in the posterior parietal cortex",
  2838. author = "Nitz, Douglas A",
  2839. abstract = "Quick and efficient traversal of learned routes is critical to
  2840. the survival of many animals. Routes can be defined by both the
  2841. ordering of navigational epochs, such as continued forward motion
  2842. or execution of a turn, and the distances separating them. The
  2843. neural substrates conferring the ability to fluidly traverse
  2844. complex routes are not well understood, but likely entail
  2845. interactions between frontal, parietal, and rhinal cortices and
  2846. the hippocampus. This paper demonstrates that posterior parietal
  2847. cortical neurons map both individual and multiple navigational
  2848. epochs with respect to their order in a route. In direct contrast
  2849. to spatial firing patterns of hippocampal neurons, parietal
  2850. neurons discharged in a place- and direction-independent fashion.
  2851. Parietal route maps were scalable and versatile in that they were
  2852. independent of the size and spatial configuration of navigational
  2853. epochs. The results provide a framework in which to consider
  2854. parietal function in spatial cognition.",
  2855. journal = "Neuron",
  2856. volume = 49,
  2857. number = 5,
  2858. pages = "747--756",
  2859. month = mar,
  2860. year = 2006,
  2861. keywords = "navigation;Spatial Navigation;Authors/Nitz",
  2862. language = "en",
  2863. issn = "0896-6273",
  2864. pmid = "16504949",
  2865. doi = "10.1016/j.neuron.2006.01.037"
  2866. }