Hypothalamic Control of Forelimb Motor Adaptation

Dane Donegan, Christoph M Kanzler, Julia Büscher, Paulius Viskaitis, Ed F Bracey, Olivier Lambercy, Denis Burdakov, Dane Donegan, Christoph M Kanzler, Julia Büscher, Paulius Viskaitis, Ed F Bracey, Olivier Lambercy, Denis Burdakov

Abstract

The ability to perform skilled arm movements is central to everyday life, as limb impairments in common neurologic disorders such as stroke demonstrate. Skilled arm movements require adaptation of motor commands based on discrepancies between desired and actual movements, called sensory errors. Studies in humans show that this involves predictive and reactive movement adaptations to the errors, and also requires a general motivation to move. How these distinct aspects map onto defined neural signals remains unclear, because of a shortage of equivalent studies in experimental animal models that permit neural-level insights. Therefore, we adapted robotic technology used in human studies to mice, enabling insights into the neural underpinnings of motivational, reactive, and predictive aspects of motor adaptation. Here, we show that forelimb motor adaptation is regulated by neurons previously implicated in motivation and arousal, but not in forelimb motor control: the hypothalamic orexin/hypocretin neurons (HONs). By studying goal-oriented mouse-robot interactions in male mice, we found distinct HON signals occur during forelimb movements and motor adaptation. Temporally-delimited optosilencing of these movement-associated HON signals impaired sensory error-based motor adaptation. Unexpectedly, optosilencing affected neither task reward or execution rates, nor motor performance in tasks that did not require adaptation, indicating that the temporally-defined HON signals studied here were distinct from signals governing general task engagement or sensorimotor control. Collectively, these results reveal a hypothalamic neural substrate regulating forelimb motor adaptation.SIGNIFICANCE STATEMENT The ability to perform skilled, adaptable movements is a fundamental part of daily life, and is impaired in common neurologic diseases such as stroke. Maintaining motor adaptation is thus of great interest, but the necessary brain components remain incompletely identified. We found that impaired motor adaptation results from disruption of cells not previously implicated in this pathology: hypothalamic orexin/hypocretin neurons (HONs). We show that temporally confined HON signals are associated with skilled movements. Without these newly-identified signals, a resistance to movement that is normally rapidly overcome leads to prolonged movement impairment. These results identify natural brain signals that enable rapid and effective motor adaptation.

Keywords: hypothalamus; motor adaptation; motor learning; movement; orexin/hypocretin; upper limb.

Copyright © 2022 the authors.

Figures

Figure 1.
Figure 1.
Mouse-robot interactions reveal forelimb motor adaptations to defined force fields. A, Left, Diagram of the ETH Pattus robotic manipulandum with top-down view of a head-fixed mouse, pulling the robot handle. Right, Diagram of the skilled motor task. Mice had to pull 9 mm in the y-direction within a ±5 mm x “tunnel” to obtain a reward. On trials 51–200, a velocity-dependent force (cyan arrows) pushed the animal's forelimb in the x direction. B, Top row, Diagram of the sequence of a single trial. Bottom row, Diagram of a single force field session. The mice performed 50 trials with no force field to get a baseline (black bar), then 150 trials of force field exposure (cyan to magenta bar), followed by 50 trials of no force field to determine the washout (red bar). C, Graphical explanation of end x-position and steering angle. D, Left panel, Typical example of trajectory evolution across a force field session. The force field was present in trials 51–200, individual lines represent average of 25 blocks. Right panel, Magnification of the first 0.5 mm of the left trajectories, demonstrating preemptive steering adaptation. E, Averages of five trajectories taken from preforce field (black), force field (blue, purple, magenta), and postforce field (washout, red) trial blocks. Data from 33 sessions from 12 mice. A, Left panel, Lateral displacement at the end of pulls (end x-position), expressed as mean ± SEM, versus trials (n = 33 sessions from 12 mice). The shared x-axis for panels F, H is given in panel H. Right panel, End x-position within indicated trials (mean ± SEM). RM ANOVA F(4,128) = 86.28, p < 0.0001; Sidak's multiple comparison tests ****p < 0.0001, ***p < 0.001. G, Left panel, Same as in F, left panel, for predictive steering angle. Right panel, Same as in F, right panel, for predictive steering angle. RM ANOVA F(4,128) = 18.07, p < 0.0001; Sidak's multiple comparison tests: ****p < 0.0001, ns = p > 0.05. H, Left panel, Same as in F, left panel, showing max y-velocity, ns = p > 0.05. Right panel, Same as in F, right panel, for max y-velocity. RM ANOVA F(4,128) = 3.578, p < 0.01; Sidak's multiple comparison tests: ns = p > 0.05. I, Left panel, Same as in F, left panel, for percentage of successful trials. Right panel, Same as in F, right panel, for percentage of successful trials. RM ANOVA F(4,128) = 5.532, p < 0.001; Sidak's multiple comparison tests: ns = p > 0.05, ****p < 0.0001.
Figure 2.
Figure 2.
HON signals associated with skilled forelimb movements and motor adaptation. A, Targeting schematic (left) and typical expression (right) of AAV1-hORX-GCaMP6s in HONs, and example fiber placements in two additional mice (bottom). Dashed yellow box indicates fiber placement. V3, third ventricle; VMH, ventromedial nucleus of the hypothalamus; LH, lateral hypothalamus. B, Typical example of raw Ca2+-dependent (470 = 470 nm-excited fluorescence) and Ca2+-independent (405 = 405 nm-excited fluorescence) components of HON-GCaMP6s signal during trials. Epochs of the single trial are labeled: light blue line indicates the cue onset, blue bar indicates the pull, and magenta line indicates reward delivery. C, Z-scored HON-GCaMP6s activity of left (black) and right hemispheres (blue) during nonforce field trials with an intertrial interval of >15 s. Cue bar (gray) indicates the range of trial cue onsets across animals. Blue bar indicates the range of pull durations across animals, magenta bar indicates the range of reward delivery times. Data are mean ± SEM for n = 50 trials from 4 mice. D, Top, Diagram explaining the window of analysis of premotor and motor-related HON-GCaMP6s signals of a single trial (green bar). Bottom, Evolution of pull-associated HON-GCaMP6s signals across control nonforce field trials (left), and force field trials (right), aligned to pull completion. Traces are baseline subtracted from −2 to −1 s before pull start and averaged across sessions (NForce field = 15 sessions, NNo Force field = 9 sessions, from 4 mice). E, Average GCaMP6s signal amplitude, expressed as Δ Z – score, within indicated trial blocks from D of no force field (black) and force field (green) sessions. Two-way mixed-effects model analysis, interaction F(4,56) = 8.082, p < 0.0001; Sidak's multiple comparison tests: ****p < 0.0001, ns = p > 0.05. Signal trends across the force field trials (cyan shading) and equivalent no force field trials (yellow shading) are compared in Results. F, Average GCaMP6s signal amplitude versus trial number across force field (green) and no force field sessions (black), Δ Z – score represents signal amplitude (see Materials and Methods; NForce field = 15 sessions, NNo Force field = 9 sessions, from 4 mice).
Figure 3.
Figure 3.
HON optosilencing impairs motor adaptation. A, Top, Targeting schematic (left) and typical expression of ArchT (right) in HONs. Bottom, Single trial diagram depicting when laser stimulation was applied (magenta bar). B, Whole-cell patch-clamp brain slice recording confirming optosilencing of an HON-ArchT neuron by green laser (representative example of n = 20 cells). C, Average trajectories of baseline, force field, and washout trajectories in control and optosilencing sessions. Blocks of 10 trials were averaged within a session and then averaged across sessions. Means ± SEM of 19 control and 14 optosilencing sessions from 7 mice. D, Average end x-position value versus trial number. Means ± SEM of 19 control (black) and 14 optosilencing (magenta) sessions from 7 mice. Inset, Exponential fits to the data (thick lines) and their confidence intervals (thin lines) demonstrating reduced adaptation rate in control versus optosilenced sessions (extra sum-of-squares F test of whether k fit parameters are different between datasets, F(1,4801) = 14.08, p < 0.001). Smoothed means are shown for visualization only, the fit was performed using individual trial points of unsmoothed data. E, For data shown in D, initial adaptation rates from individual sessions, obtained using least-squares linear fits of the first 20 force field trials of each session. Violin plot of density, dashed line at median and dotted line at the quartiles, two-tailed Mann–Whitney test, U(Ncontrol = 19, Noptosilenced = 14) = 73, *p < 0.05. F, Average end x-position value versus trial number of light control mice. Means ± SEM of 10 control (black) and 13 light on (orange) sessions from 6 mice. Inset, Exponential fits to the data (thick lines) and their confidence intervals (thin lines), demonstrating no difference in adaptation rate because of brain heating (extra sum-of-squares F test of whether k fit parameters are different between datasets, F(1,3378) = 3.330, p > 0.05) G. Same as in E, for light controls. Violin plot of density, dashed line at median and dotted line at the quartiles, two-tailed Mann–Whitney test, U(NLightOff = 10, NLightOn = 13) = 58, p > 0.05.
Figure 4.
Figure 4.
HON optosilencing does not impair reward rates or motivational metrics. A, Average percentage of successful trials versus trial number of control and HON optosilencing sessions. Means ± SEM of 19 control and 14 optosilencing sessions from 7 mice. B, For data shown in A, initial adaptation rates from individual sessions, obtained using least-squares linear fits of the first 20 force field trials of each session. Violin plot of density, dashed line at median and dotted line at the quartiles, Mann–Whitney test, U(Ncontrol = 19, Noptosilenced = 14) = 131, ns = p > 0.05. C, As in A for reaction time, calculated as the time between cue onset and pull onset. D, As in B for data in C, Mann–Whitney test, U(Ncontrol = 19, Noptosilenced = 14) = 105, ns = p > 0.05.
Figure 5.
Figure 5.
Forelimb adaptation and effects of HON optosilencing follow a sensory prediction error-based model. A, Orange line shows the average end x-position value versus trial number for control force field data from 19 sessions. The shaded orange area depicts the SEM. Black line depicts a sensory prediction error-based model that was fitted to the force field data of each session. The blue line depicts the reward prediction error-based model. The average of 50 realizations of the model with different random initial parameters is visualized. The sensory prediction error-based model shows population-level root-mean-square error of the fit is 1.12 cm, and the reward prediction error-based model shows population-level root-mean-square error of the fit is 1.98 cm. B, Normalized command strength contributions of the reward prediction error-based model (blue) and the sensory prediction error-based model (black) of control sessions. The command strength of the sensory and reward-based models were normalized with respect to the sensory and reward prediction error, respectively. Violin plot of density, dashed line at median and dotted line at the quartiles, two-tailed Mann–Whitney test, U(Ncontrol = 152, NOptosilenced = 156) = 737, ****p < 0.0001. C, As in A for only the sensory prediction error-based model for both control and optosilencing sessions. D, As in B for the normalized contributions of sensory prediction error-based model for both control and optosilencing sessions. Violin plot of density, dashed line at median and dotted line at the quartiles, two-tailed Mann–Whitney test, U(Ncontrol = 152, NOptosilenced = 156) = 6295, ****p < 0.0001.
Figure 6.
Figure 6.
HON optosilencing disrupts sensory prediction error-based model through impaired sensorimotor integration. A, Left panel, Average max y-velocity versus trial number. Means ± SEM of n = 19 force field sessions and 14 optosilencing sessions from 7 mice. Right panel, Initial adaptation rates from individual sessions for data shown in left panel, obtained using least-squares linear fits of the first 20 force field trials of each session. Violin plot of density, dashed line at median and dotted line at the quartiles, two-tailed Mann–Whitney test, U(NControl = 19, NOptosilenced = 14) = 129, ns = p > 0.05. B, Average probability of a trial having a counterforce feedback correction versus trial number. Inset describes an example trial with counterforce feedback correction (for calculation, see Materials and Methods). C, Left panel, Same as A, but for steering angle. Inset, Exponential fits to the data (thick lines) and confidence intervals (thin lines), illustrating reduced adaptation rate (extra sum-of-squares F test of whether k fit parameters are different between datasets, F(1,4797) = 13.49, p < 0.001). Smoothed means are shown for visualization only, the fit was performed using individual trial points of unsmoothed data. Right panel, Initial adaptation rates from individual sessions for data shown in A, obtained using least-squares linear fits of the first 20 force field trials of each session. Violin plot of density, dashed line at median and dotted line at the quartiles, two-tailed Mann–Whitney test, U(Ncontrol = 19, Noptosilenced = 14) = 70, *p < 0.05. D, Left panel, Same as in A, left panel, for light control mice, demonstrating no difference in adaptation rate because of tissue heating (extra sum-of-squares F test of whether k fit parameters are different between datasets, F(1,3372) = 0.9682, p > 0.05). Right panel, Same as (A, right panel) for data shown in left panel, two-tailed Mann–Whitney test, U(NLightOff = 10, NLightOff = 13) = 61, ns = p > 0.05. E, Left panel, Average steering angle value versus trial number, without the force field. Means ± SEM of 11 control and 13 optosilencing sessions from 7 mice. Right panel, As in A, right panel, for data in left panel, two-tailed Mann–Whitney test, U(NControl = 11, NOptosilenced = 13) = 64, ns = p > 0.05. F, Left panel, Average end x-position value versus trial number, without the force field. Means ± SEM of 11 control and 13 optosilencing sessions from 7 mice. Right panel, As in A, right panel, for data in left panel, two-tailed Mann–Whitney test, U(NControl = 11, NOptosilenced = 13) = 53, ns = p > 0.05.

References

    1. Arber S, Costa RM (2018) Connecting neuronal circuits for movement. Science 360:1403–1404. 10.1126/science.aat5994
    1. Bassetti CLA, Adamantidis A, Burdakov D, Han F, Gay S, Kallweit U, Khatami R, Koning F, Kornum BR, Lammers GJ, Liblau RS, Luppi PH, Mayer G, Pollmächer T, Sakurai T, Sallusto F, Scammell TE, Tafti M, Dauvilliers Y (2019) Narcolepsy - clinical spectrum, aetiopathophysiology, diagnosis and treatment. Nat Rev Neurol 15:519–539. 10.1038/s41582-019-0226-9
    1. Bayer L, Serafin M, Eggermann E, Saint-Mleux B, Machard D, Jones BE, Mühlethaler M (2004) Exclusive postsynaptic action of hypocretin-orexin on sublayer 6b cortical neurons. J Neurosci 24:6760–6764. 10.1523/JNEUROSCI.1783-04.2004
    1. Bear MF, Connors BW, Paradiso MA (2015) Neuroscience: exploring the brain. Burlington: Jones & Bartlett Publ Inc.
    1. Berniker M, Kording K (2008) Estimating the sources of motor errors for adaptation and generalization. Nat Neurosci 11:1454–1461. 10.1038/nn.2229
    1. Biswabharati S, Jean-Xavier C, Eaton SEA, Lognon AP, Brett R, Hardjasa L, Whelan PJ (2018) Orexinergic modulation of spinal motor activity in the neonatal mouse spinal cord. eNeuro 5:ENEURO.0226-18.2018. 10.1523/ENEURO.0226-18.2018
    1. Boutrel B, Kenny PJ, Specio SE, Martin-Fardon R, Markou A, Koob GF, de Lecea L (2005) Role for hypocretin in mediating stress-induced reinstatement of cocaine-seeking behavior. Proc Natl Acad Sci U S A 102:19168–19173. 10.1073/pnas.0507480102
    1. Butcher PA, Ivry RB, Kuo SH, Rydz D, Krakauer JW, Taylor JA (2017) The cerebellum does more than sensory prediction error-based learning in sensorimotor adaptation tasks. J Neurophysiol 118:1622–1636. 10.1152/jn.00451.2017
    1. Butler EG, Bourke DW, Horne MK (2000) The activity of primate ventrolateral thalamic neurones during motor adaptation. Exp Brain Res 133:514–531. 10.1007/s002210000355
    1. Chemelli RM, Willie JT, Sinton CM, Elmquist JK, Scammell T, Lee C, Richardson JA, Williams SC, Xiong Y, Kisanuki Y, Fitch TE, Nakazato M, Hammer RE, Saper CB, Yanagisawa M (1999) Narcolepsy in orexin knockout mice: molecular genetics of sleep regulation. Cell 98:437–451. 10.1016/s0092-8674(00)81973-x
    1. Cutler DJ, Morris R, Sheridhar V, Wattam TA, Holmes S, Patel S, Arch JR, Wilson S, Buckingham RE, Evans ML, Leslie RA, Williams G (1999) Differential distribution of orexin-A and orexin-B immunoreactivity in the rat brain and spinal cord. Peptides 20:1455–1470. 10.1016/s0196-9781(99)00157-6
    1. da Silva JA, Tecuapetla F, Paixão V, Costa RM (2018) Dopamine neuron activity before action initiation gates and invigorates future movements. Nature 554:244–248. 10.1038/nature25457
    1. Date Y, Mondal MS, Matsukura S, Nakazato M (2000) Distribution of orexin-A and orexin-B (hypocretins) in the rat spinal cord. Neurosci Lett 288:87–90. 10.1016/s0304-3940(00)01195-2
    1. de Lecea L, Kilduff TS, Peyron C, Gao X, Foye PE, Danielson PE, Fukuhara C, Battenberg EL, Gautvik VT, Bartlett FS 2nd, Frankel WN, van den Pol AN, Bloom FE, Gautvik KM, Sutcliffe JG (1998) The hypocretins: hypothalamus-specific peptides with neuroexcitatory activity. Proc Natl Acad Sci U S A 95:322–327. 10.1073/pnas.95.1.322
    1. de Lecea L, Jones BE, Boutrel B, Borgland SL, Nishino S, Bubser M, DiLeone R (2006) Addiction and arousal: alternative roles of hypothalamic peptides. J Neurosci 26:10372–10375. 10.1523/JNEUROSCI.3118-06.2006
    1. Feng H, Wen SY, Qiao QC, Pang YJ, Wang SY, Li HY, Cai J, Zhang KX, Chen J, Hu ZA, Luo FL, Wang GZ, Yang N, Zhang J (2020) Orexin signaling modulates synchronized excitation in the sublaterodorsal tegmental nucleus to stabilize REM sleep. Nat Commun 11:3661. 10.1038/s41467-020-17401-3
    1. Franklin DW, Wolpert DM (2011) Computational mechanisms of sensorimotor control. Neuron 72:425–442. 10.1016/j.neuron.2011.10.006
    1. Gandolfo F, Li C, Benda BJ, Schioppa CP, Bizzi E (2000) Cortical correlates of learning in monkeys adapting to a new dynamical environment. Proc Natl Acad Sci U S A 97:2259–2263. 10.1073/pnas.040567097
    1. Garau C, Blomeley C, Burdakov D (2020) Orexin neurons and inhibitory Agrp–>orexin circuits guide spatial exploration in mice. J Physiol 598:4371–4383. 10.1113/JP280158
    1. González JA, Iordanidou P, Strom M, Adamantidis A, Burdakov D (2016a) Awake dynamics and brain-wide direct inputs of hypothalamic MCH and orexin networks. Nat Commun 7:11395. 10.1038/ncomms11395
    1. González JA, Jensen LT, Iordanidou P, Strom M, Fugger L, Burdakov D (2016b) Inhibitory interplay between orexin neurons and eating. Curr Biol 26:2486–2491. 10.1016/j.cub.2016.07.013
    1. Hara J, Beuckmann CT, Nambu T, Willie JT, Chemelli RM, Sinton CM, Sugiyama F, Yagami KI, Goto K, Yanagisawa M, Sakurai T (2001) Genetic ablation of orexin neurons in mice results in narcolepsy, hypophagia, and obesity. Neuron 30:345–354. 10.1016/S0896-6273(01)00293-8
    1. Harris GC, Aston-Jones G (2006) Arousal and reward: a dichotomy in orexin function. Trends Neurosci 29:571–577. 10.1016/j.tins.2006.08.002
    1. Harris GC, Wimmer M, Aston-Jones G (2005) A role for lateral hypothalamic orexin neurons in reward seeking. Nature 437:556–559. 10.1038/nature04071
    1. Hu B, Yang N, Qiao QC, Hu ZA, Zhang J (2015) Roles of the orexin system in central motor control. Neurosci Biobehav Rev 49:43–54. 10.1016/j.neubiorev.2014.12.005
    1. Inagaki HK, Chen S, Daie K, Finklestein A, Fontolan L, Romani S, Svoboda K (2022) Neural algorithms and circuits for motor planning. Annu Rev Neurosci. Available at 10.1146/annurev-neuro-092021-121730.
    1. Izawa J, Shadmehr R (2011) Learning from sensory and reward prediction errors during motor adaptation. PLoS Comput Biol 7:e1002012. 10.1371/journal.pcbi.1002012
    1. Jin J, Chen Q, Qiao Q, Yang L, Xiong J, Xia J, Hu Z, Chen F (2016) Orexin neurons in the lateral hypothalamus project to the medial prefrontal cortex with a rostro-caudal gradient. Neurosci Lett 621:9–14. 10.1016/j.neulet.2016.04.002
    1. Kalman RE (1960) A new approach to linear filtering and prediction problems. Transactions of the ASME-Journal of Basic Engineering 82:35–45.
    1. Kanzler CM, Rinderknecht MD, Schwarz A, Lamers I, Gagnon C, Held JPO, Feys P, Luft AR, Gassert R, Lambercy O (2020) A data-driven framework for selecting and validating digital health metrics: use-case in neurological sensorimotor impairments. NPJ Digit Med 3:80. 10.1038/s41746-020-0286-7
    1. Karnani MM, Schöne C, Bracey EF, González JA, Viskaitis P, Li HT, Adamantidis A, Burdakov D (2020) Role of spontaneous and sensory orexin network dynamics in rapid locomotion initiation. Prog Neurobiol 187:101771. 10.1016/j.pneurobio.2020.101771
    1. Kawato M (1999) Internal models for motor control and trajectory planning. Curr Opin Neurobiol 9:718–727. 10.1016/s0959-4388(99)00028-8
    1. Kim CK, Yang SJ, Pichamoorthy N, Young NP, Kauvar I, Jennings JH, Lerner TN, Berndt A, Lee SY, Ramakrishnan C, Davidson TJ, Inoue M, Bito H, Deisseroth K (2016) Simultaneous fast measurement of circuit dynamics at multiple sites across the mammalian brain. Nat Methods 13:325–328. 10.1038/nmeth.3770
    1. Kiyashchenko LI, Mileykovskiy BY, Lai YY, Siegel JM (2001) Increased and decreased muscle tone with orexin (hypocretin) microinjections in the locus coeruleus and pontine inhibitory area. J Neurophysiol 85:2008–2016. 10.1152/jn.2001.85.5.2008
    1. Korotkova TM, Eriksson KS, Haas HL, Brown RE (2002) Selective excitation of GABAergic neurons in the substantia nigra of the rat by orexin/hypocretin in vitro. Regul Pept 104:83–89. 10.1016/s0167-0115(01)00323-8
    1. Kosse C, Burdakov D (2019) Natural hypothalamic circuit dynamics underlying object memorization. Nat Commun 10:2505. 10.1038/s41467-019-10484-7
    1. Krakauer JW (2006) Motor learning: its relevance to stroke recovery and neurorehabilitation. Curr Opin Neurol 19:84–90.
    1. Lambercy O, Schubring-Giese M, Vigaru B, Gassert R, Luft AR, Hosp JA (2015) Sub-processes of motor learning revealed by a robotic manipulandum for rodents. Behav Brain Res 278:569–576. 10.1016/j.bbr.2014.10.047
    1. Lee MG, Hassani OK, Jones BE (2005) Discharge of identified orexin/hypocretin neurons across the sleep-waking cycle. J Neurosci 25:6716–6720. 10.1523/JNEUROSCI.1887-05.2005
    1. Li CS, Padoa-Schioppa C, Bizzi E (2001) Neuronal correlates of motor performance and motor learning in the primary motor cortex of monkeys adapting to an external force field. Neuron 30:593–607. 10.1016/S0896-6273(01)00301-4
    1. Mahler SV, Moorman DE, Smith RJ, James MH, Aston-Jones G (2014) Motivational activation: a unifying hypothesis of orexin/hypocretin function. Nat Neurosci 17:1298–1303. 10.1038/nn.3810
    1. Mathis MW, Mathis A, Uchida N (2017) Somatosensory cortex plays an essential role in forelimb motor adaptation in mice. Neuron 93:1493–1503.e6. 10.1016/j.neuron.2017.02.049
    1. Mazzoni P, Krakauer JW (2006) An implicit plan overrides an explicit strategy during visuomotor adaptation. J Neurosci 26:3642–3645. 10.1523/JNEUROSCI.5317-05.2006
    1. Mileykovskiy BY, Kiyashchenko LI, Siegel JM (2002) Muscle tone facilitation and inhibition after orexin-a (hypocretin-1) microinjections into the medial medulla. J Neurophysiol 87:2480–2489. 10.1152/jn.2002.87.5.2480
    1. Mileykovskiy BY, Kiyashchenko LI, Siegel JM (2005) Behavioral correlates of activity in identified hypocretin/orexin neurons. Neuron 46:787–798. 10.1016/j.neuron.2005.04.035
    1. Mochizuki T, Crocker A, McCormack S, Yanagisawa M, Sakurai T, Scammell TE (2004) Behavioral state instability in orexin knock-out mice. J Neurosci 24:6291–6300. 10.1523/JNEUROSCI.0586-04.2004
    1. Nambu T, Sakurai T, Mizukami K, Hosoya Y, Yanagisawa M, Goto K (1999) Distribution of orexin neurons in the adult rat brain. Brain Res 827:243–260. 10.1016/s0006-8993(99)01336-0
    1. Nishino S, Ripley B, Overeem S, Lammers GJ, Mignot E (2000) Hypocretin (orexin) deficiency in human narcolepsy. Lancet 355:39–40. 10.1016/S0140-6736(99)05582-8
    1. Oh Y, Schweighofer N (2019) Minimizing precision-weighted sensory prediction errors via memory formation and switching in motor adaptation. J Neurosci 39:9237–9250. 10.1523/JNEUROSCI.3250-18.2019
    1. Perich MG, Gallego JA, Miller LE (2018) A neural population mechanism for rapid learning. Neuron 100:964–976.e7. 10.1016/j.neuron.2018.09.030
    1. Peyron C, Tighe DK, van den Pol AN, de Lecea L, Heller HC, Sutcliffe JG, Kilduff TS (1998) Neurons containing hypocretin (orexin) project to multiple neuronal systems. J Neurosci 18:9996–10015.
    1. Peyron C, et al. . (2000) A mutation in a case of early onset narcolepsy and a generalized absence of hypocretin peptides in human narcoleptic brains. Nat Med 6:991–997. 10.1038/79690
    1. Sakurai T (2014) The role of orexin in motivated behaviours. Nat Rev Neurosci 15:719–731. 10.1038/nrn3837
    1. Sakurai T, et al. . (1998) Orexins and orexin receptors: a family of hypothalamic neuropeptides and G protein-coupled receptors that regulate feeding behavior. Cell 92:573–585. 10.1016/S0092-8674(00)80949-6
    1. Schaefer SY, Shelly IL, Thoroughman KA (2012) Beside the point: motor adaptation without feedback-based error correction in task-irrelevant conditions. J Neurophysiol 107:1247–1256. 10.1152/jn.00273.2011
    1. Scheidt RA, Stoeckmann T (2007) Reach adaptation and final position control amid environmental uncertainty after stroke. J Neurophysiol 97:2824–2836. 10.1152/jn.00870.2006
    1. Schwarz A, Kanzler CM, Lambercy O, Luft AR, Veerbeek JM (2019) Systematic review on kinematic assessments of upper limb movements after stroke. Stroke 50:718–727. 10.1161/STROKEAHA.118.023531
    1. Shadmehr R, Mussa-Ivaldi FA (1994) Adaptive representation of dynamics during learning of a motor task. J Neurosci 14:3208–3224.
    1. Shadmehr R, Smith MA, Krakauer JW (2010) Error correction, sensory prediction, and adaptation in motor control. Annu Rev Neurosci 33:89–108. 10.1146/annurev-neuro-060909-153135
    1. Smith MA, Shadmehr R (2005) Intact ability to learn internal models of arm dynamics in Huntington's disease but not cerebellar degeneration. J Neurophysiol 93:2809–2821. 10.1152/jn.00943.2004
    1. Takahashi CD, Reinkensmeyer DJ (2003) Hemiparetic stroke impairs anticipatory control of arm movement. Exp Brain Res 149:131–140. 10.1007/s00221-002-1340-1
    1. Thannickal TC, Moore RY, Nienhuis R, Ramanathan L, Gulyani S, Aldrich M, Cornford M, Siegel JM (2000) Reduced number of hypocretin neurons in human narcolepsy. Neuron 27:469–474. 10.1016/s0896-6273(00)00058-1
    1. Thoroughman KA, Shadmehr R (2000) Learning of action through adaptive combination of motor primitives. Nature 407:742–747. 10.1038/35037588
    1. Todorov E, Jordan MI (2002) Optimal feedback control as a theory of motor coordination. Nat Neurosci 5:1226–1235. 10.1038/nn963
    1. van den Pol AN (1999) Hypothalamic hypocretin (orexin): robust innervation of the spinal cord. J Neurosci 19:3171–3182. 10.1523/JNEUROSCI.19-08-03171.1999
    1. Vidal PP, Degallaix L, Josset P, Gasc JP, Cullen KE (2004) Postural and locomotor control in normal and vestibularly deficient mice. J Physiol 559:625–638. 10.1113/jphysiol.2004.063883
    1. Vigaru B, Lambercy O, Graber L, Fluit R, Wespe P, Schubring-Giese M, Luft A, Gassert R (2011) A small-scale robotic manipulandum for motor training in stroke rats. IEEE Int Conf Rehabil Robot 2011:5975349. 10.1109/ICORR.2011.5975349
    1. Vigaru BC, Lambercy O, Schubring-Giese M, Hosp JA, Schneider M, Osei-Atiemo C, Luft A, Gassert R (2013) A robotic platform to assess, guide and perturb rat forelimb movements. IEEE Trans Neural Syst Rehabil Eng 21:796–805. 10.1109/TNSRE.2013.2240014
    1. Wayner MJ (1970) Motor control functions of the lateral hypothalamus and adjunctive behavior. Physiol Behav 5:1319–1325. 10.1016/0031-9384(70)90048-x
    1. Wolpert DM, Miall RC (1996) Forward models for physiological motor control. Neural Netw 9:1265–1279. 10.1016/s0893-6080(96)00035-4
    1. Wolpert DM, Kawato M (1998) Multiple paired forward and inverse models for motor control. Neural Netw 11:1317–1329. 10.1016/s0893-6080(98)00066-5
    1. Wolpert DM, Flanagan JR (2016) Computations underlying sensorimotor learning. Curr Opin Neurobiol 37:7–11. 10.1016/j.conb.2015.12.003
    1. Wolpert DM, Miall RC, Kawato M (1998) Internal models in the cerebellum. Trends Cogn Sci 2:338–347. 10.1016/s1364-6613(98)01221-2
    1. Wolpert DM, Diedrichsen J, Flanagan JR (2011) Principles of sensorimotor learning. Nat Rev Neurosci 12:739–751. 10.1038/nrn3112
    1. Yamuy J, Fung SJ, Xi M, Chase MH (2004) Hypocretinergic control of spinal cord motoneurons. J Neurosci 24:5336–5345. 10.1523/JNEUROSCI.4812-03.2004
    1. Zhang J, Li B, Yu L, He YC, Li HZ, Zhu JN, Wang JJ (2011) A role for orexin in central vestibular motor control. Neuron 69:793–804. 10.1016/j.neuron.2011.01.026

Source: PubMed

3
Subscribe