Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Preparatory activity and the expansive null-space

Abstract

The study of the cortical control of movement experienced a conceptual shift over recent decades, as the basic currency of understanding shifted from single-neuron tuning towards population-level factors and their dynamics. This transition was informed by a maturing understanding of recurrent networks, where mechanism is often characterized in terms of population-level factors. By estimating factors from data, experimenters could test network-inspired hypotheses. Central to such hypotheses are ‘output-null’ factors that do not directly drive motor outputs yet are essential to the overall computation. In this Review, we highlight how the hypothesis of output-null factors was motivated by the venerable observation that motor-cortex neurons are active during movement preparation, well before movement begins. We discuss how output-null factors then became similarly central to understanding neural activity during movement. We discuss how this conceptual framework provided key analysis tools, making it possible for experimenters to address long-standing questions regarding motor control. We highlight an intriguing trend: as experimental and theoretical discoveries accumulate, the range of computational roles hypothesized to be subserved by output-null factors continues to expand.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Preparatory and execution activity in empirical data and an artificial spiking network.
Fig. 2: Output-null factors are hypothesized to serve multiple purposes.
Fig. 3: Illustration of the null-space hypothesis and its predictions.
Fig. 4: Single-neuron responses, during a delayed-reach task, from motor cortex and from simulated models.
Fig. 5: The null-space allows low trajectory tangling.
Fig. 6: The ability to estimate interpretable factors allowed experiments to address key questions regarding motor cortex.
Fig. 7: Data and simulations supporting the flexibility-via-subspace hypothesis.

Similar content being viewed by others

References

  1. Tanji, J. & Evarts, E. V. Anticipatory activity of motor cortex neurons in relation to direction of an intended movement. J. Neurophysiol. 39, 1062–1068 (1976).

    CAS  PubMed  Google Scholar 

  2. Weinrich, M., Wise, S. P. & Mauritz, K. H. A neurophysiological study of the premotor cortex in the rhesus monkey. Brain 107, 385–414 (1984).

    PubMed  Google Scholar 

  3. Wise, S. P. The primate premotor cortex: past, present, and preparatory. Annu. Rev. Neurosci. 8, 1–19 (1985).

    CAS  PubMed  Google Scholar 

  4. Churchland, M. M., Santhanam, G. & Shenoy, K. V. Preparatory activity in premotor and motor cortex reflects the speed of the upcoming reach. J. Neurophysiol. 96, 3130–3146 (2006).

    PubMed  Google Scholar 

  5. Messier, J. & Kalaska, J. F. Covariation of primate dorsal premotor cell activity with direction and amplitude during a memorized-delay reaching task. J. Neurophysiol. 84, 152–165 (2000).

    CAS  PubMed  Google Scholar 

  6. Churchland, M. M., Cunningham, J. P., Kaufman, M. T., Ryu, S. I. & Shenoy, K. V. Cortical preparatory activity: representation of movement or first cog in a dynamical machine? Neuron 68, 387–400 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Churchland, M. M., Afshar, A. & Shenoy, K. V. A central source of movement variability. Neuron 52, 1085–1096 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Riehle, A. & Requin, J. The predictive value for performance speed of preparatory changes in neuronal activity of the monkey motor and premotor cortex. Behav. Brain Res. 53, 35–49 (1993).

    CAS  PubMed  Google Scholar 

  9. Churchland, M. M., Yu, B. M., Ryu, S. I., Santhanam, G. & Shenoy, K. V. Neural variability in premotor cortex provides a signature of motor preparation. J. Neurosci. 26, 3697–3712 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Churchland, M. M. & Shenoy, K. V. Delay of movement caused by disruption of cortical preparatory activity. J. Neurophysiol. 97, 348–359 (2007).

    PubMed  Google Scholar 

  11. Li, N., Daie, K., Svoboda, K. & Druckmann, S. Robust neuronal dynamics in premotor cortex during motor planning. Nature 532, 459–464 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Svoboda, K. & Li, N. Neural mechanisms of movement planning: motor cortex and beyond. Curr. Opin. Neurobiol. 49, 33–41 (2018).

    CAS  PubMed  Google Scholar 

  13. Vyas, S., Golub, M. D., Sussillo, D. & Shenoy, K. V. Computation through neural population dynamics. Annu. Rev. Neurosci. 43, 249–275 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Crammond, D. J. & Kalaska, J. F. Prior information in motor and premotor cortex: activity during the delay period and effect on pre-movement activity. J. Neurophysiol. 84, 986–1005 (2000).

    CAS  PubMed  Google Scholar 

  15. Georgopoulos, A. P., Kalaska, J. F., Caminiti, R. & Massey, J. T. On the relations between the direction of two-dimensional arm movements and cell discharge in primate motor cortex. J. Neurosci. 2, 1527–1537 (1982).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Georgopoulos, A. P., Schwartz, A. B. & Kettner, R. E. Neuronal population coding of movement direction. Science 233, 1416–1419 (1986).

    CAS  PubMed  Google Scholar 

  17. Churchland, M. M. & Lisberger, S. G. Shifts in the population response in the middle temporal visual area parallel perceptual and motor illusions produced by apparent motion. J. Neurosci. 21, 9387–9402 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Takemura, A., Inoue, Y., Kawano, K., Quaia, C. & Miles, F. A. Single-unit activity in cortical area MST associated with disparity-vergence eye movements: evidence for population coding. J. Neurophysiol. 85, 2245–2266 (2001).

    CAS  PubMed  Google Scholar 

  19. Reimer, J. & Hatsopoulos, N. G. The problem of parametric neural coding in the motor system. Adv. Exp. Med. Biol. 629, 243–259 (2009).

    PubMed  PubMed Central  Google Scholar 

  20. Fetz, E. E. Are movement parameters recognizably coded in the activity of single neurons? Behav. Brain Sci. 15, 679–690 (1992).

    Google Scholar 

  21. Scott, S. H. Inconvenient truths about neural processing in primary motor cortex. J. Physiol. 586, 1217–1224 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Churchland, M. M. & Shenoy, K. V. Temporal complexity and heterogeneity of single-neuron activity in premotor and motor cortex. J. Neurophysiol. 97, 4235–4257 (2007).

    PubMed  Google Scholar 

  23. Churchland, M. M. et al. Neural population dynamics during reaching. Nature 487, 51–56 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Sussillo, D., Churchland, M. M., Kaufman, M. T. & Shenoy, K. V. A neural network that finds a naturalistic solution for the production of muscle activity. Nat. Neurosci. 18, 1025–1033 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Mante, V., Sussillo, D., Shenoy, K. V. & Newsome, W. T. Context-dependent computation by recurrent dynamics in prefrontal cortex. Nature 503, 78–84 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Robinson, D. A. Implications of neural networks for how we think about brain function. Behav. Brain Sci. 15, 644–655 (1992).

    Google Scholar 

  27. Saxena, S. & Cunningham, J. P. Towards the neural population doctrine. Curr. Opin. Neurobiol. 55, 103–111 (2019).

    CAS  PubMed  Google Scholar 

  28. Barack, D. L. & Krakauer, J. W. Two views on the cognitive brain. Nat. Rev. Neurosci. 22, 359–371 (2021).

    CAS  PubMed  Google Scholar 

  29. Druckmann, S. & Chklovskii, D. B. Neuronal circuits underlying persistent representations despite time varying activity. Curr. Biol. 22, 2095–2103 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Kaufman, M. T., Churchland, M. M., Ryu, S. I. & Shenoy, K. V. Cortical activity in the null space: permitting preparation without movement. Nat. Neurosci. 17, 440–448 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Semedo, J. D., Zandvakili, A., Machens, C. K., Yu, B. M. & Kohn, A. Cortical areas interact through a communication subspace. Neuron 102, 249–259.e4 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Dum, R. P. & Strick, P. L. Spinal cord terminations of the medial wall motor areas in macaque monkeys. J. Neurosci. 16, 6513–6525 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Fetz, E. E. & Cheney, P. D. Postspike facilitation of forelimb muscle activity by primate corticomotoneuronal cells. J. Neurophysiol. 44, 751–772 (1980).

    CAS  PubMed  Google Scholar 

  34. Griffin, D. M., Hudson, H. M., Belhaj-Saif, A., McKiernan, B. J. & Cheney, P. D. Do corticomotoneuronal cells predict target muscle EMG activity? J. Neurophysiol. 99, 1169–1986 (2008).

    CAS  PubMed  Google Scholar 

  35. Leyton, S. S. F. & Sherrington, C. S. Observations on the excitable cortex of the chimpanzee, orang-utan and gorilla. Exp. Physiol. 11, 135–222 (1917).

    Google Scholar 

  36. Penfield, W. & Boldrey, E. Somatic motor and sensory representation in the cerebral cortex of man as studied by electrical stimulation. Brain 60, 389–443 (1937).

    Google Scholar 

  37. Asanuma, H. & Sakata, H. Functional organization of a cortical efferent system examined with focal depth stimulation in cats. J. Neurophysiol. 30, 35–54 (1967).

    Google Scholar 

  38. Russo, A. A. et al. Motor cortex embeds muscle-like commands in an untangled population response. Neuron 97, 953–966.e8 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Marshall, N. J. et al. Flexible neural control of motor units. Nat. Neurosci. 25, 1492–1504 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Erlhagen, W. & Schoner, G. Dynamic field theory of movement preparation. Psychol. Rev. 109, 545–572 (2002).

    PubMed  Google Scholar 

  41. Bastian, A., Schoner, G. & Riehle, A. Preshaping and continuous evolution of motor cortical representations during movement preparation. Eur. J. Neurosci. 18, 2047–2058 (2003).

    PubMed  Google Scholar 

  42. Cisek, P. Integrated neural processes for defining potential actions and deciding between them: a computational model. J. Neurosci. 26, 9761–9770 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Lee, C., Rohrer, W. H. & Sparks, D. L. Population coding of saccadic eye movements by neurons in the superior colliculus. Nature 332, 357–360 (1988).

    CAS  PubMed  Google Scholar 

  44. Glimcher, P. W. & Sparks, D. L. Effects of low-frequency stimulation of the superior colliculus on spontaneous and visually guided saccades. J. Neurophysiol. 69, 953–964 (1993).

    CAS  PubMed  Google Scholar 

  45. Schieber, M. H. & Rivlis, G. Partial reconstruction of muscle activity from a pruned network of diverse motor cortex neurons. J. Neurophysiol. 97, 70–82 (2007).

    PubMed  Google Scholar 

  46. Omrani, M., Kaufman, M. T., Hatsopoulos, N. G. & Cheney, P. D. Perspectives on classical controversies about the motor cortex. J. Neurophysiol. 118, 1828–1848 (2017).

    PubMed  PubMed Central  Google Scholar 

  47. Morrow, M. M., Jordan, L. R. & Miller, L. E. Direct comparison of the task-dependent discharge of M1 in hand space and muscle space. J. Neurophysiol. 97, 1786–1798 (2007).

    CAS  PubMed  Google Scholar 

  48. Sergio, L. E., Hamel-Paquet, C. & Kalaska, J. F. Motor cortex neural correlates of output kinematics and kinetics during isometric-force and arm-reaching tasks. J. Neurophysiol. 94, 2353–2378 (2005).

    PubMed  Google Scholar 

  49. Todorov, E. Direct cortical control of muscle activation in voluntary arm movements: a model. Nat. Neurosci. 3, 391–398 (2000).

    CAS  PubMed  Google Scholar 

  50. Scott, S. H., Gribble, P. L., Graham, K. M. & Cabel, D. W. Dissociation between hand motion and population vectors from neural activity in motor cortex. Nature 413, 161–165 (2001).

    CAS  PubMed  Google Scholar 

  51. Lillicrap, T. P. & Scott, S. H. Preference distributions of primary motor cortex neurons reflect control solutions optimized for limb biomechanics. Neuron 77, 168–179 (2013).

    CAS  PubMed  Google Scholar 

  52. Morrow, M. M., Pohlmeyer, E. A. & Miller, L. E. Control of muscle synergies by cortical ensembles. Adv. Exp. Meb. Biol. 629, 179–199 (2009).

    Google Scholar 

  53. Oby, E. R., Ethier, C. & Miller, L. E. Movement representation in the primary motor cortex and its contribution to generalizable EMG predictions. J. Neurophysiol. 109, 666–678 (2013).

    PubMed  Google Scholar 

  54. Heming, E. A. et al. Primary motor cortex neurons classified in a postural task predict muscle activation patterns in a reaching task. J. Neurophysiol. 115, 2021–2032 (2016).

    PubMed  PubMed Central  Google Scholar 

  55. Kwan, H. C., Yeap, T. H., Jiang, B. C. & Borrett, D. Neural network control of simple limb movements. Can. J. Physiol. Pharmacol. 68, 126–130 (1990).

    CAS  PubMed  Google Scholar 

  56. Fetz, E. E. in The Neurobiology of Neural Networks (ed. Gardner, D.) 165–190 (MIT Press, 1993).

  57. Maass, W., Natschlager, T. & Markram, H. Real-time computing without stable states: a new framework for neural computation based on perturbations. Neural Comput. 14, 2531–2560 (2002).

    PubMed  Google Scholar 

  58. Jaeger, H. & Haas, H. Harnessing nonlinearity: predicting chaotic systems and saving energy in wireless communication. Science 304, 78–80 (2004).

    CAS  PubMed  Google Scholar 

  59. van Vreeswijk, C. & Sompolinsky, H. Chaos in neuronal networks with balanced excitatory and inhibitory activity. Science 274, 1724–1726 (1996).

    PubMed  Google Scholar 

  60. Churchland, M. M. et al. Stimulus onset quenches neural variability: a widespread cortical phenomenon. Nat. Neurosci. 13, 369–378 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Rajan, K., Abbott, L. & Sompolinsky, H. Stimulus-dependent suppression of chaos in recurrent neural networks. Phys. Rev. E 82, 011903 (2010).

    Google Scholar 

  62. Schaffer, E. S., Rajan, K., Churchland, M. M., Shenoy, K. V. & Abbott, L. F. Generating complex repeatable patterns of activity by gain modulating network neurons. In Soc. Neurosci. Annual Meeting 448.3 (2006).

  63. Laje, R. & Buonomano, D. V. Robust timing and motor patterns by taming chaos in recurrent neural networks. Nat. Neurosci. 16, 925–933 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Sussillo, D. & Abbott, L. F. Generating coherent patterns of activity from chaotic neural networks. Neuron 63, 544–557 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Sussillo, D. & Barak, O. Opening the black box: low-dimensional dynamics in high-dimensional recurrent neural networks. Neural Comput. 25, 626–649 (2013).

    MathSciNet  PubMed  Google Scholar 

  66. Rokni, U. & Sompolinsky, H. How the brain generates movement. In Cosyne 222 (2005).

  67. Martens, J. & Sutskever, I. Learning recurrent neural networks with hessian-free optimization. In Proc. 28th Int. Conf. Machine Learning 1033–1040 (2009).

  68. Shenoy, K. V., Kaufman, M. T., Sahani, M. & Churchland, M. M. in Progress in Brain Research: Enhancing Performance for Action and Perception (eds Green, A., Chapman, E., Kalaska, J. F. & Lepore, F.) 33–58 (Elsevier, 2011).

  69. Shenoy, K. V., Sahani, M. & Churchland, M. M. Cortical control of arm movements: a dynamical systems perspective. Annu. Rev. Neurosci. 36, 337–359 (2013).

    CAS  PubMed  Google Scholar 

  70. Truccolo, W., Hochberg, L. R. & Donoghue, J. P. Collective dynamics in human and monkey sensorimotor cortex: predicting single neuron spikes. Nat. Neurosci. 13, 105–111 (2010).

    CAS  PubMed  Google Scholar 

  71. Stevenson, I. H. et al. Functional connectivity and tuning curves in populations of simultaneously recorded neurons. PLoS Comput. Biol. 8, e1002775 (2012).

    MathSciNet  CAS  PubMed  PubMed Central  Google Scholar 

  72. Mastrogiuseppe, F. & Ostojic, S. Linking connectivity, dynamics, and computations in low-rank recurrent neural networks. Neuron 99, 609–623.e29 (2018).

    CAS  PubMed  Google Scholar 

  73. Murphy, B. K. & Miller, K. D. Balanced amplification: a new mechanism of selective amplification of neural activity patterns. Neuron 61, 635–648 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Pandarinath, C. et al. Inferring single-trial neural population dynamics using sequential auto-encoders. Nat. Methods 15, 805–815 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Yu, B. M. et al. Gaussian-process factor analysis for low-dimensional single-trial analysis of neural population activity. J. Neurophysiol. 102, 614–635 (2009).

    PubMed  PubMed Central  Google Scholar 

  76. Seely, J. S. et al. Tensor analysis reveals distinct population structure that parallels the different computational roles of areas M1 and V1. PLoS Comput. Biol. 12, e1005164 (2016).

    PubMed  PubMed Central  Google Scholar 

  77. DePasquale, B., Sussillo, D., Abbott, L. F. & Churchland, M. M. The centrality of population-level factors to network computation is demonstrated by a versatile approach for training spiking networks. Neuron 111, 631–649.e10 (2023).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Cunningham, J. P. & Yu, B. M. Dimensionality reduction for large-scale neural recordings. Nat. Neurosci. 17, 1500–1509 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Williamson, R. C., Doiron, B., Smith, M. A. & Yu, B. M. Bridging large-scale neuronal recordings and large-scale network models using dimensionality reduction. Curr. Opin. Neurobiol. 55, 40–47 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  80. Churchland, M. M., Yu, B. M., Sahani, M. & Shenoy, K. V. Techniques for extracting single-trial activity patterns from large-scale neural recordings. Curr. Opin. Neurobiol. 17, 609–618 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Zimnik, A. J. & Churchland, M. M. Independent generation of sequence elements by motor cortex. Nat. Neurosci. 24, 412–424 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. O’Shea, D. J. et al. Direct neural perturbations reveal a dynamical mechanism for robust computation. Preprint at bioRrxiv https://doi.org/10.1101/2022.12.16.520768 (2022).

  83. Kalaska, J. F. Emerging ideas and tools to study the emergent properties of the cortical neural circuits for voluntary motor control in non-human primates. F1000Research 8, https://doi.org/10.12688/f1000research.17161.1 (2019).

  84. Brown, T. G. On the nature of the fundamental activity of the nervous centres; together with an analysis of the conditioning of rhythmic activity in progression, and a theory of the evolution of function in the nervous system. J. Physiol. 48, 18–46 (1914).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Yuste, R., MacLean, J. N., Smith, J. & Lansner, A. The cortex as a central pattern generator. Nat. Rev. Neurosci. 6, 477–483 (2005).

    CAS  PubMed  Google Scholar 

  86. Goldreich, D., Krauzlis, R. J. & Lisberger, S. G. Effect of changing feedback delay on spontaneous oscillations in smooth pursuit eye movements of monkeys. J. Neurophysiol. 67, 625–638 (1992).

    CAS  PubMed  Google Scholar 

  87. Lisberger, S. G. & Sejnowski, T. J. Motor learning in a recurrent network model based on the vestibulo-ocular reflex. Nature 360, 159–161 (1992).

    CAS  PubMed  Google Scholar 

  88. Churchland, M. M. & Lisberger, S. G. Experimental and computational analysis of monkey smooth pursuit eye movements. J. Neurophysiol. 86, 741–759 (2001).

    CAS  PubMed  Google Scholar 

  89. Kelso, J. A. Multistability and metastability: understanding dynamic coordination in the brain. Philos. Trans. R. Soc. Lond. B Biol. Sci. 367, 906–918 (2012).

    PubMed  PubMed Central  Google Scholar 

  90. van Gelder, T. The dynamical hypothesis in cognitive science. Behav. Brain Sci. 21, 615–665 (1998).

    PubMed  Google Scholar 

  91. Ajemian, R. et al. Assessing the function of motor cortex: single-neuron models of how neural response is modulated by limb biomechanics. Neuron 58, 414–428 (2008).

    CAS  PubMed  Google Scholar 

  92. Scott, S. H. Optimal feedback control and the neural basis of volitional motor control. Nat. Rev. Neurosci. 5, 532–546 (2004).

    CAS  PubMed  Google Scholar 

  93. Todorov, E. Cosine tuning minimizes motor errors. Neural Comput. 14, 1233–1260 (2002).

    PubMed  Google Scholar 

  94. Cisek, P. Preparing for speed. Focus on: “Preparatory activity in premotor and motor cortex reflects the speed of the upcoming reach”. J. Neurophysiol. 96, 2842–2843 (2006).

    PubMed  Google Scholar 

  95. Bizzi, E. & Ajemian, R. From motor planning to execution: a sensorimotor loop perspective. J. Neurophysiol. 124, 1815–1823 (2020).

    PubMed  Google Scholar 

  96. Middleton, F. A. & Strick, P. L. Basal ganglia and cerebellar loops: motor and cognitive circuits. Brain Res. Brain Res. Rev. 31, 236–250 (2000).

    CAS  PubMed  Google Scholar 

  97. Sauerbrei, B. A. et al. Cortical pattern generation during dexterous movement is input-driven. Nature 577, 386–391 (2019).

    PubMed  PubMed Central  Google Scholar 

  98. Pruszynski, J. A. et al. Primary motor cortex underlies multi-joint integration for fast feedback control. Nature 478, 387–390 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Requin, J., Riehle, A. & Seal, J. Neuronal activity and information processing in motor control: from stages to continuous flow. Biol. Psychol. 26, 179–198 (1988).

    CAS  PubMed  Google Scholar 

  100. Crammond, D. J. & Kalaska, J. F. Modulation of preparatory neuronal activity in dorsal premotor cortex due to stimulus-response compatibility. J. Neurophysiol. 71, 1281–1284 (1994).

    CAS  PubMed  Google Scholar 

  101. Pruszynski, J. A. & Scott, S. H. Optimal feedback control and the long-latency stretch response. Exp. Brain Res. 218, 341–359 (2012).

    PubMed  Google Scholar 

  102. Prut, Y. & Fetz, E. E. Primate spinal interneurons show pre-movement instructed delay activity. Nature 401, 590–594 (1999).

    CAS  PubMed  Google Scholar 

  103. Kaufman, M. T. et al. Roles of monkey premotor neuron classes in movement preparation and execution. J. Neurophysiol. 104, 799–810 (2010).

    PubMed  PubMed Central  Google Scholar 

  104. Girard, B. & Berthoz, A. From brainstem to cortex: computational models of saccade generation circuitry. Prog. Neurobiol. 77, 215–251 (2005).

    CAS  PubMed  Google Scholar 

  105. Vogels, T. P., Rajan, K. & Abbott, L. F. Neural network dynamics. Annu. Rev. Neurosci. 28, 357–376 (2005).

    CAS  PubMed  Google Scholar 

  106. Heming, E. A., Cross, K. P., Takei, T., Cook, D. J. & Scott, S. H. Independent representations of ipsilateral and contralateral limbs in primary motor cortex. eLife 8, e48190 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Churchland, M. M. & Cunningham, J. P. A dynamical basis set for generating reaches. Cold Spring Harb. Symp. Quant. Biol. 79, 67–80 (2014).

    PubMed  Google Scholar 

  108. Hennequin, G., Vogels, T. P. & Gerstner, W. Optimal control of transient dynamics in balanced networks supports generation of complex movements. Neuron 82, 1394–1406 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Kao, T. C., Sadabadi, M. S. & Hennequin, G. Optimal anticipatory control as a theory of motor preparation: a thalamo-cortical circuit model. Neuron 109, 1567–1581.e12 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Michaels, J. A., Dann, B. & Scherberger, H. Neural population dynamics during reaching are better explained by a dynamical system than representational tuning. PLoS Comput. Biol. 12, e1005175 (2016).

    PubMed  PubMed Central  Google Scholar 

  111. Morrow, M. M. & Miller, L. E. Prediction of muscle activity by populations of sequentially recorded primary motor cortex neurons. J. Neurophysiol. 89, 2279–2288 (2003).

    CAS  PubMed  Google Scholar 

  112. Ames, K. C. & Churchland, M. M. Motor cortex signals for each arm are mixed across hemispheres and neurons yet partitioned within the population response. eLife 8, e46159 (2019).

    PubMed  PubMed Central  Google Scholar 

  113. Naufel, S., Glaser, J. I., Kording, K. P., Perreault, E. J. & Miller, L. E. A muscle-activity-dependent gain between motor cortex and EMG. J. Neurophysiol. 121, 61–73 (2019).

    PubMed  Google Scholar 

  114. Elsayed, G. F. & Cunningham, J. P. Structure in neural population recordings: an expected byproduct of simpler phenomena? Nat. Neurosci. 20, 1310–1318 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Lara, A. H., Cunningham, J. P. & Churchland, M. M. Different population dynamics in the supplementary motor area and motor cortex during reaching. Nat. Commun. 9, 2754 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Evarts, E. V. Relation of pyramidal tract activity to force exerted during voluntary movement. J. Neurophysiol. 31, 14–27 (1968).

    CAS  PubMed  Google Scholar 

  117. Ashe, J. & Georgopoulos, A. P. Movement parameters and neural activity in motor cortex and area 5. Cereb. Cortex 4, 590–600 (1994).

    CAS  PubMed  Google Scholar 

  118. Sanger, T. D. Theoretical considerations for the analysis of population coding in motor cortex. Neural Comput. 6, 29–37 (1994).

    Google Scholar 

  119. Hatsopoulos, N. G., Xu, Q. & Amit, Y. Encoding of movement fragments in the motor cortex. J. Neurosci. 27, 5105–5114 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  120. Rickert, J., Riehle, A., Aertsen, A., Rotter, S. & Nawrot, M. P. Dynamic encoding of movement direction in motor cortical neurons. J. Neurosci. 29, 13870–13882 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Kalidindi, H. T. et al. Rotational dynamics in motor cortex are consistent with a feedback controller. eLife 10, e67256 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Pandarinath, C. et al. Latent factors and dynamics in motor cortex and their application to brain–machine interfaces. J. Neurosci. 38, 9390–9401 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  123. Saxena, S., Russo, A. A., Cunningham, J. & Churchland, M. M. Motor cortex activity across movement speeds is predicted by network-level strategies for generating muscle activity. eLife 11, e67620 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Foster, J. D. et al. A freely-moving monkey treadmill model. J. Neural Eng. 11, 046020 (2014).

    PubMed  Google Scholar 

  125. Rush, E. R., Jayaram, K. & Humbert, J. S. From data-fitting to discovery: interpreting the neural dynamics of motor control through reinforcement learning. Preprint at arxiv https://doi.org/10.48550/arXiv.2305.11107 (2023).

  126. Linden, H., Petersen, P. C., Vestergaard, M. & Berg, R. W. Movement is governed by rotational neural dynamics in spinal motor networks. Nature 610, 526–531 (2022).

    CAS  PubMed  Google Scholar 

  127. Suresh, A. K. et al. Neural population dynamics in motor cortex are different for reach and grasp. eLife 9, e58848 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Yanai, Y., Adamit, N., Israel, Z., Harel, R. & Prut, Y. Coordinate transformation is first completed downstream of primary motor cortex. J. Neurosci. 28, 1728–1732 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Dacre, J. et al. A cerebellar–thalamocortical pathway drives behavioral context-dependent movement initiation. Neuron 109, 2326–2338.e8 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Gao, Z. et al. A cortico-cerebellar loop for motor planning. Nature 563, 113–116 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Machens, C. K., Romo, R. & Brody, C. D. Functional, but not anatomical, separation of “What” and “When” in prefrontal cortex. J. Neurosci. 30, 350–360 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  132. Raposo, D., Kaufman, M. T. & Churchland, A. K. A category-free neural population supports evolving demands during decision-making. Nat. Neurosci. 17, 1784–1792 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Jun, J. K. et al. Heterogenous population coding of a short-term memory and decision task. J. Neurosci. 30, 916–929 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Elsayed, G. F., Lara, A. H., Kaufman, M. T., Churchland, M. M. & Cunningham, J. P. Reorganization between preparatory and movement population responses in motor cortex. Nat. Commun. 7, 13239 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  135. Lara, A. H., Elsayed, G. F., Zimnik, A. J., Cunningham, J. P. & Churchland, M. M. Conservation of preparatory neural events in monkey motor cortex regardless of how movement is initiated. eLife 7, e31826 (2018).

    PubMed  PubMed Central  Google Scholar 

  136. Ames, K. C., Ryu, S. I. & Shenoy, K. V. Neural dynamics of reaching following incorrect or absent motor preparation. Neuron 81, 438–451 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Rosenbaum, D. A. Human movement initiation: specification of arm, direction, and extent. J. Exp. Psychol. Gen. 109, 444–474 (1980).

    CAS  PubMed  Google Scholar 

  138. Riehle, A. & Requin, J. Monkey primary motor and premotor cortex: single-cell activity related to prior information about direction and extent of an intended movement. J. Neurophysiol. 61, 534–549 (1989).

    CAS  PubMed  Google Scholar 

  139. Ghez, C. et al. Discrete and continuous planning of hand movements and isometric force trajectories. Exp. Brain Res. 115, 217–233 (1997).

    CAS  PubMed  Google Scholar 

  140. Franks, I. M., Nagelkerke, P., Ketelaars, M. & Van Donkelaar, P. Response preparation and control of movement sequences. Can. J. Exp. Psychol. 52, 93–102 (1998).

    CAS  PubMed  Google Scholar 

  141. Ames, K. C., Ryu, S. I. & Shenoy, K. V. Simultaneous motor preparation and execution in a last-moment reach correction task. Nat. Commun. 10, 2718 (2019).

    PubMed  PubMed Central  Google Scholar 

  142. Stavisky, S. D., Kao, J. C., Ryu, S. I. & Shenoy, K. V. Motor cortical visuomotor feedback activity is initially isolated from downstream targets in output-null neural state space dimensions. Neuron 95, 195–208.e9 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  143. Rizzolatti, G. et al. Functional organization of inferior area 6 in the macaque monkey. II. Area F5 and the control of distal movements. Exp. Brain Res. 71, 491–507 (1988).

    CAS  PubMed  Google Scholar 

  144. Haith, A. M., Pakpoor, J. & Krakauer, J. W. Independence of movement preparation and movement initiation. J. Neurosci. 36, 3007–3015 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  145. Haith, A. M., Huberdeau, D. M. & Krakauer, J. W. The influence of movement preparation time on the expression of visuomotor learning and savings. J. Neurosci. 35, 5109–5117 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Kaufman, M. T., Churchland, M. M., Ryu, S. I. & Shenoy, K. V. Vacillation, indecision and hesitation in moment-by-moment decoding of monkey motor cortex. eLife 4, e04677 (2015).

    PubMed  PubMed Central  Google Scholar 

  147. Mirabella, G., Pani, P. & Ferraina, S. Neural correlates of cognitive control of reaching movements in the dorsal premotor cortex of rhesus monkeys. J. Neurophysiol. 106, 1454–1466 (2011).

    CAS  PubMed  Google Scholar 

  148. Kaufman, M. T. et al. The largest response component in the motor cortex reflects movement timing but not movement type. eNeuro 3, https://doi.org/10.1523/ENEURO.0085-16.2016 (2016).

  149. Inagaki, H. K. et al. A midbrain–thalamus–cortex circuit reorganizes cortical dynamics to initiate movement. Cell 185, 1065–1081.e23 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  150. Chabrol, F. P., Blot, A. & Mrsic-Flogel, T. D. Cerebellar contribution to preparatory activity in motor neocortex. Neuron 103, 506–519.e4 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  151. Bastian, A., Riehle, A., Erlhagen, W. & Schoner, G. Prior information preshapes the population representation of movement direction in motor cortex. Neuroreport 9, 315–319 (1998).

    CAS  PubMed  Google Scholar 

  152. Miri, A. et al. Behaviorally selective engagement of short-latency effector pathways by motor cortex. Neuron 95, 683–696.e11 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  153. Warriner, C. L., Fageiry, S., Saxena, S., Costa, R. M. & Miri, A. Motor cortical influence relies on task-specific activity covariation. Cell Rep. 40, 111427 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Briggman, K. L. & Kristan, W. B. Multifunctional pattern-generating circuits. Annu. Rev. Neurosci. 31, 271–294 (2008).

    CAS  PubMed  Google Scholar 

  155. Kurtzer, I., Herter, T. M. & Scott, S. H. Random change in cortical load representation suggests distinct control of posture and movement. Nat. Neurosci. 8, 498–504 (2005).

    CAS  PubMed  Google Scholar 

  156. Donchin, O. et al. Single-unit activity related to bimanual arm movements in the primary and supplementary motor cortices. J. Neurophysiol. 88, 3498–3517 (2002).

    CAS  PubMed  Google Scholar 

  157. Cisek, P., Crammond, D. J. & Kalaska, J. F. Neural activity in primary motor and dorsal premotor cortex in reaching tasks with the contralateral versus ipsilateral arm. J. Neurophysiol. 89, 922–942 (2003).

    PubMed  Google Scholar 

  158. Cross, K. P., Heming, E. A., Cook, D. J. & Scott, S. H. Maintained representations of the ipsilateral and contralateral limbs during bimanual control in primary motor cortex. J. Neurosci. 40, 6732–6747 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  159. Yang, G. R., Joglekar, M. R., Song, H. F., Newsome, W. T. & Wang, X. J. Task representations in neural networks trained to perform many cognitive tasks. Nat. Neurosci. 22, 297–306 (2019).

    CAS  PubMed  Google Scholar 

  160. Driscoll, L., Shenoy, K. & Sussillo, D. Flexible multitask computation in recurrent networks utilizes shared dynamical motifs. Preprint at bioRxiv https://doi.org/10.1101/2022.08.15.503870 (2022).

  161. Logiaco, L., Abbott, L. F. & Escola, S. Thalamic control of cortical dynamics in a model of flexible motor sequencing. Cell Rep. 35, 109090 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  162. Sabatini, D. A. & Kaufman, M. T. Reach-dependent reorientation of rotational dynamics in motor cortex. Preprint at bioRxiv https://doi.org/10.1101/2021.09.09.459647 (2023).

  163. Sadtler, P. T. et al. Neural constraints on learning. Nature 512, 423–426 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  164. Gallego, J. A., Perich, M. G., Miller, L. E. & Solla, S. A. Neural manifolds for the control of movement. Neuron 94, 978–984 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Gallego, J. A., Perich, M. G., Chowdhury, R. H., Solla, S. A. & Miller, L. E. Long-term stability of cortical population dynamics underlying consistent behavior. Nat. Neurosci. 23, 260–270 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  166. Gallego, J. A. et al. Cortical population activity within a preserved neural manifold underlies multiple motor behaviors. Nat. Commun. 9, 4233 (2018).

    PubMed  PubMed Central  Google Scholar 

  167. Gao, P. & Ganguli, S. On simplicity and complexity in the brave new world of large-scale neuroscience. Curr. Opin. Neurobiol. 32, 148–155 (2015).

    CAS  PubMed  Google Scholar 

  168. Gao, P. et al. A theory of multineuronal dimensionality, dynamics and measurement. Preprint at bioRxiv https://doi.org/10.1101/214262 (2017).

  169. Perkins, S. M., Cunningham, J. P., Wang, Q. & Churchland, M. M. Simple decoding of behavior from a complicated neural manifold. Preprint at bioRxiv https://doi.org/10.1101/2023.04.05.535396 (2023).

  170. Schneider, S., Lee, J. H. & Mathis, M. W. Learnable latent embeddings for joint behavioural and neural analysis. Nature 617, 360–368 (2023).

    CAS  PubMed  PubMed Central  Google Scholar 

  171. Chaudhuri, R., Gercek, B., Pandey, B., Peyrache, A. & Fiete, I. The intrinsic attractor manifold and population dynamics of a canonical cognitive circuit across waking and sleep. Nat. Neurosci. 22, 1512–1520 (2019).

    CAS  PubMed  Google Scholar 

  172. Gardner, R. J. et al. Toroidal topology of population activity in grid cells. Nature 602, 123–128 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  173. Hatsopoulos, N. G. Encoding in the motor cortex: was evarts right after all? Focus on “motor cortex neural correlates of output kinematics and kinetics during isometric-force and arm-reaching tasks”. J. Neurophysiol. 94, 2261–2262 (2005).

    PubMed  Google Scholar 

  174. Scott, S. H. Population vectors and motor cortex: neural coding or epiphenomenon? Nat. Neurosci. 3, 307–308 (2000).

    CAS  PubMed  Google Scholar 

  175. Herter, T. M., Korbel, T. & Scott, S. H. Comparison of neural responses in primary motor cortex to transient and continuous loads during posture. J. Neurophysiol. 101, 150–163 (2009).

    PubMed  Google Scholar 

  176. Griffin, D. M., Hoffman, D. S. & Strick, P. L. Corticomotoneuronal cells are “functionally tuned”. Science 350, 667–670 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  177. Griffin, D. M. & Strick, P. L. The motor cortex uses active suppression to sculpt movement. Sci. Adv. 6, eabb8395 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Shalit, U., Zinger, N., Joshua, M. & Prut, Y. Descending systems translate transient cortical commands into a sustained muscle activation signal. Cereb. Cortex 22, 1904–1914 (2012).

    PubMed  Google Scholar 

  179. Albert, S. T. et al. Postural control of arm and fingers through integration of movement commands. eLife 9, e52507 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Fink, A. J. et al. Presynaptic inhibition of spinal sensory feedback ensures smooth movement. Nature 509, 43–48 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  181. Soechting, J. F. & Flanders, M. Moving in three-dimensional space: frames of reference, vectors, and coordinate systems. Annu. Rev. Neurosci. 15, 167–191 (1992).

    CAS  PubMed  Google Scholar 

  182. Burnod, Y. et al. Visuomotor transformations underlying arm movements toward visual targets: a neural network model of cerebral cortical operations. J. Neurosci. 12, 1435–1453 (1992).

    CAS  PubMed  PubMed Central  Google Scholar 

  183. Wang, T., Chen, Y. & Cui, H. From parametric representation to dynamical system: shifting views of the motor cortex in motor control. Neurosci. Bull. 38, 796–808 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  184. Fu, Q. G., Flament, D., Coltz, J. D. & Ebner, T. J. Temporal encoding of movement kinematics in the discharge of primate primary motor and premotor neurons. J. Neurophysiol. 73, 836–854 (1995).

    CAS  PubMed  Google Scholar 

  185. Schwartz, A. B. Direct cortical representation of drawing. Science 265, 540–542 (1994).

    CAS  PubMed  Google Scholar 

  186. Moran, D. W. & Schwartz, A. B. Motor cortical representation of speed and direction during reaching. J. Neurophysiol. 82, 2676–2692 (1999).

    CAS  PubMed  Google Scholar 

  187. Schroeder, K. E., Perkins, S. M., Wang, Q. & Churchland, M. M. Cortical control of virtual self-motion using task-specific subspaces. J. Neurosci. 42, 220–239 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  188. Yang, L., Michaels, J. A., Pruszynski, J. A. & Scott, S. H. Rapid motor responses quickly integrate visuospatial task constraints. Exp. Brain Res. 211, 231–242 (2011).

    PubMed  Google Scholar 

  189. Russo, A. A. et al. Neural trajectories in the supplementary motor area and motor cortex exhibit distinct geometries, compatible with different classes of computation. Neuron 107, 745–758.e6 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  190. Hennig, J. A. et al. Learning is shaped by abrupt changes in neural engagement. Nat. Neurosci. 24, 727–736 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  191. Smoulder, A. L. et al. A neural basis of choking under pressure. Preprint at bioRxiv https://doi.org/10.1101/2023.04.16.537007 (2023).

  192. Perich, M. G., Gallego, J. A. & Miller, L. E. A neural population mechanism for rapid learning. Neuron 100, 964–976.e7 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  193. Vyas, S. et al. Neural population dynamics underlying motor learning transfer. Neuron 97, 1177–1186.e3 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  194. Sun, X. et al. Cortical preparatory activity indexes learned motor memories. Nature 602, 274–279 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  195. Vyas, S., O’Shea, D. J., Ryu, S. I. & Shenoy, K. V. Causal role of motor preparation during error-driven learning. Neuron 106, 329–339.e4 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  196. Sheahan, H. R., Franklin, D. W. & Wolpert, D. M. Motor planning, not execution, separates motor memories. Neuron 92, 773–779 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  197. Versteeg, C. & Miller, L. E. Dynamical feedback control: motor cortex as an optimal feedback controller based on neural dynamics. Preprint at https://doi.org/10.20944/preprints202201.0428.v1 (2022).

  198. Serruya, M. D., Hatsopoulos, N. G., Paninski, L., Fellows, M. R. & Donoghue, J. P. Instant neural control of a movement signal. Nature 416, 141–142 (2002).

    CAS  PubMed  Google Scholar 

  199. Taylor, D. M., Tillery, S. I. & Schwartz, A. B. Direct cortical control of 3D neuroprosthetic devices. Science 296, 1829–1832 (2002).

    CAS  PubMed  Google Scholar 

  200. Carmena, J. M. et al. Learning to control a brain–machine interface for reaching and grasping by primates. PLoS Biol. 1, E42 (2003).

    PubMed  PubMed Central  Google Scholar 

  201. Georgopoulos, A. P. & Carpenter, A. F. Coding of movements in the motor cortex. Curr. Opin. Neurobiol. 33C, 34–39 (2015).

    Google Scholar 

  202. Musallam, S., Corneil, B. D., Greger, B., Scherberger, H. & Andersen, R. A. Cognitive control signals for neural prosthetics. Science 305, 258–262 (2004).

    CAS  PubMed  Google Scholar 

  203. Santhanam, G., Ryu, S. I., Yu, B. M., Afshar, A. & Shenoy, K. V. A high-performance brain–computer interface. Nature 442, 195–198 (2006).

    CAS  PubMed  Google Scholar 

  204. Yu, B., Ryu, S., Santhanam, G., Churchland, M. & Shenoy, K. Improving neural prosthetic system performance by combining plan and peri-movement activity. In IEEE EMBS 26th Annual Meeting 4516–4519 (2004).

  205. Kao, J. C. et al. Single-trial dynamics of motor cortex and their applications to brain–machine interfaces. Nat. Commun. 6, 7759 (2015).

    CAS  PubMed  Google Scholar 

  206. Jarosiewicz, B. et al. Functional network reorganization during learning in a brain–computer interface paradigm. Proc. Natl Acad. Sci. USA 105, 19486–19491 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  207. Chase, S. M., Kass, R. E. & Schwartz, A. B. Behavioral and neural correlates of visuomotor adaptation observed through a brain–computer interface in primary motor cortex. J. Neurophysiol. 108, 624–644 (2012).

    PubMed  PubMed Central  Google Scholar 

  208. Golub, M. D. et al. Learning by neural reassociation. Nat. Neurosci. 21, 607–616 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  209. Oby, E. R. et al. New neural activity patterns emerge with long-term learning. Proc. Natl Acad. Sci. USA 116, 15210–15215 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  210. Peters, A. J., Chen, S. X. & Komiyama, T. Emergence of reproducible spatiotemporal activity during motor learning. Nature 510, 263–267 (2014).

    CAS  PubMed  Google Scholar 

  211. Losey, D. M. et al. Learning alters neural activity to simultaneously support memory and action. Preprint at bioRxiv https://doi.org/10.1101/2022.07.05.498856 (2022).

  212. Ethier, C., Oby, E. R., Bauman, M. J. & Miller, L. E. Restoration of grasp following paralysis through brain-controlled stimulation of muscles. Nature 485, 368–371 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  213. Willett, F. R., Avansino, D. T., Hochberg, L. R., Henderson, J. M. & Shenoy, K. V. High-performance brain-to-text communication via handwriting. Nature 593, 249–254 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  214. Wong, A. L., Goldsmith, J., Forrence, A. D., Haith, A. M. & Krakauer, J. W. Reaction times can reflect habits rather than computations. eLife 6, e28075 (2017).

    PubMed  PubMed Central  Google Scholar 

  215. Wong, A. L., Haith, A. M. & Krakauer, J. W. Motor planning. Neuroscientist. https://doi.org/10.1177/1073858414541484 (2014).

  216. Cisek, P. & Kalaska, J. F. Neural mechanisms for interacting with a world full of action choices. Annu. Rev. Neurosci. 33, 269–298 (2010).

    CAS  PubMed  Google Scholar 

  217. Stringer, C. et al. Spontaneous behaviors drive multidimensional, brainwide activity. Science 364, 255 (2019).

    PubMed  PubMed Central  Google Scholar 

  218. Yoo, S. B. M. & Hayden, B. Y. The transition from evaluation to selection involves neural subspace reorganization in core reward regions. Neuron 105, 712–724.e4 (2020).

    CAS  PubMed  Google Scholar 

  219. Herzfeld, D. J., Kojima, Y., Soetedjo, R. & Shadmehr, R. Encoding of action by the Purkinje cells of the cerebellum. Nature 526, 439–442 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank members of the Churchland laboratory for reviewing and editing this manuscript. M.M.C. thanks Y. Pavlova for laboratory management. K.V.S. thanks B. Davis for administrative support and S. Kosasih for laboratory management. M.M.C is supported by the Grossman Center for the Statistics of Mind, the Simons Foundation Collaboration on the Global Brain, the Kavli Institute for Brain Science and the Zuckerman Mind Brain Behaviour Institute. K.V.S. was supported by National Institute of Neurological Disorders and Stroke (NINDS) U01NS123101, U19NS112954 and R01NS116623; National Institute on Deafness and Other Communication Disorders (NIDCD) R01NS121097, R01DC014034, U01DC019430 and U01DC017844; National Institutes of Health (NIH) National Institute of Mental Health (NIMH) R01MH086373; the Simons Foundation Collaboration on the Global Brain; L. and P. Garlick; S. and B. Reeves; the Wu Tsai Neurosciences Institute; the Bio-X Institute at Stanford University; The Hong Seh and Vivian W. M. Lim Professorship at Stanford University; and the Howard Hughes Medical Institute (HHMI) at Stanford University.

Author information

Authors and Affiliations

Authors

Contributions

The authors contributed equally to all aspects of the article.

Corresponding author

Correspondence to Mark M. Churchland.

Ethics declarations

Competing interests

M.M.C is an inventor of the MINT decoding approach, which has been licensed to Blackrock Neurotech. K.V.S. served on the Scientific Advisory Boards (SABs) of MIND-X Inc. (acquired by Blackrock Neurotech), Inscopix Inc. (merged with Brucker Nano) and Heal Inc.; served as a consultant/adviser for CTRL-Labs (on the founding SAB; acquired by Facebook Reality Labs in Fall 2019, which is now Meta Platform Reality Labs); and served as a consultant/adviser for and was a co-founder (2016–2023) of Neuralink.

Peer review

Peer review information

Nature Reviews Neuroscience thanks Rune Berg, Matthew Perich and the other, anonymous, reviewer for their contribution to the peer review of this work.

Additional information

Dedication

I dedicate this review to my friend and colleague Krishna Shenoy, who died as we were completing the writing. That act of writing brought our long intellectual relationship full circle. Our final days together were spent much like our initial days: sitting in a room talking about preparatory activity. Pondering the nature of preparatory activity had sent us down the long path of trying to understand the chain of neural events that produces voluntary movement. We saw preparation as the ‘first cog’ in that causal chain. Preparatory activity must somehow have, wrapped up in it, the information necessary to generate time-varying patterns of muscle activity that will soon occur. Trying to figure out how that could work was the first cog in our personal journey — one whose various branches trace back to the initial question that consumed us: what does preparatory activity do? Our adoption and promotion of other ideas — dynamics, null spaces, factors — were a straightforward consequence of our desire to answer that question. We hope that this review brings readers on a similar journey.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Glossary

Single-neuron response

The response of a real (or simulated) neuron is notated as \({r}_{n}(t)\), where t indicates time and n indicates we are considering the nth neuron.

Population response

The responses of \(N\) neurons, constituting an \(N\)-dimensional vector \({\boldsymbol{r}}(t)\). Each vector element, \({r}_{n}(t)\), describes the activity of one neuron.

Weight

A number indicating the degree to which one thing (for example, a neuron) influences something else (for example, another neuron). Weights come in many flavours and can summarize many things, including synaptic strengths and the impact of neural activity on muscle activity.

Neural dimension

An N-dimensional vector containing one weight per neuron. Just like the weights themselves, neural dimensions come in multiple flavours.

Weighted sum

Weighted sums of neural activity are central to many analyses that seek to understand network function (real or simulated). If the vector \({\boldsymbol{v}}\) is a neural dimension, one computes \(\mathop{\sum }\limits_{n=1}^{N}{{\boldsymbol{v}}}_{n}{{\boldsymbol{r}}}_{n}(t)\). Using vector notation, the weighted sum \(a(t)={{\boldsymbol{v}}}^{{\rm{T}}}{\boldsymbol{r}}(t)\) is referred to as ‘activity in dimension \({\boldsymbol{v}}\)’.

Subspace

A set of one or more orthogonal neural dimensions. Dimensions are collected into columns of a matrix \({\boldsymbol{V}}\). The vector of weighted sums \({\boldsymbol{a}}(t)={{\boldsymbol{V}}}^{{\rm{T}}}{\boldsymbol{r}}(t)\) is referred to as ‘activity in the subspace of \({\boldsymbol{V}}\) ’.

Activity subspace

A subspace that fully captures neural responses. Suppose we computed a(t) as above (see subspaces). If we can successfully invert this relationship using r(t) ≈ Va(t), then the dimensions in V span the activity subspace.

Dimensionality reduction

A method, such as principal component analysis (PCA), for estimating dimensions that constitute the activity subspace.

Readout

A signal that exits the neural population, such as a descending command for muscle activity (see Fig. 1c). Assuming linearity, the readout is a weighted sum of population responses: \(z(t)={{\boldsymbol{b}}}^{{\rm{T}}}{\boldsymbol{r}}(t)\) (see readout dimensions for how \({\boldsymbol{b}}\) is defined).

Readout dimensions

Each readout has an associated neural dimension, \({\boldsymbol{b}}\), containing weights (one per neuron) used for the readout. Analysis and visualization often employ a lower-dimensional \({{\boldsymbol{b}}}^{{\rm{F}}{\rm{a}}{\rm{c}}}\), with one weight per factor.

Output-potent space

The subspace, spanned by the readout dimensions, where neural activity impacts those readouts. Dimension \({\boldsymbol{v}}\) is ‘output-potent’ if it overlaps with a readout dimension: \({{\boldsymbol{b}}}^{{\rm{T}}}{\boldsymbol{v}}\ne 0\).

Null-space

The subspace, orthogonal to the readout dimensions, where neural activity has no impact on those readouts. Dimension \({\boldsymbol{v}}\) is ‘output-null’ if it is orthogonal to each readout dimension: \({{\boldsymbol{b}}}^{{\rm{T}}}{\boldsymbol{v}}=0\).

Factors

For many recurrent networks, computation can be summarized by considering K factors rather than N neurons (see Fig. 1c), with K \(\ll \) N. Each factor is a weighted sum of activity: xi(t) = \({{\boldsymbol{w}}}_{i}^{{\rm{T}}}\)r(t), where \({{\boldsymbol{w}}}_{i}\) is a neural dimension. Considering all factors: x(t) = WTr(t). \({\boldsymbol{W}}\) reflects network connectivity (see recurrent connectivity).

Neural state

The value of the factors, \({\boldsymbol{x}}(t)\), at a particular moment t. ‘Neural state’ may also refer to the population response, \({\boldsymbol{r}}(t)\). Assuming factors are a valid summary of the population response, these definitions are equivalent.

State space

A method for visualizing the neural state as a point in a space where each axis corresponds to a factor (or some other weighted sum of neural activity).

Neural trajectory

The value of the factors, \({\boldsymbol{x}}(t)\), across a span of time. When plotted in state space, trajectories form traces that describe how activity changes with time (see Fig. 1c, inset).

Manifold

The shape formed by the collection of neural trajectories, defining the possible locations of the neural state within the activity subspace. For example, when cycling at different speeds, the manifold has a tube-like shape (see Fig. 2d).

Tuning

A description of how a neuron’s response reflects a specific quantity, such as a stimulus. In recurrent networks, factors can be the relevant quantities.

Factor tuning

The weight un,i quantifies how the ith factor impacts the nth neuron. The neural dimension \({{\boldsymbol{u}}}_{i}\) quantifies the impact on all neurons. Given a matrix \({\boldsymbol{U}}\) of factor-tuning dimensions, population activity is \({\boldsymbol{r}}(t)\approx {\boldsymbol{U}}{\boldsymbol{x}}(t)\). The approximation reflects non-linearity and spiking variability. \({\boldsymbol{U}}\) reflects network connectivity (see recurrent connectivity), and defines the activity subspace.

Output-potent factors

Factor \({x}_{i}(t)\) is output-potent (that is, impacts the readout) if its factor-tuning dimension, \({{\boldsymbol{u}}}_{i}\), is output-potent (overlaps with one or more readout dimensions).

Output-null factors

Factor \({x}_{i}(t)\) is output-null (that is, does not impact the readout) if its factor-tuning dimension, \({{\boldsymbol{u}}}_{i}\), is output-null (lies within the null-space).

Recurrent connectivity

Synaptic connections that cause activity to flow in ‘loops’. The causal flow is simplified by considering factors (Fig. 1c) and matrices \({\boldsymbol{W}}\) and \({\boldsymbol{U}}\) (see factors and factor tuning). Factors are weighted sums of neural responses: \({\boldsymbol{x}}(t)={{\boldsymbol{W}}}^{{\rm{T}}}{\boldsymbol{r}}(t)\). Responses reflect the factors, \({\boldsymbol{r}}(t)\approx {\boldsymbol{U}}{\boldsymbol{x}}(t)\), completing the loop. In some networks, \({\boldsymbol{W}}\) and \({\boldsymbol{U}}\) can be computed directly from synaptic connectivity.

Factor-level dynamics

In recurrent networks, activity at the present moment drives activity at the next moment. Such dynamics are often best described at the factor level: \(\dot{{\boldsymbol{x}}}\) = f(x) + y. The function f defines a flow-field in state space (see Fig. 1c, inset). y captures inputs from outside the observed system.

Rotational dynamics

Dynamics linking two factors, \({x}_{1}\) and \({x}_{2}\), such that neural states flow from one to the other while preserving their order (see Fig. 2a). Suppose that, for three conditions during preparation, factor \({x}_{{\rm{prep}}}\) takes values \(\{-1,1,2\}\) while factor \({x}_{{\rm{exec}}}\) takes values \(\{0,0,0\}\). After a 90° rotation, activity has left the preparatory subspace, \({x}_{{\rm{prep}}}=\{0,0,0\}\), and entered the execution subspace, \({x}_{{\rm{exec}}}=\{-1,1,2\}\). Depending on the network, rotations might continue or might end after ~90°.

Factor estimation

To estimate factors from measurements of neural activity alone, without knowledge of connectivity, experimenters use dimensionality reduction to estimate the activity subspace and invert the relationship r(t) ≈ Ux(t). For example, one may use x(t) ≈ UTr(t), assuming the estimated U is orthonormal.

Motor neurons

Neurons in the spinal cord that connect directly to a muscle. Each motor neuron spike produces one spike in the muscle fibres it innervates. Motor neurons receive direct (monosynaptic) and indirect (polysynaptic) inputs from motor-cortex neurons.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Churchland, M.M., Shenoy, K.V. Preparatory activity and the expansive null-space. Nat. Rev. Neurosci. 25, 213–236 (2024). https://doi.org/10.1038/s41583-024-00796-z

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41583-024-00796-z

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing