Comprehensive Physiology Wiley Online Library

Sensory Systems in the Control of Movement

Full Article on Wiley Online Library



Abstract

Animal movement is immensely varied, from the simplest reflexive responses to the most complex, dexterous voluntary tasks. Here, we focus on the control of movement in mammals, including humans. First, the sensory inputs most closely implicated in controlling movement are reviewed, with a focus on somatosensory receptors. The response properties of the large muscle receptors are examined in detail. The role of sensory input in the control of movement is then discussed, with an emphasis on the control of locomotion. The interaction between central pattern generators and sensory input, in particular in relation to stretch reflexes, timing, and pattern forming neuronal networks is examined. It is proposed that neural signals related to bodily velocity form the basic descending command that controls locomotion through specific and well‐characterized relationships between muscle activation, step cycle phase durations, and biomechanical outcomes. Sensory input is crucial in modulating both the timing and pattern forming parts of this mechanism. © 2012 American Physiological Society. Compr Physiol 2:2615‐2627, 2012.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1.

Ensemble cycle averages of the firing of γs and γd motoneurons (A and B), recorded in the common peroneal nerve innervating the ankle flexor tibialis anterior (TA) during spontaneous locomotion in the high decerebrate cat. (A) Three simultaneously recorded γs motoneurons in two cats (panels a and b), in each case an average of 20 step cycles aligned to TA length minima (thick vertical dashed line) and normalized in time. (i) TA electromyogram (EMG: continuous line), medial gastrocnemius (MG) EMG (dotted line), (ii) ankle angle corresponding to TA shortening upward, and (iii) mean firing rate of the γs motoneurons. Mean cycle times in (a) 640 ms and in (b) 800 ms. The three thin vertical dashed lines in A(a) indicate the three phases of TA muscle shortening. B(a) discharge of a γd motoneuron, average of 9 step cycles aligned to TA length minima in each cycle and normalized in time, mean cycle duration 740 ms, B(b) similar data from a γd motoneuron in another cat, average of 12 step cycles with mean duration 735 ms. Note the sudden onset of γd firing at the onset of TA shortening, and the cessation of firing shortly after the start of lengthening. Adapted, with permission, from Figures 3 and 7 in Taylor et al. 171.

Figure 2. Figure 2.

Ensemble averages of firing rates of group Ia, II, and Ib afferents in ankle extensors (left) and knee flexors (right), recorded during overground locomotion in normal cats. Traces from top to bottom: electromyogram (EMG) and length of receptor‐bearing muscles (lengthening upwards), firing rates of group Ia, II, and Ib afferents. The number of afferents contributing to each average is shown on the right of each firing rate plot. Step cycles were aligned to peaks in either the ankle extensor (triceps surae) or knee flexor (posterior biceps) length signals. The length signals were also used to estimate stance‐swing and swing‐stance transitions in the step cycle (vertical dashed lines). Note the high mean firing rates of Ia and II afferents, indicating high levels of γs drive and the increase in the ankle extensor Ia firing rate prior to the onset of lengthening at the stance to swing transition, compatible with increased γs drive. Derived, with permission, from Figure 6 137.

Figure 3. Figure 3.

Estimated time course of joint angle variations computed from the firing rates of 47 sensory afferents recorded simultaneously during treadmill locomotion with a microelectrode array implanted in the L7 dorsal root of a cat. This group included five spindle 10 and five spindle 20 endings, one Golgi tendon organ, four glabrous cutaneous receptors and two hair follicle receptors. Step‐cycle averages of the actual and estimated (A) position, (B) velocity, and (C) acceleration in joint‐angle coordinates. Each plot shows the mean of 162 steps (toe‐off to toe‐off). The thin lines represent ±1 s.d. from the mean of the actual trajectories. The up and down arrows indicate onset of the swing and stance phases, respectively. Reproduced, with permission, from reference 182.

Figure 4. Figure 4.

Descending control of the locomotor step cycle. (A) Increments in the intensity of stimulation in the midbrain locomotor region (MLR) in the high decerebrate cat (lower trace) increases the cadence of locomotion (upper traces) (adapted, with permission, from reference 158). (B) Schematic summarizing the velocity command hypothesis: a command signal specifying desired body velocity descends from brainstem and drives the timing element of the locomotor central pattern generator (CPG) to generate cadences with flexor and extensor phase durations that depend in a specific way on cycle duration. The velocity signal also drives the pattern formation network (PFN) to modulate the amplitudes of activation of the flexor and extensor muscles according to a square law relationship. Muscle displacement automatically modulates muscle force through the intrinsic length‐tension properties. Muscle force and displacement sensed by spindle and tendon organ afferents elicit continuous stretch reflexes as well as modulating or overriding phase transitions via the CPG timer. Presented at the Society of Experimental Biology Annual General Meeting in 2009 139.



Figure 1.

Ensemble cycle averages of the firing of γs and γd motoneurons (A and B), recorded in the common peroneal nerve innervating the ankle flexor tibialis anterior (TA) during spontaneous locomotion in the high decerebrate cat. (A) Three simultaneously recorded γs motoneurons in two cats (panels a and b), in each case an average of 20 step cycles aligned to TA length minima (thick vertical dashed line) and normalized in time. (i) TA electromyogram (EMG: continuous line), medial gastrocnemius (MG) EMG (dotted line), (ii) ankle angle corresponding to TA shortening upward, and (iii) mean firing rate of the γs motoneurons. Mean cycle times in (a) 640 ms and in (b) 800 ms. The three thin vertical dashed lines in A(a) indicate the three phases of TA muscle shortening. B(a) discharge of a γd motoneuron, average of 9 step cycles aligned to TA length minima in each cycle and normalized in time, mean cycle duration 740 ms, B(b) similar data from a γd motoneuron in another cat, average of 12 step cycles with mean duration 735 ms. Note the sudden onset of γd firing at the onset of TA shortening, and the cessation of firing shortly after the start of lengthening. Adapted, with permission, from Figures 3 and 7 in Taylor et al. 171.



Figure 2.

Ensemble averages of firing rates of group Ia, II, and Ib afferents in ankle extensors (left) and knee flexors (right), recorded during overground locomotion in normal cats. Traces from top to bottom: electromyogram (EMG) and length of receptor‐bearing muscles (lengthening upwards), firing rates of group Ia, II, and Ib afferents. The number of afferents contributing to each average is shown on the right of each firing rate plot. Step cycles were aligned to peaks in either the ankle extensor (triceps surae) or knee flexor (posterior biceps) length signals. The length signals were also used to estimate stance‐swing and swing‐stance transitions in the step cycle (vertical dashed lines). Note the high mean firing rates of Ia and II afferents, indicating high levels of γs drive and the increase in the ankle extensor Ia firing rate prior to the onset of lengthening at the stance to swing transition, compatible with increased γs drive. Derived, with permission, from Figure 6 137.



Figure 3.

Estimated time course of joint angle variations computed from the firing rates of 47 sensory afferents recorded simultaneously during treadmill locomotion with a microelectrode array implanted in the L7 dorsal root of a cat. This group included five spindle 10 and five spindle 20 endings, one Golgi tendon organ, four glabrous cutaneous receptors and two hair follicle receptors. Step‐cycle averages of the actual and estimated (A) position, (B) velocity, and (C) acceleration in joint‐angle coordinates. Each plot shows the mean of 162 steps (toe‐off to toe‐off). The thin lines represent ±1 s.d. from the mean of the actual trajectories. The up and down arrows indicate onset of the swing and stance phases, respectively. Reproduced, with permission, from reference 182.



Figure 4.

Descending control of the locomotor step cycle. (A) Increments in the intensity of stimulation in the midbrain locomotor region (MLR) in the high decerebrate cat (lower trace) increases the cadence of locomotion (upper traces) (adapted, with permission, from reference 158). (B) Schematic summarizing the velocity command hypothesis: a command signal specifying desired body velocity descends from brainstem and drives the timing element of the locomotor central pattern generator (CPG) to generate cadences with flexor and extensor phase durations that depend in a specific way on cycle duration. The velocity signal also drives the pattern formation network (PFN) to modulate the amplitudes of activation of the flexor and extensor muscles according to a square law relationship. Muscle displacement automatically modulates muscle force through the intrinsic length‐tension properties. Muscle force and displacement sensed by spindle and tendon organ afferents elicit continuous stretch reflexes as well as modulating or overriding phase transitions via the CPG timer. Presented at the Society of Experimental Biology Annual General Meeting in 2009 139.

References
 1. Ach N. Über die Willenstätigkeit und das Denken. Göttingen, 1905.
 2. af Klint R, Mazzaro N, Nielsen JB, Sinkjaer T, Grey MJ. Load rather than length sensitive feedback contributes to soleus muscle activity during human treadmill walking. J Neurophysiol 103: 2747‐2756, 2010.
 3. al‐Falahe NA, Nagaoka M, Vallbo AB. Response profiles of human muscle afferents during active finger movements. Brain 113: 325‐346, 1990.
 4. al‐Falahe NA, Nagaoka M, Vallbo AB. Dual response from human muscle spindles in fast voluntary movements. Acta Physiol Scand 141: 363‐371, 1991.
 5. Allum JH, Oude Nijhuis LB, Carpenter MG. Differences in coding provided by proprioceptive and vestibular sensory signals may contribute to lateral instability in vestibular loss subjects. Exp Brain Res 184: 391‐410, 2008.
 6. Amis A, Prochazka A, Short D, Trend PS, Ward A. Relative displacements in muscle and tendon during human arm movements. J Physiol 389: 37‐44, 1987.
 7. Aniss AM, Diener HC, Hore J, Gandevia SC, Burke D. Behavior of human muscle receptors when reliant on proprioceptive feedback during standing. J Neurophysiol 64: 661‐670, 1990.
 8. Appenteng K, Prochazka A. Tendon organ firing during active muscle lengthening in awake, normally behaiving cats. J Physiol 353: 81‐92, 1984.
 9. Arshavsky YI, Deliagina TG, Orlovsky GN. Pattern generation. Curr Opin Neurobiol 7: 781‐789, 1997.
 10. Arshavsky YI, Gelfand IM, Orlovsky GN. Cerebellum and Rhythmical Movements. Berlin: Springer, 1986.
 11. Barker D, Ip MC, Adal MN. A correlation between the receptor population of the cat's soleus muscle and the afferent fibre diameter spectrum of the nerve supplying it. In: Barker D, editor. Symposium on Muscle Receptors. Hong Kong: Hong Kong University Press, 1962, pp. 257‐261.
 12. Bastian HC. The “muscular sense”: Its nature and localisation. Brain 10: 1‐136, 1888.
 13. Bell C. The Hand. Its Mechanism and Vital Endowments as Evincing Design. London: William Pickering, 1834.
 14. Beloozerova IN, Sirota MG. The role of the motor cortex in the control of accuracy of locomotor movements in the cat. J Physiol 461: 1‐25, 1993.
 15. Bennett DJ. Torques generated at the human elbow joint in response to constant position errors imposed during voluntary movements. Exp Brain Res 95: 488‐498, 1993.
 16. Bennett DJ. Stretch reflex responses in the human elbow joint during a voluntary movement. J Physiol (Lond) 474: 339‐351, 1994.
 17. Bennett DJ, Gorassini M, Prochazka A. Catching a ball: Contributions of intrinsic muscle stiffness, reflexes, and higher order responses. Can J Physiol Pharmacol 72: 525‐534, 1994.
 18. Bergenheim M, Johansson H, Pedersen J, Ohberg F, Sjolander P. Ensemble coding of muscle stretches in afferent populations containing different types of muscle afferents. Brain Res 734: 157‐166, 1996.
 19. Bernstein N. The Coordination and Regulation of Movements. Oxford: Pergamon, 1967.
 20. Bloedel JR. Task‐dependent role of the cerebellum in motor learning. Prog Brain Res 143: 319‐329, 2004.
 21. Bosco G, Eian J, Poppele RE. Kinematic and non‐kinematic signals transmitted to the cat cerebellum during passive treadmill stepping. Exp Brain Res 167: 394‐403, 2005.
 22. Bosco G, Poppele RE. Proprioception from a spinocerebellar perspective. Physiol Rev 81: 539‐568, 2001.
 23. Boyd IA, Roberts TDM. Proprioceptive discharges from stretch receptors in the knee joint of the cat. J Physiol 122: 38‐58, 1953.
 24. Brown TG. The intrinsic factors in the act of progression in the mammal. Proc R Soc Lond, Series B 84: 308‐319, 1911.
 25. Burgess PR, Clark FJ. Characteristics of knee joint receptors in the cat. J Physiol 203: 317‐335, 1969.
 26. Burke D, Hagbarth KE, Lofstedt L. Muscle spindle activity in man during shortening and lengthening contractions. J Physiol 277: 131‐142, 1978.
 27. Cabelguen JM. Static and dynamic fusimotor action on the response of spindle primary endings to sinusoidal stretches in the cat. Brain Res 169: 45‐54, 1979.
 28. Capaday C, Stein RB. Amplitude modulation of the soleus H‐reflex in the human during walking and standing. J Neurosci 6: 1308‐1313, 1986.
 29. Carli G, Farabollini F, Fontani G, Meucci M. Slowly adapting receptors in cat hip joint. J Neurophysiol 42: 767‐778, 1979.
 30. Clark FJ, Burgess RC, Chapin JW, Lipscomb WT. Role of intramuscular receptors in the awareness of limb position. J Neurophysiol 54: 1529‐1540, 1985.
 31. Cleland CL, Hayward L, Rymer WZ. Neural mechanisms underlying the clasp‐knife reflex in the cat. II. Stretch‐sensitive muscular‐free nerve endings. J Neurophysiol 64: 1319‐1330, 1990.
 32. Cleland CL, Rymer WZ. Neural mechanisms underlying the clasp‐knife reflex in the cat. I. Characteristics of the reflex. J Neurophysiol 64: 1303‐1318, 1990.
 33. Collins DF, Prochazka A. Movement illusions evoked by ensemble cutaneous input from the dorsum of the human hand. J Physiol 496(Pt 3): 857‐871, 1996.
 34. Conway BA, Hultborn H, Kiehn O. Proprioceptive input resets central locomotor rhythm in the spinal cat. Brain Res 68: 643‐656, 1987.
 35. Cordo PJ, Flores‐Vieira C, Verschueren SM, Inglis JT, Gurfinke lV. Position sensitivity of human muscle spindles: Single afferent and population representations. J Neurophysiol 87: 1186‐1195, 2002a.
 36. Cordo PJ, Flores‐Vieira C, Verschueren SM, Inglis JT, Gurfinkel V. Position sensitivity of human muscle spindles: Single afferent and population representations. J Neurophysiol 87: 1186‐1195, 2002b.
 37. Cruse H. What mechanisms coordinate leg movement in walking arthropods? Trends Neurosci 13: 15‐21, 1990.
 38. Day BL, Fitzpatrick RC. The vestibular system. Curr Biol 15: R583‐R586, 2005.
 39. Dimitriou M, Edin BB. Discharges in human muscle receptor afferents during block grasping. J Neurosci 28: 12632‐12642, 2008a.
 40. Dimitriou M, Edin BB. Discharges in human muscle spindle afferents during a key‐pressing task. J Physiol 586: 5455‐5470, 2008b.
 41. Dimitriou M, Edin BB. Human muscle spindles act as forward sensory models. Curr Biol 20: 1763‐1767, 2010.
 42. Donelan JM, McVea DA, Pearson KG. Force regulation of ankle extensor muscle activity in freely walking cats. J Neurophysiol 101: 360‐371, 2009.
 43. Donelan JM, Pearson KG. Contribution of force feedback to ankle extensor activity in decerebrate walking cats. J Neurophysiol 92: 2093‐2104, 2004.
 44. Drew T. Motor cortical activity during voluntary gait modifications in the cat. I. Cells related to the forelimbs. J Neurophysiol 70: 179‐199, 1993.
 45. Drew T, Andujar JE, Lajoie K, Yakovenko S. Cortical mechanisms involved in visuomotor coordination during precision walking. Brain Res Rev 57: 199‐211, 2008.
 46. Edin BB. Cutaneous afferents provide information about knee joint movements in humans. J Physiol 531: 289‐297, 2001.
 47. Edin BB. Quantitative analyses of dynamic strain sensitivity in human skin mechanoreceptors. J Neurophysiol 92: 3233‐3243, 2004.
 48. Ekeberg O, Pearson K. Computer simulation of stepping in the hind legs of the cat: An examination of mechanisms regulating the stance‐to‐swing transition. J Neurophysiol 94: 4256‐4268, 2005.
 49. Elek J, Prochazka A, Hulliger M, Vincent S. In‐series compliance of gastrocnemius muscle in cat step cycle: Do spindles signal origin‐to‐insertion length? J Physiol 429: 237‐258, 1990.
 50. Ellaway PH, Prochazka A, Chan M, Gauthier MJ. The sense of movement elicited by transcranial magnetic stimulation in humans is due to sensory feedback. J Physiol 556: 651‐660, 2004.
 51. Emonet‐Denand F, Jami L, Laporte Y. Skeleto‐fusimotor axons in the hind‐limb muscles of the cat. J Physiol 249: 153‐166, 1975.
 52. Feldman AG. Functional tuning of the nervous system with control of movement and maintenance of steady posture. II. Controllable parameters of the muscle. Biophysics 11: 565‐578, 1966.
 53. Ferrell WR. The adequacy of stretch receptors in the cat knee joint for signalling joint angle throughout a full range of movement. J Physiol 299: 85‐100, 1980.
 54. Ferrell WR, Gandevia SC, McCloskey DI. The role of joint receptors in human kinaesthesia when intramuscular receptors cannot contribute. J Physiol 386: 63‐71, 1987.
 55. Forssberg H. Stumbling corrective reaction: A phase‐dependent compensatory reaction during locomotion. J Neurophysiol 42: 936‐953, 1979.
 56. Freusberg A. Reflexbewegungen beim Hunde. Pflug Archiv Physiol 9: 358‐391, 1874
 57. Fukunaga T, Kubo K, Kawakami Y, Fukashiro S, Kanehisa H, Maganaris CN. In vivo behaviour of human muscle tendon during walking. Proc R Soc Lond B Biol Sci 268: 229‐233, 2001.
 58. Gandevia S. Kinesthesia: Roles for afferent signals and motor commands. In: Rowell L, Sheperd JT, editors. Handbook of Physiology. Section 12. Exercise: Regulation and Integration of Multiple Systems. New York: American Physiological Society, 1996, pp. 128‐172.
 59. Geyer H, Seyfarth A, Blickhan R. Positive force feedback in bouncing gaits? Proc Biol Sci 270: 2173‐2183, 2003.
 60. Gibson JJ. A critical review of the concept of set in contemporary experimental psychology. Psychol Bull 38: 781‐817, 1941.
 61. Godwin‐Austen RB. The mechanoreceptors of the costo‐vertebral joints. J Physiol 202: 737‐753, 1969.
 62. Goodwin GM, Hulliger M, Matthews PB. Studies on muscle spindle primary endings with sinusoidal stretching. Prog Brain Res 44: 89‐98, 1976.
 63. Gorassini M, Prochazka A, Taylor JL. Cerebellar ataxia and muscle spindle sensitivity. J Neurophysiol 70: 1853‐1862, 1993.
 64. Gorassini MA, Prochazka A, Hiebert GW, Gauthier MJ. Corrective responses to loss of ground support during walking. I. Intact cats. J Neurophysiol 71: 603‐610, 1994.
 65. Goslow GE, Jr., Reinking RM, Stuart DG. The cat step cycle: Hind limb joint angles and muscle lengths during unrestrained locomotion. J Morphol 141: 1‐41, 1973.
 66. Goslow GE, Stauffer EK, Nemeth WC, Stuart DG. The cat step cycle: responses of muscle spindles and tendon organs to passive stretch within the locomotor range. Brain Res 60: 35‐54, 1973b.
 67. Gregory JE, McIntyre AK, Proske U. Tendon organ afferents in the knee joint nerve of the cat. Neurosci Lett 103: 287‐292, 1989.
 68. Grey M, Ladouceur M, Andersen JB, Nielsen JB, Sinkjaer T. Contribution of group II muscle afferents to the medium latency soleus stretch reflex during walking in man. J Physiol 534: 925‐933, 2001.
 69. Grey MJ, Nielsen JB, Mazzaro N, Sinkjaer T. Positive force feedback in human walking. J Physiol 581: 99‐105, 2007.
 70. Griffiths RI. Shortening of muscle fibres during stretch of the active cat medial gastrocnemius muscle: The role of tendon compliance. J Physiol (Lond) 436: 219‐236, 1991.
 71. Grigg P, Greenspan BJ. Response of primate joint afferent neurons to mechanical stimulation of knee joint. J Neurophysiol 40: 1‐8, 1977.
 72. Grillner S. Control of locomotion in bipeds, tetrapods, and fish. In: Handbook of physiology.The nervous system. Bethesda: American Physiological Society, 1981, sect. 2, pp. 1179‐1236.
 73. Grillner S, Cangiano L, Hu G, Thompson R, Hill R, Wallen P. The intrinsic function of a motor system–from ion channels to networks and behavior. Brain Res 886: 224‐236, 2000.
 74. Grillner S, Zangger P. How detailed is the central pattern generation for locomotion? Brain Res 88: 367‐371, 1975.
 75. Gritsenko V, Kalaska JF. Rapid online correction is selectively suppressed during movement with a visuomotor transformation. J Neurophysiol 2010.
 76. Gritsenko V, Mushahwar V, Prochazka A. Adaptive changes in locomotor control after partial denervation of triceps surae muscles in the cat. J Physiol 533: 299‐311, 2001.
 77. Halbertsma JM. The stride cycle of the cat: The modelling of locomotion by computerized analysis of automatic recordings. Acta Physiol Scand Suppl 521: 1‐75, 1983.
 78. Haridas C, Zehr EP, Misiaszek JE. Adaptation of cutaneous stumble correction when tripping is part of the locomotor environment. J Neurophysiol 99: 2789‐2797, 2008.
 79. Herbert RD, Moseley AM, Butler JE, Gandevia SC. Change in length of relaxed muscle fascicles and tendons with knee and ankle movement in humans. J Physiol 539: 637‐645, 2002.
 80. Hermer‐Vazquez L, Hermer‐Vazquez R, Chapin JK. The reach‐to‐grasp‐food task for rats: A rare case of modularity in animal behavior? Behav Brain Res 177: 322‐328, 2007.
 81. Hiebert GW, Whelan PJ, Prochazka A, Pearson KG. Contribution of hind limb flexor muscle afferents to the timing of phase transitions in the cat step cycle. J Neurophysiol 75: 1126‐1137, 1996.
 82. Hoang PD, Herbert RD, Todd G, Gorman RB, Gandevia SC. Passive mechanical properties of human gastrocnemius muscle tendon units, muscle fascicles and tendons in vivo. J Exp Biol 210: 4159‐4168, 2007.
 83. Hoffer JA, Andreassen S. Regulation of soleus muscle stiffness in premammillary cats: Intrinsic and reflex components. J Neurophysiol 45: 267‐285, 1981.
 84. Hoffer JA, Caputi AA, Pose IE, Griffiths RI. Roles of muscle activity and load on the relationship between muscle spindle length and whole muscle length in the freely walking cat. Prog Brain Res 80: 75‐85, 1989.
 85. Horch KW, Tuckett RP, Burgess PR. A key to the classification of cutaneous mechanoreceptors. J Invest Dermatol 69: 75‐82, 1977.
 86. Horn KM, Pong M, Gibson AR. Discharge of inferior olive cells during reaching errors and perturbations. Brain Res 996: 148‐158, 2004.
 87. Hospod V, Aimonetti JM, Roll JP, Ribot‐Ciscar E. Changes in human muscle spindle sensitivity during a proprioceptive attention task. J Neurosci 27: 5172‐5178, 2007.
 88. Houk J, Henneman E. Responses of Golgi tendon organs to active contractions of the soleus muscle of the cat. J Neurophysiol 30: 466‐481, 1967.
 89. Houk JC, Singer JJ, Goldman MR. An evaluation of length and force feedback to soleus muscles of decerebrate cats. J Neurophysiol 33: 784‐811, 1970.
 90. Hulliger M. The mammalian muscle spindle and its central control. Rev Physiol Biochem Pharmacol 101: 1‐110, 1984a.
 91. Hulliger M. The mammalian muscle spindle and its central control. [Review]. Rev Physiol Biochem Pharmacol 101: 1‐110, 1984b.
 92. Iles JF, Stokes M, Young A. Reflex actions of knee joint afferents during contraction of the human quadriceps. Clin Physiol 10: 489‐500, 1990.
 93. James W. The Principles of Psychology. New York: Henry Holt, 1890.
 94. Jami L. Golgi tendon organs in mammalian skeletal muscle: Functional properties and central actions. Physiol Rev 72: 623‐666, 1992.
 95. Johansson H, Sjolander P, Sojka P. Receptors in the knee joint ligaments and their role in the biomechanics of the joint. CRC Crit Rev Biomed Eng 18: 341‐368, 1991.
 96. Johansson RS, Vallbo AB. Tactile sensibility in the human hand: Relative and absolute densities of four types of mechanoreceptive units in glabrous skin. J Physiol (Lond) 286: 283‐300, 1979.
 97. Jones KE, Wessberg J, Vallbo AB. Directional tuning of human forearm muscle afferents during voluntary wrist movements. J Physiol 536: 635‐647, 2001.
 98. Kakuda N, Vallbo AB, Wessberg J. Fusimotor and skeletomotor activities are increased with precision finger movement in man. J Physiol 492(Pt 3): 921‐929, 1996.
 99. Kennedy WR. Innervation of normal human muscle spindles. Neurology 20: 463‐475, 1970.
 100. Kiehn O. Locomotor circuits in the mammalian spinal cord. Annu Rev Neurosci 29: 279‐306, 2006.
 101. Kniffki KD, Schomburg ED, Steffens H. Effects from fine muscle and cutaneous afferents on spinal locomotion in cats. J Physiol 319: 543‐554, 1981.
 102. Kowalczewski J, Prochazka A. Interactive Receptor Model. University of Alberta Libraries, Free Internet Access, www.library.ualberta.ca. 2006.
 103. Lafreniere‐Roula M, McCrea DA. Deletions of rhythmic motoneuron activity during fictive locomotion and scratch provide clues to the organization of the mammalian central pattern generator. J Neurophysiol 94: 1120‐1132, 2005.
 104. Loeb GE. Somatosensory unit input to the spinal cord during normal walking. Can J Physiol Pharmacol 59: 627‐635, 1981.
 105. Loeb GE, Levine WS, He J. Understanding sensorimotor feedback through optimal control. Cold Spring Harb Symp Quant Biol 55: 791‐803, 1990.
 106. Loram ID, Lakie M, Di Giulio I, Maganaris CN. The consequences of short‐range stiffness and fluctuating muscle activity for proprioception of postural joint rotations: The relevance to human standing. J Neurophysiol 102: 460‐474, 2009.
 107. Lund JP, Matthews B. Responses of temporomandibular joint afferents recorded in the Gasserian ganglion of the rabbit to passive movements of the mandible. In: Kawamura Y, editor. Oral‐facial Sensory and Motor Functions. Tokyo: Quintessence, 1981, pp. 153‐160.
 108. Maganaris CN, Paul JP. In vivo human tendinous tissue stretch upon maximum muscle force generation. J Biomech 33: 1453‐1459, 2000.
 109. Maganaris CN, Paul JP. Tensile properties of the in vivo human gastrocnemius tendon. J Biomech 35: 1639‐1646, 2002.
 110. Matthews PBC. Mammalian Muscle Receptors and Their Central Actions. London: Arnold, 1972.
 111. McCrimmon DR, Ramirez JM, Alford S, Zuperku EJ. Unraveling the mechanism for respiratory rhythm generation. Bioessays 22: 6‐9, 2000.
 112. McIntyre AK, Proske U, Tracey DJ. Afferent fibres from muscle receptors in the posterior nerve of the cat's knee joint. Exp Brain Res 33: 415‐424, 1978.
 113. Merton PA. How we control the contraction of our muscles. Sci Am 226: 30‐37, 1972.
 114. Miall C. Motor control: Correcting errors and learning from mistakes. Curr Biol 20: R596‐598, 2010.
 115. Miall RC. The cerebellum, predictive control and motor coordination. Novartis Found Symp 218: 272‐284; discussion 284‐290, 1998.
 116. Miall RC, Weir DJ, Wolpert DM, Stein JF. Is the cerebellum a Smith predictor? J Mot Behav 25: 203‐216, 1993.
 117. Murphy PR, Martin HA. Fusimotor discharge patterns during rhythmic movements. Trends Neurosci 16: 273‐278, 1993.
 118. Murphy PR, Stein RB, Taylor J. Phasic and tonic modulation of impulse rates in gamma‐motoneurons during locomotion in premammillary cats. J Neurophysiol 52: 228‐243, 1984.
 119. Mushahwar VK, Gillard DM, Gauthier MJ, Prochazka A. Intraspinal micro stimulation generates locomotor‐like and feedback‐controlled movements. IEEE Trans Neural Syst Rehabil Eng 10: 68‐81, 2002.
 120. Newsom Davis J. The response to stretch of human intercostal muscle spindles stduied in vitro. J Physiol 249: 561‐579, 1975.
 121. Nichols R, Ross KT. The implications of force feedback for the lambda model. Adv Exp Med Biol 629: 663‐679, 2009.
 122. Nichols TR, Houk JC. Improvement in linearity and regulation of stiffness that results from actions of stretch reflex. J Neurophysiol 39: 119‐142, 1976.
 123. Patla AE, Niechwiej E, Racco V, Goodale MA. Understanding the contribution of binocular vision to the control of adaptive locomotion. Exp Brain Res 142: 551‐561, 2002.
 124. Patla AE, Prentice SD, Rietdyk S, Allard F, Martin C. What guides the selection of alternate foot placement during locomotion in humans. Exp Brain Res 128: 441‐450, 1999.
 125. Patla AE, Vickers JN. Where and when do we look as we approach and step over an obstacle in the travel path? Neuroreport 8: 3661‐3665, 1997.
 126. Pearson K, Ekeberg O, Buschges A. Assessing sensory function in locomotor systems using neuro‐mechanical simulations. Trends Neurosci 29: 625‐631, 2006.
 127. Pearson KG. Role of sensory feedback in the control of stance duration in walking cats. Brain Res Rev 57: 222‐227, 2008.
 128. Pearson KG, Collins DF. Reversal of the influence of group Ib afferents from plantaris on activity in medial gastrocnemius muscle during locomotor activity. J Neurophysiol 70: 1009‐1017, 1993.
 129. Perret C. Centrally generated pattern of motoneuron activity during locomotion in the cat. Symp Soc Exp Biol 37: 405‐422, 1983.
 130. Perret C, Buser P. Static and dynamic fusimotor activity during locomotor movements in the cat. Brain Res 40: 165‐169, 1972.
 131. Perret C, Cabelguen JM. Main characteristics of the hindlimb locomotor cycle in the decorticate cat with special reference to bifunctional muscles. Brain Res 187: 333‐352, 1980.
 132. Poppele RE, Kennedy WR. Comparison between behavior of human and cat muscle spindles recorded in vitro. Brain Res 75: 316‐319, 1974.
 133. Poppele RE, Terzuolo CA. Myotatic reflex: Its input‐output relation. Science 159: 743‐745, 1968.
 134. Prochazka A. Comparison of natural and artificial control of movement. IEEE Trans Rehab Eng 1: 7‐17, 1993.
 135. Prochazka A. Proprioceptive feedback and movement regulation. In: Rowell L, Sheperd JT, editors. Handbook of Physiology. Exercise: Regulation and Integration of Multiple Systems. New York: American Physiological Society, 1996, sect. 12, pp. 89‐127.
 136. Prochazka A. Quantifying proprioception. Prog Brain Res 123: 133‐142, 1999.
 137. Prochazka A, Gillard D, Bennett DJ. Implications of positive feedback in the control of movement. J Neurophysiol 77: 3237‐3251, 1997.
 138. Prochazka A, Gorassini M. Ensemble firing of muscle afferents recorded during normal locomotion in cats. J Physiol 507(Pt 1): 293‐304, 1998a.
 139. Prochazka A, Gorassini M. Ensemble firing of muscle afferents recorded during normal locomotion in cats. J Physiol 507: 293‐304, 1998b.
 140. Prochazka A, Hulliger M, Zangger P, Appenteng K. ‘Fusimotor set’: New evidence for alpha‐independent control of gamma‐motoneurones during movement in the awake cat. Brain Res 339: 136‐140, 1985.
 141. Prochazka A, Sorensen C. Biomechanical imperatives in the neural control of locomotion. Comp Biochem Physiol 153: S135‐S136, 2009.
 142. Prochazka A, Yakovenko S. The neuromechanical tuning hypothesis. In: Cisek P, Drew, T, Kalaska, J, editors. Progress in Brain Research Computational Neuroscience: Theoretical Insights into Brain Function. NY: Elsevier, 2007, pp. 255‐265.
 143. Proske U. The Golgi tendon organ. Properties of the receptor and reflex action of impulses arising from tendon organs. In: Porter R, editor. MTP International Review of Physiology, Neurophysiology IV. Baltimore: MTP University Park Press, 1981, pp. 127‐171.
 144. Proske U, Gregory JE. Signalling properties of muscle spindles and tendon organs. Adv Exp Med Biol 508: 5‐12, 2002.
 145. Proske U, Morgan DL. Stiffness of cat soleus muscle and tendon during activation of part of muscle. J Neurophysiol 52: 459‐468, 1984.
 146. Rack PM, Ross HF, Thilmann AF, Walters DK. Reflex responses at the human ankle: The importance of tendon compliance. J Physiol 344: 503‐524, 1983.
 147. Ribot‐Ciscar E, Hospod V, Roll JP, Aimonetti JM. Fusimotor drive may adjust muscle spindle feedback to task requirements in humans. J Neurophysiol 101: 633‐640, 2009.
 148. Ribot‐Ciscar E, Rossi‐Durand C, Roll JP. Increased muscle spindle sensitivity to movement during reinforcement manoeuvres in relaxed human subjects. J Physiol 523(Pt 1): 271‐282, 2000.
 149. Rigosa J, Weber D, Prochazka A, Stein R, Micera S. Neuro‐fuzzy decoding of sensory information from ensembles of simultaneously recorded dorsal root ganglion neurons for FES applications. J Neural Eng 8(4): 046019, 2011.
 150. Rossignol S, Dubuc R, Gossard JP. Dynamic sensorimotor interactions in locomotion. Physiol Rev 86: 89‐154, 2006.
 151. Rybak IA, Shevtsova NA, Lafreniere‐Roula M, McCrea DA. Modelling spinal circuitry involved in locomotor pattern generation: Insights from deletions during fictive locomotion. J Physiol 577: 617‐639, 2006.
 152. Rybak IA, Stecina K, Shevtsova NA, McCrea DA. Modelling spinal circuitry involved in locomotor pattern generation: Insights from the effects of afferent stimulation. J Physiol 577: 641‐658, 2006.
 153. Sechenov IM. Reflexes of the Brain, (Refleksy Golovnogo Mozga). In: Subkov AA, editor. I.M. Sechenov, Selected Works. Moscow: State Publishing House, 1863, pp. 264‐322.
 154. Selverston AI. Modeling of neural circuits: What have we learned? Annu Rev Neurosci 16: 531‐546, 1993.
 155. Shadmehr R, Krakauer JW. A computational neuroanatomy for motor control. Exp Brain Res 185: 359‐381, 2008.
 156. Sherrington CS. On the proprio‐ceptive system, especially in its reflex aspects. Brain 29: 467‐482, 1906.
 157. Shemmell J, Krutky MA, Perreault EJ. Stretch sensitive reflexes as an adaptive mechanism for maintaining limb stability. Clin Neurophysiol 121: 1680‐1689, 2010.
 158. Sherrington CS. Flexion‐reflex of the limb, crossed extension‐reflex, and reflex stepping and standing. J Physiol (London) 40: 28‐121, 1910.
 159. Sherrington CS. Further observations on the production of reflex stepping by combination of reflex excitation with reflex inhibition. J Physiol 47: 196‐214, 1914.
 160. Shik ML, Severin FV, Orlovsky GN. Control of walking and running by means of electrical stimulation of the mid‐brain. Biophysics 11: 756‐765, 1966.
 161. Shimansky Y, Wang JJ, Bauer RA, Bracha V, Bloedel JR. On‐line compensation for perturbations of a reaching movement is cerebellar dependent: Support for the task dependency hypothesis. Exp Brain Res 155: 156‐172, 2004.
 162. Sinkjaer T, Andersen JB, Ladouceur M, Christensen LO, Nielsen JB. Major role for sensory feedback in soleus EMG activity in the stance phase of walking in man. J Physiol 523: 817‐827, 2000.
 163. Sinkjaer T, Toft E, Andreassen S, Hornemann BC. Muscle stiffness in human ankle dorsiflexors: Intrinsic and reflex components. J Neurophysiol 60: 1110‐1121, 1988.
 164. Smith OJM. A controller to overcome dead time. Instrum Soc Am J 6: 28‐33, 1959.
 165. Spencer H. The principles of psychology. London: Longman, Brown, Green, and Longmans, 1855.
 166. St George RJ, Fitzpatrick RC. The sense of self‐motion, orientation and balance explored by vestibular stimulation. J Physiol 589: 807‐813, 2011.
 167. Stein RB, Misiaszek JE, Pearson KG. Functional role of muscle reflexes for force generation in the decerebrate walking cat. J Physiol 525: 781‐791, 2000.
 168. Stephens JA, Reinking RM, Stuart DG. Tendon organs of cat medial gastrocnemius: Responses to active and passive forces as a function of muscle length. J Neurophysiol 38: 1217‐1231, 1975.
 169. Taga G. A model of the neuro‐musculo‐skeletal system for human locomotion. II Real‐time adaptability under various constraints. Biol Cybern 73: 113‐121, 1995.
 170. Taga G, Yamaguchi Y, Shimizu H. Self‐organized control of bipedal locomotion by neural oscillators in unpredictable environment. Biol Cybern 65: 147‐159, 1991.
 171. Takakusaki K, Oohinata‐Sugimoto J, Saitoh K, Habaguchi T. Role of basal ganglia‐brainstem systems in the control of postural muscle tone and locomotion. Prog Brain Res 143: 231‐237, 2004.
 172. Taylor A, Cody FW. Jaw muscle spindle activity in the cat during normal movements of eating and drinking. Brain Res 71: 523‐530, 1974.
 173. Taylor A, Durbaba R, Ellaway PH, Rawlinson S. Patterns of fusimotor activity during locomotion in the decerebrate cat deduced from recordings from hindlimb muscle spindles. J Physiol 522: 515‐532, 2000.
 174. Taylor A, Durbaba R, Ellaway PH, Rawlinson S. Static and dynamic gamma‐motor output to ankle flexor muscles during locomotion in the decerebrate cat. J Physiol 571: 711‐723, 2006.
 175. Tomovic R, Anastasijevic R, Vuco J, Tepavac D. The study of locomotion by finite state models. Biol Cybern 63: 271‐276, 1990.
 176. Tomovic R, McGhee R. A finite state approach to the synthesis of control systems. IEEE Trans Hum Fac Electron 7: 122‐128, 1966.
 177. Tracey DJ. Characteristics of wrist joint receptors in the cat. Brain Res 34: 165‐176, 1979.
 178. van Beers RJ, Wolpert DM, Haggard P. When feeling is more important than seeing in sensorimotor adaptation. Curr Biol 12: 834‐837, 2002.
 179. Vaziri S, Diedrichsen J, Shadmehr R. Why does the brain predict sensory consequences of oculomotor commands? Optimal integration of the predicted and the actual sensory feedback. J Neurosci 26: 4188‐4197, 2006.
 180. von Holst E. Relations between the central nervous system and the peripheral organs. Br J Anim Behav 2: 89‐94, 1954.
 181. von Holst E, Mittelstaedt H. Das Reafferenzprincip. Naturwissenschaften 37: 464‐476, 1950.
 182. Voss H. Tabelle der absoluten und relativen Muskelspindelzahlen der menschlichen Skelettmuskulatur. Anatomische Anzeiger 129: 5562‐5572, 1971.
 183. Watt HJ. Experimentelle Beiträge zu einer Theorie des Denkens. Archiv gesamter Psychologie 4: 289‐436, 1905.
 184. Weber DJ, Stein RB, Everaert DG, Prochazka A. Decoding sensory feedback from firing rates of afferent ensembles recorded in cat dorsal root ganglia in normal locomotion. IEEE Trans Neural Syst Rehabil Eng 14: 240‐243, 2006.
 185. Weber DJ, Stein RB, Everaert DG, Prochazka A. Limb‐state feedback from ensembles of simultaneously recorded dorsal root ganglion neurons. J Neural Eng 4: S168‐S180, 2007.
 186. Widajewicz W, Kably B, Drew T. Motor cortical activity during voluntary gait modifications in the cat. II. Cells related to the hindlimbs. J Neurophysiol 72: 2070‐2089, 1994.
 187. Willis WD, Coggeshalll RE. Sensory Mechanisms of the Spinal Cord. N.Y.: Plenum, 1991.
 188. Wilson LR, Gandevia SC, Burke D. Discharge of human muscle spindle afferents innervating ankle dorsiflexors during target isometric contractions. J Physiol 504: 221‐232, 1997.
 189. Witney AG, Goodbody SJ, Wolpert DM. Predictive motor learning of temporal delays. J Neurophysiol 82: 2039‐2048, 1999.
 190. Wolpert DM, Ghahramani Z. Computational principles of movement neuroscience. Nat Neurosci 3: 1212‐1217, 2000.
 191. Wolpert DM, Miall RC. Forward Models for Physiological Motor Control. Neural Netw 9: 1265‐1279, 1996.
 192. Yakovenko S, Drew T. A motor cortical contribution to the anticipatory postural adjustments that precede reaching in the cat. J Neurophysiol 102: 853‐874, 2009.
 193. Yakovenko S, Gritsenko V, Prochazka A. Contribution of stretch reflexes to locomotor control: A modeling study. Biol Cybern 90: 146‐155, 2004.
 194. Yakovenko S, McCrea DA, Stecina K, Prochazka A. Control of locomotor cycle durations. J Neurophysiol 94: 1057‐1065, 2005.
 195. Zalkind VI. Method for an adequate stimulation of receptors of the cat carpo‐radialis joint. Sechenov Physiol J USSR 57: 1123‐1127, 1971.
 196. Zehr EP, Stein RB. What functions do reflexes serve during human locomotion? Progress in Neurobiology 58: 185‐205, 1999.
 197. Zelenin PV, Deliagina TG, Grillner S, Orlovsky GN. Postural control in the lamprey: A study with a neuro‐mechanical model. J Neurophysiol 84: 2880‐2887, 2000.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Arthur Prochazka, Peter Ellaway. Sensory Systems in the Control of Movement. Compr Physiol 2012, 2: 2615-2627. doi: 10.1002/cphy.c100086