Skip to main content
Log in

Hypothetical neural control of human bipedal walking with voluntary modulation

  • Original Article
  • Published:
Medical & Biological Engineering & Computing Aims and scope Submit manuscript

Abstract

A hypothetical neuromusculoskeletal model is developed to simulate human normal walking and its modulated behaviors. A small set of neural periodic patterns drive spinal muscle synergies which in turn lead to specific pattern of muscle activation and supraspinal feedback systems maintain postural balance during walking. Then, the model demonstrates modulated behaviors by superimposing voluntary perturbations on the underlying walking pattern. Motions of kicking a ball and obstacle avoidance during walking are simulated as examples. The superposition of the new pulse command to a set of invariant pulses representing spino-locomotor is sufficient to achieve the coordinated behaviors. Also, forward bent walking motion is demonstrated by applying similar superposition. The composition of activations avoids a complicated computation of motor program for a specific task and presents a simple control scheme for different walking patterns.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

References

  1. Amstrong DM, Edgley SA (1988) Discharges of interpositus and Purkinje cells of the cat cerebellum during locomotion under different conditions. J Physiol 400:425–445

    Google Scholar 

  2. Brand RA, Pedersen DR, Friederich JA et al (1986) The sensitivity of muscle force predictions to changes in physiological cross-sectional area. J Biomech 8:589–596

    Article  Google Scholar 

  3. Brooke JD, Cheng J, Collins DF, et al (1997) Sensori-sensory afferent conditioning with leg movement: gain control in spinal reflex and ascending paths. Prog Neurobiol 51:393–421

    Article  Google Scholar 

  4. Cajigas González I (2003) Linear control model of the spinal processing of descending neural signals, Master’s thesis, Massachusetts Institute of Technology, Cambridge, MA

  5. Calancie B, Needham-Shropshire B, Jacobs P et al (1994) Involuntary stepping after chronic spinal cord injury: evidence for a central rhythm generator for locomotion in man. Brain 117:1143–1159

    Article  Google Scholar 

  6. Collins JJ, Stewart I (1994) A group-theoretic approach to rings of coupled biological oscillators. Biol Cybern 71:95–103

    Article  MATH  Google Scholar 

  7. Crespi A, Badertscher A, Guignard A, Ijspeert AJ (2005) AmphiBot I: an amphibious snake-like robot. Robot Auto Sys 50:163–175

    Article  Google Scholar 

  8. d’Avella A, Satiel P, Bizzi E (2003) Combinations of muscle synergies in the construction of a natural motor behavior. Nat Neurosci 6(3):300–308

    Article  Google Scholar 

  9. Dimitrijevic MR, Gerasimenko Y, Pinter MM (1998) Evidence for a spinal central pattern generator in humans. Ann NY Acad Sci 860:360–376

    Article  Google Scholar 

  10. Dietz V (1992) Human neuronal control of automatic functional movements: interaction between central programs and afferent input. Physiol Rev 72(1):33–69

    MathSciNet  Google Scholar 

  11. Dietz V, Harkema J (2004) Locomotor activity in spinal cord-injured persons. J Appl Physiol 96:1954–1960

    Article  Google Scholar 

  12. Flash T (1987) The control of hand equilibrium trajectories in multi-joint arm movements. Biol Cybern 57:257–274

    Article  MATH  Google Scholar 

  13. Fuglevand AJ, Winter DA (1993) Models of recruitment and rate coding organization in motor-unit pools. J Neurophysiol 70(6):2470–2488

    Google Scholar 

  14. Fujita K, Sato H (1998) Intrinsic viscoelasticity of ankle joint during standing. In: Proceedings of the 20th annual international conference of the IEEE engineering in medicine and biology society 20(5):2343–2345

  15. Fukuoka Y, Kimura H, Cohen AH (2003) Adaptive dynamic walking of a quadruped robot on irregular terrain based on biological concepts. Int J Robot Res 22(3–4):187–202

    Article  Google Scholar 

  16. Georgopoulos AP (1988) Neural integration of movement: role of motor cortex in rearching. FASEB J 2:2849–2857

    Google Scholar 

  17. Grasso R, Zago M, Lacquaniti F (2000) Interactions between posture and locomotion: motor patterns in humans walking with bent posture versus erect posture. J Neurophysiol 83:288–300

    Google Scholar 

  18. Huffman KJ, Krubitzer L (2001) Thalamo-cortical connections of area 3a and M1 in marmoset monkeys. J Comp Neurol 435(3):291–310

    Article  Google Scholar 

  19. Inman VT, Ralston HJ, Todd F (1981) Human walking. In: Lieberman JC (ed) Williams & Wilkins, Baltimore, pp 41–55

  20. Ito M (1997) Cerebellar microcomplexes. In: Schmahmann JD (ed) The cerebellum and cognition, vol 41. Academic, New York, pp 475–487

  21. Ivanenko YP, Cappellini G, Dominici N et al (2005) Coordination of locomotion with voluntary movements in humans. J Neurosci 25(31):7238–7253

    Article  Google Scholar 

  22. Ivanenko YP, Poppele RE, Lacquaniti F (2006) Spinal cord maps of spatiotemporal alpha-motoneuron activation in humans walking at different speeds. J Neurophysiol 95:602–618

    Article  Google Scholar 

  23. Iwasaki T, Zhen M (2006) Sensory feedback mechanism underlying entrainment of central pattern generator to mechanical resonance. Biol Cybern 94(4):245–261

    Article  MATH  Google Scholar 

  24. Jo S, Massaquoi SG (2004) A model of cerebellum stabilized and scheduled hybrid long-loop control of upright balance. Biol Cybern 91:188–202

    Article  MATH  Google Scholar 

  25. Jo S, Massaquoi SG (2007) A model of cerebrocerebello-spinomuscular interaction in the sagittal control of human walking. Biol Cybern 96

  26. Karameh FN, Massaquoi SG (2005) A model of nonlinear motor cortical integration and its relation to movement speed profile control. In: Proceedings 2005 IEEE Engineering in Medicine and Biology, 27th Annual conference, Shanghai, China, Sept 1–4

  27. Katayama M, Kawato M (1993) Virtual trajectory and stiffness ellipse during multijoint arm movement predicted by neural inverse models. Biol Cybern 69:353–362

    MATH  Google Scholar 

  28. Lacquaniti F, Soechting JF (1986) Simulation studies on the control of posture and movement in a multi-jointed limb. Biol Cybern 54:367–378

    Article  Google Scholar 

  29. Massaquoi SG (1999) Modelling the function of the cerebellum in scheduled linear servo control of simple horizontal planar arm movements. Electrical Engineering & Computer Science, Massachusetts Institute of Technology, Cambridge, MA

  30. Massaquoi SG, Topka H (2002) Models of cerebellar function. In: Pandolfo M, Manto M (eds) The cerebellum and its disorders. Cambridge University Press, Cambridge, pp 69–94

    Google Scholar 

  31. Matarić MJ (2002) Sensory-motor primitives as a basis for imitation: linking perception to action and biology to robotics. In: Dautenhahn K, Nehaniv C (eds) Imitation in animals and artifacts. MIT, Cambridge, pp 392–422

  32. Mori S, Nakajima K, Mori F, Matsuyama K (2004) Integration of multiple motor segments for the elaboration of locomotion: role of the fastigial nucleus of the cerebellum. Prog Brain Res 143:341–351

    Article  Google Scholar 

  33. Nathan PW (1994) Effects on movement of surgical incisions into the human spinal cord. Brain 117(Pt2):337–346

    Article  Google Scholar 

  34. Ogihara N, Yamazaki N (2001) Generation of human bipedal locomotion by a bio-mimetic neuro-musculo-skeletal model. Biol Cybern 84:1–11

    Article  Google Scholar 

  35. Rossignol S, Dubbue R, Gossard J-P (2006) Dynamic sensorimotor interaction in locomotion. Physiol Rev 86:89–154

    Article  Google Scholar 

  36. Rudomin P, Schmidt RF (1999) Presynaptic inhibition in the vertebrate spinal cord revisited. Exp Brain Res 129(1):1–37

    Article  Google Scholar 

  37. Saltiel P, Wyler-Duda K, d’Avella A, et al (2001) Muscle synergies encoded within the spinal cord: evidence from focal intraspinal NMDA iontophoresis in the frog. J Neurophysiol 85:605–619

    Google Scholar 

  38. Shik ML, Orlovsky GN (1976) Neurophysiology of locomotor automatism. Phsyiol Rev 56(3):465–501

    Google Scholar 

  39. Taga G (1995) A model of the neuro-musculo-skeletal system for human locomotion I. Emergence of basic gait. Biol Cybern 73:97–111

    Article  MATH  Google Scholar 

  40. Taga G (1998) A model of the neuro-musculo-skeletal system for anticipatory adjustment of human locomotion during obstacle avoidance. Biol Cybern 78:9–17

    Article  MATH  Google Scholar 

  41. Thach WT, Goodlin HP, Keating JG (1992) The cerebellum and the adaptive coordination of movement. Annu Rev Neurosci 15:403–442

    Article  Google Scholar 

  42. Tresch MC, Salteil P, Bizzi E (1999) The construction of movement by the spinal cord. Nat Neurosci 2:162–167

    Article  Google Scholar 

  43. Winter DA (1990) Biomechanics and motor control of human movement, 2nd edn. Wiley, New York

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sungho Jo.

Appendices

Appendix A: Anthropometry used to model a human body

The length of each segment is represented with respect to the total body’s height h t (=180 cm) and mass m t (=80 kg) based on [41]. trunk, thigh, shank, and foot masses are respectively 0.678 m t, 0.1 m t, 0.047 m t, and 0.015 m t; their moments of inertia are respectively 0.031 m t h 2t , 6.262 × 10−4 m t h 2t , 2.566 × 10−4 m t h 2t , and 4.976 × 10−6 m t h 2t ; their lengths are respectively 0.47 h t, 0.245 h t, and 0.246 h t; their COM distances from lower end are respectively 0.235h t, 0.1389 h t, and 0.1395 h t; foot is modeled as a triangle with height 0.039 h t and length 0.152 h t; foot COM is located at 0.0195 h t high from bottom and 0.0304 h t ahead from heel.

Appendix B: Foot interaction with the ground

The vertical ground reaction force is modeled by:

$$ F_{gy}^{i}=(K_{gy}(f_{gy}(x^{i})-y^{i})- B_{gy}\dot{y}^{i})\cdot1\left[f_{gy}(x^{i})-y^{i},0\right]_{+} $$
(29)

where (x i, y i) indicates the positions of either heel or toe with i = heel, toe. f gy (x i) represents the ground profile as a function of x i.

If the toe or heel reaches zero horizontal velocity, the horizontal reaction force is modeled by a spring and damper system as long as the horizontal reaction force is smaller than the maximal friction force.

$$ F_{gx}^{i}=(K_{gx}(x_{o}^{i}-x^{i})- B_{gx}\dot{x}^{i})\cdot1\left[f_{gy}(x^{i})-y^{i},0\right]_{+}\,\, \hbox{if}\left|F_{gx}^{i}\right|\leq\left|\mu_{s}F_{gy}^{i}\right|,\quad i=\hbox{heel, toe} $$
(30)

where x i o is a location where either heel or toe touches the ground initially and μ s is the static frictional coefficient.

Otherwise, the horizontal reaction force is modeled as a dynamic friction force.

$$ F_{gx}^{i}=-\mu_{k}F_{gy}^{i}\hbox{sgn}(\dot{x}^{i}) $$
(31)

where μ k is the dynamic frictional coefficient.

Appendix C: Simulation parameter values

1.1 C.1. Normal walking

  • Transmission neural delays

Closed-loop transmission delays are conservatively taken to be 60, 70, and 80 ms for long-loop response to and from the hip, knee, ankle, respectively based on 50 m/s neural conduction velocity, and five synaptic delays of less than 1 ms. Therefore, \({\mathbf{T}}_{\rm spr} + {\mathbf{T}}_{\rm sp} + {\mathbf{T}}_{\rm pr} = {\left[{\begin{array}{*{20}c} {80}& {70}& {60}\\ \end{array}} \right]}^{\rm T}.\) For simulation, T spr = T sp + T pr, T sp = T pr are assumed. EC(s) also lags neural signals.

  • Foot interaction to the ground

$$K_{gy} = 30,000, \enspace B_{{gy}} = 500, \enspace K_{{gx}} = 10,000, \enspace B_{{gx}} = 1,000; \enspace \mu_{k} = 0.6, \enspace \mu_{s} = 1.2.$$
  • Spinal pattern generator

$$ f_{{\rm PG}} = 1.3\quad \hbox{and} \quad m_{{\rm PG}} = 1.2. $$
$$ {\mathbf{W}}_{\rm PG} = {\left[{\begin{array}{*{20}c} {0.3}& {0}& {0}& {0}& {0.8}& {0}& {0}& {0}& {0}\\ {0}& {0.29}& {0}& {0}& {0}& {0}& {0}& {0}& {0}\\ {0.4}& {0}& {0.64}& {0}& {0.9}& {0}& {0.35}& {0}& {0}\\ {0}& {0}& {0}& {0}& {0}& {0.4}& {0}& {0}& {0.3}\\ {0}& {0.1}& {0}& {0.8}& {0}& {0}& {0.4}& {0.1}& {0}\\ \end{array}} \right]}^{\rm T} $$

(Table 2).

Table 2 Parameters for periodic pattern generation
  • Spinal segmental inhibition

$$ \theta_{{\rm th,a}} = 0.35; \quad \theta_{{\rm th,k}} = - 0.35; \quad \theta_{{\rm th,h}} = 0.55. $$
$$ {\mathbf{W}}_{{\rm reflex}} = \rho {\left[{\begin{array}{*{20}c} {0}& {0}& {0}& {0}& {1}& {0}& {0}& {0}& {0}\\ {0}& {0}& {1}& {0}& {0}& {0}& {0}& {0}& {0}\\ {1}& {0}& {0}& {0}& {0}& {0}& {0}& {0}& {0} \\\end{array}} \right]}^{\rm T} $$

where ρ is a sufficient large number (ρ > m PG).

  • Supraspinal system

Cerebro-cerebellar feedback control i a  = 0.2, i r  = 100,f = 0.6, g b  = 0, g k  = 3,

$$ {\mathbf{W}}_{\rm C} = {\left[{\begin{array}{*{20}c} {0}& {0}& {2}& {- 5}& {6}& {- 1}& {3}& {- 1}& {- 3}\\ \end{array}} \right]}^{\rm T} $$

Estimate of COM \((\hat{x}_{\rm com}): p_{1} = 0.97,\; p_{2} = 0.53,\; p_{3} = {0.14}.\)

  • Vestibulospinal reflex feedback control

$$ k_{\rm p} = 3, \quad b_{\rm p} = 0.9. $$
$$ {\mathbf{W}}_{\rm ves} = {\left[{\begin{array}{*{20}c} {0.132}& {- 0.092}& {0}& {0}& {0}& {0}& {0.049}& {- 0.054}& {0}\\ \end{array}} \right]}^{\rm T} $$
  • Initial positions

$$ \theta_{\rm a} = 0.2, \; \theta_{\rm k} = 0, \; \theta_{\rm h} = - 0.2\quad \hbox{for\; right\; leg};\, \theta_{\rm a} = 0, \; \theta_{\rm k} = - 0.1, \; \theta_{\rm h} = 0.4 \quad \hbox{for\; left\; leg}. $$
  • Initial velocities

$$ \dot{\theta}_{\rm a} = (f_{{\rm PG}} + 1)/2, \dot{\theta}_{\rm k} = - (f_{{\rm PG}} + 1)/2, \dot{\theta}_{\rm h} = (f_{{\rm PG}} + 1)/2 \quad \hbox{for \; right \; leg}; $$
$$ \dot{\theta}_{\rm a} = - (f_{{\rm PG}} + 1)/2, \dot{\theta}_{\rm k} = (f_{{\rm PG}} + 1)/2, \dot{\theta}_{\rm h} = - (f_{{\rm PG}} + 1)/2 \quad \hbox{for\; left \; leg}. $$
  • Reference signals

$$ u_{\rm ref} = 0.25; \quad \theta_{\rm tr,ref} = 0. $$

1.2 C.2. Forward bent walking

Other parameters are the same as in two except the followings:

  • Initial positions

$$ \begin{aligned} \theta_{\rm a} &= 0.2, \; \theta_{\rm k} = 0, \; \theta_{\rm h} = 0.7 \quad\hbox{for\; right\; leg};\\ \theta_{\rm a} &= 0, \; \theta_{\rm k} = - 0.1, \; \theta_{\rm h} = 1.3 \quad \hbox{for\; left\; leg}. \end{aligned} $$
  • Reference signals

$$ \theta_{{\rm tr,ref}} = 0.7. $$
  • Neuronal network of spinal pattern generation

$$ {\mathbf{W}}_{{\rm PG}} = {\left[{\begin{array}{*{20}c} {0.4}& {0}& {0}& {0}& {0}& {0.1}& {0}& {0.4}& {0.2}\\ {0}& {0.3}& {0.5}& {0}& {0}& {0}& {0}& {0.3}& {0.2}\\ {0.5}& {0}& {0.64}& {0}& {1.0}& {0}& {0.35}& {0}& {0}\\ {0}& {0}& {0}& {0}& {0}& {0.1}& {0}& {0.2}& {0}\\ {0.4}& {0.1}& {0}& {0.8}& {0}& {0}& {0.4}& {0.1}& {0}\\ \end{array}} \right]}^{\rm T} $$

Rights and permissions

Reprints and permissions

About this article

Cite this article

Jo, S. Hypothetical neural control of human bipedal walking with voluntary modulation. Med Bio Eng Comput 46, 179–193 (2008). https://doi.org/10.1007/s11517-007-0277-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11517-007-0277-8

Keywords

Navigation