Skip to main content
Log in

Dynamical consequences of sensory feedback in a half-center oscillator coupled to a simple motor system

  • Original Article
  • Published:
Biological Cybernetics Aims and scope Submit manuscript

Abstract

We investigate a simple model for motor pattern generation that combines central pattern generator (CPG) dynamics with a sensory feedback (FB) mechanism. Our CPG comprises a half-center oscillator with conductance-based Morris–Lecar model neurons. Output from the CPG drives a push–pull motor system with biomechanics based on experimental data. A sensory feedback conductance from the muscles allows modulation of the CPG activity. We consider parameters under which the isolated CPG system has either “escape” or “release” dynamics, and we study both inhibitory and excitatory feedback conductances. We find that increasing the FB conductance relative to the CPG conductance makes the system more robust against external perturbations, but more susceptible to internal noise. Conversely, increasing the CPG conductance relative to the FB conductance has the opposite effects. We find that the “closed-loop” system, with sensory feedback in place, exhibits a richer repertoire of behaviors than the “open-loop” system, with motion determined entirely by the CPG dynamics. Moreover, we find that purely feedback-driven motor patterns, analogous to a chain reflex, occur only in the inhibition-mediated system. Finally, for pattern generation systems with inhibition-mediated sensory feedback, we find that the distinction between escape- and release-mediated CPG mechanisms is diminished in the presence of internal noise. Our observations support an anti-reductionist view of neuromotor physiology: Understanding mechanisms of robust motor control requires studying not only the central pattern generator circuit in isolation, but the intact closed-loop system as a whole.

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
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22
Fig. 23
Fig. 24
Fig. 25
Fig. 26
Fig. 27

Similar content being viewed by others

Notes

  1. Note that [75] used normalized muscle lengths (cf. eqn. (6) of that paper). Here, we fix a length scale \(\ell =1\)mm to give a physiologically plausible range of motion.

  2. We leave implicit a conversion factor from mV to Hz.

References

  1. Anderson DF, Ermentrout B, Thomas PJ (2015) Stochastic representations of ion channel kinetics and exact stochastic simulation of neuronal dynamics. J Comput Neurosci 38(1):67–82

    Article  PubMed  Google Scholar 

  2. Bässler U (1986) On the definition of central pattern generator and its sensory control. Biol Cybern 54(1):65–69

    Article  Google Scholar 

  3. Ben-Tal A, Smith JC (2008) A model for control of breathing in mammals: coupling neural dynamics to peripheral gas exchange and transport. J Theor Biol 251(3):480–497

    Article  PubMed  Google Scholar 

  4. Ben-Tal A, Tawhai MH (2013) Integrative approaches for modeling regulation and function of the respiratory system. Wiley Interdiscip Rev Syst Biol Med 5(6):687–699

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Ben-Tal A, Wang Y, Leite MCA (2019) The logic behind neural control of breathing pattern. Sci Rep 9(1):1–19

    Article  CAS  Google Scholar 

  6. Borisyuk GN, Borisyuk RM, Khibnik AI, Roose D (1995) Dynamics and bifurcations of two coupled neural oscillators with different connection types. Bull Math Biol 57(6):809–840

    Article  CAS  PubMed  Google Scholar 

  7. Brown TG (1914) 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(1):18–46

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Brown TG (1911) The intrinsic factors in the act of progression in the mammal. Proc R Soc Lond Ser B 84(572):308–319

    Article  Google Scholar 

  9. Butera RJ Jr, Rinzel J, Smith JC (1999) Models of respiratory rhythm generation in the pre-Bötzinger complex I Bursting pacemaker neurons. J Neurophysiol 82(1):382–397

    Article  PubMed  Google Scholar 

  10. Butera RJ Jr, Rinzel J, Smith JC (1999) Models of respiratory rhythm generation in the pre-Bötzinger complex. II. Populations of coupled pacemaker neurons. J Neurophysiol 82(1):398–415

  11. Chiel HJ (2007) Aplysia feeding biomechanics. Scholarpedia 2(9):4165

  12. Chiel HJ, Beer RD (1997) The brain has a body: adaptive behavior emerges from interactions of nervous system, body and environment. Trends Neurosci 20(12):553–557

    Article  CAS  PubMed  Google Scholar 

  13. Crespi A, Lachat D, Pasquier A, Ijspeert AJ (2008) Controlling swimming and crawling in a fish robot using a central pattern generator. Autonomous Robots 25(1–2):3–13

    Article  Google Scholar 

  14. Daun S, Rubin JE, Rybak IA (2009) Control of oscillation periods and phase durations in half-center central pattern generators: a comparative mechanistic analysis. J Comput Neurosci 27(1):3

    Article  PubMed  PubMed Central  Google Scholar 

  15. Dhooge A, Willy Govaerts YA, Kuznetsov HGEM, Sautois B (2008) New features of the software matcont for bifurcation analysis of dynamical systems. Math Comput Model Dyn Syst 14(2):147–175

    Article  Google Scholar 

  16. Diekman CO, Thomas PJ, Wilson CG (2017) Eupnea, tachypnea, and autoresuscitation in a closed-loop respiratory control model. J Neurophysiol 118(4):2194–2215

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Diekman CO, Thomas PJ, Wilson CG (2018) Experimental validation of a closed-loop respiratory control model using dynamic clamp. In: 2018 40th annual international conference of the IEEE engineering in medicine and biology society (EMBC). IEEE, pp 5273–5276

  18. Dietz V, Zijlstra W, Duysens J (1994) Human neuronal interlimb coordination during split-belt locomotion. Exp Brain Res 101(3):513–520

    Article  CAS  PubMed  Google Scholar 

  19. Drushel RF, Neustadter DM, Hurwitz I, Crago PE, Chiel HJ (1998) Kinematic models of the buccal mass of Aplysia californica. J Exp Biol 201(10):1563–1583

    Article  CAS  PubMed  Google Scholar 

  20. Drushel RF, Sutton GP, Neustadter DM, Mangan EV, Adams BW, Crago PE, Chiel HJ (2002) Radula-centric and odontophore-centric kinematic models of swallowing in Aplysia californica. J Exp Biol 205(14):2029–2051

    Article  PubMed  Google Scholar 

  21. Ermentrout GB, Terman DH (2010) Mathematical foundations of neuroscience, vol 35. Springer, Berlin

    Google Scholar 

  22. Ferree TC, Lockery SR (1999) Computational rules for chemotaxis in the nematode C. elegans. J Comput Neurosci 6(3):263–277

  23. Flor KC, Barnett WH, Karlen-Amarante M, Molkov Y, Zoccal DB (2019) Inhibitory control of active expiration by the bötzinger complex in rats. bioRxiv

  24. Frigon A, Desrochers É, Thibaudier Y, Hurteau M-F, Dambreville C (2017) Left-right coordination from simple to extreme conditions during split-belt locomotion in the chronic spinal adult cat. J Physiol 595(1):341–361

    Article  CAS  PubMed  Google Scholar 

  25. Gesslbauer B, Hruby LA, Roche AD, Farina D, Blumer R, Aszmann OC (2017) Axonal components of nerves innervating the human arm. Ann Neurol 82(3):396–408

    Article  CAS  PubMed  Google Scholar 

  26. Gill Jeffrey P, Chiel Hillel J (2020) Rapid adaptation to changing mechanical load by ordered recruitment of identified motor neurons. Eneuro. https://doi.org/10.1523/ENEURO.0016-20.2020

  27. Golubitsky M, Stewart I (2018) Homeostasis with multiple inputs. SIAM J Appl Dyn Syst 17(2):1816–1832

    Article  Google Scholar 

  28. Golubitsky M, Stewart I, Buono P-L, Collins JJ (1999) Symmetry in locomotor central pattern generators and animal gaits. Nature 401(6754):693–695

    Article  CAS  PubMed  Google Scholar 

  29. Grodins FS, June B, Bart AJ (1967) Mathematical analysis and digital simulation of the respiratory control system. J Appl Physiol 22(2):260–276

    Article  CAS  PubMed  Google Scholar 

  30. Guertin PA (2009) The mammalian central pattern generator for locomotion. Brain Res Rev 62(1):45–56

    Article  PubMed  Google Scholar 

  31. Harris-Warrick RM, Cohen AH (1985) Serotonin modulates the central pattern generator for locomotion in the isolated lamprey spinal cord. J Exp Biol 116(1):27–46

    Article  CAS  PubMed  Google Scholar 

  32. Hurwitz I, Neustadter D, Morton DW, Chiel HJ, Susswein AJ (1996) Activity patterns of the B31/B32 pattern initiators innervating the I2 muscle of the buccal mass during normal feeding movements in Aplysia californica. J Neurophysiol 75(4):1309–1326

    Article  CAS  PubMed  Google Scholar 

  33. Ijspeert AJ (2008) Central pattern generators for locomotion control in animals and robots: a review. Neural Netw 21(4):642–653

    Article  PubMed  Google Scholar 

  34. Ijspeert AJ, Crespi A (2007) Online trajectory generation in an amphibious snake robot using a lamprey-like central pattern generator model. In: Proceedings 2007 IEEE international conference on robotics and automation. IEEE, pp 262–268

  35. Izquierdo EJ, Beer RD (2013) Connecting a connectome to behavior: an ensemble of neuroanatomical models of C. elegans klinotaxis. PLoS Comput Biol 9(2):e1002890

  36. Kehl CE, Wu J, Lu S, Neustadter DM, Drushel RF, Smoldt RK, Chiel HJ (2019) Soft-surface grasping: radular opening in Aplysia californica. J Exp Biol 222(16):jeb191254

  37. Kuo AD (2002) The relative roles of feedforward and feedback in the control of rhythmic movements. Mot Control 6(2):129–145

    Article  Google Scholar 

  38. Kuznetsov Y (2004) Elements of Applied Bifurcation Theory, vol 112. Springer, New York

    Google Scholar 

  39. Lu H, McManus JM, Cullins MJ, Chiel HJ (2015) Preparing the periphery for a subsequent behavior: motor neuronal activity during biting generates little force but prepares a retractor muscle to generate larger forces during swallowing in Aplysia. J Neurosci 35(12):5051–5066

    Article  PubMed  PubMed Central  Google Scholar 

  40. Lyttle DN, Gill JP, Shaw KM, Thomas PJ, Chiel HJ (2017) Robustness, flexibility, and sensitivity in a multifunctional motor control model. Biol Cybern 111(1):25–47

    Article  PubMed  Google Scholar 

  41. Manfredi L, Assaf T, Mintchev S, Marrazza S, Capantini L, Orofino S, Ascari L, Grillner S, Wallén P, Ekeberg Ö et al (2013) A bioinspired autonomous swimming robot as a tool for studying goal-directed locomotion. Biol Cybern 107(5):513–527

    Article  CAS  PubMed  Google Scholar 

  42. Maturana HR, Varela FJ (1980) Autopoiesis and cognition: The realization of the living, volume 42 of Boston Studies in the Philosophy of Science. D. Reidel Publishing Company, Boston

  43. McManus Jeffrey M, Hui Lu, Chiel Hillel J (2012) An in vitro preparation for eliciting and recording feeding motor programs with physiological movements in Aplysia californica. JoVE 70:e4320

    Google Scholar 

  44. Meng L, Macleod CA, Porr B, Gollee H (2018) Bipedal robotic walking control derived from analysis of human locomotion. Biol Cybern 112(3):277–290

    Article  PubMed  PubMed Central  Google Scholar 

  45. Molkov YI, Shevtsova NA, Park C, Ben-Tal A, Smith JC, Rubin JE, Rybak IA (2014) A closed-loop model of the respiratory system: focus on hypercapnia and active expiration. PloS one 9(10):e109894

    Article  PubMed  PubMed Central  Google Scholar 

  46. Morris C, Lecar H (1981) Voltage oscillations in the barnacle giant muscle fiber. Biophys J 35(1):193–213

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Morton DW, Chiel HJ (1993) The timing of activity in motor neurons that produce radula movements distinguishes ingestion from rejection in Aplysia. J Comp Physiol A 173(5):519–536

    Article  CAS  PubMed  Google Scholar 

  48. Neustadter DM, Drushel RF, Chiel HJ (2002) Kinematics of the buccal mass during swallowing based on magnetic resonance imaging in intact, behaving Aplysia californica. J Exp Biol 205(7):939–958

    Article  PubMed  Google Scholar 

  49. Neustadter DM, Drushel RF, Crago PE, Adams BW, Chiel HJ (2002) A kinematic model of swallowing in Aplysia californica based on radula/odontophore kinematics and in vivo magnetic resonance images. J Exp Biol 205(20):3177–3206

    Article  PubMed  Google Scholar 

  50. Neustadter David M, Herman Robert L, Drushel Richard F, Chestek David W, Chiel Hillel J (2007) The kinematics of multifunctionality: comparisons of biting and swallowing in Aplysia californica. J Exp Biol 210(2):238–260

    Article  CAS  PubMed  Google Scholar 

  51. Novakovic VA, Sutton GP, Neustadter DM, Beer RD, Chiel HJ (2006) Mechanical reconfiguration mediates swallowing and rejection in Aplysia californica. J Comp Physiol A 192(8):857

    Article  Google Scholar 

  52. Pearson KG (1985) Are there central pattern generators for walking and flight in insects? In: Feedback and motor control in invertebrates and vertebrates. Springer, pp 307–315

  53. Shusen Pu, Thomas Peter J (2020) Fast and accurate langevin simulations of stochastic Hodgkin-Huxley dynamics. Neural Comput 32(10):1775–1835

    Article  Google Scholar 

  54. Rinzel J, Ermentrout GB (1998) Analysis of neural excitability and oscillations. Methods Neuronal Model 2:251–292

    Google Scholar 

  55. Schroer RT, Boggess MJ, Bachmann RJ, Quinn RD, Ritzmann RE (2004) Comparing cockroach and whegs robot body motions. In:IEEE International conference on robotics and automation, 2004. Proceedings. ICRA’04. 2004, vol 4. IEEE, pp 3288–3293

  56. Shaw Kendrick M, Lyttle David N, Gill Jeffrey P, Cullins Miranda J, McManus Jeffrey M, Hui L, Thomas Peter J, Chiel Hillel J (2015) The significance of dynamical architecture for adaptive responses to mechanical loads during rhythmic behavior. J Comput Neurosci 38(1):25–51

    Article  CAS  PubMed  Google Scholar 

  57. Skinner FK, Kopell N, Marder E (1994) Mechanisms for oscillation and frequency control in reciprocally inhibitory model neural networks. J Comput Neurosci 1(1–2):69–87

    Article  CAS  PubMed  Google Scholar 

  58. Skinner FK, Turrigiano GG, Marder E (1993) Frequency and burst duration in oscillating neurons and two-cell networks. Biol Cybern 69(5–6):375–383

    Article  CAS  PubMed  Google Scholar 

  59. Smith JC, Ellenberger HH, Ballanyi K, Richter DW, Feldman JL (1991) Pre-Botzinger complex: a brainstem region that may generate respiratory rhythm in mammals. Science 254(5032):726–729

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Somers D, Kopell N (1993) Rapid synchronization through fast threshold modulation. Biol Cybern 68(5):393–407

    Article  CAS  PubMed  Google Scholar 

  61. Song S, Geyer H (2015) A neural circuitry that emphasizes spinal feedback generates diverse behaviours of human locomotion. J Physiol 593(16):3493–3511

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Song S, Geyer H (2017) Evaluation of a neuromechanical walking control model using disturbance experiments. Frontiers Comput Neurosci 11:15

    Article  Google Scholar 

  63. Song S, Geyer H (2018) Predictive neuromechanical simulations indicate why walking performance declines with ageing. J Physiol 596(7):1199–1210

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Spardy LE, Markin SN, Shevtsova NA, Prilutsky BI, Rybak IA, Rubin JE (2011) A dynamical systems analysis of afferent control in a neuromechanical model of locomotion: I. Rhythm generation. J Neural Eng 8(6):065003

    Article  PubMed  Google Scholar 

  65. Spardy LE, Markin SN, Shevtsova NA, Prilutsky BI, Rybak IA, Rubin JE (2011) A dynamical systems analysis of afferent control in a neuromechanical model of locomotion: II. Phase asymmetry. J Neural Eng 8(6):065004

    Article  PubMed  Google Scholar 

  66. Sutton GP, Macknin JB, Gartman SS, Sunny GP, Beer RD, Crago PE, Neustadter DM, Chiel HJ (2004) Passive hinge forces in the feeding apparatus of Aplysia aid retraction during biting but not during swallowing. J Comp Physiol A 190(6):501–514

    Article  CAS  Google Scholar 

  67. Sutton GP, Mangan EV, Neustadter DM, Beer RD, Crago PE, Chiel HJ (2004) Neural control exploits changing mechanical advantage and context dependence to generate different feeding responses in Aplysia. Biol Cybern 91(5):333–345

    Article  PubMed  Google Scholar 

  68. Wang X-J, Rinzel J (1992) Alternating and synchronous rhythms in reciprocally inhibitory model neurons. Neural Comput 4(1):84–97

    Article  Google Scholar 

  69. Wang Y, Gill JP, Chiel HJ, Thomas PJ (2021) Shape versus timing: linear responses of a limit cycle with hard boundaries under instantaneous and static perturbation. SIAM J Appl Dyn Syst (in press). arXiv preprint arXiv:1906.04387

  70. Webster-WVA, Gill JP, Thomas PJ, Chiel HJ (2020) Control for multifunctionality: Bioinspired control based on feeding in Aplysia californica. arXiv preprint arXiv:2008.04978

  71. White JA, Rubinstein JT, Kay AR (2000) Channel noise in neurons. Trends Neurosci 23(3):131–137

    Article  CAS  PubMed  Google Scholar 

  72. Norbert W (1948) Cybernetics or control and communication in the animal and the machine. Technology Press

  73. Williams ND, Mehlsen J, Tran HT, Olufsen MS (2019) An optimal control approach for blood pressure regulation during head-up tilt. Biol Cybern 113(1–2):149–159

    Article  PubMed  Google Scholar 

  74. Ye H, Morton DW, Chiel HJ (2006) Neuromechanics of multifunctionality during rejection in Aplysia californica. J Neurosci 26(42):10743–10755

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Sung Nien Yu, Crago PE, Chiel HJ (1999) Biomechanical properties and a kinetic simulation model of the smooth muscle I2 in the buccal mass of Aplysia. Biol Cybern 81(5–6):505–513

    Google Scholar 

  76. Zhang C, Lewis TJ (2013) Phase response properties of half-center oscillators. J Comput Neurosci 35(1):55–74

    Article  CAS  PubMed  Google Scholar 

  77. Zoccal DB, Silva JN, Barnett WH, Lemes EV, Falquetto B, Colombari E, Molkov YI, Moreira TS, Takakura AC (2018) Interaction between the retrotrapezoid nucleus and the parafacial respiratory group to regulate active expiration and sympathetic activity in rats. Am J Physiol Lung Cellular Mol Physiol 315(5):L891–L909

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The authors thank Profs. H. Chiel and D. Durand (CWRU), as well as J. Gill, Ph.D. (CWRU), for illuminating discussions of neurophysiology and motor control. This work was supported in part by National Institutes of Health BRAIN Initiative grant R01 NS118606. PT thanks the Oberlin College Department of Mathematics for research support.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zhuojun Yu.

Additional information

Communicated by Benjamin Lindner.

Publisher's Note

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

Appendices

Table of Parameter Values

Table 2 specifies the parameter values used for simulations

Table 2 Parameter Values

Model equations

For completeness, we list here the full equations of the model, as introduced in Sect.  2. The simulation codes are available from https://github.com/zhuojunyu-appliedmath/CPG-FB. Instructions for reproducing each figure and table in the paper are provided (see the README file at the github site).

For cell i (\(\ne j\)),

$$\begin{aligned} C\frac{dV_i}{dt}&=I_\text {ext}-g_\text {L}(V_i-E_\text {L})-g_\text {Ca}M_\infty (V_i)(V_i-E_\text {Ca})\\&\quad -g_\text {K}N_i(V_i-E_\text {K})\\&\quad -g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}(V_j)\left( V_i-E_\text {syn}^\text {CPG}\right) \\&\quad -g_\text {syn}^\text {FB}S_\infty ^\text {FB}(L_j)(V_i-E_\text {syn}^\text {FB}),\\ \frac{dN_i}{dt}&=\lambda _N(V_i)(N_\infty (V_i)-N_i), \end{aligned}$$

and

$$\begin{aligned} M_\infty (V_i)&=\frac{1}{2}\left( 1+\tanh \left( \frac{V_i-E_1}{E_2}\right) \right) ,\\ N_\infty (V_i)&=\frac{1}{2}\left( 1+\tanh \left( \frac{V_i-E_3}{E_4}\right) \right) ,\\ \lambda _N(V_i)&=\phi _N\left( \cosh \left( \frac{V_i-E_3}{2E_4}\right) \right) ,\\ S_\infty ^\text {CPG}(V_j)&=\frac{1}{2}\left( 1+\tanh \left( \frac{V_j-E_\text {thresh}}{E_\text {slope}}\right) \right) ,\\ S_\infty ^\text {FB}(L_j)&=1-\frac{1}{2}\tanh \left( \frac{L_j-x_0\ell }{L_\text {slope}}\right) . \end{aligned}$$

For muscle i,

$$\begin{aligned} F_i(t)=F_0\cdot a_i(t)\cdot LT(L_i(t)). \end{aligned}$$

Here,

$$\begin{aligned}&LT(L_i(t))=-\left( 5.27\times 10^{-4}\right) (L_i(t)/\ell )^2\\&\quad +0.1054(L_i(t)/\ell )-4.27,\\&L_i(t)=(50\pm 0.8x(t))\ell ,\\&\quad L_j(t)=(50\mp 0.8x(t))\ell ,\quad -50<x<50, \end{aligned}$$

and

$$\begin{aligned} a_i(t)=g[A_i(t)-a_0],\quad \quad 0<a_i\le 1, \end{aligned}$$

where

$$\begin{aligned} dA_i/dt= & {} \tau ^{-1}\{U_i-[\beta +(1-\beta )U_i]A_i\},\\ U_i= & {} 1.03-4.31e^{(-0.198u_i)},\quad \quad 0\le U_i\le 1,\\ u_i= & {} (1/2)V_i. \end{aligned}$$

The position of the pendulum is determined by

$$\begin{aligned} \frac{dx}{dt}=\frac{1}{b\ell }(F_2-F_1). \end{aligned}$$

Flow-invariant set

Here we show that the state space of our system has a a compact flow-invariant domain, that is, a subset \(\Omega \subset \mathbb {R}^7\) that is invariant under the flow generated by the model, forward in time. We describe the set as follows:

$$\begin{aligned}&(V_1,V_2,N_1,N_2,A_1,A_2,x)\in \Omega \nonumber \\&\quad = [V_\text {min},V_\text {max}]^2 \times [0,1]^2 \nonumber \\&\qquad \times [0,A_\text {max}]^2 \times [-x_\text {max},x_\text {max}]. \end{aligned}$$
(23)

For the specific parameters, we use \(V_\text {min}\approx -76.63\) mV, \(V_\text {max}= 110\) mV, \(A_\text {max}\approx 1.021\), and \(x_\text {max}\approx 8.05\).

For the range of \(N_i\), since \(N_\infty \in [0,1]\), \(N_i\in [0,1]\) for all time as in Eq. (3). The voltage \(V_i\in [V_\text {min},V_\text {max}]\) because if \(V<V_\text {min}\), \(\frac{dV}{dt}>0\) for all values of the other variables’ ranges, and if \(V>V_\text {max}\), \(\frac{dV}{dt}<0\) always holds. Considering the reversal potentials of all conductances, when \(V\ge E_\text {Ca}\), the maximal possible inward current would occur if the potassium, CPG and FB conductances were shut off (\(N=0, S_\infty ^\text {CPG}=S_\infty ^\text {FB}=0\)), with extremal value of the calcium gate (\(M_\infty \)) to be determined. Thus,

$$\begin{aligned} \frac{dV_i}{dt}&\le I_\text {ext}-g_\text {L}(V_i-E_\text {L})-g_\text {Ca}M_{\infty }(V_i-E_\text {Ca})\\&=(I_\text {ext}+g_\text {L}E_\text {L}+g_\text {Ca}M_{\infty }E_\text {Ca})-(g_\text {L}+g_\text {Ca}M_\infty )V_i. \end{aligned}$$

To make \(\frac{dV_i}{dt}<0\) hold \(\forall \, V_i>V_\text {max}\), then,

$$\begin{aligned} V_\text {max}= & {} \max _{M_\infty \in [0,1]} \left\{ \frac{I_\text {ext}+g_\text {L}E_\text {L}+g_\text {Ca}M_{\infty }E_\text {Ca}}{g_\text {L}+g_\text {Ca}M_\infty }\right\} \\= & {} \max _{M_\infty \in [0,1]}\left\{ \frac{I_\text {ext}+g_\text {L}E_\text {L}+E_\text {Ca}(g_\text {L}+g_\text {Ca}M_\infty )-E_\text {Ca}g_\text {L}}{g_\text {L}+g_\text {Ca}M_\infty }\right\} \\= & {} E_\text {Ca}+\max _{M_\infty \in [0,1]}\left\{ \frac{I_\text {ext}+g_\text {L}E_\text {L}-E_\text {Ca}g_\text {L}}{g_\text {L}+g_\text {Ca}M_\infty }\right\} . \end{aligned}$$

As \(M_\infty \) takes its minimum at 0, \(V_\text {max}=110\) mV for parameters we choose in Table 2.

In order to find \(V_\text {min}\), we turn off the \(\text {Ca}^{2+}\) channel and allow for maximal inhibitory feedback:

$$\begin{aligned} \frac{dV_i}{dt}&\ge I_\text {ext}-g_\text {L}(V_i-E_\text {L})-g_\text {K}N_i(V_i-E_\text {K})\\&\quad -g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}\left( V_i-E_\text {syn}^\text {CPG}\right) -g_\text {syn}^\text {FB}S_\infty ^\text {FB}(V_i-E_\text {syn}^\text {FB})\\&=(I_\text {ext}+g_\text {L}E_\text {L}+g_\text {K}N_iE_\text {K}\\&\quad +g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}E_\text {syn}^\text {CPG}+g_\text {syn}^\text {FB} S_\infty ^\text {FB}E_\text {syn}^\text {FB})\\&\quad -(g_\text {L}+g_\text {K}N_i+g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}+g_\text {syn}^\text {FB}S_\infty ^\text {FB})V_i. \end{aligned}$$

If \(\frac{dV_i}{dt}>0\) always holds for arbitrary \(V_i<V_\text {min}\), then

$$\begin{aligned} V_\text {min}&=\min _{N_i,S^\text {CPG}_\infty ,S_\infty ^\text {FB}\in [0,1]}\\&\quad \left\{ \frac{I_\text {ext}+g_\text {L}E_\text {L}+g_\text {K}N_iE_\text {K}+g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}E_\text {syn}^\text {CPG} +g_\text {syn}^\text {FB}S_\infty ^\text {FB}E_\text {syn}^\text {FB}}{g_\text {L}+g_\text {K}N_i+g_\text {syn}^\text {CPG}S_\infty ^\text {CPG} +g_\text {syn}^\text {FB}S_\infty ^\text {FB}}\right\} \\&=\min _{N_i,S^\text {CPG}_\infty ,S_\infty ^\text {FB}\in [0,1]}\\&\quad \left\{ \frac{I_\text {ext}+g_\text {L}E_\text {L}-80(g_\text {L}+g_\text {K}N_i+g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}+g_\text {syn}^\text {FB}S_\infty ^\text {FB}) +80g_\text {L}}{g_\text {L}+g_\text {K}N_i+g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}+g_\text {syn}^\text {FB}S_\infty ^\text {FB}}\right\} \\&=-80+\min _{N_i,S^\text {CPG}_\infty ,S_\infty ^\text {FB}\in [0,1]}\left\{ \frac{I_\text {ext}+g_\text {L}E_\text {L}+80g_\text {L}}{g_\text {L}+g_\text {K}N_i+g_\text {syn}^\text {CPG}S_\infty ^\text {CPG}+ g_\text {syn}^\text {FB}S_\infty ^\text {FB}}\right\} . \end{aligned}$$

When \(N_i\), \(S_\infty ^\text {CPG}\) and \(S_\infty ^\text {FB}\) simultaneously reach their maximum, \(V_\text {min}\approx -76.63\) mV. For comparison, the voltage range in Fig. 2 roughly corresponds to \(V_\text {min}<V<V_\text {max}\).

For the variables related to the biomechanics, as shown in Eq. (12), the range of \(U_i\) is

$$\begin{aligned} {[}0,U_\text {max}]=[0,f(V_\text {max}/2)]\subset [0,1.03). \end{aligned}$$

The U-A relation, Eq. (13), gives

$$\begin{aligned} A_i\in [0,A_\text {max}]=\left[ 0,\frac{U_\text {max}}{\beta +(1-\beta )U_\text {max}}\right] \subset [0,1.021). \end{aligned}$$

Since \(L_1=(50+0.8x)\ell \) and \(L_2=(50-0.8x)\ell \), then as in Eq. (11),

$$\begin{aligned} LT(L_1)= & {} -(5.27\times 10^{-4})(50+0.8x)^2\\&+0.1054(50+0.8x)-4.27,\\ LT(L_2)= & {} -(5.27\times 10^{-4})(50-0.8x)^2\\&+0.1054(50-0.8x)-4.27. \end{aligned}$$

For convenience, we drop the length scale \(\ell \,(=1\text {mm})\) since it does not affect the calculation result. Equation (14) defines the range for \(a_i\): \(a_i\in (0,a_\text {max}]=(0,g(A_\text {max}-a_0)]\subset (0,1.712)\). Because \(F_i\) is determined by \(a_i\) and x (Eq. (10)), and x is in turn determined by \(F_i\) (Eq. (16)), we can compute \(x_\text {min}\) and \(x_\text {max}\) through the ODE \(dx/dt=\frac{1}{b}(F_2-F_1)\), i.e.,

$$\begin{aligned} \frac{dx}{dt}&=\frac{1}{b}(F_2-F_1)\\&=\frac{1}{b}(F_0\cdot a_2\cdot LT(L_2)-F_0\cdot a_1\cdot LT(L_1))\\&=\frac{F_0}{b}\{a_2[-(5.27\times 10^{-4})(50-0.8x)^2\\&\quad +0.1054(50-0.8x)-4.27]\\&\quad -a_1[-(5.27\times 10^{-4})(50+0.8x)^2\\&\quad +0.1054(50+0.8x)-4.27]\}\\&=\frac{F_0}{b}[(3.3728\times 10^{-4})(a_1-a_2)x^2\\&\quad -0.04216(a_1+a_2)x+0.3175(a_1-a_2)]\\&:=\frac{F_0}{b}[W(a_1-a_2)x^2-Y(a_1+a_2)x+Z(a_1-a_2)]. \end{aligned}$$

On the one hand, if \(x>x_\text {max}\), then \(\frac{dx}{dt}<0\) for all other variables within their respective ranges. Thus, we aim to find \(x_\text {max}\in (-50,50)\) such that for all \(x_\text {max}<x<50\)\(W(a_1-a_2)x^2-Y(a_1+a_2)x+Z(a_1-a_2)<0\) for arbitrary \(a_1,\, a_2\in (0,1.712)\). If \(a_1\le a_2\), the inequality holds for all \(x>0\). When \(a_1>a_2\), \(x=50\) always satisfies the inequality, indicating that \(x_\text {max}(<50)\) exists and should be the maximal abscissa of the left intersection of the quadratic function and horizontal axis, denoted by \(x_\text {L}\):

$$\begin{aligned} x_\text {L}&=\frac{Y(a_1+a_2)-\sqrt{Y^2(a_1+a_2)^2-4WZ(a_1-a_2)^2}}{2W(a_1-a_2)}\\&=\frac{Y}{2W}\left( \frac{1+\frac{a_2}{a_1}}{1-\frac{a_2}{a_1}}\right) -\frac{1}{2W}\sqrt{Y^2\left( \frac{1+\frac{a_2}{a_1}}{1-\frac{a_2}{a_1}}\right) ^2-4WZ}. \end{aligned}$$

Let \(t=\frac{a_2}{a_1}\in (0,1)\) and let \(r=\frac{1+t}{1-t}=1+\frac{2}{\frac{1}{t}-1}\). Then r is a monotonically increasing function from 1 to \(\infty \). Hence,

$$\begin{aligned}x_\text {L}=\frac{Y}{2W}r-\frac{1}{2W}\sqrt{Y^2r^2-4WZ}.\end{aligned}$$

It is not hard to find that \(x_\text {L}\) monotonically decreases by calculating its derivative, and thus,

$$\begin{aligned}x_\text {max}=(x_\text {L})_\text {max}=x_\text {L}\big |_{r=1}=\frac{Y-\sqrt{Y^2-4WZ}}{2W}\approx 8.05.\end{aligned}$$

On the other hand, if \(x<x_\text {min}\), then \(\frac{dx}{dt}>0\) for all other variables within their respective ranges. Given the symmetry of our model, \(x_\text {min}=-x_\text {max}\approx -8.05\). Therefore, the flow-invariant set for the solutions of our HCO–muscle model has the form given by (23), with \(V_\text {min}\approx -76.63\) mV, \(V_\text {max}= 110\) mV, \(A_\text {max}\approx 1.021\) and \(x_\text {max}\approx 8.05\) for the specific parameters used in this paper.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yu, Z., Thomas, P.J. Dynamical consequences of sensory feedback in a half-center oscillator coupled to a simple motor system. Biol Cybern 115, 135–160 (2021). https://doi.org/10.1007/s00422-021-00864-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00422-021-00864-y

Navigation