Skip to main content
Log in

Control of oscillation periods and phase durations in half-center central pattern generators: a comparative mechanistic analysis

  • Published:
Journal of Computational Neuroscience Aims and scope Submit manuscript

Abstract

Central pattern generators (CPGs) consisting of interacting groups of neurons drive a variety of repetitive, rhythmic behaviors in invertebrates and vertebrates, such as arise in locomotion, respiration, mastication, scratching, and so on. These CPGs are able to generate rhythmic activity in the absence of afferent feedback or rhythmic inputs. However, functionally relevant CPGs must adaptively respond to changing demands, manifested as changes in oscillation period or in relative phase durations in response to variations in non-patterned inputs or drives. Although many half-center CPG models, composed of symmetric units linked by reciprocal inhibition yet varying in their intrinsic cellular properties, have been proposed, the precise oscillatory mechanisms operating in most biological CPGs remain unknown. Using numerical simulations and phase-plane analysis, we comparatively investigated how the intrinsic cellular features incorporated in different CPG models, such as subthreshold activation based on a slowly inactivating persistent sodium current, adaptation based on slowly activating calcium-dependent potassium current, or post-inhibitory rebound excitation, can contribute to the control of oscillation period and phase durations in response to changes in excitatory external drive to one or both half-centers. Our analysis shows that both the sensitivity of oscillation period to alterations of excitatory drive and the degree to which the duration of each phase can be separately controlled depend strongly on the intrinsic cellular mechanisms involved in rhythm generation and phase transitions. In particular, the CPG formed from units incorporating a slowly inactivating persistent sodium current shows the greatest range of oscillation periods and the greatest degree of independence in phase duration control by asymmetric inputs. These results are explained based on geometric analysis of the phase plane structures corresponding to the dynamics for each CPG type, which in particular helps pinpoint the roles of escape and release from synaptic inhibition in the effects we find.

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

Similar content being viewed by others

References

  • Brown, T. (1911). The intrinsic factors in the act of progression in the mammal. Proceedings of the Royal Society of London Series B, 84, 308–319.

    Article  Google Scholar 

  • Brown, T. (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. Journal of Physiology, 48, 18.

    PubMed  CAS  Google Scholar 

  • Butera, R., Rinzel, J., & Smith, J. (1999). Models of respiratory rhythm generation in the pre-Bötzinger complex. I. Bursting pacemaker neurons. Journal of Neurophysiology, 81, 382–397.

    Google Scholar 

  • Calabrese, R. (1995). Half-center oscillators underlying rhythmic movements. In M. Arbib (Ed.), The handbook of brain theory and neural networks (pp. 444–447). Cambridge: MIT.

    Google Scholar 

  • Curtu, R., Shpiro, A., Rubin, N., & Rinzel, J. (2008). Mechanisms for frequency control in neuronal competition models. SIAM Journal on Applied Dynamical Systems, 7, 609–649.

    Article  Google Scholar 

  • Grillner, S. (1985). Neurobiological bases of rhythmic motor acts in vertebrates. Science, 228, 143–149.

    Article  PubMed  CAS  Google Scholar 

  • Grillner, S. (2006). Biological pattern generation: The cellular and computational logic of networks in motion. Neuron, 52, 751–766.

    Article  PubMed  CAS  Google Scholar 

  • Halbertsma, J. (1983). The stride cycle of the cat: The modelling of locomotion by computerized analysis of automatic recordings. Acta Physiologica Scandinavica Supplementum, 521, 1–75.

    PubMed  CAS  Google Scholar 

  • Harris-Warrick, R. (1993). Pattern generation. Current Opinion in Neurobiology, 3, 982–988.

    Article  PubMed  CAS  Google Scholar 

  • Izhikevich, E. (2006). Dynamical systems in neuroscience: The geometry of excitability and bursting. Cambridge: MIT.

    Google Scholar 

  • Lafreniere-Roula, M., & McCrea, D. (2005). Deletions of rhythmic motoneuron activity during fictive locomotion and scratch provide clues to the organization of the mammalian central pattern generator. Journal of Neurophysiology, 94, 1120–1132.

    Article  PubMed  Google Scholar 

  • Lundberg, A. (1981). Half-centres revisited. In J. Szentagothai, M. Palkovits, & J. Hamori (Eds.), Regulatory functions of the CNS. Motion and organization principles (pp. 155–167). Budapest: Pergamon Akadem Kiado.

    Google Scholar 

  • Marder, E. (2000). Motor pattern generation. Current Opinion in Neurobiology, 10, 691–698.

    Article  PubMed  CAS  Google Scholar 

  • Marder, E., & Bucher, D. (2001). Central pattern generators and the control of rhythmic movements. Current Biology, 11, 986–996.

    Article  Google Scholar 

  • Marder, E., Bucher, D., Schulz, D., & Taylor, A. (2005). Invertebrate central pattern generation moves along. Current Biology, 15, 685–699.

    Article  CAS  Google Scholar 

  • Marder, E., & Calabrese, R. (1996). Principles of rhythmic motor pattern generation. Physiological Reviews, 76, 687–717.

    PubMed  CAS  Google Scholar 

  • Matsuoka, K. (1987). Mechanisms of frequency and pattern control in the neural rhythm generators. Biological Cybernetics, 56, 345–353.

    Article  PubMed  CAS  Google Scholar 

  • McCrea, D., & Rybak, I. (2007). Modeling the mammalian locomotor CPG: Insights from mistakes and perturbations. Progress in Brain Research, 165, 235–253.

    Article  PubMed  Google Scholar 

  • Mischenko, E., Kolesov, Y., Kolesov, A., & Rozov, N. (1994). Asymptotic methods in singularly perturbed systems. New York: Plenum.

    Google Scholar 

  • Olypher, A., Cymbalyuk, G., & Calabrese, R. (2006). Hybrid systems analysis of the control of burst duration by ed calcium current in leech heart interneurons. Journal of Neurophysiology, 96, 2857–2867.

    Article  PubMed  Google Scholar 

  • Orlovskiĭ, G., Deliagina, T., & Grillner, S. (1999). Neuronal control of locomotion: From mollusc to man. Oxford: Oxford University Press.

    Google Scholar 

  • Rowat, P., & Selverston, A. (1993). Modeling the gastric mill central pattern generator of the lobster with a relaxation-oscillator network. Journal of Neurophysiology, 70, 1030–1053.

    PubMed  CAS  Google Scholar 

  • Rubin, J. (2006). Bursting induced by excitatory synaptic coupling in non-identical conditional relaxation oscillators or square-wave bursters. Physical Review E, 74, 021917.

    Article  CAS  Google Scholar 

  • Rubin, J., & Terman, D. (2000). Analysis of clustered firing patterns in synaptically coupled networks of oscillators. Journal of Mathematical Biology, 41, 513–545.

    Article  PubMed  CAS  Google Scholar 

  • Rubin, J., & Terman, D. (2002). Geometric singular perturbation analysis of neuronal dynamics. In B. Fiedler (Ed.), Handbook of dynamical systems. Towards applications (vol. 2). Amsterdam: Elsevier.

    Google Scholar 

  • Rubin, J., & Terman, D. (2004). High frequency stimulation of the subthalamic nucleus eliminates pathological thalamic rhythmicity in a computational model. Journal of Computational Neuroscience, 16, 211–235.

    Article  PubMed  Google Scholar 

  • Rybak, I., Shevtsova, N., Lafreniere-Roula, M., & McCrea, D. (2006). Modelling spinal circuitry involved in locomotor pattern generation: Insights from deletions during fictive locomotion. Journal of Physiology, 577, 617–639.

    Article  PubMed  CAS  Google Scholar 

  • Selverston, A., & Moulins, M. (1985). Oscillatory neural networks. Annual Review of Physiology 47, 29–48.

    Article  PubMed  CAS  Google Scholar 

  • Shpiro, A., Curtu, R., Rinzel, J., & Rubin, N. (2007). Dynamical characteristics common to neuronal competition models. Journal of Neurophysiology, 97, 462–473.

    Article  PubMed  Google Scholar 

  • Sirota, M., & Shik, M. (1973). The cat locomotion elicited through the electrode implanted in the mid-brain. Sechenov Physiological Journal of the USSR, 59, 1314–1321.

    PubMed  CAS  Google Scholar 

  • Skinner, F., Kopell, N., & Marder, E. (1994). Mechanisms for oscillation and frequency control in reciprocally inhibitory model neural networks. Journal of Computational Neuroscience, 1, 69–87.

    Article  PubMed  CAS  Google Scholar 

  • Skinner, F., Turrigiano, G., & Marder, E. (1993). Frequency and burst duration in oscillating neurons and two-cell networks. Biological Cybernetics, 69, 375–383.

    PubMed  CAS  Google Scholar 

  • Sohal, V., & Huguenard, J. (2002). Reciprocal inhibition controls the oscillatory state in thalamic networks. Neurocomputers, 4, 653–659.

    Article  Google Scholar 

  • Somers, D., & Kopell, N. (1993). Rapid synchronization through fast threshold modulation. Biological Cybernetics, 68, 393–407.

    Article  PubMed  CAS  Google Scholar 

  • Sorensen, M., DeWeerth, S., Cymbalyuk, G., & Calabrese, R. (2004). Using a hybrid neural system to reveal regulation of neuronal network activity by an intrinsic current. Journal of Neuroscience, 24, 5427–5438.

    Article  PubMed  CAS  Google Scholar 

  • Tabak, J., O’Donovan, M., & Rinzel, J. (2006). Differential control of active and silent phases in relaxation models of neuronal rhythms. Journal of Computational Neuroscience, 21, 307–328.

    Article  PubMed  Google Scholar 

  • Taylor, A., Cottrell, G., & Kristan, Jr., W. (2002). Analysis of oscillations in a reciprocally inhibitory network with synaptic depression. Neural Computation, 14, 561–581.

    Article  PubMed  Google Scholar 

  • Tazerart, S., Viemari, J., Darbon, P., Vinay, L., & Brocard, F. (2007). Contribution of persistent sodium current to locomotor pattern generation in neonatal rats. Journal of Neurophysiology, 98, 613.

    Article  PubMed  CAS  Google Scholar 

  • Wang, X., & Rinzel, J. (1992). Alternating and synchronous rhythms in reciprocally inhibitory model neurons. Neural Computation, 4, 84–97.

    Article  Google Scholar 

  • Yakovenko, S., McCrea, D., Stecina, K., & Prochazka, A. (2005). Control of locomotor cycle durations. Journal of Neuroscience, 94, 1057–1065.

    CAS  Google Scholar 

  • Zhong, G., Masino, M., & Harris-Warrick, R. (2007) Persistent sodium currents participate in fictive locomotion generation in neonatal mouse spinal cord. Journal of Neuroscience, 27, 4507.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

S. Daun received support from NIH grant R01 HL76137. J.E. Rubin received support from NSF award DMS 0716936. I.A. Rybak was supported by NIH grants RO1 NS048844 and RO1 NS057815.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jonathan E. Rubin.

Additional information

Action Editor: Frances K. Skinner

Appendix

Appendix

We summarize some central notation used in this paper in Table 4. In the following subsections, we list the auxiliary functions and parameter values used in the three example systems that we consider, as introduced in Section 2.2. The XPPAUT code for these systems is available at www.math.pitt.edu/~rubin/pub/CPGFILES/.

Table 4 Some notation used in the paper. Note that h is replaced by Ca for our adaptation model

1.1 A.1 Model featuring the persistent sodium current (Butera et al. 1999)

The ordinary differential equations for the model featuring the persistent sodium current are

$$\begin{array}{rll} \label{persistent} C_mv' & = & -I_{NaP}-I_L-I_{syn}-I_{app},\\ h' & = & (h_{\infty}(v)-h)/\tau_h(v),\\ s' & = & \alpha(1-s)s_{\infty}(v)-\beta s, \end{array}$$
(21)

with associated functions

$$\begin{array}{lll} I_{syn} & = &g_{syn}s(v-e_{syn}),\\ I_{NaP} & = & g_{nap} m_{\infty}(v) h (v-e_{na}),\\ I_L & = & g_l (v-e_l),\\ I_{app} & = & g_{app} v,\\ h_{\infty}(v)& = & 1/(1+\exp((v-\theta_h)/\sigma_h)),\\ s_{\infty}(v) & = & 1/(1+\exp((v-\theta_{syn})/\sigma_{syn})),\\ \tau_h(v) & = & \epsilon \cosh((v-\theta_h)/\sigma_h/2), \; \mbox{and}\\ m_{\infty}(v) & = & 1/(1+\exp((v-\theta_m)/\sigma_m)), \end{array}$$

where C m  = 0.21, g nap  = 10, g l  = 2.8, e na  = 50, e l  = − 65, e syn  = − 80, θ m  = − 37, σ m  = − 6, θ h  = − 30, σ h  = 6, ε = 0.01, θ syn  = − 43, σ syn  = − 0.1, g syn  = 1, g app  = 0.19, α = 1, and β = 1 (fast decay) or β = 0.08 (slow decay).

1.2 A.2 Model featuring postinhibitory rebound (Rubin and Terman 2004; Sohal and Huguenard 2002)

For the model featuring postinhibitory rebound, the relevant differential equations are

$$\begin{array}{lll} C_mv' & = & -I_T -I_{L}-I_{syn}-I_{app},\\ h' & = & (h_{\infty}(v)-h)/\tau_{h}(v),\\ s' & = & \alpha(1-s)s_{\infty}(v)-\beta s,\\ \end{array}$$

with associated functions

$$\begin{array}{lll} I_{syn} & = & g_{syn}s(v-e_{syn}), \\ I_{app} & = & g_{app}v,\\ I_{L} & = & g_{l}(v-v_{l}),\\ I_{T} & = & g_{T}m_{\infty}(v)h(v-v_{ca}),\\ m_{\infty}(v) & = & 1/(1+\exp(-(v-\theta_m)/\sigma_m)),\\ h_{\infty}(v) & = & 1/(1+\exp(-(v-\theta_h)/\sigma_h)),\\ \tau_{h}(v) & = & t_{0}+t_{1}/(1+\exp(-(v-\theta_{ht})/\sigma_{ht})), \; \mbox{and}\\ s_{\infty}(v) & = & 1/(1+\exp((v-\theta_{syn})/\sigma_{syn})), \end{array}$$
(22)

where g T  = 4, v ca  = 90, θ m  = − 40, σ m  = 7.4, θ h  = − 70, σ h  = − 4, θ ht  = − 50, σ ht  = − 3, g l  = 0.4, v l  = − 70, t 0 = 30, t 1 = 200, g syn  = 1.4, v syn  = − 85, θ syn  = − 35, σ syn  = − 0.1, g app  = 0.01, α = 1, and β = 1 (fast decay) or β = 0.05 (slow decay).

1.3 A.3 Model featuring adaptation (modified from Izhikevich 2006)

The model featuring adaptation is given by the differential equations

$$\begin{array}{lll}\label{adaptationeqn} C_mv' & = & -I_{Ca}-I_{ahp}-I_{L}-I_{syn}-I_{app},\\ Ca' & = & \epsilon(-g_{ca}I_{Ca}(v)-k_{ca}(Ca-ca_{base})),\\ s' & = & ((1-s)s_{\infty}(v)-ks)/\tau_s, \end{array}$$

with associated functions

$$\begin{array}{lll} m_{\infty}(v) & = & 1/(1+\exp((v-\theta_m)/\sigma_m)), \\ s_{\infty}(v) & = & 1/(1+\exp((v-\theta_{syn})/\sigma_{syn})),\\ Ca_{\infty}(v) & = & 1/(1+\exp((v-\theta_{ca})/\sigma_{ca})),\\ I_L & = & g_l(v-e_l),\\ I_{app} & = & g_{app}v,\\ I_{syn} & = & g_{syn}s(v-e_{syn}),\\ I_{ahp} & = & g_{ahp}(v-e_k)(Ca^2)/(Ca^2+k_{ahp}^2), \; \mbox{and}\\ I_{ca} & = & \bar{g}_{ca}((Ca_{\infty}(v))^2)(v-v_{ca}). \end{array}$$
(23)

In case 1, C m  = 21, g syn  = 2, g app  = 0.7, e l  = − 55, θ m  = − 34, σ m  = − 5, e k  = − 85, g l  = 1, e syn  = − 70, k = 1, τ s  = 1 (fast decay) or τ s  = 400 (slow decay), σ syn  = − 5, θ syn  = − 20, θ ca  = − 34, σ ca  = − 8.0, k ca  = 22.5, ε = 5e − 05, k ahp  = 0.7, v ca  = 140, g ahp  = 7, g ca  = 0.05, g ca1 = 1, and ca base  = 0.08. In case 2, the same parameters are used except g syn  = 6 (fast decay) or g syn  = 4 (slow decay), e syn  = 50, k = 0.02 (fast decay) or k = 0.002 (slow decay), σ syn  = − 4, and θ syn  = 20.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Daun, S., Rubin, J.E. & Rybak, I.A. Control of oscillation periods and phase durations in half-center central pattern generators: a comparative mechanistic analysis. J Comput Neurosci 27, 3–36 (2009). https://doi.org/10.1007/s10827-008-0124-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-008-0124-4

Keywords

Navigation