Skip to main content
Log in

Multivariable harmonic balance analysis of the neuronal oscillator for leech swimming

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

Abstract

Biological systems, and particularly neuronal circuits, embody a very high level of complexity. Mathematical modeling is therefore essential for understanding how large sets of neurons with complex multiple interconnections work as a functional system. With the increase in computing power, it is now possible to numerically integrate a model with many variables to simulate behavior. However, such analysis can be time-consuming and may not reveal the mechanisms underlying the observed phenomena. An alternative, complementary approach is mathematical analysis, which can demonstrate direct and explicit relationships between a property of interest and system parameters. This paper introduces a mathematical tool for analyzing neuronal oscillator circuits based on multivariable harmonic balance (MHB). The tool is applied to a model of the central pattern generator (CPG) for leech swimming, which comprises a chain of weakly coupled segmental oscillators. The results demonstrate the effectiveness of the MHB method and provide analytical explanations for some CPG properties. In particular, the intersegmental phase lag is estimated to be the sum of a nominal value and a perturbation, where the former depends on the structure and span of the neuronal connections and the latter is roughly proportional to the period gradient, communication delay, and the reciprocal of the intersegmental coupling strength.

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

Similar content being viewed by others

Notes

  1. When the neuronal interconnections have a certain structure, the MHB prediction can be exact (Chen and Iwasaki 2008).

  2. Throughout the paper, mathematical expression A: = B means that A is defined as B.

  3. The function max (x,x th) is equal to x th when x is less than the threshold value x th, and otherwise, it is equal to x.

  4. The maximal eigenvalue is defined to be the one with the largest imaginary part among those with the largest real part.

References

  • Berns, D. W., Moiola, J. L., & Chen, G. (1998). Feedback control of limit cycle amplitudes from a frequency domain approach. Automatica, 34(12), 1567–1573.

    Article  Google Scholar 

  • Blechman, I. I. (1971). Synchronization of dynamical systems. Moscow: Science.

    Google Scholar 

  • Braun, G., & Mulloney, B. (1995). Coordination in the crayfish swimmeret system: Differential excitation causes changes in intersegmental phase. Journal of Neurophysiology, 73, 880–885.

    PubMed  CAS  Google Scholar 

  • Cang, J., & Friesen, W. O. (2002). Model for intersegmental coordination of leech swimming: Central sensory mechanisms. Journal of Neurophysiology, 87, 2760–2769.

    PubMed  Google Scholar 

  • Chen, Z., & Iwasaki, T. (2006). Analysis and synthesis of weakly coupled oscillators by multivariable harmonic balance approach. IEEE Conference on Decision & Control, 661–666.

  • Chen, Z., & Iwasaki, T. (2008). Circulant synthesis of central pattern generators with application to control of rectifier systems. IEEE Transactions on Automatic Control, 53(3), 273–286.

    Article  Google Scholar 

  • Cohen, A. H., Ermentrout, G. B., Kiemel, T., Kopell, N., Sigvardt, K. A., & Williams, T. L. (1992). Modelling of intersegmental coordination in the lamprey central pattern generator for locomotion. Trends in Neurosciences, 15(11), 434–438.

    Article  PubMed  CAS  Google Scholar 

  • Cohen, A. H., Holmes, P. J., & Rand, R. H. (1982). The nature of the coupling between segmental oscillators of the lamprey spinal generator for locomotion: A mathematical model. Journal of Mathematical Biology, 13, 345–369.

    Article  PubMed  CAS  Google Scholar 

  • Collado, A., Ramirez, F., & Pascual, J. P. (2004). Harmonic-balance analysis and synthesis of coupled-oscillator arrays. IEEE Microwave and Wireless Components Letters, 14(5), 192–194.

    Article  Google Scholar 

  • Dale, N. (1995). Experimentally derived model for the locomotor pattern generator in the xenopus embryo. Journal of Physiology, 489, 489–510.

    PubMed  CAS  Google Scholar 

  • Debski, E. A., & Friesen, W. O. (1986). Role of central interneurons in habituation of swimming activity in the medicinal leech. Journal of Neurophysiology, 55(5), 977–994.

    PubMed  CAS  Google Scholar 

  • Ekeberg, O., & Grillner, S. (1999). Simulations of neuromuscular control in lamprey swimming. Philosophical Transactions of the Royal Society of London B, 354, 895–902.

    Article  CAS  Google Scholar 

  • Ekeberg, O., Wallen, P., Lansner, A., Traven, H., Brodin, L., & Grillner, S. (1991). A computer based model for realistic simulations of neural networks. I. The single neuron and synaptic interaction. Biological Cybernetics, 65, 81–90.

    Article  PubMed  CAS  Google Scholar 

  • Ermentrout, G. B., & Chow, C. C. (2002). Modeling neural oscillations. Physiology and Behavior, 77, 629–633.

    Article  PubMed  CAS  Google Scholar 

  • Ermentrout, G. B., & Kopell, N. (1984). Frequency plateaus in a chain of weakly coupled oscillators, I. SIAM Journal on Mathematical Analysis, 15(2), 215–237 (March).

    Article  Google Scholar 

  • Farkas, M. (1994). Periodic motions. New York: Springer.

    Google Scholar 

  • Friesen, W. O., & Hocker, C. G. (2001). Functional analyses of the leech swim oscillator. Journal of Neurophysiology, 86, 824–835.

    PubMed  CAS  Google Scholar 

  • Friesen, W. O., Poon, M., & Stent, G. S. (1978). Neuronal control of swimming in the medicinal leech IV. Identification of a network of oscillatory interneurones. Journal of Experimental Biology, 75, 25–43.

    PubMed  CAS  Google Scholar 

  • Friesen, W. O. (1989). Neuronal control of leech swimming movements II. Motor neuron feedback to oscillator cells 115 and 28. Journal of Comparative Physiology A, 166, 205–215.

    Article  Google Scholar 

  • Friesen, W. O., & Stent, G. S. (1977). Generation of a locomotory rhythm by a neural network with recurrent cyclic inhibition. Biological Cybernetics, 28, 27–40.

    Article  PubMed  CAS  Google Scholar 

  • Futakata, Y., & Iwasaki, T. (2008). Formal analysis of resonance entrainment by central pattern generator. Journal of Mathematical Biology, 57(2),183–207.

    Article  PubMed  CAS  Google Scholar 

  • Glad, T., & Ljung, L. (2000). Control theory—multivariable and nonlinear methods. London: Taylor & Francis.

    Google Scholar 

  • Grillner, S., & Wallen, P. (2002). Cellular bases of a vertebrate locomotor system—steering, intersegmental and segmental co-ordination and sensory control. Brain Research Reviews, 40, 92–106.

    Article  PubMed  Google Scholar 

  • Guan, L., Kiemel, T., & Cohen, A. H. (2001). Impact of movement and movement-related feedback on the lamprey central pattern generator for locomotion. Journal of Experimental Biology, 204, 2361–2370.

    PubMed  CAS  Google Scholar 

  • Guckenheimer, J., & Holmes, P. (1983). Nonlinear oscillations, dynamical systems, and bifurcations of vector fields. New York: Springer.

    Google Scholar 

  • Hocker, C. G., Yu, X., & Friesen, W. O. (2000). Functionally heterogeneous segmental oscillators generate swimming in the medicinal leech. Journal of Comparative Physiology, 186(9), 871–883.

    Article  PubMed  CAS  Google Scholar 

  • Hoppensteadt, F. C., & Izhikevich, E. M. (1997). Weakly connected neural networks. New York: Springer.

    Google Scholar 

  • Iwasaki, T. (2006). Analysis and synthesis of central pattern generators via multivariable harmonic balance. American Control Conference (pp. 4106–4111). Minneapolis, MN (June 14–16).

  • Iwasaki, T. (2008). Multivariable harmonic balance for central pattern generators. Automatica (in press).

  • Iwasaki, T., & Zheng, M. (2006). Sensory feedback mechanism underlying entrainment of central pattern generator to mechanical resonance. Biological Cybernetics, 94(4), 245–261.

    Article  PubMed  CAS  Google Scholar 

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

    Google Scholar 

  • Khalil, H. K. (1996). Nonlinear systems. Englewood Cliffs: Prentice Hall.

    Google Scholar 

  • Kiemel, T., Gormley, K. M., Guan, L., Williams, T. L., & Cohen, A. H. (2003). Estimating the strength and direction of functional coupling in the lamprey spinal cord. Journal of Computational Neuroscience, 15, 233–245.

    Article  PubMed  Google Scholar 

  • Kristan, W. B. Jr., Calabrese, R., & Friesen, W. O. (2005). Neuronal control of leech behavior. Progress in Neurobiology, 76, 279–327.

    Article  PubMed  Google Scholar 

  • Koch, C., & Segev, I. (1989). Methods in neuronal modeling: From synapses to networks. Cambridge: MIT.

    Google Scholar 

  • Kotaleski, J. H., Grillner, S., & Lansner, A. (1999). Neural mechanisms potentially contributing to the intersegmental phase lag in lamprey. Biological Cybernetics, 81, 317–330.

    Article  PubMed  CAS  Google Scholar 

  • Linkens, D. A. (1974). Analytical solution of large numbers of mutually coupled nearly sinusoidal oscillators. IEEE Transactions on Circuits and Systems, 21(2), 294–300.

    Article  Google Scholar 

  • Matsuoka, K. (1985). Sustained oscillations generated by mutually inhibiting neurons with adaptation. Biological Cybernetics, 52, 367–376.

    Article  PubMed  CAS  Google Scholar 

  • Marder, E., & Calabrese, R. L. (1996). Principles of rhythmic motor pattern generation. Physiological Reviews, 76(3), 687–717.

    PubMed  CAS  Google Scholar 

  • McMillen, T., & Holmes, P. (2006). An elastic rod model for anguilliform swimming. Journal of Mathematical Biology, 53(5), 843–886.

    Article  PubMed  CAS  Google Scholar 

  • Mulloney, B., & Hall, W. M. (2007). Local and intersegmental interactions of coordinating neurons and local circuits in the swimmeret system. Journal of Neurophysiology, 98, 405–413.

    Article  PubMed  Google Scholar 

  • Mulloney, B., Harness, P. I., & Hall, W. M. (2006). Bursts of information: Coordinating interneurons encode multiple parameters of a periodic motor pattern. Journal of Neurophysiology, 95, 850–861.

    Article  PubMed  Google Scholar 

  • Nusbaum, M. P., Friesen, W. O., Kristan, W. B. Jr., & Pearce, R. A. (1987). Neural mechanisms generating the leech swimming rhythm: Swim-initiator neurons excite the network of swim oscillator neurons. Journal of Comparative Physiology A, 161, 355–366.

    Article  CAS  Google Scholar 

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

    Google Scholar 

  • Pearce, R. A., & Friesen, W. O. (1984). Intersegmental coordination of leech swimming: comparison of in situ and isolated nerve cord activity with body wall movement. Brain Research, 299, 363–366.

    Article  PubMed  CAS  Google Scholar 

  • Pearce, R. A., & Friesen, W. O. (1988). A model for intersegmental coordination in the leech nerve cord. Biological Cybernetics, 58, 301–311.

    Article  PubMed  CAS  Google Scholar 

  • Pearce, R. A., & Friesen, W. O. (1985a). Intersegmental coordination of the leech swimming rhythm. I. Roles of cycle period gradient and coupling strength. Journal of Neurophysiology, 54, 1444–1459.

    PubMed  CAS  Google Scholar 

  • Pearce, R. A., & Friesen, W. O. (1985b). Intersegmental coordination of the leech swimming rhythm. II. Comparison of long and short chains of ganglia. Journal of Neurophysiology, 54, 1460–1472.

    PubMed  CAS  Google Scholar 

  • Sigvardt, K. A., & Williams, T. L. (1996). Effects of local oscillator frequency on intersegmental coordination in the lamprey locomotor CPG: Theory and experiment. Journal of Neurophysiology, 76(6), 4094–4103.

    PubMed  CAS  Google Scholar 

  • Skinner, F. K., Kopell, N., & Mulloney, B. (1997). How does the crayfish swimmeret system work? Insights from nearest-neighbor coupled oscillator models. Journal of Computational Neuroscience, 4, 151–160.

    Article  PubMed  CAS  Google Scholar 

  • Skinner, F. K., & Mulloney, B. (1988). Intersegmental coordination of limb movements during locomotion: Mathematical models predict circuits that drive swimmeret beating. Journal of Neuroscience, 18, 3831–3842.

    Google Scholar 

  • Skinner, F. K., & Mulloney, B. (1998a). Intersegmental coordination in invertebrates and vertebrates. Current Opinion in Neurobiology, 8, 725–732.

    Article  PubMed  CAS  Google Scholar 

  • Skinner, F. K., & Mulloney, B. (1998b). Intersegmental coordination of limb movements during locomotion: Mathematical models predict circuits that drive swimmeret beating. Journal of Neuroscience, 18, 3831–3842.

    PubMed  CAS  Google Scholar 

  • Soffe, S. R., Zhao, F.-Y., & Roberts, A. (2001). Functional projection distances of spinal interneurons mediating reciprocal inhibition during swimming in xenopus tadpoles. European Journal of Neuroscience, 13, 617–627.

    Article  PubMed  CAS  Google Scholar 

  • Taylor, A., Cottrell, G. W., & Kristan, W. B. Jr. (2000). A model of the leech segmental swim central pattern generator. Neurocomputing, 32–33, 573–584.

    Article  Google Scholar 

  • Tesi, A., Abed, E. H., Genesio, R., & Wang, H. O. (1996). Harmonic balance analysis of period-doubling bifurcations with implications for control of nonlinear dynamics. Automatica, 32(9), 1255–1271.

    Article  Google Scholar 

  • Tschuluun, N., Hall, W. M., & Mulloney, B. (2001). Limb movements during locomotion: Tests of a model of an intersegmental coordinating circuit. Journal of Neuroscience, 21, 7859–7869.

    PubMed  CAS  Google Scholar 

  • Varkonyi, P. L., Kiemel, T., Hoffman, K., Cohen, A. H., & Holmes, P. (2008). On the derivation and tuning of phase oscillator models for lamprey central pattern generators. Journal of Computational Neuroscience. doi:10.1007/s10827-008-0076-8.

  • Verdaasdonk, B. W., Koopman, H. F., & Van der Helm, F. C. (2006). Energy efficient and robust rhythmic limb movement by central pattern generators. Neural Network, 19(4), 388–400.

    Article  CAS  Google Scholar 

  • Wallen, P., Ekeberg, O., Lansner, A., Brodin, L., Traven, H., & Grillner, S. (1992). A computer-based model for realistic simulations of neural networks. II. Simulation of the segmental network generating locomotor rhythmicity in the lamprey. Journal of Neurophysiology, 68, 1939–1950.

    PubMed  CAS  Google Scholar 

  • Weeks, J. C. (1982). Synaptic basis of swim initiation in the leech I Connections of a swim-initiating neuron (cell 204) with motor neurons and pattern-generating “oscillator” neurons. Journal of Comparative Physiology A, 148, 253–263.

    Article  Google Scholar 

  • Weeks, J. C., & Kristan, W. B. Jr. (1978). Initiation, maintenance and modulation of swimming in the medicinal leech by the activity of a single neuron. Journal of Experimental Biology, 77, 71–88.

    Google Scholar 

  • Williams, T. L. (1992). Phase coupling by synaptic spread in chains of coupled neuronal oscillators. Science, 258(5082), 662–665.

    Article  PubMed  CAS  Google Scholar 

  • Williamson, M. M. (1998). Neural control of rhythmic arm movements. Neural Networks, 11, 1379–1394.

    Article  PubMed  Google Scholar 

  • Zheng, M., Friesen, W. O., & Iwasaki, T. (2007). Systems-level modeling of neuronal circuits for leech swimming. Journal of Computational Neuroscience, 22(1), 21–38.

    Article  PubMed  CAS  Google Scholar 

  • Zheng, M., Iwasaki, T., & Friesen, W. O. (2004). Systems approach to modeling the neuronal CPG for leech swimming. In Proc. 26th annual international conference of the engineering in medicine and biology society (EMBC) (Vol. 1, pp. 703–706). (September)

Download references

Acknowledgements

This work is supported by NIH/NINDS 1 R01 NS46057-01, as part of the NSF/NIH Collaborative Research in Computational Neuroscience Program, and by NSF 0237708.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tetsuya Iwasaki.

Additional information

Action Editor: Nancy Kopell

Appendices

Appendix

1 Notation

The imaginary unit is denoted by \({\text{j}}\). Let e be a column vector with all entries being 1, and I is an identity matrix. A scalar function f acts elementwise when its argument is a matrix or vector. For instance, when X is a matrix, Y = f(X) means that Y kl  = f(X kl ) for all k and l where Y kl and X kl are the (k,l) entries of the matrices. The set of n dimensional real (complex) vectors is denoted by ℝn (ℂn). For a complex number x ∈ ℂ, the imaginary part is denoted by \(\Im(x)\), and \(\angle x\) is the polar angle of x in the complex plane, i.e., if \(x=a+{\text{j}} b\), then \(\angle x=\theta\) with tan(θ) = b/a. For a vector x ∈ ℂn, we define y = |x| and \(y=\angle x\) by y k  = |x k | and \(y_k=\angle x_k\), respectively, for k = 1,...,n, and \({\mbox{diag}}(x)\) is the diagonal matrix with entries x k . For a scalar x ∈ ℂ and a vector y ∈ ℂn, we write z = x + y to mean z k  = x + y k for k = 1,...,n. The Kronecker product of vectors (or matrices) x and y is denoted by x ⊗ y.

2 Formulae for describing functions

The following are formulae for describing functions of the threshold nonlinearity ϕ(x): =  max (x,0):

$$\begin{array}{lll} \kappa_1(\mathfrak{a},\mathfrak{b})&:=&\frac{1}{\mathfrak{a}\pi}\int_{-\pi}^\pi \phi(\mathfrak{a}[\sin t+\mathfrak{b}]) \sin t dt \\ &=&\left\{\begin{array}{ll} (1/\pi)[\mathfrak{b}\sqrt{1-\mathfrak{b}^2}+\pi/2+\sin^{-1}\mathfrak{b}], & (|\mathfrak{b}|\leq 1)\\ 1, & (\mathfrak{b}>1) \\ 0, & (\mathfrak{b}<-1) \end{array}\right. \\ \kappa_2(\mathfrak{a},\mathfrak{b})&:=& \frac{1}{2\mathfrak{a}\pi}\int_{-\pi}^\pi \phi(\mathfrak{a}[\sin t+\mathfrak{b}]) dt \\ &=&\left\{\begin{array}{ll} (1/\pi)[\mathfrak{b}(\pi/2+\sin^{-1}\mathfrak{b})+\sqrt{1-\mathfrak{b}^2}], & (|\mathfrak{b}|\leq 1)\\ \mathfrak{b}, & (\mathfrak{b}>1) \\ 0, & (\mathfrak{b}<-1) \end{array}\right. \\ \end{array} $$

3 The leech CPG model

A systems-level model for leech swimming CPG has been developed in Zheng et al. (2007). The following is a slightly modified version of the model, where pure delays are used for the intersegmental coupling dynamics instead of first-order time lags:

$$\begin{array}{rcl} v_k&=&{\beta}+\sum\limits_{l=1}^m{\text{M}}_{kl}(s)\phi(v_l), {} (k=1,\ldots,m), \\ {\text{M}}_{kl}(s)&=&\left\{\begin{array}{ll} \mu f(\tau_k s) M & (l=k), \\[4pt] {\epsilon} e^{-|k-l|\tau_ds}M_{kl} & (l\neq k), \end{array}\right. \\ f(s)&=&\frac{1-r}{1+(1-r)s}, {} \phi(x):=\max(x,0), \\ M&=&\left[ \begin{array}{rrr} 0 & -1 & 0 \\ 0 & 0 & -1 \\ -1 & 0 & 0 \end{array}\right], {} M_A=\left[ \begin{array}{rrr} 0 & -1 & 0 \\ 0 & 0 & -1 \\ 0 & 0 & 0 \end{array} \right], {}\\ M_D&=&\left[ \begin{array}{rrr} 2 & 0 & 0 \\ 0 & 0 & 0 \\ 0 & 0 & 0 \end{array}\right], \\ M_{kl}&=&\left \{\begin{array}{llll} M_A (k<l\leq k+q), \\[4pt] M_D (k-q\leq l<k), \\[4pt] 0 (\mbox{otherwise}), \end{array} \right. \end{array}$$
$$\begin{array}{l}{m = 17,\,\,\,\,\tau _o = 200\,\text{ms,}\,\,\,\,r = 0.3,} \\ \,\,{q = 5,\,\,\,\,\,\,\,\tau _d = 15\,\text{ms,}\,\,\,\,\,\,\,\mu = 6,} \\ {\qquad\qquad\,\,\,\beta = 30r\,\text{mV}\,\,\,\sigma: = {\varepsilon \mathord{\left/ {\vphantom {\varepsilon {\mu = 0.015.}}} \right. \kern-\nulldelimiterspace} {\mu = 0.015.}}} \end{array}$$

The nominal value τ o can be chosen equal to either one of the τ k values or their average value. Let Δτ o be a nominal size of the variations in time constants; for instance,

$$ {\Delta}\tau_o:=\left[\max\limits_k(\tau_k)-\min\limits_k(\tau_k)\right]/(m-1). $$

We assume that the variation is small, i.e., ε: = Δτ o /τ o  ≪ 1. To the first-order approximation in ε, we have

$$ f(\tau_k s)\approx f(\tau_o s)+{\varepsilon} g_k(s), g_k(s):=\frac{\tau_k-\tau_o}{{\Delta}\tau_o}\tau_o sf'(\tau_o s). $$
(20)

As a result, the leech swimming CPG can be put in the form of Eq. (4) with

$$ M_o(s)=f(\tau_o s)M,{} \Delta_{kl}(s)=\left\{\begin{array}{ll} ({\varepsilon}/\sigma) g_k( s) M & (l=k), \\[4PT] e^{-|k-l|\tau_ds} M_{kl} & (l\neq k), \end{array}\right. $$

In Eq. (20), note that τ k  − τ o is determined by the time constant gradient Δτ ∈ ℝm − 1 defined by its kth entry

$$ {\Delta}\tau_k:=\tau_{k+1}-\tau_k $$

for k = 1,...,m − 1. Thus, the shape of the time constant variation along the body is encoded into g k (s), while the magnitude of the variation is embedded in ε. For instance, if the nominal value τ o is chosen equal to τ p , then

$$ g_k(s)=\left(\sum\limits_{i=p}^{k-1}\zeta_i-\sum\limits_{i=k}^{p-1}\zeta_i\right) \tau_o sf'(\tau_o s) $$

where the summation from i = i 1 to i 2 is considered to be zero if i 1 > i 2, and ζ: = Δτ/Δτ o represents the shape of the time constant variation. If, in particular, we consider the uniformly increasing time constant gradient, we have ζ = e.

4 Derivation of the MHB analysis condition for the nominal segmental oscillator

Let (ω, a,b,ψ) be the oscillation profile of the nominal segmental oscillator Eq. (3) with ε = 0. We define \(\hbar:=Ae^{{\text{j}}\psi}\) the phasor of v(t) with \(A:={\mbox{diag}}(a)\). We have the so-called harmonic balance equations, with \(B:={\mbox{diag}}(b)\), shown as follows:

$$ 0 = \left[I- \mu M_o({\text{j}}\omega) \kappa_1(B)\right]\hbar, {} \hbar=Ae^{{\text{j}}\psi}, $$
(21)
$${\beta} = \left[ B- \mu M_o(0)\kappa_2(B)\right]a. $$
(22)

We see that the matrix \(\mu M_o({\text{j}}\omega) \kappa_1(B)\) has an eigenvalue 1 and an associated eigenvector \(\hbar\). For further analysis later, we denote the normalized left eigenvector by ℓ ∈ ℂn:

$$ 0 = \ell^*\left[I- \mu M_o({\text{j}}\omega) \kappa_1(B)\right], {} \ell^*\hbar=1. $$

Under the uniformity assumptions \(a=\mathfrak{a} e\) and \(b= \mathfrak{b} e\), Eqs. (21) and (22) reduce Eqs. (8) and (9), respectively, and ℓ becomes a left eigenvector of M.

5 Derivation of the MHB analysis condition for weakly coupled oscillators

In this section, we will derive condition Eq. (11). For the uncoupled oscillators, we have their MHB equations as follows:

$$\begin{array}{lll} 0&=&[I-\mu{{\cal M}}_o({\text{j}}\omega)\kappa_1({{\cal B}}_o)]{\text{h}}_o, {} {\text{h}}_o:=\alpha\otimes\hbar, {} |{\text{h}}_o|={\text{a}}_o \\ {\beta} &=&[{{\cal B}}_o-\mu{{\cal M}}_o(0)\kappa_2({{\cal B}}_o)]{\text{a}}_o, {} {{\cal B}}_o:={\rm diag}(B,\ldots,B) \end{array}$$

for an arbitrary vector α ∈ ℂm satisfying |α| = e. The MHB equations for the coupled oscillators are given by

$$\begin{array}{lll} 0&=&[I-{{\cal M}}({\text{j}}\varpi)\kappa_1({{\cal B}})]{\text{h}}, {} |{\text{h}}|={\text{a}} \\ {\beta} &=&[{{\cal B}}-{{\cal M}}(0)\kappa_2({{\cal B}})]{\text{a}}. \end{array}$$

Let the variables after the weak coupling be approximated by

$$ \varpi\approx\omega, {} {\text{h}}\approx {\text{h}}_o+{\epsilon}\tilde{{\text{h}}},{} {{\cal B}}\approx {{\cal B}}_o+{\epsilon}\tilde{{{\cal B}}} $$

where we assume that the frequency perturbation is small and can be neglected.

Then, noting that

$$ \kappa_i({{\cal B}})\approx\kappa_i({{\cal B}}_o)+{\epsilon}\kappa_i'({{\cal B}}_o)\tilde{{\cal B}}, $$

the first MHB equation for the coupled oscillator becomes

$$ \begin{array}{lll} &&{\kern-6pt}[I-\mu{{\cal M}}_o({\text{j}}\omega)\kappa_1({{\cal B}}_o)]\tilde{\text{h}}\\ &&{\kern4pt} = [\mu{{\cal M}}_o({\text{j}}\omega)\kappa_1'({{\cal B}}_o)\tilde{{\cal B}}+ \Delta({\text{j}}\omega)\kappa_1({{\cal B}}_o)] (\alpha\otimes\hbar), \end{array} $$
(23)

where we used the MHB equations for the uncoupled oscillators and neglected the O(ε 2) terms. By the definitions of L and H, we have

$$ L^*(I-\mu{{\cal M}}_o({\text{j}}\omega)\kappa_1({{\cal B}}_o))=0, {} L^*H=I. $$

Multiplying Eq. (23) by L * from the left,

$$ (L^*\Upsilon H-\nabla)\alpha=0, $$

where

$$ \Upsilon:=-\kappa_1({{\cal B}}_o)^{-1}\kappa_1'({{\cal B}}_o)\tilde{{\cal B}}, {} \nabla:=L^*\Delta({\text{j}}\omega)\kappa_1({{\cal B}}_o)H. $$

We consider a chain of oscillators where the intersegmental coupling is uniform over the chain. That is, the coupling matrix from the (k − l)th oscillator to the kth is the same as that from the kth oscillator to the (k + l)th, so that the overall intersegmental coupling matrix has the block-Toeplitz structure. In this case, the perturbation vectors for all oscillators may be assumed approximately parallel to each other, i.e.,

$$ \tilde{{{\cal B}}} \approx \Gamma\otimes\tilde{B}, $$

where Γ is a real diagonal matrix, and we have

$$ L^*\Upsilon H=\rho\Gamma, {} \rho:=-\ell^*\kappa_1(B)^{-1}\kappa_1'(B)\tilde B\hbar. $$

Thus, the MHB equation implies

$$ (\rho\Gamma-\nabla)\alpha=0. $$

6 Numerical algorithm for solving the MHB condition

We propose the following algorithm for solving Eq. (11):

  1. 0.

    Let ε 1, ε 2 be small positive numbers and initialize k = 1 and \(\gamma_k=[1\;\cdots\; 1]{^{\mbox{\tt T}}}/\sqrt{m}\in{\mathbb{R}}^m\).

  2. 1.

    Let \(\Gamma_k=\mbox{diag}(\gamma_k)\). Let (ρ k ,q k ) be the maximal eigenvalue/eigenvector pair of \(\nabla\Gamma_k^{-1}\) with ||q k || = 1.

  3. 2.

    Update \(\gamma_{k+1}:={\bar\gamma}_k/\|{\bar\gamma}_k\|\) where \({\bar\gamma}_k:=\gamma_k+{\epsilon}_1|q_k|\).

  4. 3.

    If ||γ k + 1 − γ k || ≤ ε 2, then stop; otherwise, increment k and go to step 1 to iterate.

From \((\rho_kI-\nabla\Gamma_k^{-1})q_k=0\), we have \((\rho_k\Gamma_k-\nabla)\Gamma_k^{-1}q_k=0\). Thus, the first equation in Eq. (11) holds for Γ = Γ k , ρ = ρ k , and \(\alpha=\Gamma_k^{-1}q_k\) at every step k. If the algorithm stops for a sufficiently small ε 2, we have γ k + 1 ≈ γ k , implying that γ k  ≈ |q k |. Hence, we have |α| ≈ e on exit, satisfying the second equation in Eq. (11) approximately.

7 MHB analysis of the intersegmental phase lag

We shall derive the formula Eq. (18) for the average value of the intersegmental phase lags for a general chain of weakly coupled oscillators, where the coupling structures M, M A , and M D are arbitrary. We first consider the case where the segmental oscillators are identical (no period gradient), the biases within each oscillator are uniform (\(b=\mathfrak{b} e\)), and the intersegmental coupling is uniform over the chain. Then, the basic coupling matrix \(\nabla\) is given by

$$ \nabla={\text{k}}\left[\begin{array}{cccccc} 0 & \nabla_1 & \cdots & \nabla_{{q_{\mbox{$A$}}}} && 0 \\[-1pt] \nabla_{-1} & 0 & \nabla_1 & \ddots & \ddots & \\[-1pt] \vdots & \nabla_{-1} & 0 & \nabla_1 & \ddots & \nabla_{{q_{\mbox{$A$}}}} \\[-1pt] \nabla_{-{q_{\mbox{$D$}}}} & \ddots & \nabla_{-1} & 0 & \ddots & \vdots \\[-1pt] & \ddots & \ddots & \ddots & \ddots & \nabla_1 \\[-1pt] 0 && \nabla_{-{q_{\mbox{$D$}}}} & \cdots & \nabla_{-1} & 0 \end{array}\right] $$

where \({q_{\mbox{$A$}}}\) and \({q_{\mbox{$D$}}}\) are the intersegmental projection spans in the ascending and descending directions, respectively; \({\text{k}}\) is defined by \({\text{k}}:=\kappa_1(\mathfrak{b})\); and

$$ \nabla_{ k}:={r_{\mbox{$A$}}} e^{{\text{j}}(\eta_A-k\omega\tau_d)}, {} {r_{\mbox{$A$}}} e^{{\text{j}}\eta_A}:=\ell^*M_A\hbar,\\ \nabla_{-k}:={r_{\mbox{$D$}}} e^{{\text{j}}(\eta_D-k\omega\tau_d)}, {} {r_{\mbox{$D$}}} e^{{\text{j}}\eta_D}:=\ell^*M_D\hbar, $$

for integers k ≥ 1. For the leech CPG, we have \({r_{\mbox{$A$}}}={r_{\mbox{$D$}}}=2/3\), η A  = π/3, and η D  = 0. Note that \(\nabla\) is m×m, where m is the number of segmental oscillators. To estimate the average phase lag, let us consider the limiting case where the oscillator chain is infinitely long (m = ∞ ). In this case, a generic row of the MHB equation \((\rho I-\nabla)\alpha=0\) takes the following form:

$$ \sum\limits_{k=-{q_{\mbox{$D$}}}}^{{q_{\mbox{$A$}}}}\nabla_k\alpha_k=0, {} \nabla_0:=-\rho/{\text{k}}. $$
(24)

Due to the uniformity of the intersegmental connections over the chain (that makes \(\nabla\) a Toeplitz matrix), the eigenvector α has the structure such that α k + 1/α k is constant over k. Consequently, the intersegmental phase lag is uniform over the chain, and we may let

$$ \alpha_k=re^{-{\text{j}} k\eta_o} $$
(25)

where η o is the intersegmental phase lag per segment. Substituting Eq. (25) into Eq. (24) and solving for ρ, we have

$$ \rho={\text{k}}{r_{\mbox{$A$}}}\sum\limits_{k=1}^{{q_{\mbox{$A$}}}}e^{{\text{j}}(\eta_A-k(\eta_o+\omega\tau_d))} +{\text{k}}{r_{\mbox{$D$}}}\sum\limits_{k=1}^{{q_{\mbox{$D$}}}}e^{{\text{j}}(\eta_D+k(\eta_o-\omega\tau_d))}.\\[-6pt] $$
(26)

This is a parametrization of the set of infinitely many eigenvalues of the infinite matrix \(\nabla\) in terms of η o  ∈ ℝ. The profile of a stable oscillation can be estimated from the maximal eigenvalue, and hence, we are interested in finding η o such that \(f(\eta_o):=\Re[\rho]\) takes its maximum value. Taking the derivative and setting it to zero, we have

$$\begin{array}{lll} f'(\eta_o)/{\text{k}} & = & {r_{\mbox{$A$}}}\sum\limits_{k=1}^{{q_{\mbox{$A$}}}}k\sin(\eta_A-k(\eta_o+\omega\tau_d))\\[-2pt] &&- {r_{\mbox{$D$}}}\sum\limits_{k=1}^{{q_{\mbox{$D$}}}}k\sin(\eta_D+k(\eta_o-\omega\tau_d)) \\[-2pt] & \approx & \left[ {r_{\mbox{$A$}}}\sum\limits_{k=1}^{{q_{\mbox{$A$}}}}(k\eta_A-k^2\omega\tau_d)- {r_{\mbox{$D$}}}\sum\limits_{k=1}^{{q_{\mbox{$D$}}}}(k\eta_D-k^2\omega\tau_d) \right]\\[-2pt] &&- \left[ {r_{\mbox{$A$}}}\sum\limits_{k=1}^{{q_{\mbox{$A$}}}}k^2+ {r_{\mbox{$D$}}}\sum\limits_{k=1}^{{q_{\mbox{$D$}}}}k^2 \right]\eta_o=0 \end{array}$$

where we used the approximation sin(x) ≈ x for small x. Solving this equation for η o , we obtain the formula in Eq. (18). It can be verified that the η o thus found approximates the global maximizer of f(η o ).

Next, we consider the case where the intrinsic periods are not uniform. Assume that each segmental oscillator has the nominal time constant τ o except that the pth segment has the time constant τ p  ≠ τ o . With this change, the basic coupling matrix, denoted by \(\bar \nabla\), is slightly modified from \(\nabla\) by

$$ \bar\nabla=\nabla+\xi_pe_pe_p{^{\mbox{\tt T}}} $$

where e p is the vector having zero entries except for the pth entry, which is one, and ξ p is defined using g k (s) in Eq. (20) as

$$ \xi_p:={\text{k}}({\varepsilon}/\sigma)(\ell^*M\hbar)g_p({\text{j}}\omega). $$
(27)

Note that \(\ell^*M\hbar=e^{{\text{j}}\pi/3}\) for the leech CPG. Now, the MHB equation becomes \((\bar\rho I-\bar\nabla)\bar\alpha=0\). We assume that the effect of the time constant perturbation is local in the sense that \(\bar\alpha\approx\alpha+e_p\delta_p\) for some δ p  ∈ ℂ, where α is the maximal eigenvector of \(\nabla\) satisfying \(\nabla\alpha=\rho\alpha\) for the eigenvalue ρ defined in Eq. (26). We further assume that the perturbation in the eigenvalue is small; \(\bar\rho\approx\rho\). With these approximations, we have

$$ (\rho I-\nabla-\xi_pe_pe_p{^{\mbox{\tt T}}})(\alpha+e_p\delta_p)\approx0. $$
(28)

Multiplying the equation by \(e_p{^{\mbox{\tt T}}}\) from the left and using \(\nabla\alpha=\rho\alpha\), we obtain

$$ \delta_p\approx\varsigma_p\alpha_p, {} \varsigma_p:=\xi_p/\rho $$

where |ς p | is assumed small. As a result, we have

$$ \theta_p=-p\eta_o+\Im(\varsigma_p), {} \theta_k=-k\eta_o \mbox{~~for~}k\neq p. $$
(29)

To derive a formula for the phase lags in the case of an arbitrary period gradient, we consider the time constant perturbation in a single oscillator as before and start with Eq. (28). Taking the summation of Eq. (28) over p = 1,..., ∞, we have

$$ (\rho I-\nabla-\Xi)\delta-\Xi\alpha\approx0 $$

where Ξ is the (infinite) diagonal matrix having entries ξ p with p = 1,... ∞, and δ and α are (infinite) vectors with entries δ p and α p , respectively. This equation implies

$$ (\rho I-\nabla-\Xi)(\alpha+\delta)\approx(\rho I-\nabla)\alpha=0. $$

Hence, if the coupling matrix \(\nabla\) is perturbed to \(\nabla+\Xi\) due to the period gradient, the resulting phases are perturbed from \(\angle\alpha\) to \(\angle(\alpha+\delta)\) or \(\angle\alpha+\Im(\varsigma)\). In summary, the intersegmental phases for general period gradients, perturbed from a nominal τ o , are estimated as

$$ \theta_k=-k\eta_o+\Im(\varsigma_k), {} \varsigma_k:=\xi_k/\rho, $$

for all k where η o is given by Eq. (16) and ρ and ξ k are defined by Eqs. (26) and (27), respectively. Using the definition of g k (s), the intersegmental phase lags are given by

$$ \begin{array}{lll} \eta_k&=&\eta_o+({\varepsilon}/\sigma)({\Delta}\tau_k/{\Delta}\tau_o)\varphi, \\ \varphi&:=&-\Im[{\text{k}}(\ell^*M\hbar){\text{j}}\tau_o\omega f'({\text{j}}\tau_o\omega)/\rho] \end{array} $$
(30)

where φ depends on the communication delay τ d through ρ.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Chen, Z., Zheng, M., Friesen, W.O. et al. Multivariable harmonic balance analysis of the neuronal oscillator for leech swimming. J Comput Neurosci 25, 583–606 (2008). https://doi.org/10.1007/s10827-008-0105-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-008-0105-7

Keywords

Navigation