1:
2: %
3: \documentclass[aps,twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
4: %
5: %
6: %
7: \usepackage{colordvi}
8: %
9: %
10: \input{epsf}
11:
12: \def\bc{\begin{center}}
13: \def\ec{\end{center}}
14:
15:
16:
17: \def\be{\begin{equation}}
18: \def\ee{\end{equation}}
19:
20: \def\bi{\begin{itemize}}
21: \def\ei{\end{itemize}}
22:
23: \def\bea{\begin{eqnarray}}
24: \def\eea{\end{eqnarray}}
25:
26: \def\bR{{\sf R}}
27: \def\bv{{\bf v}}
28: \def\br{{\bf r}}
29: \def\bs{{\bf s}}
30: \def\bV{{\bf V}}
31: \def\bE{{\bf E}}
32: \def\bL{{\bf L}}
33: \def\bC{{\bf C}}
34: \def\bK{{\sf K}}
35: \def\bQ{{\sf Q}}
36: \def\bA{{\sf A}}
37: \def\bB{{\sf B}}
38: \def\bW{{\sf W}}
39: \def\bJ{{\sf J}}
40: \def\bH{{\sf H}}
41: \def\bT{{\sf T}}
42: \def\i{{\it i}}
43: \def\ri{{\rm i}}
44: \def\cc{\it c.c.}
45: \def\Re {{\it Re}}
46: \def\Im{{\it Im}}
47: \def\bI{{\bf I}}
48: \def\bJall{{\cal J}}
49: \def\bWall{{\cal W}}
50: \def\cH{{\cal H}}
51:
52: \def\wasred{}
53: \def\wasblue{}
54:
55:
56: \begin{document}
57: \title{Mathematical analysis and simulations of the neural circuit for
58: %
59: locomotion in lamprey}
60: \author{Li Zhaoping$^1$, Alex Lewis$^1$, and Silvia Scarpetta$^2$}
61: \affiliation{$^1$University College, London, UK\\
62: $^2$INFM \& Dept.of Physics, University of Salerno, Italy}
63: \begin{abstract}
64: %
65: %
66: %
67: %
68: %
69: %
70: %
71: %
72: %
73: %
74: %
75: %
76: %
77: %
78: %
79: %
80: We analyze the dynamics of the neural circuit of the
81: {lamprey} central pattern generator (CPG)
82: %
83: This analysis provides insights into how neural interactions
84: form oscillators and enable spontaneous oscillations in a network
85: of damped oscillators, which were not apparent in previous simulations
86: or abstract phase oscillator models.
87: We also show how the different behaviour regimes (characterized by phase and amplitude
88: relationships between oscillators) of forward/backward swimming,
89: and turning, can be controlled using the neural connection
90: strengths and external inputs.
91:
92: \pacs{{87.19.La, 87.19.St}}
93: \end{abstract}
94: \maketitle
95:
96: %
97: \wasred{Locomotion in verbebrates {(walking},
98: swimming, etc.) is}
99: generated by central pattern generators (CPGs) in the spinal cord.
100: The CPG for {swimming in lamprey}
101: is \wasred{one of the best known
102: \cite{Grillner2002,Buchanan2001}}, and
103: has been a model system for
104: %
105: \wasred{investigations.}
106: {It}
107: produces left-right {anti-phase} oscillatory neural and motor activities
108: propagating along a body composed of around 100 segments.
109: \wasred{A head-to-tail negative or positive oscillation phase
110: gradient,
111: of about
112: 1\% of an oscillation cycle per segment, gives forward or background
113: swimming respectively, and
114: one wavelength from head to tail.}
115: %
116: %
117: %
118: %
119: \wasred{External inputs from the brain stem {}{switch} the CPG between
120: %
121: } forward and backward swimming of
122: various speeds and turning.
123: Since {isolated sections} of the spinal cord,
124: \wasred{ down to 2-3 segments long \cite{Buchanan2001}},
125: can produce swimming-like {activity},
126: %
127: the
128: oscillations are thought to be generated by the neurons
129: within the \wasred{CPG}.
130: %
131: %
132: The neural circuit responsible is shown topologically
133: in Fig. \ref {fig:circuit}. It has
134: \wasred{ipsilaterally projecting
135: %
136: excitatory ($\bE$) neurons and
137: inhibitory ($\bL$) neurons,
138: and contralaterally projecting
139: %
140: inhibitory $(\bC$) neurons, and
141: provides output to motor neurons via the $\bE$ neurons.}
142: %
143: {All {neurons} project both intra- and inter-segmentally.
144: %
145: %
146: \wasred{The projection distances are {}{mainly} within a few
147: segments, especially from $\bE$ and $\bC$ neurons, and
148: are longer, and possibly stronger, in the head-to-tail
149: or descending direction} \cite{Grillner2002}-\cite{McClellan1998}.
150: \begin{figure}[t]
151: \setlength{\unitlength}{0.6pt}
152: \begin{center}
153: %
154: %
155: %
156: \mbox{\epsfxsize=250pt \epsfbox{fig1.eps}}
157: \end{center}
158: \caption{\label {fig:circuit} The lamprey CPG circuit. The solid and dashed lines
159: denote intra- and inter-segment connections respectively. }
160: \end{figure}
161: %
162:
163: %
164: %
165: Previous analytical work \cite{CohenEtal82, E&K}
166: {mainly} treated the CPG as a chain of coupled phase oscillators
167: \wasred{in a general form $\dot\theta_i = \omega_i +
168: \sum_jf_{ij}(\theta_i,\theta_j)$.
169: Here $\theta_i$ is oscillation phase and $\omega_i$
170: {is} intrinsic frequency,
171: modelling the
172: behaviour of {one} segment, and
173: $f_{ij}(\theta_i,\theta_j)$
174: {models} inter-segmental coupling.}
175: This approach \wasred{provided important insights into
176: the conditions {for}
177: %
178: phase-locked solutions
179: applicable} to \wasred{various} systems of coupled oscillators.
180: However,
181: %
182: \wasred{its generality} obscures the
183: %
184: {roles of specific} neural types and their
185: \wasred{connections}
186: in generating and controlling behaviour.
187: \wasblue{More recently, %
188: {bifurcation} analysis of the dynamics
189: of a single segment was carried out,} for a
190: phase oscillator model derived from a kinetic (Hodgkin Huxley)
191: equation for neurons \cite{Taylor1998}, and
192: for a neural circuit model similar to the one used in this paper
193: \cite{Jung}.}
194: {Extensive simulations}, including all neural types
195: {}{and detailed neural} properties,
196: have reproduced many features of experimental data \cite{Grillner2002},
197: though the {model's} complexity limits further
198: understanding.
199:
200:
201: In all previous approaches, it is assumed that a single segment in the
202: {}{CPG can oscillate} spontaneously, contrary to experimental evidence
203: that at least 2-3 segments are needed for
204: oscillations \cite{Buchanan2001}.
205: We present \wasred{an} analytical study, confirmed by simulations, of
206: a model of the CPG neural circuit in which {}{an isolated} single
207: segment
208: has a stable
209: fixed point, with spontaneous oscillations occurring only in
210: chains of coupled segments.
211: The phase oscillator approach is not
212: applicable here since it assumes spontaneously {{oscillating
213: individual segments perturbed by inter-segment coupling}.
214: Including specific cell types and their {connections} enables us to
215: analyse the role of each of them in generating and controlling
216: swimming.
217: We
218: show how external inputs select forward
219: and backward swimming, by {}{controlling
220: the relative strengths of connections
221: between various neurons}, and produces
222: turning, by additional input to one
223: side of the {CPG} only.
224: We also {analyse} behaviour near the {body ends}.
225:
226:
227: We model the CPG neural circuit
228: %
229: \wasred{with} $N = 100$ segments denoted by {$i=1,
230: .,N$.}
231: The vector states
232: \wasred{($\bE_l,\bL_l,\bC_l$) and
233: \wasred{$(\bE_r,\bL_r,\bC_r)$}, modelling the membrane potentials of the local populations of neurons \wasred{at the left and
234: right side of the body respectively},
235: with $\bE \wasred{_l} = (E^1 \wasred{_l},E^2 \wasred{_l},\cdots,E^N \wasred{_l})$ etc.,
236: are modelled as leaky integrators of their inputs:}
237: \bea
238: \dot \bE_{ \wasred{l}} &=& - \bE_{ \wasred{l}} - \bK^0 g_C(\bC_{ \wasred{r}}) + \bJ^0
239: g_E(\bE_{ \wasred{l}}) +\bI_{E, \wasred{l}}
240: \nonumber \\
241: \dot \bL_{ \wasred{l}} &=& - \bL_{ \wasred{l}} - \bA^0 g_C(\bC_{ \wasred{r}}) + \bW^0
242: g_E(\bE_{ \wasred{l}}) +\bI_{L, \wasred{l}}
243: {\label {uno}}\\
244: \dot \bC_{ \wasred{l}} &=& - \bC_{ \wasred{l}} - \bB^0 g_C(\bC_{ \wasred{r}}) + \bQ^0 g_E(\bE_{ \wasred{l}}) -
245: \bH^0 g_L(\bL_{ \wasred{l}}) +\bI_{C, \wasred{l}} \nonumber
246: \eea
247: with the same equation for
248: swapped subscripts
249: {$(l\leftrightarrow
250: r)$.}
251: {$g_E(\bE_l) = (g_E(E^1_l),\dots,g_E(E^N_l))$}
252: are the neural activities or firing rates,
253: %
254: \wasred{as non-negative (sigmoid-like) activation functions of $\bE_l$,}
255: and likewise for \wasred{$g_C(\bC_l)$ and $g_L(\bL_l)$}. $\bK^0$,
256: $\bJ^0$,
257: $\bA^0$, $\bW^0$, $\bB^0$, $\bQ^0$, and $\bH^0$
258: are
259: $N\times N$
260: matrices \wasred{of} non-negative elements
261: %
262: \wasred{modeling} the synaptic strengths between neurons.
263: %
264: %
265: $\bI_{E, \wasred{l}}$, $\bI_{L, \wasred{l}}$, and $\bI_{C, \wasred{l}}$ are
266: external inputs, \wasred{including those from the brain stem},
267: assumed to be {static}.
268: %
269: \wasred{A left-right symmetric
270: fixed point $(\bar\bE,\bar\bL,\bar\bC)$ where
271: $(\dot\bE,\dot\bL,\dot\bC)=0$ exists by setting
272: external inputs to $\bI_{E,l} = \bar \bE_l + \bK^0g_C(\bar \bC_{r})
273: - \bJ^0 g_E(\bar \bE_{l})$ (and analogously for other $\bI$'s).
274: %
275: %
276: Dynamics for small deviations from $(\bar\bE,\bar\bL,\bar\bC)$
277: can be approximated linearly,
278: {and,}
279: %
280: with a coordinate rotation
281: {$(\bE_\pm,\bL_\pm,\bC_\pm) {\equiv }
282: [(\bE_l,\bL_l,\bC_l) -(\bar\bE,\bar\bL,\bar\bC)]
283: \pm
284: [(\bE_r,\bL_r,\bC_r)-(\bar\bE,\bar\bL,\bar\bC)]$,}
285: transformed into two decoupled modes --
286: the left-right synchronous mode
287: $(\bE_+, \bL_+, \bC_+)$
288: and the antiphase mode $(\bE_-, \bL_-, \bC_-)$}.
289: Swimming requires oscillations, with wavelength of one body length, in
290: the anti-phase mode while the synchronous mode is damped. The
291: linearised equations are
292: \bea
293: \dot \bE_\pm &=& - \bE_\pm \mp \bK \bC_\pm + \bJ \bE_\pm \nonumber \\
294: \dot \bL_\pm &=& -\bL_\pm \mp \bA \bC_\pm + \bW \bE_\pm \nonumber \\
295: \dot \bC_\pm &=& - \bC_\pm \mp \bB \bC_\pm + \bQ\bE_\pm -\bH
296: \bL_\pm \label{plus_minus} \eea
297: where $\bK \equiv \bK^0 g'_C(\bar \bC )$,
298: $\bA \equiv \bA^0 g'_C(\bar \bC )$, $\bB \equiv \bB^0 g'_C(\bar \bC )$,
299: $\bJ \equiv \bJ^0 g'_E(\bar \bE )$,
300: $\bW \equiv \bW^0 g'_E(\bar \bE )$, $\bQ \equiv \bQ^0 g'_E(\bar \bE )$, and
301: $\bH \equiv \bH^0 g'_L(\bar \bL)$
302: are effective connection matrices,
303: \wasred{and
304: the $g'(.)$'s denote derivatives}.
305: The $\bC$ neurons thus become effectively excitatory in the
306: \wasred{anti-phase mode.}
307: Noting that the lengths of the neural connections are much
308: shorter {than}
309: %
310: the body,
311: \wasred{and that
312: isolated sections of spinal cord from any part
313: of the body generate oscillations with similar amplitude
314: and phase relationships \cite{Grillner2002,Buchanan2001}},
315: we make the approximation of translation invariance,
316: so that matrix elements such as $\bJ_{ij}$
317: %
318: depend only on ($i-j$),
319: and impose the periodic boundary condition,
320: %
321: {$\bJ_{ij}= \bJ(x)$,} where $x= (i-j) \bmod N$. This is adequate
322: \wasred{when
323: behaviour near body ends is not considered.}
324: Then all connection matrices {}{commute with each other},
325: with common eigenvectors (expressed as functions of segment number $x$)
326: \wasred{$(\bE(x), \bL(x), \bC(x)) \propto e^{i (2\pi m/N)x}$ for
327: {integer} eigenmode $-N/2< m \le N/2$.
328: The system solutions are thus combinations of modes
329: $(\bE_\pm (x,t), \bL_\pm (x,t), \bC_\pm (x,t)) \propto
330: e^{\lambda^\pm _m t + i (2\pi m/N)x}$ where $\lambda^\pm _m$ is
331: eigenvalue of eq. (\ref{plus_minus})
332: for mode $m$. Forward swimming results if
333: the real part $\Re(\lambda^\pm _m ) < 0$ for all modes except
334: the antiphase mode with $m=1$, ie $Re(\lambda_1^-)>0$. Then this
335: mode dominates the
336: solution (whose growing amplitude
337: will be constrained by nonlinearity)
338: $(\bE (x,t), \bL (x,t), \bC (x,t))
339: \propto e^{ \Re (\lambda^-_1 ) t -i (\omega t - kx)}$,
340: with oscillation frequency $\omega \equiv \left|\Im (\lambda^-_1)\right|$ and
341: wave number $k = 2\pi /N$.
342: Using the convention $e^{-i\omega t}$ for oscillations, we omitted
343: the solution $\propto e^{\Re(\lambda^-_1 ) t +i \omega t}$ in
344: the conjugate pair of eigenvalues.}
345: %
346: %
347: %
348: %
349: %
350: %
351: %
352: To simplify our system, {we} note from experimental data that in forward swimming,
353: {$\bE$} and {$\bL$} oscillate roughly in phase within {a}
354: segment, while {$\bC$}
355: leads them
356: %
357: \cite{Buchanan2001}.
358: We scale our \wasred{variable definitions} {so that $\bE_-=\bL_-$
359: in forward swimming. Then }
360: {eq. (\ref
361: {plus_minus})} implies
362: %
363: %
364: %
365: that
366: %
367: $(\bK -\bA ) \bC_- = -(\bJ - \bW ) \bE_-$ in forward swimming.
368: \wasred{Since $\bE$ and $\bC$ have much shorter connections than
369: the wavelength of oscillations during swimming, the connection
370: matrices have zero elements far from the diagonal, making
371: $(\bK -\bA ) \bC_-$ and $(\bJ - \bW ) \bE_-$ roughly either in
372: phase or in
373: anti-phase
374: with $\bC_-$ and $\bE_-$ respectively.
375: As $\bC_-$ phase leads $\bE_-$, $(\bK -\bA ) \bC_- = -(\bJ - \bW ) \bE_-$
376: is impossible unless $(\bJ -\bW)\bE_-=(\bK - \bA)\bC_-=0$.}
377: For simplicity we henceforth assume $\bJ =\bW$ and $\bK =
378: \bA$, since non-swimming modes do not concern us. Consequently
379: $\bE_\pm=\bL_\pm$ and
380: %
381: \be
382: \begin{pmatrix}
383: \dot\bE_\pm \\
384: \dot\bC_\pm \\
385: \end{pmatrix}
386: =
387: \begin{pmatrix}
388: \bJ-1 &~~& \mp\bK \\
389: -(\bH-\bQ) &~~& \mp\bB-1
390: \end{pmatrix}
391: \begin{pmatrix}
392: \bE_\pm \\
393: \bC_\pm\\
394: \end{pmatrix}
395: \label {eqv} \ee
396: \wasred{where $\bL$ and $\bE$ {are} treated as a single population
397: inhibiting or exciting $\bC$ via connections $\bH-\bQ$.}
398: %
399: {The eigenvalues for mode $m$ are}
400: {
401: \begin{align}
402: %
403: \lambda^+_m &= \left[-2 + J_m - B_m \pm \sqrt{R_m + 2(B_m^2+J_m^2)}\right]/2
404: \nonumber \\
405: \lambda^-_m &=
406: \left[-2 + J_m + B_m - i \sqrt{R_m }\right]/2.
407: %
408: \label {lams}
409: \end{align}}
410: {$J_m \equiv \sum_x \bJ (x) e^{ -\ri (2\pi m/N) x} $} is the
411: eigenvalue of $\bJ$ (and analogously for other matrices),
412: and {}{$R_m$} is the eigenvalue {of}
413: %
414: $\bR\equiv 4\bK (\bH-\bQ) - (\bB-\bJ)^2$.
415: %
416: %
417: %
418: %
419: %
420: %
421: %
422: %
423:
424: %
425: %
426: %
427: %
428:
429: To elucidate {}{the}
430: conditions {}{needed} for {}{the} {antiphase}
431: %
432: {mode with}
433: $m=\pm 1$
434: for forward or backward swimming to dominate, we {}{analyse the
435: bifurcations which occur
436: as $\lambda^\pm_m$ for each mode ($m$, $\pm$)
437: changes as the effective neural connections are varied, either directly or via
438: the external inputs.}
439: First, we focus on the left-right mode space
440: {(as in \cite{Taylor1998,Jung}
441: %
442: for a single segment)}
443: of $+$ and $-$, i.e., the synchronous
444: and antiphase modes, by simply taking $m =0$.
445: Then, $J_0$, $B_0$,
446: $H_0$, $K_0$, and $Q_0$,
447: are all real and
448: non-negative, each {}{being} the total connection
449: strength on a postsynaptic cell from all cells of a particular
450: type.
451: Oscillation in the antiphase mode requires $R_0>0$, necessitating
452: $H_0>Q_0$, or that in the AC component of interactions above the background
453: DC level, $\bC$ neurons receive stronger inhibition from $\bL$ neurons
454: than excitation from $\bE$ neurons.
455: Consequently, $\lambda^+_0$ is real and
456: the synchronous mode is non-oscillatory.
457: As neural connections increase from zero,
458: the {antiphase}
459: %
460: mode undergoes a Hopf bifurcation
461: when {$\Re (\lambda^-_0)=0$},
462: %
463: at {}{$J_0+B_0=2$},
464: {}{and} the {synchronous}
465: %
466: mode undergoes a
467: pitchfork bifurcation when {}{$\lambda^+_0=0$, which occurs when
468: $(B_0+1)(1-J_0) = K_0(H_0 - Q_0)$}.
469: {Oscillations result} if
470: the Hopf bifurcation has occurred but the
471: pitchfork bifurcation has not,
472: i.e., $Re(\lambda^-_0)>0> \lambda^+_0$. The condition
473: $\lambda^+_0<0$ implies
474: $(B_0+1)(1-J_0) > K_0(H_0
475: - Q_0)$, necessitating $J_0<1$. Meanwhile,
476: $Re(\lambda^-_0)>\lambda^+_0$
477: leads to
478: $B_0 > \sqrt {J_0^2 +R_0/2}>J_0$, meaning that there must be
479: sufficient {inhibitory}
480: connections between left and right $\bC$ cells.
481: %
482: %
483: %
484: %
485: %
486: The $\bJ$, $\bW$, and
487: $\bQ$ {connections from} $\bE$ cells have to be relatively
488: weak,
489: consistent with the findings {of \cite{Jung}}.
490: %
491: (If $R_0 < 0$, the antiphase mode will undergo {}{a pitchfork}
492: bifurcation, {}{and the synchronous {mode}
493: either a pitchfork or Hopf bifurcation}.
494: These {}{regimes} are
495: less relevant to modeling the {lamprey}.)
496: %
497:
498: Assuming the synchronous mode {}{is damped},
499: we focus now on {the} antiphase mode in the $m$ mode space.
500: {}{Hopf bifurcations occur} sequentially in various modes $m$
501: in the order of descending $\Re (\lambda^-_m)$.
502: Taylor expanding
503: $J_m$ (and similarly {$B_m$, $R_m$}) for small
504: wave number $k = 2\pi m/N$ {}{as is relevant
505: for {swimming},
506: %
507: $J_m = j_0 -i k j_1 - k^2 j_2 + {\mathcal{O} }
508: (k^3)$
509: with $ j_n = \sum_x \bJ(x)\frac{x^n}{n!}$}, we have
510: \bea
511: \!2\!\Re {(\!\lambda^-(k))\!} &=& \! - \!2 \! +\! j_0 \! +\!\! b_0\!\!-\!\! k
512: r_1/(2\sqrt{r_0})\! -\!\! k^2\!
513: (j_2\!\!+\!b_2)\!\! + \!{\mathcal{O}}\!(k^3\!) \nonumber \, \\
514: \!2\!\Im {(\!\lambda^-(k))}\! &=& \sqrt{r_0} + {\mathcal{O} } (k)
515: \eea
516: making the mode with
517: {}{$k \approx -r_1/[4\sqrt{r_0} (j_2+b_2)] $}, {}{which has the largest
518: $Re(\lambda^-(k))$,} dominant.
519: From the definition, {}{$(r_0,j_2, b_2) \ge 0$}, {}{while}
520: stronger and/or longer connections in the descending
521: direction {}{imply $(j_1,b_1)>0$}.
522: Simply, $\bJ$ (and similarly for other matrices)
523: is said to be descending (or ascending), if {}{$j_1>0$ (or $j_1<0$).}
524: Hence, {}{if} $\bR$ {}{is ascending}, i.e., {}{$r_1<0$},
525: $\Re ({\lambda^- }(k))$ {}{increases with} increasing $k$, and the
526: dominant wave number can be set to $k=2\pi/{N}$ for mode $m=1$
527: by {tuning} the values of $\bR$, $\bJ$, and $\bB$.
528: If the connection strengths are such that
529: only the $m=1$ mode undergoes the Hopf bifurcation,
530: forward swimming emerges spontaneously.
531: Switching $\bR$ to descending {leads to} backward swimming.
532: Note that $\bJ$, $\bB$, $\bH$, $\bK$, and $\bQ$ are all descending,
533: {multiplications} and summations of {}{descending} connections
534: are still {}{descending}, {and} negating a descending
535: connection makes
536: it ascending.
537: Since $\bB$ and $\bH$ have to dominate $\bJ$ and $\bQ$ respectively,
538: $\bR$ is composed of an ascending term $-(\bB-\bJ)^2$
539: and a descending term $4\bK(\bH-\bQ)$.
540: {}{Depending on} the relative strengths of {}{these two terms},
541: $\bR $ can be {}{made} ascending or descending to achieve forward
542: or backward swimming.
543: This could be achieved by changing the
544: {static inputs}
545: %
546: to shift the fixed point $(\bar{\bE},\bar{\bL},\bar{\bC})$ of the
547: system to a different gain regime
548: $g'_E(\bar \bE), g'_L(\bar \bL), g'_C(\bar \bC)$, and
549: thus different effective connection strengths $\bH =
550: \bH^0g'_L(\bar \bL)$, etc. without
551: changing the underlying connection structure $\bH^0$.
552: Alternatively, the \wasred{external} inputs {might} recruit
553: %
554: %
555: {extra functional cells} to alter the {effective}
556: %
557: connection strengths
558: \cite{Kozlov2002}.
559:
560: When connections are such that additional modes
561: satisfy $\Re (\lambda^-_m)>0$, the {}{resulting behavior
562: depends on the nonlinear coupling between modes}.
563: For illustration, consider nonlinearity
564: only in $g_C(\bC)$.
565: \bea
566: \dot \bE_{\pm} &=& - \bE_{\pm} \mp \bK^0 g_\pm(\bC) + \bJ \bE_{\pm} \nonumber \\
567: \dot \bL_{\pm} &=& -\bL_{\pm} \mp \bA^0 g_\pm(\bC) + \bW \bE_{\pm} \nonumber \\
568: \dot \bC_{\pm} &=& - \bC_{\pm} \mp \bB^0 g_\pm(\bC) +
569: \bK'\bE_{\pm} -\bH \bL_{\pm}
570: \eea
571: where $g_\pm(\bC) = g_C(\bC_l) \pm
572: g_C(\bC_r)$. If the nonlinearity is of the form $g_C(x) = x + ax^2 - b
573: x^3 + \mathcal{O}(x^4)$, {we have}
574: %
575: {\begin{align}
576: g_-(C) &\approx C_- + a C_+C_- -
577: bC_-^3/4 -
578: 3bC_-C_+^2/4 \\
579: g_+(C) &\approx C_+ + a C_+^2/2 +
580: aC_-^2/2 - bC_+^3/4 - 3bC_+C_-^2/4
581: \nonumber \end{align}}
582: Hence, when $C_- = 0$,
583: $C_+$ cannot excite it since $g_-(\bC) = 0$.
584: {However}, if $a\ne 0$, the \wasred{synchronous mode},
585: will be excited {passively}
586: %
587: by the \wasred{antiphase} mode through the quadratic coupling
588: term $aC_-^2/2$, responding with double frequency, as could be easily tested.
589:
590: To analyse coupling between {}{the}
591: antiphase modes, we {assume for simplicity that $g_c(\bC)$ is odd, so
592: $\bC_+=0$ since the synchronous mode is damped,
593: and $g_-(C_-)=2g_C(C_-/2)$.}
594: Consider a small perturbation, in the $m'$ mode direction,
595: to the $m=1$ cycle (the final orbit resulting from a
596: small deviation from the fixed point in the $m=1$ mode, with
597: a fundamental harmonic in the $m=1$ mode)
598: %
599: {such} that
600: $\bC_-(x) \approx C_1 \cos (2\pi x/N) + C_{m'} \cos(2\pi m' x/N)$
601: with $C_{m'} \ll C_1$. Expressing
602: $g_-(\bC)$ as $g_-(\bC) = \sum_n g_n e^{\ri 2\pi n x/N}$, it can be
603: shown that {for
604: large $N$},
605: $g_{m'} \approx C_{m'} \bar g'_C$ where $\bar g'_C$ is the derivative
606: of $g_C$ averaged over the unperturbed cycle.
607: (More
608: detailed analysis will be given in a future paper.)
609: {Because of the sigmoid form of $g_C(C)$, $\bar g'_C<g'_C(\bar \bC)$.}
610: Then $C_{m'} \propto e^{\rm\lambda^-_{m'}t}$
611: with $\lambda^-_{m'}$ as in equation (\ref {lams}) except that
612: $(B_{m'}, K_{m'})$, values derived from connections from $\bC$
613: cells, are rescaled by a factor {$\bar g'_C /g'_C(\bar \bC)<1$.}
614: %
615: %
616: {}{Thus the swimming cycle at large amplitude always
617: remains stable against
618: perturbation in other modes, even when the fixed point is unstable
619: against these perturbations}.
620:
621: {}{
622: If $Re(\lambda_1^-) \gg Re(\lambda_{m'}^-) \gtrsim 0$,
623: the $m'$ cycle will have a small amplitude
624: and hence $\bar g'(C) \sim g'(\bar{C})$ and it will be unstable
625: against perturbation in the $m=1$ mode. For larger
626: $Re(\lambda_{m'}^-)$
627: the amplitude of the cycle is larger and either cycle will be stable.
628: } Suppose the neural connections are such that {}{ the $m=\pm 1$
629: cycles, giving forward or backward swimming, are
630: both stable}. {}{The system would then display hysteresis, with
631: the final behaviour depending on the initial conditions.}
632: {}{Forward or backward swimming could then be selected by}
633: transient inputs from the brain stem,
634: rather than by
635: %
636: {setting constant inputs}
637: as described above.
638: This {}{seems less likely to be} the actual selection mechanism since
639: experiments on fictive swimming
640: %
641: (presumably with random initial conditions)
642: seldom observe {spontaneous} backward swimming. {}{However, the
643: forward swimming could simply have
644: a larger basin of attraction than backward
645: swimming.}
646:
647: When the lamprey turns, neural activities {on left and right sides are
648: unequal.} This is realizable by adding an additional
649: constant {input} to one side in the animal and
650: in models \cite{McClellanHagevik1997, Kozlov2002}, leading to {}{unequal
651: mean} {activities}
652: %
653: without
654: disrupting the oscillations,
655: provided that the gains $g'(.)$ are
656: roughly constant \wasred{near} the fixed points.
657: {}{Simulations (Fig.2) confirm the
658: analysis above.}
659:
660:
661:
662:
663: \begin{figure}[!!!hhhhhhhh!]
664: \setlength{\unitlength}{1.0pt}
665: \begin{center}
666: \begin{picture}(350, 170)
667: \put(0,80){{\epsfxsize=240pt \epsfbox{fig2abc.eps}}}
668: \put(0,-15){{\epsfxsize=240pt \epsfbox{fig2def.eps}}}
669: \end{picture}
670: %
671: %
672: \end{center}
673: \caption{Simulations. A: Membrane potentials of $\bE$
674: population on {either} side of one segment during forward
675: swimming and turning. The oscillations are in anti-phase between
676: the two sides. Turning is induced by an additional constant input
677: to one side only, {starting} at the time indicated by the dashed line. B:
678: $\bC$ slightly phase leads $\bE$ during forward swimming.
679: C \& D: {Waveform} of $\bE$ along the body
680: in forward and backward swimming, at consecutive times increasing
681: in the direction indicated
682: by the arrows, in the translational invariant model.
683: The switch from forward to backward swimming is achieved by increasing
684: the strength of $\bH$ and $\bQ$.
685: E: Oscillation waveforms (note different amplitudes) in body
686: segments at head, tail and centre of the body, {without}
687: %
688: translational invariance. F: Forward and backward swimming from
689: different initial conditions, with the same connection strengths and inputs.
690: }
691: \end{figure}
692:
693: To study behaviour {at the body ends}
694: %
695: or in {}{short sections of spinal cord \cite{Grillner2002}}, or
696: equivalently to see the effects of longer connections,
697: we abandon
698: {}{translation invariance}.
699: %
700: {}{Eliminating
701: $\bC$ in eq. (\ref {eqv})},
702: the minus mode has:
703: \be
704: \ddot \bE + (2-\bJ -\bB) \dot \bE + [1-\bJ - \bB +\bB\bJ + \bK(
705: \bH-\bQ)]\bE = 0, \nonumber
706: \ee
707: or, oscillator $i$ is driven by \wasred{force $F_i$} from {other} oscillators
708: %
709: \be
710: \begin{split}
711: \ddot \bE_i &+ (2-\bJ_{ii}-\bB_{ii}) \dot\bE_{i} + [1-
712: \tilde\bR_{ii}] \bE_{i}
713: \label {osci_couple} \\
714: &= F_i \equiv \textstyle{\sum_{j \ne i} F_{ij} \equiv \sum_{j \ne i}
715: (\bJ_{ij} +
716: \bB_{ij}) \dot \bE_j +\sum_{j\ne i} \tilde\bR_{ij} \bE_j}
717: \end{split}\nonumber\ee
718: where $\tilde\bR = \bB+\bJ -\bB\bJ
719: -\bK(\bH-\bQ)$.
720: {}{The intrinsic oscillation},
721: $\bE_i \sim e^{\lambda t}$,
722: {}{is damped, $\Re(\lambda ) = -1 +(\bJ+\bB)_{ii}/2
723: <0$},
724: as
725: indicated by {experiments} \cite{Grillner2002,Buchanan2001}.
726: %
727: %
728: We estimate $F_i$
729: %
730: {using} the approximation that oscillators $j\ne i$ still
731: {behave as} $\bE_j \propto e^{-{\rm i }(\omega t -kj)}$.
732: {}{We then have}
733: $F_i = \alpha_i \dot \bE_i + \beta _i\bE_i$, where
734: \begin{align}
735: \alpha_i &\equiv
736: \textstyle{\sum_{j\ne i} \left[(\bJ + \bB)_{ij} \cos (k(j-i))
737: -\tilde\bR_{ij} \sin (k(j-i)) /\omega \right]}
738: \nonumber \\
739: \beta _i &\equiv
740: \textstyle{\sum_{j\ne i} \left[(\bJ + \bB)_{ij} \omega \sin (k(j-i))
741: + \tilde\bR_{ij} \cos (k(j-i))\right]}
742: \nonumber \end{align}
743: The term $\alpha _i\dot \bE_i$ when $\alpha_i>0$ feeds oscillation
744: energy into the {$i^{th}$ (receiving)} oscillator,
745: \wasred{causing {}{emergent} oscillations
746: in {}{ coupled damped} oscillators.}
747: We divide {}{$\alpha_i =\alpha_{i,\text{desc}}+\alpha_{i,\text{asc}}$}
748: into the {}{descending and ascending}
749: parts, with summations over {}{$\sum_{j<i}$
750: and $\sum_{j>i}$} respectively.
751: Hence, for $i=1$, $\alpha_1 = \alpha_{i,\text{asc}}$; for $i=N$,
752: $\alpha_N = \alpha_{i,\text{desc}}$, {and for $1\ll i \ll N$},
753: $\alpha_{i} = \alpha_{1} + \alpha_{N}$. Since the first
754: and last {}{segments}
755: oscillate due to the driving force from other
756: oscillators, $\alpha_{1}>0$ and $\alpha_{N}>0$. Consequently,
757: $\alpha_{1}< \alpha_{N/2}$ and $\alpha_{N}< \alpha_{N/2}$.
758: {}{Further}, since {}{descending connections
759: are stronger}, \wasred{it is most
760: likely that, for $1\ll i \ll N$}, $\alpha_{i,\text{desc}}>
761: \alpha_{i,\text{asc}}$. Consequently, $\alpha _{1} < \alpha _{N}$.
762: {Hence, the rostral oscillator has a smaller amplitude than the caudal one,
763: which in turn has a smaller amplitude than the central one
764: {(Fig 2(E))}.
765: Firing rate saturation and variations of the fixed point along the
766: body may obscure {}{this pattern} in experimental data,
767: although body movements are indeed smallest near the head \cite{Paggett1998}.
768: \wasred{Similarly, oscillation amplitudes {}{will}
769: be reduced in sections of
770: spinal cords shorter than the typical lengths of inter-segment
771: connections, and
772: will eventually be zero in ever shorter sections,
773: as observed in experiments \cite{Buchanan2001}.}
774:
775:
776: %
777: %
778: In summary, {analysis of} a model of the
779: CPG neural circuit in lampreys {has given}
780: {}{new insights into} the neural connection structures {needed}
781: to
782: generate and control the swimming behaviour.
783: In particular, we predict that the contra-lateral connections between $\bC$
784: must be stronger than the self-excitatory connection strength of
785: the $\bE$ neurons; that the
786: $\bC$ neurons are more inhibited
787: (in their AC components) by $\bL$ neurons than excited by
788: the $\bE $ neurons;
789: and {have shown} how different swimming regimes can be selected by
790: scaling the strengths of the various neural connections without changing
791: \wasred{the} connection patterns.
792: Our framework should help to provide further
793: insights into CPGs of animal locomotion.
794: %
795:
796: {\bf Acknowledgements}:
797: We thank Peter Dayan for \wasred{comments, Carl van Vreeswijk for
798: {}{help with literature}, and
799: the Gatsby Charitable Foundation for support.}
800: %
801: %
802: \begin{thebibliography}{apssamp}
803: \small{
804: %
805: \bibitem{Grillner2002}Grillner S, Wallen P.
806: %
807: Brain Res Rev.2002;40(1-3):92-106.
808: \bibitem{Buchanan2001} J.T.Buchanan
809: %
810: Progr. of Neurobiology 63,
811: 2001,p.441-446.
812: and J. Neurophysiol. 1999, 81(5), P2037-45.
813: %
814: \bibitem{McClellan1998}McClellan AD, Hagevik A.
815: %
816: %
817: %
818: Exp.Brain Res.1998,126:93-108
819: %
820: %
821: %
822: %
823: %
824: %
825: \bibitem{CohenEtal82}
826: \wasred{Cohen AH, Holmes PJ, and Rand RH
827: %
828: %
829: J. Math. Biol. (1982) 13(3):345-69.}
830: \bibitem{E&K} Ermentrout G B, Kopell N.
831: %
832: %
833: Comm. Pure
834: Appl. Math.
835: (1986), 39(5), 623-660; Kopell N, Ermentrout G B.
836: %
837: %
838: SIAM J. Appl. Math.
839: (1994), 54(2), 478-507; Kopell N, Ermentrout G B.
840: %
841: %
842: SIAM
843: J. Appl. Math. 50(4),1014-10522.
844: \bibitem{Taylor1998} Taylor D, Holmes P, J. Math. Biol. (1998) 37, 419-446.
845: \bibitem{Jung} Jung R, Kiemel T, \& Cohen A H. J. Neurophysiol. (1996)
846: 75(3). 1074-86
847: %
848: %
849: \bibitem{Kozlov2002}
850: Kozlov AK, Ullen F, Fagerstedt P, Aurell E, Lansner A, Grillner S
851: %
852: Biol Cybern 2002 86(1):1-14
853: \wasred{\bibitem{McClellanHagevik1997}
854: McClellan AD, Hagevik A. J Neurophysiol. 1997 78(1): 214-28.}
855: \bibitem{Miller1998} Miller W L, Sigvardt K A.
856: %
857: %
858: J. Neurosci. Methods 1998, 80:113-128.
859: \bibitem{Paggett1998} Paggett K C, Gupta V, McClellan A D. Exp. Brain Res. 1998 119:213-223.
860: }
861:
862: \end{thebibliography}
863: %
864:
865:
866:
867: \end{document}
868:
869:
870:
871:
872: