cs0202034/CoRR.tex
1: %format=latex
2: 
3: \documentstyle[12pt]{article}
4: \renewcommand{\baselinestretch}{1.5}
5: \oddsidemargin=0cm
6: \textwidth=16cm
7: \textheight=23cm
8: \topmargin=0cm
9: \headheight=0cm
10: \headsep=0cm
11: \topskip=0cm
12: 
13: \newcommand{\ee }{^{\mbox{\scriptsize EE}}}
14: \newcommand{\ei }{^{\mbox{\scriptsize EI}}}
15: \newcommand{\ie }{^{\mbox{\scriptsize IE}}}
16: \newcommand{\ii }{^{\mbox{\scriptsize II}}}
17: 
18: \newcommand{\E }{^{\mbox{\scriptsize E}}}
19: \newcommand{\I }{^{\mbox{\scriptsize I}}}
20: 
21: \newcommand{\low }{_{\mbox{\scriptsize low}}}
22: \newcommand{\sn }{_{\mbox{\scriptsize sn}}}
23: \newcommand{\hopf }{_{\mbox{\scriptsize hopf}}}
24: 
25: \newcommand{\eqdef}{\stackrel{\rm def}{=}}
26: 
27: \begin{document}
28: 
29: \input{psfig}
30: 
31: \begin{titlepage}
32: \centerline{\huge COVARIANCE PLASTICITY}
33: \vglue .5cm
34: \centerline{\huge AND REGULATED CRITICALITY}
35: \footnotetext
36: {Supported by the Jean and H\'{e}l\`{e}ne Alfassa fund
37: for research in Artificial Intelligence,
38: Office of Naval Research contract N00014-91-J-1021,
39: National Science Foundation contract DMS-9217655,
40: and ARL contract MDA972-93-1-0012.}
41: \vglue 1cm
42: \centerline{\bf Elie Bienenstock}
43: \vglue .2cm
44: \centerline{Division of Applied Mathematics}
45: \centerline{Brown University}
46: \centerline{Providence RI 02912}
47: \centerline{USA}
48: \centerline{and CNRS, Paris, FRANCE}
49: \vglue .2cm
50: \centerline{elie@dam.brown.edu}
51: \vglue .5cm
52: \centerline{\bf Daniel Lehmann}
53: \vglue .2cm
54: \centerline{Department of Computer Science}
55: \centerline{Hebrew University}
56: \centerline{Jerusalem}
57: \centerline{ISRAEL}
58: \vglue .2cm
59: \centerline{lehmann@cs.huji.ac.il}
60: \vglue .5cm
61: \centerline{January 1995}
62: 
63: \end{titlepage}
64: 
65: \vglue 1cm
66: \begin{abstract}
67: We propose that a regulation mechanism
68: based on Hebbian covariance plasticity
69: may cause the brain to operate near criticality.
70: We analyze the effect of such a regulation
71: on the dynamics of a network with excitatory and inhibitory neurons
72: and uniform connectivity within and across the two populations.
73: We show that, under broad conditions,
74: the system converges to a critical state
75: lying at the common boundary of three regions in parameter space;
76: these correspond to three modes of behavior:
77: high activity, low activity, oscillation.
78: \end{abstract}
79: 
80: \newpage
81: 
82: \section{Introduction}
83: \label{intro}
84: 
85: That evolved brains are highly sensitive organs is an everyday observation.
86: Viewed as a dynamical system,
87: a brain may be said to be {\em unusually} susceptible to perturbations
88: and initial conditions.
89: This leads one to ask whether
90: brains may be operating near some form of instability, or criticality,
91: a hypothesis related to the notions
92: of computation at the edge of chaos (Langton 1990)
93: and self-organized criticality (Bak et al. 1987).
94: In this paper we propose that
95: while most regulation mechanisms at work in the brain
96: act according to a classical homeostasis schema,
97: i.e., have a stabilizing effect,
98: an opposite effect could result from the regulation of synaptic weights
99: by a specific form of Hebbian covariance plasticity.
100: Such a regulation may bring the system near criticality.
101: We suggest that regulated criticality may be the mechanism
102: whereby sensitivity is {\em maintained} throughout life
103: in the face of ongoing changes in brain connectivity.
104: 
105: Hebbian synaptic plasticity (Hebb 1949) plays an important role
106: in the development of the nervous system,
107: and is also believed to underlie many instances of learning in the adult.
108: A {\em covariance rule} of Hebbian plasticity roughly states
109: that the change in the efficacy of a given synapse
110: varies in proportion to the covariance
111: between the presynaptic and postsynaptic activities.
112: As noted by many authors
113: (e.g. Sejnowski 1977a, 1977b; Bienenstock et al. 1982;
114: Linsker 1986; Sejnowski et al. 1988),
115: a covariance-type rule is preferable to a rule that uses the mere product
116: of pre- and post-synaptic activities
117: because the covariance rule predicts not only weight increases
118: but also activity-related weight decreases,
119: and as a consequence allows convergence to non-trivial connectivity states.
120: Some forms of covariance plasticity have been shown to be optimal
121: for information storage
122: (Willshaw and Dayan 1990; Dayan and Willshaw 1991; Dayan and Sejnowski 1993).
123: Also, evidence for Hebbian plasticity of the covariance type
124: has been reported in several preparations
125: (Fr\'egnac et al. 1988, 1992; Stanton and Sejnowski 1989;
126: Artola et al. 1990; Dudek and Bear 1992).
127: 
128: We shall investigate,
129: in a simple network including excitatory and inhibitory neurons,
130: the effect of covariance plasticity
131: acting as a mechanism of {\em regulation,} rather than supervised learning.
132: Synaptic modification results
133: in changes---quantitative or qualitative---in the activity
134: that reverberates in the network;
135: these changes in turn cause further modification of the weights,
136: thereby creating a feedback loop between activity and connectivity.
137: Studying this loop as such,
138: i.e., independently from any input and output,
139: we demonstrate that, under fairly general conditions,
140: it causes the network to converge to a critical surface in parameter space,
141: the locus of an abrupt transition between different activity modes.
142: In Metzger and Lehmann (1990, 1994) a similar Hebbian rule
143: has been studied in the context of supervised learning of temporal sequences.
144: 
145: Schematically, the convergence to a critical state can be explained as follows.
146: One mode of behavior of a network
147: including excitatory and inhibitory neurons is oscillation;
148: such behavior takes place if the synaptic weights
149: linking excitatory neurons to each other---we
150: will refer to these as E-to-E weights---are
151: high enough but not too high.
152: Oscillation entails high covariance values,
153: hence, according to the covariance rule,
154: results in further increase of the E-to-E weights.
155: If however these weights are higher than a certain critical value---which
156: depends on other parameters of the system---oscillatory
157: behavior is impossible,
158: hence covariance is low or zero,
159: hence, in accord with the covariance rule used,
160: the E-to-E weights {\em decrease.}
161: As a result, the E-to-E weights stabilize around
162: the critical surface that separates the region of oscillation
163: from the region(s) of steady firing.
164: 
165: Our study is conducted in the simplest type of network
166: that will support oscillatory activity:
167: all synaptic weights of a given type are given identical values,
168: and so are all firing thresholds of a given type.
169: This results in a system with just six parameters---four
170: synaptic weights and two thresholds---and a limited range of behaviors.
171: Essentially, all neurons fire uniformly, either at a constant rate
172: (the number of possible rates of firing is one or two, depending on parameters)
173: or periodically in time.
174: In the {\em thermodynamic,} i.e., large-size, limit,
175: the dynamics of the network is adequately described
176: by a system of differential equations
177: obtained through a classical mean-field approximation.
178: 
179: We first perform a simple bifurcation analysis of this differential system
180: (Guckenheimer and Holmes 1983).
181: We then show that the effect of covariance regulation
182: is to stabilize the parameter state
183: at a surface of transition,
184: where the dynamics exhibits an instability.
185: Such a critical parameter state for a dynamical system
186: may be characterized as {\em degenerate.}
187: A generic, i.e., non-exceptional, state is one
188: where one expects to find the system
189: in the absence of further assumptions.
190: Mathematically, a generic state of a dynamical system is in the {\em interior}
191: of a parameter region corresponding to a given behavior,
192: and the system in such a parameter state
193: is said to be {\em structurally stable;}
194: the set of non-generic parameter states has measure zero.
195: We shall show that a state of higher degeneracy,
196: characterized as a point of intersection of {\em several} critical surfaces,
197: can be achieved by the simultaneous regulation of {\em several} parameters.
198: In the vicinity of that highly degenerate state,
199: the system displays a range of behaviors, including chaos.
200: 
201: The plan of the paper is as follows.
202: In the next section we study the dynamical properties
203: of our simple network---in the differential-equation formulation---with
204: {\em fixed} parameters (synaptic weights and firing thresholds).
205: We characterize the bifurcations
206: which take place at the boundaries between domains
207: corresponding to different modes of behavior.
208: This study is conducted for a {\em reduced} system,
209: where the thresholds are eliminated in such a way
210: as to render the dynamics symmetric about the origin.
211: Section \ref{regulation} describes the regulation equations.
212: Section \ref{reduced} describes the behavior of the regulated reduced system,
213: and Section \ref{full} that of the regulated full system.
214: 
215: \section{The fixed-parameter model}
216: \label{model}
217: 
218: This section describes the dynamics of the model with fixed parameters.
219: We first briefly describe a network consisting of a large number ($2N$)
220: of binary-valued neurons operating under a stochastic dynamics.
221: However, rather than using this network for our study of plasticity,
222: we make a number of simplifications and approximations,
223: leading to a deterministic two-variable differential system
224: with just six parameters.
225: The two variables are the excitatory and inhibitory
226: population averages of cell activity
227: in the $2N$-dimensional model;
228: the six parameters include the four average weights of the synapses
229: within and between these two populations,
230: as well as the average firing thresholds for the two populations.
231: We then study the asymptotic behavior
232: of this differential system for various parameter values.
233: Different types of asymptotic behavior,
234: in different regions of the parameter space,
235: correspond to different {\em phases} of the stochastic system,
236: and we pay particular attention to the {\em bifurcations} of the solutions,
237: where the bifurcation parameters are the synaptic weights---see
238: Schuster and Wagner (1990) and Borisyuk and Kirillov (1992)
239: for a related bifurcation analysis.
240: Bifurcations correspond to {\em phase transitions}
241: in the statistical-physics formulation
242: (the original $2N$-dimensional model).
243: 
244: We consider a fully-connected network
245: of $N$ excitatory and $N$ inhibitory linear-sigmoidal
246: $\{0,1\}$-valued neurons,\footnote
247: {It is not essential
248: that the numbers of excitatory and inhibitory neurons be the same.}
249: operating under a stochastic dynamics.
250: We denote the activity of the $i$-th excitatory,
251: resp. inhibitory, neuron by
252: $x_i\E (t)$, resp. $x_i\I (t)$,
253: with $x_i\E (t), x_i\I (t)\in\{0,1\}$, $i=1\ldots N$,
254: and we denote the synaptic weights by
255: $w_{ij}\ee , w_{ij}\ei , w_{ij}\ie , w_{ij}\ii $, $i,j=1\ldots N$,
256: where $i$ is postsynaptic and $j$ presynaptic,
257: and the superscripts indicate,
258: for each of the two neurons, whether it is excitatory or inhibitory.
259: Thus, for all $i$ and $j$, $w_{ij}\ee $ and $w_{ij}\ie $ are positive or zero,
260: whereas $w_{ij}\ei $ and $w_{ij}\ii $ are negative or zero.
261: 
262: The {\em local field} on excitatory neuron $i$,
263: i.e., the difference between its membrane potential
264: and its firing threshold $h_i\E $,
265: is $g_i\E (t)=\sum_j w_{ij}\ee  x_j\E (t) +
266: \sum_j w_{ij}\ei  x_j\I (t) - h_i\E $.  
267: Similarly, the local field on inhibitory neuron $j$ is
268: $g_i\I (t)=\sum_j w_{ij}\ie  x_j\E (t) +
269: \sum_j w_{ij}\ii  x_j\I (t) - h_i\I $,
270: where $h_i\I $ is the threshold of inhibitory neuron $i$.
271: The network dynamics is defined by:
272: (i) selecting at random, with uniform probability,
273: one of the $2N$ neurons;
274: (ii) computing its local field $g(t)$,
275: of the form $g_i\E (t)$ or $g_i\I (t)$; and
276: (iii) defining the state of the network at time $t+\delta t$
277: to be equal to the state at time $t$
278: except, possibly, for the selected neuron, whose state becomes---or stays---1
279: with probability $\frac{1}{2}(1+\tanh(\beta g(t)))$.
280: Parameter $\beta$ is a fixed non-negative number,
281: an {\em inverse temperature}.
282: The temperature $T=1/\beta$ measures the amount of noise in the system:
283: the higher the temperature, the noisier the dynamics.
284: The update interval is $\delta t = 1/(2N)$,
285: so that each neuron is updated on average once every time unit.
286: This {\em asynchronous} dynamics, of the Glauber type (Glauber 1963),
287: is widely used in statistical-mechanics models;
288: it lends itself to a convenient mean-field approximation (see below).
289: 
290: A system such as the one just described
291: will exhibit a highly diverse range of behaviors,
292: depending on the values of the synaptic weights and firing thresholds.
293: But we now make the much simplifying assumption that synaptic weights
294: and firing thresholds are {\em uniform} across each class.
295: Specifically, for all $i,j=1,\ldots,N$,
296: we assume that $h_i\E =h\E $, $h_i\I =h\I $,
297: $w_{ij}\ee =w\ee /N$, $w_{ij}\ei =-w\ei /N$,
298: $w_{ij}\ie =w\ie /N$, and $w_{ij}\ii =-w\ii /N$,
299: where $h\E $, $h\I $, $w\ee $, $w\ei $, $w\ie $ and $w\ii $ are fixed parameters,
300: and $w\ee $, $w\ei $, $w\ie $ and $w\ii $ are non-negative.
301: The dynamics is thus parameterized by six constants,
302: four synaptic weights and two thresholds;
303: $\beta$ is a mere multiplicative factor common to all six parameters,
304: yet it is convenient to use it as a seventh parameter.
305: Unless otherwise mentioned, $\beta$ will be 1.
306: 
307: Due to this uniformity assumption,
308: all neurons in any of the two populations experience the same field
309: at any given time.
310: This system exhibits a limited number of fairly simple behaviors,
311: of which Figure 1 is an example.
312: This figure shows the time variation of
313: $\langle x_i\E (t)\rangle$ and $\langle x_i\I (t)\rangle$,
314: the {\em average} activation levels across
315: the excitatory and inhibitory populations.
316: In this example, parameters are:
317: $N=70$, $w\ee =12$, $w\ie =8$, $w\ei =10$, $w\ii =2$,
318: $h\E = 1$, $h\I = 3$.
319: One unit on the time axis corresponds to $2N$ updates,
320: so that each neuron is updated, on average, once every time unit.
321: For these parameter values, the system {\em oscillates.}
322: Note that the oscillation is not perfectly regular, a finite-size effect.
323: Note also that the inhibitory activity lags somewhat behind the excitatory activity:
324: the excitatory neurons first trigger the inhibitory ones,
325: which in turn extinguish, for a while, the excitatory population.
326: \vglue .4cm
327: \centerline{\em (Insert Figure 1 around here)}
328: \vglue .4cm
329: The presence of oscillations and the amplitude and shape of the waveform
330: depend on the various parameters.
331: However, rather than pursuing this study of the stochastic system,
332: we shall consider the approximation that obtains
333: in the {\em thermodynamic limit},
334: that is, when $N\rightarrow\infty$.
335: The update interval $\delta t=1/(2N)$ then goes to 0
336: and so does each individual synaptic weight.
337: Straightforward approximations (Rubin 1988; Schuster and Wagner 1990)
338: then lead to a continuous-time differential system
339: for the population averages of the excitatory and inhibitory activation levels,
340: which we denote, respectively, by $s$ and $\sigma$:
341: \begin{equation} \left\{ \begin{array}{l}
342: \dot{s}(t) = .5 - s(t) + .5\tanh[\beta(w\ee s(t)-w\ei \sigma(t)-h\E )]\\
343: \dot{\sigma}(t)=.5- \sigma(t) +.5\tanh[\beta(w\ie s(t)-w\ii \sigma(t)-h\I )].
344: \end{array} \right. \label{sys_full} \end{equation}
345: Note that the variables $s(t)$ and $\sigma(t)$
346: remain at all $t$ within the interval [0,1].
347: When $\beta=0$
348: system \ref{sys_full} has a unique attractor, $(s,\sigma)=(.5,.5)$.
349: Indeed, in the high-temperature limit,
350: all neurons act independently of each other
351: and fire with probability .5 at each time.
352: 
353: We shall now make a last simplification,
354: whose purpose it is to render $(.5,.5)$ a fixed point---though
355: not necessarily stable---at {\em all} temperatures
356: and for all values of the synaptic weights.
357: This is easily achieved by letting the thresholds $h\E $ and $h\I $
358: be determined by the synaptic weights as follows:
359: \begin{equation} \begin{array}{l}
360: h\E  = .5 (w\ee -w\ei )\\
361: h\I  = .5 (w\ie -w\ii ).
362: \end{array} \label{chvar} \end{equation}
363: It is then convenient to adopt the change of variables:
364: $s\mapsto s-.5$, $\sigma\mapsto \sigma-.5$,
365: and system \ref{sys_full} becomes:
366: \begin{equation} \left\{ \begin{array}{l}
367: \dot{s}(t) = -s(t) + .5\tanh[\beta(w\ee s(t)-w\ei \sigma(t))]\\
368: \dot{\sigma}(t) = -\sigma(t) + .5\tanh[\beta(w\ie s(t)-w\ii \sigma(t))].
369: \end{array} \right. \label{sys_red} \end{equation}
370: In \ref{sys_red}, the variables $s$ and $\sigma$
371: are in the interval $[-.5,+.5]$,
372: and the only parameters left are the four synaptic weights
373: and the inverse temperature.
374: For all parameter values, the origin is a fixed point of system \ref{sys_red}.
375: A different position for the fixed point
376: could be obtained with an appropriate modification of equations \ref{chvar},
377: yet in the current version the fixed point is also a center of symmetry.
378: For the moment,
379: this hard-wired symmetry should be regarded as an {\em ad-hoc} device,
380: whose purpose is to make the analysis more convenient.
381: We shall refer to system \ref{sys_full} as the {\em full} system,
382: and to system \ref{sys_red} as the {\em reduced} system.
383: We shall see in Section \ref{full}
384: that under appropriate regulation
385: the two systems behave similarly.
386: 
387: We now discuss some important properties of the reduced system,
388: system \ref{sys_red} (see also Rubin 1988).
389: Consider first Figure 2a (phase diagram), which shows
390: four trajectories of the state $(s(t),\sigma(t))$;
391: the starting points of these trajectories are indicated by triangles.
392: The parameters (synaptic weights) used in this example are identical
393: to those used in Figure 1, i.e.,
394: $w\ee =12$, $w\ie =8$, $w\ei =10$, $w\ii =2$.
395: As expected, the asymptotic behavior is {\em periodic;}
396: there is a limit cycle which attracts all points of the square $[-.5,.5]^2$,
397: except the unstable equilibrium $(0,0)$.
398: Motion is counterclockwise,
399: for, as mentioned above, $\sigma(t)$ lags behind $s(t)$.
400: 
401: In addition to these four orbits, Figure 2a shows two curves,
402: the $s$- and $\sigma$-{\em nullclines} for system \ref{sys_red}.
403: These are the loci of the points $(s,\sigma)$ such that $ds/dt$,
404: resp. $d\sigma/dt$, vanish.
405: The equations for the $s$- and $\sigma$-nullclines
406: are easily seen to be, respectively:
407: \begin{equation}
408: \sigma = \frac{1}{w\ei }(w\ee s - T\tanh^{-1}(2s)),
409: \label{snull}\end{equation}
410: \begin{equation}
411: s = \frac{1}{w\ie }(w\ii \sigma + T\tanh^{-1}(2\sigma)).
412: \label{signull}\end{equation}
413: The $\sigma$-nullcline is an increasing sigmoid-shaped curve,
414: whereas the $s$-nullcline generally has the shape of an `S' lying on its side.
415: Of particular interest are the intersection points of the two nullclines;
416: these are the {\em fixed points} of the dynamics.
417: In the case illustrated in Figure 2a, the only intersection is $(0,0)$,
418: an unstable equilibrium.
419: Trajectories intersect the $s$-, resp. $\sigma$-, nullcline
420: in a direction parallel to the $\sigma$-, resp. $s$-, axis.
421: 
422: The study of the nullclines is of interest
423: because it is often possible
424: to predict how a parameter change will affect the dynamics of the system
425: by reasoning about how the nullcline diagram will change;
426: the bifurcation we shall be mostly interested in
427: is associated with a conspicuous change in this diagram.
428: Note that the $s$-nullcline is affected by parameters $w\ee $ and $w\ei $,
429: whereas the $\sigma$-nullcline is affected by parameters $w\ii $ and $w\ie $.
430: \vglue .4cm
431: \centerline{\em (Insert Figure 2 around here)}
432: \vglue .4cm
433: Let us consider first the changes brought about
434: by letting parameter $w\ee $ grow,
435: starting from the point $w\ee  = 12$
436: for which the system oscillates;
437: other parameters are unchanged.
438: When $w\ee $ grows, the slope of the central,
439: quasi-linear, part of the $s$-nullcline increases
440: (see equation \ref{snull});
441: that part of the curve rotates about the symmetry center (0,0).
442: As a result, the peak of the $s$-nullcline to the right approaches
443: the upper part of the sigmoid-shaped $\sigma$-nullcline,
444: and the minimum of the $s$-nullcline to the left approaches
445: the lower part of the $\sigma$-nullcline.
446: Eventually, at a certain critical value $\hat w\ee \sn $
447: (subscript `sn' stands for `saddlenode'---see below),
448: the two curves become tangent to each other.
449: This happens in two points at once,
450: near the upper right-hand corner and near the lower left-hand corner,
451: due to the symmetry of the system.
452: This situation is depicted in Figure 2b:
453: $w\ee $ is exactly equal to the critical value $\hat w\ee \sn $
454: (with parameters as above, $\hat w\ee \sn \approx 14.22$),
455: and the nullclines are just tangent to each other.
456: 
457: When $w\ee $ grows a little further,
458: each point of contact splits into two intersection points,
459: of which one is an attractor.
460: Figure 2c shows this situation, with $w\ee =15$,
461: somewhat above the critical value $\hat w\ee \sn $.
462: Four trajectories are shown, in addition to the two nullclines.
463: The system has five fixed points,
464: three unstable ones and two stable ones (attractors).
465: Only the stable fixed points are of interest to us;
466: they are very near the upper right-hand
467: and lower left-hand corners of the square,
468: corresponding to high,
469: respectively low, excitatory and inhibitory activities.
470: 
471: The bifurcation occurring at $\hat w\ee \sn $
472: is of the {\em saddlenode} type.
473: It results in a drastic change of behavior of the system:
474: the periodic attractor disappears
475: and is `siphoned' into the two new point attractors.
476: These two points attract the entire square,
477: except a set of measure 0
478: which includes the three unstable fixed points.
479: Thus, altough this bifurcation
480: is caused by a mere {\em local} change,
481: namely the intersection of the nullclines,
482: it results in a reorganization of the dynamics
483: that is both abrupt and {\em global.}\footnote
484: {As mentioned, {\em two} distinct saddlenode bifurcations
485: take place simultaneously.
486: Such a double bifurcation is not generic;
487: it occurs here due to the symmetry that we introduced
488: when reducing system \ref{sys_full} into system \ref{sys_red}.}
489: 
490: Having described the breakdown of oscillations
491: when parameter $w\ee $ is increased,
492: we now consider the opposite change, that is,
493: we let $w\ee $ decrease.
494: This results in a decrease of the slope of the central,
495: increasing, portion of the $s$-nullcline (equation \ref{snull}).
496: Eventually, the curve becomes monotonically decreasing;
497: this does not alter the number of intersections of the nullclines,
498: point $(0,0)$ remaining the sole equilibrium.
499: However, the amplitude of the limit cycle decreases along with $w\ee $.
500: The cycle eventually collapses to a point;
501: the equilibrium $(0,0)$ has then become stable.
502: This can be seen in a linear stability analysis
503: of system \ref{sys_red} around point $(0,0)$.
504: It is easily shown that,
505: in case there are two complex conjugate eigenvalues,\footnote
506: {The condition for this is $4w\ei w\ie >(w\ee +w\ii )^2$.}
507: the real part of these eigenvalues is negative if and only if $w\ee <w\ii +4T$.
508: Thus, $w\ii +4T$ is a critical value for parameter $w\ee $.
509: We define $\hat w\ee \hopf \eqdef w\ii +4T$
510: (with the current parameter setting, $\hat w\ee \hopf =6$).
511: The change of behavior occurring at $\hat w\ee \hopf $
512: is a {\em normal}\footnote
513: {That is, supercritical.
514: However, for very large values of $w\ie $,
515: the bifurcation is subcritical---see footnote 5.}
516: Hopf bifurcation.
517: 
518: So far, we studied the behavior of system \ref{sys_red}
519: for different values of parameter $w\ee $, all other parameters being fixed.
520: In other words, we described the system's behavior
521: on a particular 1-dimensional subspace
522: of the 4-dimensional parameter space.
523: We now extend this study to a 2-dimensional subspace, the $(w\ee ,w\ie )$ plane.
524: Figure 3a is the bifurcation diagram of system \ref{sys_red} in that plane,
525: with other parameters as before ($w\ei =10$, $w\ii =2$).
526: This diagram shows three distinct regions,
527: corresponding to three different attractor configurations;
528: unstable fixed points and unstable limit cycles are ignored in this diagram.
529: In the middle region---which we call region $\cal P$,
530: for {\em Periodic}---the system oscillates.
531: The boundary of this region to the right
532: is the saddlenode bifurcation curve, which we denote $\cal S$;
533: as discussed above,
534: the rightmost region has {\em two} point attractors,
535: and we call it region $\cal T$.
536: The leftmost region, which we call $\cal O$,
537: has only one point attractor, the center of symmetry $(0,0)$;
538: it is separated from region $\cal P$ by the Hopf bifurcation curve,
539: a vertical line of equation $w\ee =\hat w\ee \hopf $.
540: The curve in the lower left of the diagram,
541: separating region $\cal O$ from region $\cal T$,
542: is the locus of a {\em pitchfork} bifurcation.
543: This bifurcation diagram, obtained for one particular set of values
544: of the parameters $w\ei , w\ii $ and $\beta$,
545: is representative of the general case.\footnote
546: {It is however simplified in two ways.
547: First, the transition from region $\cal P$ to region $\cal T$
548: is of the saddlenode type only for large enough values of $w\ie $;
549: this range of values corresponds roughly
550: to the straight portion of curve $\cal S$ (Figure 3a).
551: To see why this is so, consider again Figure 2b,
552: the nullcline diagram at the bifurcation, with $w\ie = 8$.
553: Note that the points of contact between the nullclines
554: appear near the corners of the square, far from the origin;
555: this is due to the fact that $w\ie $ is relatively large,
556: hence the slope of the $\sigma$-nullcline at the origin
557: is larger than the slope of the $s$-nullcline.
558: The bifurcation is then of the saddlenode type, as described.
559: If however $w\ie $ is small,
560: hence so is the slope of the $\sigma$-nullcline at the origin, 
561: the transition from $\cal P$ to $\cal T$ as $w\ee $ is increased
562: takes place differently.
563: A pair of intersection points between the nullclines
564: first split off {\em from the origin;}
565: these are unstable equilibria.
566: As $w\ee $ increases,
567: these two equilibria move away from the origin,
568: while remaining inside the large stable limit cycle.
569: At a certain critical value for $w\ee $ they become stable---a
570: (double) subcritical Hopf bifurcation---and
571: almost immediately thereafter the large limit cycle disappears.
572: Thus, the transition from region $\cal P$ to region $\cal T$
573: really takes place in two steps,
574: giving rise to a {\em three-attractor} behavior:
575: the system has one large limit-cycle attractor
576: {\em as well as} two point attractors,
577: the latter being inside the cycle.
578: The region of the $(w\ee ,w\ie )$ plane where this behavior takes place
579: is a strip extending along the lower, curved,
580: part of the $\cal P/\cal T$ boundary;
581: it is too narrow to be seen in Figure 3a.
582: (With parameters $w\ei $ and $w\ii $ as above and $w\ie =2.75$,
583: the three-attractor behavior occurs for $w\ee $ between $8.993$ and $9.030$.
584: For some other values of $w\ei $ and $w\ii $ this behavior does not occur at all,
585: and the transition from $\cal P$ to $\cal T$ is always of the saddlenode type.)
586: For the purpose of this paper (see footnotes 8 and 11)
587: it is important to note that the point attractors
588: appear either {\em exactly} or {\em almost} at the same time
589: as the periodic attractor disappears.
590: The second approximation in the bifurcation diagram,
591: mentioned only for the sake of completeness,
592: concerns the $\cal O$-to-$\cal P$ transition.
593: This is generally a smooth, supercritical, Hopf bifurcation.
594: However, as mentioned in footnote 4, this Hopf bifurcation becomes subcritical
595: for very large values of $w\ie $.
596: There is thus a narrow region to the left of the bifurcation line $w\ee =\hat w\ee \hopf $
597: where the limit-cycle attractor coexists with the point attractor (0,0);
598: for instance, at $w\ie =100$, the width of this region is $\approx 0.63$.}
599: \vglue .4cm
600: \centerline{\em (Insert Figure 3 around here)}
601: \vglue .4cm
602: In sum, the $(w\ee ,w\ie )$ bifurcation diagram for system \ref{sys_red}
603: is characterized by a central periodic-attractor region,
604: a large vertical patch extending to $+\infty$ in the $w\ie $ direction
605: (phase $\cal P$),
606: flanked by point-attractor regions on each side (phases $\cal O$ and $\cal T$).
607: The transition from $\cal P$ to $\cal T$ is abrupt ($\cal S$ line),
608: while the transition from $\cal O$ to $\cal P$ is smooth.
609: As mentioned in the Introduction, system \ref{sys_full}---the full system---is
610: not amenable to such a thorough analysis;
611: however, we shall see in Section \ref{full}
612: that the two systems behave in much the same way
613: under the plasticity rules that we shall now introduce.
614: 
615: \section{The regulation equations}
616: \label{regulation}
617: 
618: Whereas in the previous section the synaptic weights $w\ee $ and $w\ie $
619: were fixed parameters, they will now be made to evolve.
620: Their evolution will obey a Hebbian covariance rule, hence be a function
621: of second-order temporal averages of the dynamic variables $s$ and $\sigma$.
622: Synaptic plasticity creates a {\em regulation loop:}
623: changing the parameters affects the dynamics of the system,
624: which in turn alters the second-order moments of $s$ and $\sigma$.
625: Formally, the regulation is implemented by
626: introducing additional differential equations, coupled to system \ref{sys_red}
627: (or to system \ref{sys_full}---see Section \ref{full}).
628: The rate of change of $w\ee $ and $w\ie $ will typically be
629: several orders of magnitude slower than that of $s$ and $\sigma$.
630: 
631: Let us first define, for any function of time $r(t)$,
632: a moving time average:
633: \begin{equation}
634: \bar{r}(t) = \rho \int_{-\infty}^t r(u) e^{\rho (u-t)} du.
635: \nonumber\end{equation}
636: Parameter $\rho$ is a positive constant, physically an inverse time;
637: the larger $\rho$, the narrower the averaging kernel.
638: Equivalently, $\bar{r}(t)$ may be defined by a differential equation,
639: more convenient for simulation purposes:
640: \begin{equation}
641: \frac{d\bar{r}(t)}{dt} = \rho (r(t) - \bar{r}(t)).
642: \nonumber\end{equation}
643: Consider now, with reference to the original stochastic model (Section \ref{model}),
644: the {\em instantaneous covariance} between two excitatory neurons $i$ and $j$,
645: defined as:
646: $c_{ij}\ee (t) \eqdef (x_i\E (t)- \bar x_i\E (t))(x_j\E (t) - \bar x_j\E (t))$.
647: If we take the {\em population average} $\langle c_{ij}\ee (t)\rangle$
648: of this instantaneous covariance,
649: we obtain, in the thermodynamic limit $N\rightarrow\infty$,
650: the instantaneous variance of $s(t)$:
651: \begin{equation}
652: c\ee (t) \eqdef (s(t) -\bar{s}(t))^2.
653: \label{covee}\end{equation}
654: 
655: It is this quantity $c\ee $ that we use to regulate
656: the excitatory-to-excitatory synaptic weight $w\ee $.
657: The regulation equation is linear in $c\ee $:
658: \begin{equation}
659: \frac{dw\ee (t)}{dt} = \varepsilon\ee  ( c\ee (t) - \theta\ee  ).
660: \label{regee}\end{equation}
661: Parameters $\varepsilon\ee $ and $\theta\ee $ are positive.
662: Note that the quantity $c\ee (t)$ is always non-negative;
663: the term $-\theta\ee$ is therefore necessary
664: to allow for decreases of $w\ee $.
665: 
666: We shall also consider a regulation for $w\ie $,
667: the synaptic weight from excitatory to inhibitory neurons,
668: although this regulation will play a less important role than that of $w\ee $. 
669: The modification rule for $w\ie $ has the same form as equation \ref{regee},
670: yet it uses the excitatory-to-inhibitory instantaneous covariance,
671: defined as:
672: \begin{equation}
673: c\ie (t) \eqdef (s(t) -\bar{s}(t)) (\sigma(t) -\bar{\sigma}(t)).
674: \label{covie}\end{equation}
675: The regulation equation for $w\ie $ then reads:
676: \begin{equation}
677: \frac{dw\ie (t)}{dt} = \varepsilon\ie  ( c\ie (t) - \theta\ie  ).
678: \label{regie}\end{equation}
679: In equation \ref{regie}, $\theta\ie $ is a positive constant,
680: as $\theta\ee $ in equation \ref{regee}.
681: However, the modification rate constant $\varepsilon\ie $ is negative.
682: The main reason for this will be given in the next section;
683: for now, note that this choice is consistent with the spirit of Hebb's principle,
684: for, when considered {\em postsynaptically} to the target neuron,
685: the effect of synapse reinforcement if that target neuron is inhibitory
686: is the opposite of the effect obtained if the target neuron is excitatory.
687: 
688: \section{Behavior of the regulated reduced system}
689: \label{reduced}
690: 
691: This section describes the behavior of the regulated reduced system.
692: We demonstrate that each of the two regulation loops
693: introduced in Section \ref{regulation},
694: when acting separately,
695: brings the system to the critical surface $\cal S$,
696: the locus of an abrupt phase transition (saddlenode bifurcation).
697: We then examine the behavior of the system
698: with the two regulation loops active simultaneously;
699: we show that under some conditions
700: the state converges to a point on $\cal S$
701: with a remarkable nullcline configuration.
702: 
703: Before we consider the regulation proper,
704: let us examine how the covariances
705: change across the $(w\ee ,w\ie )$ plane.
706: Figure 3b shows the values of $\bar c\ee $,
707: the time average of the instantaneous variance of $s(t)$,\footnote
708: {This corresponds, in the original system,
709: to the population- {\em and} time-average of the covariance,
710: $\langle \bar c_{ij}\ee \rangle$;
711: the latter becomes $\bar c\ee $
712: in the thermodynamic limit $N\rightarrow\infty$.
713: In the regulation equation,
714: we use the {\em instantaneous} covariance $c\ee (t)$
715: rather than its time average $\bar c\ee $ (see Discussion).
716: The time-averaged variance $\bar c\ee $
717: is used here for illustration purposes only.
718: In order to obtain an essentially constant value for $\bar c\ee $
719: rather than an oscillating function of time,
720: different values of $\rho$ are used for the two averaging operations:
721: the kernel used to average $c\ee $ into $\bar c\ee $
722: is ten times broader than the kernel used to compute $\bar s$ from $s$.}
723: along several horizontal lines in the $(w\ee ,w\ie )$ plane.
724: As expected, $\bar c\ee $ is positive only in region $\cal P$,
725: where the dynamics is periodic;\footnote
726: {In general, positive average covariance across a neuronal population
727: indicates collective fluctuations;
728: in our simplified two-dimensional system,
729: the only possible nontrivial asymptotic behavior
730: is periodic oscillation.}
731: although not shown, the same is true of $\bar c\ie $,
732: the time average of the E-to-I covariance.
733: Note that as $w\ee $ crosses the $\cal O$-to-$\cal P$ boundary
734: (Hopf bifurcation) from left to right,
735: $\bar c\ee $ increases {\em smoothly} from 0 to positive values:
736: as discussed above, the amplitude of the limit cycle
737: at this bifurcation is infinitesimal.
738: In contrast, the change in $\bar c\ee $ and in $\bar c\ie $
739: at $\cal S$ (saddlenode bifurcation) is a sharp one,
740: as the system undergoes there a transition from a {\em large} limit-cycle regime
741: to a fixed-point attractor.
742: 
743: We now start our study of covariance plasticity
744: by regulating parameter $w\ee $ in system \ref{sys_red}
745: while all other parameters, including $w\ie $, remain fixed.
746: The system under study then consists
747: of coupled equations \ref{sys_red}, \ref{covee}, \ref{regee}.
748: Equation \ref{regee} prescribes an increase of $w\ee $ when $c\ee > \theta\ee $,
749: and a decrease when $c\ee < \theta\ee $.
750: Referring to Figure 3b, we see that to the left of $\cal S$,
751: where $c\ee $ is high, the first of the two conditions applies;
752: in this region $w\ee $ increases.
753: To the right of $\cal S$ the covariance vanishes, and $w\ee $ decreases.
754: Therefore, $w\ee (t)$ is attracted to the transition line $\cal S$.\footnote
755: {The control parameter $\theta\ee $ should be smaller
756: than the value of $\bar c\ee $ immediately to the left of $\cal S$.
757: The portion of the boundary line
758: where the bifurcation is a subcritical Hopf rather than a saddlenode
759: (footnote 5) yields similar behavior,
760: since the disruption of the large-amplitude limit cycle
761: occurs very near the emergence of point attractors (see also footnote 11).}
762: \vglue .4cm
763: \centerline{\em (Insert Figure 4 around here)}
764: \vglue .4cm
765: The behavior of this $w\ee $ regulation loop is illustrated in Figure 4a
766: for the following setting of parameters:
767: $w\ei = 10$, $w\ii = 6$, $\rho=.1$, $\theta\ee =.01$, $\varepsilon\ee =.01$.
768: This figure focuses on a small region of the $(w\ee ,w\ie )$ plane,
769: and shows the projection of the trajectory of $(s, \sigma, w\ee ,w\ie )$.
770: Several trajectories are shown, all horizontal since $w\ie $ is a constant,
771: These trajectories terminate on the critical line $\cal S$,
772: and the behavior of the $s$ and $\sigma$ components on them is as follows.
773: On the trajectories coming from the left, in the $\cal P$ region,
774: $(s,\sigma)$ moves along a cyclic orbit,
775: whose amplitude grows as $w\ee $ increases
776: and approaches the bifurcation line.
777: On the trajectories coming from the right, in the $\cal T$ region,
778: $(s, \sigma)$ stays in one of the two point attractors while $w\ee $ decreases
779: until it reaches the bifurcation curve.
780: When $\cal S$ is reached, either from the left or from the right,
781: motion does not really stop.
782: Rather, $w\ee $ sets in a periodic oscillation of small amplitude
783: synchronized with a large-amplitude periodic motion of $(s,\sigma)$;
784: the frequency of this oscillation 
785: is several orders of magnitude slower than in $\cal P$,
786: hence covariance is small---it
787: matches, on average, the control parameter $\theta\ee $.
788: When in this regime, the system spends a long time in one of the two 
789: almost-attracting corners of the $[-.5,+.5]^2$ box before leaving it
790: and moving rapidly to the other corner.
791: This results in an almost-square wave,
792: a behavior that is intermediate between the fast periodic motion observed
793: in $\cal P$ and the bistable situation prevailing in $\cal T$.
794: The period of this oscillation and the amplitude
795: of the oscillation of $w\ee $
796: depend on parameters $\rho$, $\theta\ee $, and $\varepsilon\ee $.\footnote
797: {Not shown on Figure 4a is the leftmost part of region $\cal P$,
798: near the Hopf bifurcation,
799: where the limit cycle is of small amplitude
800: hence the condition $\bar c\ee  > \theta\ee $ is not realized.
801: When initialized there,
802: the system does not converge to $\cal S$.
803: However, in both the $w\ee $ and the $w\ie $ directions,
804: the domain of attraction of $\cal S$ extends to $+\infty$.}
805: 
806: We next consider the $w\ie $-regulated system,
807: where $w\ee $ and all other parameters remain fixed.
808: This system consists of coupled equations \ref{sys_red}, \ref{covie}, \ref{regie}.
809: As noted, the E-to-I covariance $c\ie $ vanishes
810: outside region $\cal P$, just like $c\ee $;
811: within $\cal P$ it varies, in a first approximation, like $c\ee $.
812: Since we chose $\varepsilon\ie $ to be {\em negative,}
813: $w\ie $ decreases in $\cal P$ and increases in $\cal T$,
814: whereas the opposite was true of $w\ee $
815: when it was regulated.
816: Figure 4b shows this $w\ie $ dynamics
817: in the same region of the $(w\ee ,w\ie )$ plane as before.
818: Parameters are $w\ei = 10$, $w\ii = 6$, $\rho=.1$,
819: $\theta\ie =.01$ and $\varepsilon\ie =-.01$.
820: The trajectories are now parallel to the $w\ie $ axis,
821: and $(w\ee ,w\ie )$ is again attracted to the critical line $\cal S$
822: separating region $\cal P$ from region $\cal T$.
823: This is true only to the left of the vertical asymptote of that curve;
824: trajectories to the right of that line go to $+\infty$.
825: 
826: In sum, regulation of either one of the two parameters $w\ee $, $w\ie $
827: has the effect of bringing the system
828: to the critical surface $\cal S$ separating the region of oscillation
829: from the region of bistable steady firing;
830: the nullcline diagram is then as in Figure 2b.
831: Note that when the system is on $\cal S$,
832: a small perturbation in the weights will elicit either oscillation,
833: constant firing at near-maximum rate,
834: or constant firing at near-minimum rate.
835: 
836: We now turn to the behavior of the system
837: when the two regulation loops act simultaneously;
838: we thus study the system of coupled equations
839: \ref{sys_red}, \ref{covee}, \ref{regee}, \ref{covie}, \ref{regie}.
840: Figure 4c shows the $(w\ee ,w\ie )$ dynamics for the same parameters as before,
841: i.e., $w\ei = 10$, $w\ii = 6$, $\rho=.1$, $\theta\ee =.01$, $\varepsilon\ee =.01$,
842: $\theta\ie =.01$ and $\varepsilon\ie =-.01$.
843: It appears from this diagram that the evolution proceeds
844: in two clearly distinct stages.
845: In the first stage,
846: which could be predicted from the study of the regulation loops
847: acting separately,
848: $(w\ee ,w\ie )$ moves toward line $\cal S$.\footnote
849: {The direction of this linear motion is roughly parallel to the line $w\ee = -w\ie $.
850: This is because $\varepsilon\ee =-\varepsilon\ie $,
851: $\theta\ee = \theta\ie $,
852: and the two covariances $c\ee $ and $c\ie $ are nearly the same.
853: Another choice of parameters would result in a different slope,
854: but otherwise similar behavior.}
855: When this line is reached,
856: motion slows down considerably---typically
857: by several orders of magnitude---and
858: proceeds {\em along} the critical line,
859: eventually converging to a point on $\cal S$ denoted $G$ in Figure 4c.
860: As before, attractor $G$ is in reality a slow limit cycle,
861: of small amplitude in $w\ee $ and $w\ie $,
862: and large amplitude in $s$ and $\sigma$.
863: All four variables, $s(t)$, $\sigma(t)$, $w\ee (t)$, and $w\ie (t)$,
864: are now synchronized;
865: the distinction between slow and fast variables has thus vanished.
866: The basin of attraction of $G$ in the $(w\ee ,w\ie )$ plane
867: roughly consists of the {\em union} of the two domains of attraction
868: of $\cal S$ for the separate $w\ee $ and $w\ie $ regulation dynamics;
869: only the region to the left of and around the Hopf line
870: is not attracted to the saddlenode line $\cal S$ and eventually to $G$.
871: \vglue .4cm
872: \centerline{\em (Insert Figure 5 around here)}
873: \vglue .4cm
874: The location of $\cal S$ in the $(w\ee ,w\ie )$ plane depends
875: on the values of the fixed parameters $w\ei $ and $w\ii $.
876: The location of the attractor $G$ on $\cal S$
877: further depends on the control parameters $\theta\ee $ and $\theta\ie $.
878: When the latter are given identical values,
879: as in the case illustrated in Figure 4c,
880: the attractor $G$ has the remarkable property
881: that the $s$- and $\sigma$-nullclines stand in {\em near overlap}
882: over a large portion of the interval [-.5,+.5] (Figure 5);
883: the flow of the system in this configuration
884: nearly vanishes on a large one-dimensional manifold
885: in the two-dimensional phase space.
886: Further, $s(t)$ and $\sigma(t)$ remain nearly identical at all times.\footnote
887: {Giving different values to parameters $\theta\ee $ and $\theta\ie $
888: mostly affects the behavior of the system
889: after it has reached $\cal S$;
890: if $\theta\ee $ is larger, resp. smaller, than $\theta\ie $,
891: the state moves downward, resp. upward, on $\cal S$.
892: When $(w\ee ,w\ie )$ is on $\cal S$ but above point $G$,
893: the nullclines are tangent to each other but do not overlap;
894: such a situation is illustrated in Figure 2b.
895: When $(w\ee ,w\ie )$ is on $\cal S$ but below point $G$,
896: the nullclines do overlap, but over a smaller domain.
897: With $\theta\ee = .0118$ and $\theta\ie = .0100$,
898: the state stabilizes in the narrow three-attractor region described in footnote 5.
899: The state $(s,\sigma)$ then visits each of the three `attractors' in turn:
900: its motion consists of a succession
901: of large-amplitude oscillations (periodic attractor)
902: and of spiraling orbits around two symmetric points
903: in the interior of the large cycle (point attractors).
904: The amplitude of the motion of $(w\ee ,w\ie )$ remains small.
905: This is a mildly chaotic behavior;
906: a more pronounced chaotic behavior
907: will be described in the next section for the full system.}
908: 
909: \section{Behavior of the regulated full system}
910: \label{full}
911: 
912: Recall that system \ref{sys_red},
913: which we used so far,
914: was derived from system \ref{sys_full}
915: by eliminating the firing thresholds $h\E $ and $h\I $ (equations \ref{chvar})
916: in such a way as to make $(.5,.5)$---$(0,0)$ in system \ref{sys_red}---a
917: center of symmetry of the dynamics.
918: While easier to analyze,
919: the reduced system is less realistic.
920: There is no clear biological justification for this hard-wired symmetry;
921: moreover, when the system is in phase $\cal T$,
922: i.e., to the right of the critical surface $\cal S$,
923: it can stay for arbitrarily long periods of time in one
924: of the two fixed point attractors, e.g. in the high-activity one;
925: this is unrealistic.
926: 
927: In this section we consider a biologically more plausible way
928: of introducing symmetry in the dynamics.
929: Rather than eliminating the thresholds according to equations \ref{chvar},
930: we {\em regulate} them,
931: thereby implementing a form of `soft' symmetry.
932: Regulating the firing thresholds in a neural network
933: is a simple way to maintain the mean activity
934: around an intermediate, useful, value.
935: This may be viewed as a simplification of the regulation mechanisms
936: at work in real brains,
937: which, in all likelihood, involve systems of inhibitory neurons
938: acting on various time scales.
939: 
940: The simultaneous regulation of four parameters
941: results in a complex dynamics,
942: which makes a thorough analysis impractical.
943: We shall proceed as follows.
944: We first consider, in system \ref{sys_full},
945: the regulation of $w\ee $ and $h\E $
946: for a given setting of all other parameters.
947: We show that the system converges to the intersection of two critical curves,
948: each of which corresponds to the establishement
949: of one point of contact between the nullclines.
950: We next consider the system
951: with all four parameters $h\E $, $h\I $, $w\ee $ and $w\ie $ regulated,
952: and study the projection of the dynamics on the $(w\ee ,w\ie )$ plane.
953: There are again two stages;
954: the first essentially reproduces the behavior observed
955: with the sole $(w\ee ,h\E )$ regulation,
956: while the second is analogous to that observed
957: when regulating $w\ee $ and $w\ie $ in the reduced system;
958: this applies for a broad range of the remaining fixed parameters $w\ei $ and $w\ii $.
959: 
960: Figure 6a is the bifurcation diagram of system \ref{sys_full}
961: in the $(w\ee ,h\E )$ plane,
962: for the following values of the fixed parameters:
963: $w\ei =10$, $w\ie =10$, $w\ii =1$, $h\I =5$.
964: As before, we ignore unstable equilibria and unstable limit cycles.
965: As before there are three regions,
966: denoted respectively by $\cal O$, $\cal T$ and $\cal P$,
967: corresponding to three types of asymptotic behavior:
968: single fixed-point attractor;
969: two fixed-point attractors (high and low activity);
970: one periodic attractor.
971: We now however subdivide region $\cal O$---somewhat arbitrarily---according
972: to the location of the fixed-point attractor in the phase space:
973: the three subregions denoted ${\cal O}_h$, ${\cal O}_m$, and ${\cal O}_l$,
974: correspond, respectively, to high, middle, and low activity for this attractor.
975: The transition between region $\cal P$ and region ${\cal O}_m$
976: takes place through the familiar, smooth, Hopf bifurcation.
977: The transition between ${\cal P}$ and ${\cal O}_h$,
978: as well as its continuation between ${\cal O}_l$ and $\cal T$,
979: takes place through a saddlenode bifurcation.
980: We denote by ${\cal S}_h$ the locus of this transition;
981: it marks the appearance of a point of contact between the nullclines
982: near the high-activity corner,
983: and is thus similar to the $\cal S$ transition in the reduced system.
984: However, due to the symmetry of that system,
985: another point of contact appeared simultaneously near the low-activity corner,
986: giving rise to a double bifurcation.
987: In system \ref{sys_full} this is no more the case,
988: and the intersection of the nullclines near the low-activity corner
989: gives rise to a distinct saddlenode bifurcation line,
990: the transition between ${\cal P}$ and ${\cal O}_l$,
991: which we denote ${\cal S}_l$.
992: \vglue .4cm
993: \centerline{\em (Insert Figure 6 around here)}
994: \vglue .4cm
995: When regulating $w\ee $ according to equation \ref{regee}
996: and leaving all other parameters fixed,
997: the behavior of system \ref{sys_full} is as follows.
998: When starting in region $\cal P$ to the left of the critical line ${\cal S}_h$,
999: the system oscillates, covariance is high, hence $w\ee $ increases
1000: until it reaches the critical line ${\cal S}_h$.
1001: A point of contact is then established
1002: near the high-activity corner of the square.
1003: The system settles in a slow periodic attractor,
1004: of small amlitude in $w\ee $ and large amplitude in $(s,\sigma)$,
1005: whereby nearly all the time is spent in the high-activity state.
1006: 
1007: We now regulate the threshold $h\E $ as well,
1008: in such a way as to stabilize $\bar{s}$,
1009: the time average of $s$, around a given target value $\theta\E $:
1010: \begin{equation}
1011: \frac{dh\E (t)}{dt} = \varepsilon\E ( \bar{s}(t) - \theta\E  ).
1012: \label{reghe}
1013: \end{equation}
1014: The rate constant $\varepsilon\E$ is positive and small,
1015: and the control parameter $\theta\E $ is chosen
1016: well in the interior of the interval $(0,1)$,
1017: e.g. between .2 and .8
1018: (remember that in system \ref{sys_full}
1019: the activity variables $s$ and $\sigma$ lie in the interval $(0,1)$).
1020: To see how equation \ref{reghe} achieves the desired regulation,
1021: note for instance that, if $\bar{s}(t) > \theta\E $, $h\E $ will increase,
1022: which in turn will result in a decrease of $\bar{s}(t)$.
1023: 
1024: When both $w\ee $ and $h\E $ are regulated,
1025: the system converges to the {\em intersection}
1026: of the two critical lines ${\cal S}_h$ and ${\cal S}_l$.
1027: In effect, we saw that the full system,
1028: when at a generic point of ${\cal S}_h$,
1029: stays nearly all the time in the high-activity state;
1030: this results in a high value of $\bar {s}$.
1031: To achieve the condition $\bar {s} \approx \theta\E $,
1032: the equilibrium for equation \ref{reghe},
1033: the system can only be on ${\cal S}_l$ {\em as well.}
1034: 
1035: The joint $(w\ee ,h\E )$ dynamics is illustrated in Figure 6b,
1036: for parameters $w\ei $, $w\ie $, $w\ii $ and $h\I $ as above,
1037: and $\rho = .2$, $\varepsilon\E = .001$, $\theta\E = .5$,
1038: $\varepsilon\ee =.01$, $\theta\ee =.01$.
1039: The intersection of ${\cal S}_h$ and ${\cal S}_l$, denoted $F$ in Figure 6b,
1040: is reached from all directions in the $(w\ee ,h\E )$ plane.
1041: When coming from low $w\ee $ values,
1042: the system oscillates and converges to $F$ through region $\cal P$.
1043: When coming from high $w\ee $ values,
1044: the system reaches $F$ through region $\cal T$,
1045: where it bounces back and forth
1046: between the high- and low-activity point attractors
1047: (an oscillation much slower than in $\cal P$).
1048: 
1049: The nullcline diagram for point $F$ of Figure 6b is illustrated in Figure 6c.
1050: There are now two points of contact between the nullclines,
1051: a situation more degenerate than the one
1052: that obtains from regulating $w\ee $ only,
1053: but `equivalent' to the situation
1054: obtained in the reduced system 
1055: by regulating a single parameter, $w\ee $ or $w\ie $
1056: (compare Figure 6b to Figure 2b).
1057: What characterizes the dynamics at point $F$
1058: is that the system is on the verge of oscillation
1059: and on the boundary of each of the two steady-firing phases.
1060: 
1061: We finally consider the system
1062: with the four parameters $h\E $, $h\I $, $w\ee $ and $w\ie $ regulated.
1063: We thus include, in addition to equations
1064: \ref{sys_full}, \ref{covee}, \ref{regee}, \ref{covie}, \ref{regie} and \ref{reghe},
1065: a regulation equation for the inhibitory threshold $h\I $:
1066: \begin{equation}
1067: \frac{dh\I (t)}{dt} = \varepsilon\I ( \bar{\sigma}(t) - \theta\I  ).
1068: \label{reghi}
1069: \end{equation}
1070: As in equation \ref{reghe},
1071: the rate constant $\varepsilon\I$ is positive and small,
1072: and $\theta\I $ is chosen in the interval $(.2,.8)$,
1073: with $\theta\I \approx \theta\E $.
1074: The variables now include the activity state $(s,\sigma)$
1075: as well as the four regulated parameters
1076: $h\E $, $h\I $, $w\ee $ and $w\ie $.
1077: 
1078: Figure 7 illustrates the behavior of this system projected
1079: on the $(w\ee ,w\ie )$ plane,
1080: for the following parameter values:
1081: $w\ei = 10$, $w\ii = 6$, $\rho=.05$, $\theta\ee =.01$, $\varepsilon\ee =.01$,
1082: $\theta\ie =.01$, $\varepsilon\ie =-.005$,
1083: $\theta\E = .5$, $\varepsilon\E = .005$,
1084: $\theta\I = .5$, $\varepsilon\I = .002$.
1085: In the sequel, this parameter setting will be referred to as {\em standard.}
1086: In a first stage, the system converges to a doubly critical point $F$
1087: as described above;
1088: each such point $F$ belongs to the common boundary of the regions
1089: of oscillation, high steady firing, and low steady firing.
1090: Although we cannot thoroughly characterize the surface of $F$ points
1091: in the four-dimensional $(w\ee , w\ie , h\E , h\I )$ space
1092: as we did in the $(w\ee ,w\ie )$ plane for the reduced system,
1093: there is, as remarked above,
1094: a functional equivalence with the $\cal S$ surface.
1095: Note that the projection of the $F$ surface on the $(w\ee ,w\ie )$ plane
1096: has a shape quite similar to that of $\cal S$ in the reduced system.
1097: As before, when the system reaches a point $F$,
1098: all variables settle in a slow, synchronous, almost-periodic motion.
1099: The oscillation of $s$ and $\sigma $ is a nearly rectangular wave,
1100: the system spending nearly all its time in the two corners of the square,
1101: where the relative amount of time spent in each corner
1102: is determined according to the value of parameter $\theta\E (\approx \theta\I )$.
1103: As before too, the first stage,
1104: which consists of the convergence to a doubly critical point $F$,
1105: is robust against parameter changes;
1106: most parameters can be individually
1107: varied over several orders of magnitude
1108: without qualitatively affecting this part of the behavior.
1109: 
1110: The second stage, consisting of a much slower motion on the $F$ surface,
1111: depends on the values of the various parameters.
1112: For most parameter settings,
1113: including the standard set (see above),
1114: the behavior on this critical surface is a slow, {\em simple,} periodic motion,
1115: of large amplitude in $(s,\sigma)$
1116: and very small amplitude in $(w\ee ,w\ie )$.
1117: The system eventually settles in a periodic attractor of this simple type,
1118: denoted again $G$ in Figure 7.
1119: Figure 8a shows the $(s,\sigma)$ projection of this attractor
1120: for the standard parameter set;
1121: its $(w\ee ,w\ie )$ projection
1122: is a small cycle around point $G$,
1123: whose nullcline diagram
1124: is similar to the one shown in Figure 5 (largely overlapping nullclines).
1125: 
1126: There exists however a small region of parameter space,
1127: mostly around $\varepsilon\I \approx \varepsilon\E$,
1128: for which a variety of more complex behaviors
1129: are observed during the second stage.
1130: The following two cases are examples of such complex behavior.
1131: For parameters as above (standard) except that
1132: $\varepsilon\E = .0051$,
1133: $\varepsilon\I = .0046$,
1134: and $\theta\ee =.011$,
1135: the system settles in a complex quasi-periodic motion (Figure 8b).
1136: For parameters as standard except that $\varepsilon\I = \varepsilon\E = .005$,
1137: the system displays strongly chaotic behavior (Figure 8, c--e).\footnote
1138: {This behavior takes place only
1139: for {\em some} initial values in the $(w\ee ,w\ie )$ plane;
1140: other initial values converge to a point attractor.}
1141: Both of these behaviors are actually {\em attractors,}
1142: reached after considerable time,
1143: yet similar behaviors also take place
1144: while the system is still moving slowly on the critical surface.
1145: 
1146: To summarize, both in the reduced and in the full system,
1147: convergence to a doubly critical surface
1148: between the regions of fast oscillations
1149: and of high and low steady firing
1150: takes place reliably for a broad range of parameters.
1151: Once this doubly critical surface is reached,
1152: motion becomes slow,
1153: depends on parameters,
1154: and, when examined in detail,
1155: reveals a variety of behaviors,
1156: ranging from simple periodic firing to chaos.
1157: 
1158: \section{Discussion}
1159: \label{disc}
1160: 
1161: This paper proposes that a regulation mechanism
1162: underlies criticality in brain dynamics.
1163: In such a scheme, regulation stabilizes the dynamics near an instability.
1164: The force driving the system towards criticality
1165: is a covariance-governed modification of synaptic efficacies
1166: in a recurrent network.
1167: Although it has been argued that criticality
1168: in some physical systems may be self-organized (Bak et al. 1987),
1169: this phenomenon may not be very widespread.
1170: The nervous system is moreover regulated homeostatically
1171: to withstand perturbations of various sorts.
1172: It is then of interest to explain
1173: how criticality in brain dynamics may nevertheless arise,
1174: from a specific, well-documented, mechanism of synaptic plasticity.
1175: 
1176: The chief motivation for viewing the dynamics of the nervous sytem as critical
1177: is the observation that brains are very sensitive organs.
1178: Not only do our brains draw distinctions between stimuli
1179: that differ only in minute details,
1180: they also allow us to establish subtle yet clear-cut boundaries
1181: between cognitive categories at different levels.
1182: The various manifestations of hyperacuity in sensory systems
1183: may be no more than elaborate forms of signal amplification;
1184: however, in higher cognitive functions
1185: such as language, abstract thinking, or, for instance, artistic composition,
1186: the ability to create a new category by drawing a fine line---an
1187: ability which manifests itself early in life
1188: and stays with us for a long time---argues
1189: in favor of regulated criticality.
1190: Such a mechanism appears to be necessary
1191: in order to explain how sensitivity is {\em maintained}
1192: in the face of the profound changes
1193: that affect the connectivity of the brain
1194: throughout development and learning.
1195: If no such mechanism were present,
1196: one would expect that the ongoing modification
1197: of the networks which carry mental representations
1198: would soon bring these networks to generic states;
1199: as mentioned in Section \ref{intro},
1200: a dynamical system in a generic state does not show high susceptibility
1201: to external influences.
1202: 
1203: The emergence of new cognitive categories
1204: may in effect be likened to a process of morphogenesis in embryology,
1205: or differentiation in cell biology.
1206: A biological structure that is about to undergo differentiation
1207: is at that particular instant of time unstable,
1208: and, as the well-studied mechanism of {\em induction} shows,
1209: highly susceptible to external signals.
1210: From a dynamical-system perspective,
1211: the emergence of qualitatively new behavior,
1212: e.g., the splitting of one attractor into two, is a bifurcation.
1213: The complexification of an individual's cognitive apparatus
1214: in the course of his or her life
1215: may be viewed as an open-ended sequence of such bifurcations.
1216: Such an interpretation has been defended by Ren\'e Thom (1975),
1217: and related ideas have been expressed by several authors
1218: (e.g. van der Maas and Molenaar 1992).
1219: Thom (1975) also suggested that structurally stable non-generic singularities
1220: may arise from a process he termed the {\em stabilization of thresholds;}
1221: this process itself would result, in various biological contexts,
1222: from the reinforcment of homeostatic mechanisms.\footnote
1223: {We thank Jean Petitot for pointing out to us
1224: that regulated criticality as proposed here
1225: is closely related to Thom's ideas.}
1226: 
1227: The covariance plasticity rule we use is linear and straightforward.
1228: Equation \ref{regee} may be viewed
1229: as a mean-field version of the covariance rule
1230: used in the associative-memory literature
1231: (see e.g. Willshaw and Dayan 1990).
1232: However, we make a rather different use of this rule.
1233: In an associative-memory model,
1234: pre- and post-synaptic activities are generally assumed to be independent,
1235: yielding a zero expected value for the covariance.
1236: Weights are modified according to the {\em instantaneous} covariance,
1237: and, as noted in Dayan and Sejnowski (1993),
1238: storage is marked by the departures of the empirical average of this quantity
1239: from its expected value, which is zero.
1240: In our model,
1241: the expected covariance is positive in the oscillatory phase.
1242: The regulation mechanism acts on a {\em slow} time scale,
1243: and, although we use the instantaneous covariance
1244: in the modification rule (Equation \ref{regee}),
1245: we might as well have used the time-averaged covariance;
1246: fast variations of the instantaneous covariance are actually smoothed out
1247: in the integration of the differential equation.
1248: Of course by the very principle of regulation proposed,
1249: the system does not dwell in the oscillatory phase;
1250: in the regulated state, the average covariance is low.
1251: 
1252: The other major difference between the situation studied here
1253: and the associative-memory paradigm
1254: is the assumption of uniform weights.
1255: As noted in Section \ref{regulation},
1256: the covariance in our uniform-weight network
1257: is simply the variance of the population-averaged activity about its mean,
1258: and is always non-negative.
1259: This makes it necessary to subtract from it a positive constant $\theta\ee$
1260: in order to allow for {\em decreases} of the weights.
1261: Thus, whereas in associative-memory models a synaptic weight
1262: decreases as a result of {\em negative} instantaneous covariance
1263: between the pre- and post-synaptic neurons,
1264: the condition for weight decrease in our model
1265: is that the mean covariance be small or zero,
1266: which happens when the system is at rest in a point attractor,
1267: of either low or high activity.
1268: 
1269: The uniform-weight network used in the present study
1270: lends itself to a detailed mathematical/numerical analysis.
1271: We performed a bifurcation analysis of the continuous-time differential system
1272: that describes the behavior of this network in the thermodynamic limit.
1273: This analysis (Section \ref{full}) reveals, among other features,
1274: the existence of a critical surface ${\cal S}_h$ in parameter space,
1275: where the system undergoes an abrupt transition
1276: from oscillatory behavior to high-rate steady firing.
1277: We showed (Section \ref{full})
1278: that Hebbian modification of the E-to-E synaptic weights
1279: drives the system toward this surface ${\cal S}_h$;
1280: this is the main mechanism of regulated criticality proposed.
1281: 
1282: However, when the system is at a general position on ${\cal S}_h$,
1283: it spends most of its time in the high-activity state,
1284: which is undesirable.
1285: A more realistic situation results from introducing some form of symmetry
1286: between the high- and low-activity phases,
1287: making use of the firing thresholds.
1288: We investigated two ways---formally
1289: different but functionally equivalent---to do this.
1290: The mathematically simpler way
1291: is to enforce an accurate symmetry on the dynamics,
1292: by imposing an appropriate relationship between
1293: the firing thresholds and the synaptic weights (equations \ref{chvar}).
1294: This results in a {\em reduced} system, with only four parameters;
1295: in this system,
1296: there occurs a {\em double} bifurcation when the system traverses
1297: the critical surface---now denoted $\cal S$---separating
1298: the oscillatory phase from the bistable, high/low, steady-firing phase
1299: (Section \ref{model}).
1300: Regulation of the sole E-to-E weight brings the system
1301: to this doubly critical surface $\cal S$ (Section \ref{reduced}).
1302: 
1303: A biologically more satisfactory solution
1304: is to {\em regulate} one or both of the firing thresholds
1305: so as to control the mean firing rates (Section \ref{full}).
1306: Thus, when we regulate the threshold for the excitatory neurons
1307: in addition to the E-to-E weight,
1308: the system converges to the intersection of ${\cal S}_h$
1309: with another critical surface, ${\cal S}_l$,
1310: which separates the oscillatory phase from the low-activity fixed-point region.
1311: Intersection points between ${\cal S}_h$ and ${\cal S}_l$
1312: are again doubly critical,
1313: and they attract the system for a wide range of parameter values.
1314: 
1315: When the system is on this doubly critical surface,
1316: it takes only a small weight perturbation
1317: to induce either of the three behaviors:
1318: intrinsic oscillation (region $\cal P$ of section \ref{full}),
1319: high activity, quiescence.
1320: It is easily seen that, when in this state,
1321: the network can also be efficiently driven
1322: by a small-amplitude time-varying signal, i.e., an external field;
1323: it is thus highly sensitive to input.
1324: The stabilization at the boundary of a region of oscillatory behavior
1325: appears to be consistent with at least one conclusion
1326: that can be safely drawn from the recent literature on cortical oscillations,
1327: namely the fact that the precise conditions under which
1328: these oscillations occur are difficult to pin down.
1329: 
1330: We further investigated the effect of regulating
1331: the E-to-I weight in addition to the E-to-E weight,
1332: according to a similar covariance rule.
1333: We showed that regulating these two weights
1334: as well as the two firing thresholds results,
1335: under appropriate parametric conditions,
1336: in convergence to an even more degenerate state.
1337: When the system is in that state,
1338: its flow vanishes on an entire one-dimensional curve
1339: in the two-dimensional phase space,
1340: instead of on isolated points.
1341: This convergence is slow and parameter-dependent,
1342: yet it is interesting to note that
1343: when the system is in or near this highly degenerate state
1344: it exhibits a range of diverse behaviors,
1345: including chaos (Section \ref{full}).
1346: The chaotic behavior shown in Figure 8, c--e
1347: consists of an irregular sequence of spontaneous transitions
1348: between the three fundamental phases of the system:
1349: oscillatory, high-activity, low-activity.
1350: 
1351: While the uniform-weight network studied in this paper
1352: lends itself to a convenient mathematical analysis,
1353: it would be interesting to know whether critical behavior
1354: may arise from {\em local} covariance plasticity,
1355: where synaptic changes are made to depend on pre- and post-synaptic activities
1356: relative to individual synapses.
1357: This question should be focused by considerations
1358: about the elaborate forms of input sensitivity
1359: that could play a role in higher brain functions.
1360: 
1361: {\bf Acknowledgments} Important contributions in the early stages of this work
1362: were made by Howard Gutowitz.
1363: We thank Rob de Boer for making available to us the use of the software {\em GRIND,}
1364: which proved an extremely valuable tool in this study.
1365: Various suggestions were made by a number of colleagues,
1366: from whom we wish to thank in particular G\'erard Weisbuch, Jean Petitot,
1367: G\'erard Toulouse, Jean-Pierre Nadal, Claude Meunier, David Hansel,
1368: Ha\"{\i}m Sompolinsky, and Stuart Geman.
1369: 
1370: \vglue 1cm
1371: 
1372: \noindent{\bf References}
1373: 
1374: \begin{description}
1375: 
1376: \item Artola, A., Br\"ocher, S., and Singer, W. 1990.
1377: Different voltage-dependent thresholds for inducing long-term depression
1378: and long-term potentiation in slices of rat visual cortex.
1379: {\em Nature}, {\bf 347}, 69--72.
1380: 
1381: \item Bak, P., Tang, C., and Wiesenfeld, K. 1987.
1382: Self-Organized Criticality: An Explanation of $1/f$ Noise.
1383: {\em Phys. Rev. Lett.}, {\bf 59}, 381--384.
1384: 
1385: \item Bienenstock, E., Cooper, L.N., and Munro, P. 1982.
1386: Theory for the development of neuron selectivity:
1387: Orientation specificity and binocular interaction in visual cortex.
1388: {\em J. Neurosci.}, {\bf 2}, 32--48.
1389: 
1390: \item Borisyuk, R.M., and Kirillov, A.B. 1992.
1391: Bifurcation analysis of a neural network model.
1392: {\em Biol. Cybernetics} {\bf 66}, 319--325.
1393: 
1394: \item Dayan, P., and Willsahw, D. 1991.
1395: Optimising synaptic learning rules in linear associative memories.
1396: {\em Biol. Cybernetics} {\bf 65}, 253--265.
1397: 
1398: \item Dayan, P., and Sejnowski, T.J. 1993.
1399: The Variance of Covariance Rules for Associative Matrix Memories
1400: and Reinforcement Learning.
1401: {\em Neural Computation}, {\bf 5}, 205--209.
1402: 
1403: \item Dudek, S.M., and Bear, M.F. 1992.
1404: Homosynaptic long-term depression in area CA1 of hippocampus
1405: and effects of {\em N}-methyl-D-aspartate receptor blockade.
1406: {\em Proc. Natl. Acad. Sci. USA}, {\bf 89}, 4363--4367.
1407: 
1408: \item Glauber, R.J. 1963.
1409: Time-dependent Statistics of the Ising Model.
1410: {\em J. Math. Phys.} 4, 294--307.
1411: 
1412: \item Hebb D.O. 1949.
1413: {\em The Organization of Behavior.}
1414: Wiley, New York.
1415: 
1416: \item Fr\'egnac, Y., Shulz, D., Thorpe, S., and Bienenstock, E. 1988.
1417: A cellular analogue of visual cortical plasticity.
1418: {\em Nature}, {\bf 33}, 367--370.
1419: 
1420: \item Fr\'egnac, Y., Shulz, D., Thorpe, S., and Bienenstock, E. 1992.
1421: Cellular analogs of visual cortical epigenesis.
1422: I--Plasticity of orientation selectivity.
1423: {\em Journal of Neuroscience}, {\bf 12}, 1280--1300.
1424: 
1425: \item Guckenheimer, J., and Holmes, P. 1983
1426: {\em Nonlinear oscillations, dynamical systems,
1427: and bifurcations of vector fields.}
1428: Springer-Verlag, New York.
1429: 
1430: \item Langton, C.R. 1990.
1431: Computation at the edge of chaos:
1432: Phase transitions and emergent computation.
1433: {\em Physica D}, {\bf 42}, 12--37.
1434: 
1435: \item Linsker, R. 1986.
1436: From basic network principles to neural architecture:
1437: Emergence of spatial opponent cells.
1438: {\em Proc. Natl. Acad. Sci. USA}, {\bf 83}, 7508--7512.
1439: 
1440: \item Metzger, Y., and Lehmann, D. 1990.
1441: Learning Temporal Sequences by Local Synaptic Changes.
1442: {\em Network: Computation in Neural Systems}, {\bf 1}(3), 169--188.
1443: 
1444: \item Metzger, Y., and Lehmann, D. 1994.
1445: Learning Temporal Sequences by Excitatory Synaptic Changes Only.
1446: {\em Network: Computation in Neural Systems}, {\bf 5}, 89--99.
1447: 
1448: \item Rubin, N. 1988.
1449: Equilibrium and oscillations in stochastic neural networks.
1450: Master's thesis, The Hebrew University, Jerusalem, Israel (in Hebrew).
1451: 
1452: \item Schuster, H.G., and Wagner, P. 1990
1453: A model for neuronal oscillations in the visual cortex.
1454: 1. Mean-field theory and derivation of the phase equations.
1455: {\em Biological Cybernetics}, {\bf 64}, 77--82.
1456: 
1457: \item Sejnowski, T.J. 1977a.
1458: Storing covariance with nonlinearly interacting neurons.
1459: {\em J. Math. Biol.}, {\bf 4}, 303--321.
1460: 
1461: \item Sejnowski, T.J. 1977b.
1462: Statistical constraints on synaptic plasticity.
1463: {\em J. Theor. Biol.}, {\bf 69}, 385--389.
1464: 
1465: \item Stanton, P.K. and Sejnowski, T.J. 1989.
1466: Associative long-term depression in the hippocampus
1467: induced by hebbian covariance.
1468: {\em Nature}, {\bf 339}, 215, May 1989.
1469: 
1470: \item Thom, R. 1975.
1471: {\em Structural Stability and Morphogenesis:
1472: an Outline of a General Theory of Models}.
1473: W. A. Benjamin, Reading, Mass.
1474: 
1475: \item van der Maas, H.L.J., and Molenaar, P.C.M. 1992.
1476: Stagewise Cognitive Development: An Application of Catastrophe Theory.
1477: {\em Psychological Review}, {\bf 99}(3), 395--417.
1478: 
1479: \item Willshaw, D., and Dayan, P. 1990.
1480: Optimal Plasticity from Matrix Memories: What Goes Up Must Come Down.
1481: {\em Neural Computation}, {\bf 2}, 85--93.
1482: 
1483: \end{description}
1484: 
1485: 
1486: 
1487: \renewcommand{\baselinestretch}{1}
1488: \small
1489: 
1490: \newpage
1491: 
1492: \centerline{\psfig{file=fig1.ps,height=20cm}} 
1493: \begin{quote}
1494: Figure 1: Mean activities of excitatory and inhibitory populations
1495: in a moderate-size uniform-weight system
1496: exhibiting oscillatory behavior ($N=70$; Glauber dynamics).
1497: \end{quote}
1498: 
1499: 
1500: \newpage
1501: 
1502: \centerline{\psfig{file=fig2a.ps,height=6cm}} 
1503: \centerline{\psfig{file=fig2b.ps,height=6cm}} 
1504: \centerline{\psfig{file=fig2c.ps,height=6cm}} 
1505: \begin{quote}
1506: Figure 2: Behavior of mean-field system
1507: for different values of the E-to-E synaptic weight $w\ee $.
1508: Diagrams show trajectories and nullclines.
1509: (a) $w\ee = 12$ (parameters are as in Figure 1);
1510: all trajectories converge to a limit cycle.
1511: (b) $w\ee = \hat w\ee \sn \approx 14.22$;
1512: the system is at the saddlenode bifurcation:
1513: nullclines are tangent to each other
1514: (no trajectories shown).
1515: (c) $w\ee = 15$; nullclines intersect,
1516: the periodic attractor has vanished,
1517: two point attractors have appeared.
1518: \end{quote}
1519: 
1520: \newpage
1521: 
1522: \centerline{\psfig{file=fig3a.ps,height=8cm}}
1523: \centerline{\psfig{file=fig3b.ps,height=8cm}}
1524: \vglue 2cm
1525: \begin{quote}
1526: Figure 3: Bifurcation diagram in $(w\ee ,w\ie )$ plane.
1527: (a) Diagram shows three regions,
1528: characterizing different attractor configurations.
1529: Region $\cal O$: single point attractor, of intermediate activity level;
1530: region $\cal P$: periodic attractor, as depicted in Figure 2a;
1531: region $\cal T$: two point attractors,
1532: of high and low activity, as depicted in Figure 2c.
1533: Transitions between regions occur through bifurcations,
1534: of Hopf type between $\cal O$ and $\cal P$,
1535: of saddlenode type between $\cal P$ and $\cal T$ (curve $\cal S$),
1536: and of pitchfork type between $\cal O$ and $\cal T$.
1537: (b) Average covariance along four different lines of constant $w\ie $
1538: in the $(w\ee ,w\ie )$ plane.
1539: Note the sharp variation of the covariance on the critical line $\cal S$
1540: separating $\cal P$ from $\cal T$.
1541: \end{quote}
1542: 
1543: \newpage
1544: 
1545: \centerline{\psfig{file=fig4a.ps,height=6cm}} 
1546: \centerline{\psfig{file=fig4b.ps,height=6cm}} 
1547: \centerline{\psfig{file=fig4c.ps,height=6cm}} 
1548: \begin{quote}
1549: Figure 4: Regulation of system \ref{sys_red} by covariance plasticity.
1550: (a) $w\ee $ is regulated, $w\ie $ is constant:
1551: state converges to critical surface $\cal S$.
1552: (b) $w\ie $ is regulated, $w\ee $ is constant:
1553: state converges to critical surface $\cal S$.
1554: (c) both $w\ee $ and $w\ie $ are regulated:
1555: state converges to a particular point, denoted $G$,
1556: on critical surface $\cal S$.
1557: \end{quote}
1558: 
1559: \newpage \vglue 2cm
1560: 
1561: \centerline{\psfig{file=fig5.ps,height=16cm}} 
1562: \vglue 2cm
1563: \begin{quote}
1564: Figure 5: Nullcline diagram at point $G$ (see figure 4c).
1565: Nullclines overlap almost perfectly over much of the interval $[-.5,.5]$.
1566: \end{quote}
1567: 
1568: \newpage
1569: 
1570: \centerline{\psfig{file=fig6a.ps,height=6cm}} 
1571: \centerline{\psfig{file=fig6b.ps,height=6cm}} 
1572: \centerline{\psfig{file=fig6c.ps,height=6cm}} 
1573: \begin{quote}
1574: Figure 6: Regulation of two parameters in system \ref{sys_full}.
1575: (a) Bifurcation diagram in $(w\ee ,h\E )$ plane.
1576: (b) Regulation of $w\ee $ and $h\E $ causes convergence to point $F$,
1577: the intersection of critical lines ${\cal S}_h$ and ${\cal S}_l$.
1578: (c) Nullcline diagram at $F$.
1579: \end{quote}
1580: 
1581: \newpage \vglue 2cm
1582: 
1583: \centerline{\psfig{file=fig7.ps,height=16cm}} 
1584: \begin{quote}
1585: Figure 7: Behavior of full system under simultaneous regulation of four parameters.
1586: Diagram shows projection on $(w\ee ,w\ie )$ plane,
1587: illustrating the similarity of behavior with reduced system
1588: (compare with Figure 4c, but note difference of scales).
1589: Limits of the attraction basin to the left are roughly indicated
1590: by the starting points of the trajectories shown;
1591: attraction basin is unbounded in all other directions.
1592: \end{quote}
1593: 
1594: \newpage
1595: 
1596:   \begin{tabular}{cc}
1597:     \psfig{file=fig8a.ps,height=6cm} &  \psfig{file=fig8b.ps,height=6cm} \\
1598:     \psfig{file=fig8c.ps,height=6cm} &  \psfig{file=fig8d.ps,height=6cm} \\
1599:     \psfig{file=fig8e.ps,height=6cm} 
1600:   \end{tabular}
1601: \begin{quote}
1602: Figure 8: Various behaviors of regulated full system
1603: after it has reached critical surface (Figure 7).
1604: Diagrams show $s(t)$ for three slightly different parameter settings (see text);
1605: in all cases, the projection of the motion on the $(w\ee ,w\ie )$ plane
1606: remains of small amplitude.
1607: (a) Simple periodic attractor, point $G$ of Figure 7;
1608: similar periodic attractors are reached for most parameter settings.
1609: (b) Complex quasi-periodic attractor.
1610: (c--e) Chaotic attractor;
1611: for a given parameter setting,
1612: three diagrams corresponding to different instants of time
1613: and different lengths of time;
1614: characteristic are the irregular transitions
1615: between the high-activity, low-activity, and oscillatory phases.
1616: 
1617: \end{quote}
1618: 
1619: \end{document} 
1620: