q-bio0605012/main.tex
1: % Note:
2: % 1. draft option in documentclass suppresses figures from appearing
3: % 2. Using hyperref to allow active links to references turns
4: %    off the collapsing of successive references in a
5: %    single citation. So use hyperref only for
6: %    preprints that will be viewed onscreen, not
7: %    printed out.
8: 
9: % \documentclass[preprint,pre,showpacs,superscriptaddress]{revtex4}
10: %\documentclass[pre,showpacs,superscriptaddress]{revtex4}
11: \documentclass[twocolumn,showpacs,superscriptaddress]{revtex4}
12: 
13: \usepackage{amsmath}
14: \usepackage{amssymb}
15: \usepackage{graphicx}
16: %\usepackage[colorlinks=true,linkcolor=blue]{hyperref}
17: 
18: \newcommand{\hvcra}{HVC${}_\textrm{RA}$}
19: \newcommand{\cai}{\ensuremath{\left[{\rm Ca}\right]_i}}
20: 
21: \begin{document}
22: 
23: 
24: \title{Stable Propagation of a Burst Through a
25:  One-Dimensional Homogeneous Excitatory Chain
26:  Model of Songbird Nucleus HVC}
27: 
28: \author{MengRu Li}
29: \affiliation{Department of Physics, Duke University, Durham
30: NC 27708-0305}
31: 
32: \author{Henry Greenside}
33: \affiliation{Department of Physics, Duke University, Durham
34: NC 27708-0305}
35: 
36: \date{\today}
37: 
38: \begin{abstract}
39:   We demonstrate numerically that a brief burst
40:   consisting of two to six spikes can propagate in a
41:   stable manner through a one-dimensional homogeneous
42:   feedforward chain of non-bursting neurons with
43:   excitatory synaptic connections. Our results are
44:   obtained for two kinds of neuronal models, leaky
45:   integrate-and-fire (LIF) neurons and Hodgkin-Huxley
46:   (HH) neurons with five conductances.  Over a range of
47:   parameters such as the maximum synaptic conductance,
48:   both kinds of chains are found to have multiple
49:   attractors of propagating bursts, with each attractor
50:   being distinguished by the number of spikes and total
51:   duration of the propagating burst.  These results
52:   make plausible the hypothesis that sparse
53:   precisely-timed sequential bursts observed in
54:   projection neurons of nucleus~HVC of a singing zebra
55:   finch are intrinsic and causally related.
56: \end{abstract}
57: 
58: \maketitle
59: 
60: 
61: \section{Introduction}
62: \label{sec:intro}
63: 
64: A question of great interest to neurobiology is how
65: animals learn to generate temporal patterns of muscle
66: activation. An example that has been much studied in
67: recent years because of its relevance to human
68: speech~\cite{Doupe99} and because of the rich variety
69: of possible experiments is how songbirds learn to sing
70: a song by auditory-guided vocal
71: feedback~\cite{Catchpole95}. A young male bird first
72: memorizes the song of an adult male of the same
73: species. Then over many months, over many iterations
74: (more than 50,000 iterations for a zebra
75: finch~\cite{Johnson02}), and by just listening to his
76: own vocalizations, a young male learns to vary the
77: activation of its respiratory and syringeal muscles
78: until his song is able to match accurately the original
79: memorized song. This is an impressive feat given that
80: an adult song might last several seconds, that there is
81: auditory structure that lasts less than 10~ms, that
82: many syringeal and respiratory muscles need to be
83: coordinated, and that some species of songbirds are
84: able to learn and sing many different songs. Nearly all
85: details of this process remain poorly understood, in
86: particular how the young male memorizes a complex song,
87: and how auditory feedback is used to adjust the pattern
88: of muscle activation until the songbird is able to
89: reproduce accurately the memorized song.
90: 
91: Studies of songbirds have shown that certain
92: anatomically and physiologically distinct brain regions
93: called nuclei are associated with the recognition,
94: learning, and production of song (see
95: Fig.~\ref{fig:songbird-brain}). A recent experiment by
96: Hahnloser et al~\cite{Hahnloser02} recorded
97: extracellular action potentials (spikes) from neurons
98: in awake singing male zebra finches and found that the
99: neurons in the nucleus~HVC\cite{Reiner04} that project
100: to the robust nucleus of the arcopallium (abbreviated
101: as~RA~\cite{Reiner04}) have the remarkable properties
102: of firing sparsely and precisely during singing. (In
103: the following, HVC~neurons that project to~RA will be
104: abbreviated as \hvcra~neurons.)  Typically, each
105: \hvcra~neuron fires a brief burst of about 3-4 spikes
106: once per song motif with each burst lasting about 6~ms.
107: (A motif is a cluster of distinct auditory syllables
108: that is repeated as a single pattern and that lasts an
109: average of 0.6~s for zebra finch songs.)  Measurements
110: during successive motifs from a given adult male bird
111: show that each burst from a particular \hvcra~neuron is
112: aligned with certain acoustic features of the motif to
113: a precision of about 0.7~ms~\cite{Hahnloser02}. These
114: bursts are important to understand since they are
115: believed to provide the temporal framework for
116: organizing the syllables of a song~\cite{Fee04}.
117: 
118: \begin{figure}
119:   \centering
120:   \includegraphics[width=3in]{songbird-brain.eps}
121:   \caption{ Schematic anatomical diagram of a sagittal
122:     cross section of one hemisphere of a zebra finch
123:     brain, showing the major nuclei associated with the
124:     learning and production of song.  The nucleus~HVC
125:     plays a central role since it receives auditory and
126:     other input, it drives the premotor
127:     nucleus~RA~\cite{Hahnloser02}, and it sends
128:     information to a parallel anterior forebrain
129:     pathway (nuclei~X, DLM, and~LMAN) that play a role
130:     in the learning and maintenance of a song. There
131:     are two HVC~nuclei, one on the left and one on the
132:     right side of the songbird's brain.}
133:   \label{fig:songbird-brain}
134: \end{figure}
135: 
136: The observation that each \hvcra~neuron bursts just
137: once per motif and other data which show that there is
138: a reproducible temporal ordering of \hvcra~neurons
139: according to when they fire their single burst per
140: motif~\cite{Hahnloser02}, suggest the hypothesis that
141: bursts propagate along a feedforward network of
142: excitatory \hvcra~neurons~\cite{Fee04}. The idea is
143: similar to a path of dominoes such that the falling of
144: one domino causes the next domino in the path to
145: topple. Here, a given \hvcra~neuron receives one or
146: more bursts from other \hvcra~neurons that have fired
147: recently during the motif. In response, the given
148: neuron itself fires a burst that helps to activate
149: neurons that fire later during the motif. That each
150: \hvcra~neuron fires exactly once per motif implies that
151: the connections are feedforward (although this
152: assumption can be weakened if a recurrent burst arrives
153: while some earlier neuron is still in a refractory
154: state~\cite{Jin05}). The bursts continue to propagate
155: until the chain ends and the motif stops, or perhaps
156: there is a branch point to a different chain whose
157: dynamics generate a different motif.
158: 
159: While there has been much previous theoretical research
160: concerning when spikes can propagate along excitatory
161: networks of different
162: kinds~\cite{Abeles91,Golomb97,Diesmann99,Reyes03,Litvak03,Osan04,Vogels05},
163: the songbird experiments on~HVC of Hahnloser et al
164: during singing~\cite{Hahnloser02} and related
165: experiments by Mooney of bursts in HVC projection
166: neurons during audition~\cite{Mooney00} motivate asking
167: new theoretical questions about the propagation of a
168: burst through various kinds of networks.
169: 
170: In this paper, we show that a minimal feedforward
171: architecture---a one-dimensional (1d), homogeneous,
172: excitatory chain of non-bursting neurons---can support
173: the stable propagation of a brief high-frequency burst.
174: By studying two kinds of chains, one with leaky
175: integrate-and-fire neurons (abbreviated as LIF~neurons)
176: and one with more realistic single-compartment
177: conductance-based neurons (abbreviated as HH~neurons
178: for Hodgkin-Huxley~like), we show that a burst can
179: propagate in a stable manner that does not require a
180: careful choice of neuronal model nor a careful tuning
181: of model parameters.  Depending on the values of
182: parameters such as the maximum synaptic conductance, we
183: find that each kind of 1d~chain (LIF or~HH) has
184: multiple attractors that differ in the number of spikes
185: and in the total width of the propagating burst.
186: 
187: Our calculations are the first to demonstrate that a
188: brief high-frequency burst similar to those observed
189: in~HVC can propagate in a stable manner along a simple
190: excitatory chain of neurons. (A similar result was
191: announced independently by Jin and
192: collaborators~\cite{Jin05}, who studied a more
193: complicated model. We discuss this preprint briefly in
194: Section~\ref{sec:conclusions} below.) Our results make
195: plausible one of the simplest explanations of the
196: Hahnloser et al result, namely that the observed bursts
197: of \hvcra~neurons during singing are intrinsic to~HVC
198: in that external input from other brain regions is not
199: needed to generate the bursts (except the first burst),
200: and in that the bursts are causally related such that
201: one burst initiates the generation of the next
202: burst. Our results also make some predictions such as
203: the existence of multistability (different kinds of
204: propagating bursts can occur in the same chain
205: depending on how the chain is activated), and that
206: transitions from one kind of propagating burst to
207: another kind (differing in the number of spikes and
208: burst width) can occur as parameters are varied. New
209: experimental studies of~HVC, especially using optical
210: methods with high time resolution~\cite{Ohki05} that
211: can examine the spatiotemporal activity of many neurons
212: simultaneously, may help to confirm these ideas.
213: 
214: The rest of this paper is organized as follows.
215: Section~\ref{sec:methods} discusses some details of the
216: two classes of mathematical neurons that we use in the
217: 1d~chains, and how the excitatory synaptic currents are
218: modeled.  Section~\ref{sec:results-discussion} presents
219: numerical results mainly for a 1d~chain of~HH neurons
220: (as opposed to LIF~neurons), especially the existence
221: of different attractors corresponding to different
222: kinds of propagating bursts.  The paper concludes in
223: Section~\ref{sec:conclusions}, where key results are
224: summarized. We also discuss there how our results are
225: related to alternative hypotheses such as that the
226: bursting arises from propagation through synfire
227: chains~\cite{Abeles91} or by a central pattern
228: generator (abbreviated as CPG).
229: 
230: In a second paper~\cite{Li06II,Li06thesis}, we will
231: discuss how noise and network heterogeneities affect
232: the propagation of bursts through a synfire chain. This
233: second paper shows that our key conclusions still hold:
234: a brief high-frequency burst can still propagate in a
235: stable manner for a range of noise strengths and for
236: various amounts of heterogeneity. However, we also find
237: that single spikes, even if synchronized across a
238: synfire pool, do not propagate in a stable way for most
239: of the parameters studied, which suggests that short
240: bursts of several spikes are important for achieving
241: robust propagation through realistic synfire chains.
242: 
243: 
244: \section{Methods}
245: \label{sec:methods}
246: 
247: In this section, we discuss some experimental data that
248: justifies the study of a one-dimensional chain of
249: excitatory neurons, and we also point out some of the
250: experimental details that are ignored in our model. We
251: also discuss some details of the integrate-and-fire and
252: conductance-based neuronal models with excitatory
253: synapses that we use in the
254: 1d~chains. Appendix~\ref{appendix:equations-parameters}
255: provides further details of equations and parameter
256: values.
257: 
258: \subsection{A Homogeneous 1d~Chain of Excitatory Neurons}
259: 
260: \label{sec:1d-chain-motivation}
261: 
262: The physiological properties of HVC~neurons and how
263: they are interconnected within~HVC are poorly
264: understood~\cite{Kubota98,Dutar98,Mooney00,Mooney05,Wild05}
265: so that it is not possible at this time to develop a
266: quantitative model of the HVC~microcircuitry, say at
267: the level of hippocampus
268: models~\cite{Traub99}. Researchers have
269: shown~\cite{Dutar98,Mooney00} that there are at least
270: three main classes of neurons: excitatory neurons that
271: project to nucleus~RA, excitatory neurons that project
272: to area~X in the anterior forebrain pathway, and
273: inhibitory interneurons~\cite{Dutar98,Mooney00} that
274: connect only to other neurons within~HVC. Recent
275: paired-electrode recordings by Mooney and
276: Prather~\cite{Mooney05} have shown that each type
277: of~HVC neuron makes local connections with the other
278: two types of HVC~neurons but details of the connections
279: such as the number, kinds, and strengths of synapses
280: are incompletely known.
281: 
282: Of special importance for this paper is the
283: experimental observation that \hvcra~neurons synapse
284: with other \hvcra~neurons~\cite{Mooney05}. Thus it is
285: possible for a feedforward network of excitatory
286: neurons to exist in~HVC, although we emphasize that
287: there is no evidence presently for such a
288: network. Further, there are of order~40,000
289: \hvcra~neurons in HVC~\cite{Fee04}, which should be
290: sufficiently many to create a chain whose dynamics
291: spans a motif or even several motifs. (Since the
292: observed bursts in \hvcra~neurons last
293: 6~ms~\cite{Hahnloser02}, a chain of about one hundred
294: pools of neurons would suffice to span a motif
295: of~0.6~s, provided that the bursts do not overlap in
296: time.)
297: 
298: To determine in principle whether the propagation of a
299: burst can explain the experimental
300: data~\cite{Hahnloser02,Mooney00}, we study one of the
301: simplest possible feedforward networks, namely a
302: one-dimensional homogeneous feedforward chain of
303: identical neurons such that each neuron connects via a
304: single identical excitatory synapse with the next
305: neuron of the chain. The assumption of homogeneity is
306: not realistic biologically but reduces the parameter
307: space of the calculations and increases our ability to
308: understand how the existence and properties of a
309: propagating burst depend on parameters.
310: 
311: Our 1d~model leaves out many experimental details (we
312: discuss this further in
313: Section~\ref{sec:results-discussion} below), of which
314: two are especially important, the need for multiple
315: pathways and the presence of inhibitory neurons.
316: First, as pointed out by Abeles~\cite{Abeles91} and
317: others, transmission of information along a
318: one-dimensional chain is not robust since damage to any
319: neuron in the chain can alter or stop the transmission
320: of information. (That~HVC is robust is demonstrated by
321: the fact that the adult song of a zebra finch changes
322: little over the life of the bird, even though there is
323: ongoing neurogenesis and hence a steady turnover of
324: \hvcra~neurons~\cite{NottebohmJN02,Wang02}.) If some
325: kind of propagation occurs in~HVC, there must be
326: parallel pathways.
327: % (not all of which have to be strictly feedforward~\cite{Jin05}).
328: 
329: \begin{figure}[htb]
330:   \centering
331:   \includegraphics[height=2.5in]{synfireScheme.eps}
332:   \caption{ (a) Schematic diagram of a feedforward
333:   synfire chain consisting of successive pools of
334:   neurons (vertical column of circles). Each pool
335:   contains excitatory \hvcra~neurons that fire
336:   approximately synchronously to activate neurons in
337:   the next pool. The horizontal hollow arrow on the
338:   left represents input that can initiate the synfire
339:   chain. The hollow arrows at the top of each pool
340:   indicate efferents that could convey information from
341:   a given pool to other brain areas. (b) Schematic
342:   diagram of the one-dimensional homogeneous
343:   feedforward excitatory chain that we study in this
344:   paper. In the biologically unrealistic but
345:   theoretically convenient case that all neurons in the
346:   synfire chain of~(a) are identical, that all neurons
347:   in one pool connect to neurons in the next pool in
348:   the same way with identical synapses, that axonal
349:   delays between pools can be ignored, and there is no
350:   noise, the 1d-chain will have dynamics identical to a
351:   synfire chain whose spikes have become fully
352:   synchronized. }
353:   \label{fig:syn-fire-chain}
354: \end{figure}
355: 
356: A widely studied example of a feedforward neuronal
357: architecture with multiple pathways is a synfire
358: chain~\cite{Abeles91,Diesmann99,Ikegaya04}, see
359: Fig.~\ref{fig:syn-fire-chain}(a). Although there are
360: variations of this architecture, a synfire chain is
361: usually described as a strictly feedforward chain of
362: pools of neurons such that the neurons in one pool
363: project only to neurons of the next pool. Reliable
364: transmission occurs even in the presence of noise or of
365: network heterogeneities because of the multiple paths,
366: and also because each neuron receives spikes from many
367: neurons of the previous pool and the spikes
368: spontaneously become highly synchronized as propagation
369: continues~\cite{Abeles91,Diesmann99}. The arrival of
370: synchronized spikes at the many synapses of a given
371: synfire neuron causes a large postsynaptic current that
372: triggers new spikes with a probability that can
373: approach certainty.
374: 
375: The idea that the sparse brief bursts observed in
376: \hvcra~neurons are causally connected by sequential
377: firing through some kind of chain has been proposed by
378: experimentalists~\cite{Fee04,Leonardo05} and assumed in
379: several theoretical
380: studies~\cite{Doya98,Abarbanel04jn,Berger04}. In a
381: recent experiment~\cite{Ikegaya04}, calcium imaging and
382: timing studies of intracellular recordings have
383: suggested the existence of synfire chains in the visual
384: cortices of mice and cats but no similar imaging study
385: has yet been carried out for songbirds.
386: 
387: The dynamics of our one-dimensional homogeneous chain
388: is equivalent to the dynamics of a homogeneous
389: noiseless synfire chain (homogeneity here means equal
390: numbers of identical neurons per pool all connected in
391: identical ways from one pool to the next) such that all
392: spikes within a given pool have become perfectly
393: synchronized. This is an important observation since
394: the synapse that connects one neuron to the next in the
395: 1d~chain of Fig.~\ref{fig:syn-fire-chain}(b) then has
396: to be strong to correspond to the many synapses of a
397: neuron in the full synfire chain. In general, noise,
398: heterogeneities, delay times, and imperfect
399: synchronization of spikes cause the dynamics of a
400: synfire chain to differ from that of our 1d~chain so
401: that simulations of real synfire chains are needed to
402: understand these effects on the propagation of a
403: burst~\cite{Jin05,Li06II,Li06thesis}.
404: 
405: A second important experimental detail that we leave
406: out of our model is the presence of the inhibitory
407: HVC~interneurons. Our justification for this assumption
408: is simply that the function of these interneurons is
409: not known. Proposed mathematical models for sequence
410: generation and sequence recognition incorporate
411: inhibitory neurons in various
412: ways~\cite{Kleinfeld89,Drew03,Huerta04,Abarbanel04jn}
413: but there is no support yet for these models in the
414: context of the songbird system. We argue that ignoring
415: the inhibitory neurons is a useful first step toward
416: understanding the HVC~microcircuitry since the
417: calculations discussed below show that a minimal
418: idealized 1d~chain of identical excitatory neurons
419: already suffices to produce a stable propagating burst,
420: in which case other effects such as noise,
421: inhomogeneities, and interneurons could be included
422: perturbatively. Further, some experiments suggest that
423: ignoring the interneurons may not be too severe an
424: assumption.  HVC~interneurons tend to fire rapidly,
425: chronically, and in approximate synchrony throughout a
426: motif~\cite{Fee04}, and rapidly and chronically during
427: audition of a bird's own song~\cite{Mooney00}, so that
428: the interneuron spikes might not be closely correlated
429: with the brief sparse bursts of the \hvcra~neurons. In
430: this case, the main purpose of the interneurons might
431: to prevent runaway excitation of the projection
432: neurons.
433: 
434: We conclude this section with the technical observation
435: that our assumption that the 1d~chain is homogeneous
436: and that there is no delay along axons (the axons along
437: the chain have effectively zero length) allow the
438: 1d~chain to be simulated by integrating numerically a
439: \textit{single} neuron with a single afferent
440: synapse. During the integration, the output of the
441: neuron can be stored in memory and then used as input
442: to the same neuron which then represents the next
443: neuron in a chain. A related idea was used by
444: Reyes~\cite{Reyes03}, who used a single real
445: hippocampal neuron in a slice together with the dynamic
446: clamp technique to simulate a synfire chain consisting
447: of many pools of identical neurons.
448: 
449: 
450: \subsection{A Leaky Integrate-and-Fire (LIF) Model of an \hvcra~Neuron}
451: 
452: We studied propagation of bursts along 1d~chains
453: consisting of one of two types of neurons, leaky
454: integrate-and-fire (LIF) neurons that are described in
455: this section, and more realistic conductance-based
456: Hodgkin-Huxley (HH) neurons that are described in the
457: next section. Neither neuronal model bursts
458: intrinsically when subject to a direct current (DC)
459: external stimulus. Provided that an initial burst is
460: applied to the first neuron of the 1d~chain, we find
461: that successive neurons are capable of
462: generating an identical output burst when stimulated
463: itself by a burst.
464: 
465: With respect to conductance-based models, LIF
466: models~\cite{Dayan01} are attractive since they have
467: fewer parameters, lead to efficient numerical
468: simulations, and are sometimes amenable to analytical
469: methods. The $k$th~LIF neuron of the 1d~chain ($k \ge
470: 1$) was modeled as a first-order non-autonomous
471: ordinary differential equation
472: \begin{equation}
473: \label{eq:lif}
474:   \tau_m \frac{dv^k}{dt} = v_0 - v^k  
475:    +  R \left( I^k_e(t) +  I^k_S(t) \right) ,
476: \end{equation}
477: where the parameter~$\tau_m$ is the capacitive
478: $RC$-time scale of the neuronal membrane, the
479: variable~$v^k(t)$ is the potential difference across
480: the membrane of the $k$th~neuron in the chain, the
481: external current~$I^k_e(t)$ of the $k$th~neuron is a
482: specified function (in this paper applied only to the
483: first neuron to initiate activity in the chain), the
484: synaptic current~$I^k_S(t)$ arises from spikes in the
485: previous $(k-1)$st~neuron of the chain (see
486: Eqs.~(\ref{eq:total-synaptic-current})
487: and~(\ref{eq:lif-syn-current}) below), the
488: parameter~$v_0$ is the resting potential toward which
489: the potential~$v$ asymptotes in the absence of external
490: and synaptic currents ($I_e=I_S=0$), and the
491: parameter~$R$ is the total resistance of the membrane.
492: 
493: Eq.~(\ref{eq:lif}) was supplemented with the usual
494: spiking rule that, whenever the potential~$v^k(t)$ is
495: increasing and crosses a specified threshold
496: value~$v_\textrm{thresh} > v_0$, the neuron is assumed
497: to have spiked and the potential~$v^k$ is instantly and
498: discontinuously decreased to a reset value
499: $v_\textrm{reset} < v_0$. An absolute refractory period
500: was included in this LIF~model by freezing the
501: potential~$v^k$ to the value~$v_\textrm{reset}$ for a
502: time interval $t_\textrm{refract}$ after a spike.  We
503: found that including a refractory interval of
504: $t_\textrm{refract}= 1\,\rm ms$ produced only minor
505: qualitative changes to the results obtained in the
506: absence of a refractory period, for example a 1~ms
507: refractory period expanded by a modest amount the
508: basins of attraction for the different stable
509: propagating bursts discussed in
510: Section~\ref{sec:results-discussion} (see
511: Fig.~\ref{fig:steadyPlot} below). Although
512: \hvcra~neurons show some accommodation~\cite{Mooney00},
513: we did not include this detail in our LIF~model.
514: 
515: The total postsynaptic current~$I^k_S(t)$ of the
516: $k$th~neuron in the chain was assumed to be a linear
517: accumulation
518: \begin{equation}
519:    I^k_S(t) = \sum_i I_s(t-t^{k-1}_i) ,
520:    \label{eq:total-synaptic-current}
521: \end{equation}
522: of postsynaptic currents~$I_s(t-t^{k-1}_i)$
523: associated with spikes that occurred in the previous
524: ${(k-1)}$st~neuron of the chain at times~$t^{k-1}_i <
525: t$. (Eq.~(\ref{eq:total-synaptic-current}) also assumes
526: that there are no time delays arising from the time for
527: a spike to propagate from one neuron to the next.) The
528: time dependence of the synaptic conductivity associated
529: with a single spike at time~$t^k_i=0$ was modeled as a
530: double exponential~\cite{Dayan01}, leading to the
531: following expression for the postsynaptic current per
532: spike:
533: \begin{equation}
534:    I_s(t,v) = n I_0 \left( 
535:     e^{\displaystyle -t/\tau_1} -  e^{\displaystyle
536:     -t/\tau_2}
537:     \right) .
538:    \label{eq:lif-syn-current}
539: \end{equation}
540: The slow and fast time constants~$\tau_1$ and~$\tau_2$
541: had respectively the values~1.1~ms and~0.2~ms to match
542: roughly the~4.0~ms onset-to-peak time constants
543: observed in paired intracellular recordings of
544: \hvcra~neurons~\cite{Mooney05}.  The
545: value~$I_0=0.3\,\rm nA$ was chosen to give a 1.0~mV
546: peak excitatory post-synaptic potential (EPSP) which is
547: consistent with experiment~\cite{Mooney05}.  The
548: number~$n$, which varied from~1 to~32, indicated the
549: synaptic strength in terms of the number of synchronous
550: spikes that a neuron would receive if placed in a
551: uniform synfire chain with~$n$ neurons per pool and all
552: neurons of one pool connecting to all neurons in the
553: next pool (with synapses of the same form
554: Eq.~(\ref{eq:lif-syn-current})). The range~1 to~32 was
555: found empirically to span the range of stable dynamics,
556: see Fig.~\ref{fig:steadyPlot} below.
557: 
558: An homogeneous chain was then obtained by using for
559: each LIF~neuron the same ten parameters~$\tau_m$,
560: $v_0$, $R$, $v_\textrm{thresh}$, $v_\textrm{reset}$,
561: $t_{\rm refract}$, $n$, $I_0$, $\tau_1$, and $\tau_2$,
562: and the same function Eq.~(\ref{eq:lif-syn-current})
563: for the postsynaptic current per spike. A time
564: constant~$\tau_m = 15\,\rm ms$ was used to approximate
565: the {\it in vivo} response time~\cite{Mooney00}, and a
566: membrane resistance of~$R=60\,\rm M\Omega$ was chosen
567: to match the impedance of
568: \hvcra~neurons~\cite{Mooney00}.  Other parameter values
569: used were~$v_0=-70$~mV, $v_\textrm{thresh}=-55$~mV,
570: $v_\textrm{reset}=-75$~mV, and~ $t_{\rm
571: refract}=1.0$~ms.
572: 
573: The chain of LIF~neurons represented by
574: Eqs.~(\ref{eq:lif})-(\ref{eq:lif-syn-current}) was
575: integrated by using a forward-Euler
576: method~\cite{Kincaid96} with a constant time step
577: of~$\Delta{t}=0.01\,\rm ms$.  (Smaller time steps by
578: factors of four did not lead to significant changes in
579: the results.) The supplementary reset and refractory
580: rules were applied at the end of each time step. For
581: most runs, a zero external current~$I^k_e=0$ was
582: assumed for each neuron of the chain except for the
583: first neuron, for which a step function was used to
584: stimulate a burst.  The code was programmed and run
585: using the computational mathematics program Matlab
586: version~6.5~\cite{Matlab}.
587: 
588: 
589: \subsection{A  Single-Compartment Hodgkin-Huxley
590:   Model of an \hvcra~Neuron}
591: 
592: \label{sec:hh-model-of-hvcra-neuron}
593: 
594: The second neuronal model that we used to study the
595: propagation of a burst in a 1d~chain of excitatory
596: neurons was a single-compartment model based on the
597: Hodgkin-Huxley equations with five representative
598: conductances.  The evolution equation for the membrane
599: potential~$v^k(t)$ of the~$k$th neuron in the chain was
600: \begin{equation}
601:   C_m {dv^k \over dt} = \sum_{i=1}^5 g_i(t,v^k)
602:    \left( v_i - v^k \right)  +  I^k_e(t) + I^k_S(t) .
603:   \label{eq:HH-model}
604: \end{equation}
605: The symbols have the following meanings: $C_m$ is the
606: total membrane capacitance, $g_i(t,v)$~is the
607: voltage-dependent conductance of the $i$th~kind of
608: membrane channel, $v_i$~is the resting potential for
609: the $i$th~channel, and the currents~$I^k_e(t)$
610: and~$I^k_S(t)$ have the same meaning as in
611: Eq.~(\ref{eq:lif}). Appendix~\ref{appendix:equations-parameters}
612: gives the details of parameter values and other
613: evolution equations related to the conductances~$g_i$.
614: 
615: Although more realistic in terms of its time
616: dependence, the HH~model Eq.~(\ref{eq:HH-model}) is not
617: necessarily more scientifically appropriate than
618: LIF~models since the properties of the various membrane
619: conductances are only partially known for
620: HVC~neurons~\cite{Kubota91,Kubota98,Dutar98,Mooney00},
621: the spatial distribution of the channels in HVC~neurons
622: has not been determined (which would be needed to
623: construct accurate multi-compartment models), and
624: little is known about the type, strengths, and
625: locations of synapses in~HVC. We did find that a
626: single-compartment HH~neuron with five conductances is
627: only able to match some features of the
628: experimental~HVC data, and it is not clear which
629: experimental details are more important than others to
630: include in the process of fitting a mathematical model
631: to data. For these reasons, and because this paper is
632: concerned whether in principle an HVC-like burst can
633: propagate stably through an excitatory chain, the
634: conductances of the HH~model Eq.~(\ref{eq:HH-model})
635: were only loosely based on existing HVC~data.
636: 
637: Our starting point for choosing the conductances in
638: Eq.~(\ref{eq:HH-model}) was a recent paper by Prinz et
639: al~\cite{Prinz03}, which showed that an
640: eight-dimensional phase space obtained by varying the
641: maximum conductances of eight channels (whose
642: functional properties were obtained from lobster
643: stomatogastric neuronal data) contained a great
644: diversity of dynamical behavior. We chose the
645: functional forms of the leakage, sodium, and potassium
646: channels from this paper and added two other channels
647: as suggested by current-clamp data of~HVC
648: neurons~\cite{Kubota91}: a fourth channel was a
649: low-threshold transient calcium current, and a fifth
650: channel was a calcium-activated potassium current
651: (denoted in the following by the symbol~KCa).  The
652: calcium channel was chosen to activate around -50~mV
653: and inactivate around -80~mV with time scales such that
654: the channel could be activated
655: transiently~\cite{Kubota91,Huguenard92}.
656: %%% mrli: would help to explain what is meant here by
657: %%% different time scales
658: The KCa~conductance activates the potassium current
659: after intracellular calcium concentration rises, with a
660: functional form adopted from Yamada's
661: model~\cite{Yamada98}.  The maximum conductances of all
662: five conductances were adjusted iteratively by hand to
663: produce a short spike width and to generate a spike
664: adaptation with a high initial firing rate as observed
665: by Kubota~\cite{Kubota91}.
666: %%% Please be more specific: how short was the spike
667: %%% width, what was the value of the high initial firing rate?
668: A reasonably good fit to the Mooney {\it in~vivo} data
669: (see Fig.~1 of Ref.~\cite{Mooney00}) unfortunately
670: could not also match Kubota's {\it in~vitro} spike
671: profiles of Ref~\cite{Kubota98} and vice versa. (We
672: hope to obtain more complete fits to HVC~data by adding
673: more channels and more compartments when more
674: anatomical and physiological data become available.)
675: 
676: %%% MengRu: please add further details of what you
677: %%% actually did. These details should especially
678: %%% appear in your thesis.
679: 
680: The total synaptic current was also assumed to satisfy
681: the linear relation
682: Eq.~(\ref{eq:total-synaptic-current}) but an
683: alpha~function~\cite{Dayan01} was used instead of
684: Eq.~(\ref{eq:lif-syn-current}) to approximate the
685: time-dependent probability of a channel opening.  The
686: excitatory postsynaptic current (EPSC) arising from a
687: spike that occurs at time~$t=0$ had the form
688: \begin{equation}
689:    I_s(t;v) = 
690:     g_s C \left( t \over \tau_s \right) e^{\displaystyle -t/\tau_s} 
691:    \left( v_s - v \right) ,
692:   \label{eq:hh-syn-current}
693: \end{equation}
694: where the constant~$C$ with value~$e$ normalizes the
695: maximum value of the expression
696: $C(t/\tau_s)\exp(-t/\tau_s)$ to~1. This normalization
697: makes the parameter~$g_s$ the maximum conductance,
698: which we varied over the range 0 to 0.55~nS.  In
699: Eq.~(\ref{eq:hh-syn-current}), the synaptic time
700: constant~$\tau_s$ was fixed with value~7~ms and
701: %%% mrli: where did the value for tau_s come from?
702: %%% Should give reference if you have one
703: the synaptic reversal voltage~$v_s$ was set to~120~mV
704: to make the synapse excitatory.
705: Eqs.~(\ref{eq:HH-model}),
706: (\ref{eq:total-synaptic-current})
707: and~(\ref{eq:hh-syn-current}) together with the
708: evolution equations for gate variables given in
709: Appendix~\ref{appendix:equations-parameters} were
710: integrated with the Neuron simulation code
711: version~5.6~\cite{NeuronCode}, with a constant time step
712: of~$\Delta{t}=0.1\,\rm ms$.  
713: 
714: %Another simulation with
715: %$\tau_s=1.5\,\rm ms$, $v_s = 0\,\rm mV$, $g_s$ between
716: %0.4~nS and 12~nS, and time step 0.02~ms showed the same
717: %result with more compact bursts.
718: 
719: %%% MengRu: how did you figure out that this time step
720: %%% was sufficiently accurate? What is the importance
721: %%% of mentioning that the bursts were more compact?
722: 
723: %%% What value of synaptic reversal potential v_s did
724: %%% we use for HH synapses? What other details do we 
725: %%% need to add?
726: 
727: 
728: \section{Results and Discussion}
729: 
730: \label{sec:results-discussion}
731: 
732: 
733: \subsection{Results for a 1d Homogeneous Chain of HH~Neurons}
734: 
735: The calculations discussed below of one-dimensional
736: homogeneous chains using the LIF or HH~neurons of the
737: previous section with excitatory synapses show that a
738: brief high-frequency burst similar to that observed in
739: \hvcra~neurons during singing~\cite{Hahnloser02} can
740: propagate stably over a range of parameter values.
741: These results make more plausible the hypothesis that
742: an excitatory chain of \hvcra~neurons generates the
743: sparse precisely-aligned bursts observed
744: experimentally~\cite{Hahnloser02}. Our results show
745: further that different asymptotic attractors can exist
746: for given parameter values so that hysteresis can occur
747: (different initial conditions can lead to different
748: non-transient dynamics).  The fact that the bursts
749: exist as an attractor means that there is a transient
750: time during which properties of the burst such as the
751: number of spikes and burst width (alternatively, the
752: average burst frequency) evolve until the final stable
753: values are obtained. This transient time varies with
754: the initial state used to start the chain and with the
755: choice of parameters.
756: 
757: Since these qualitative conclusions turned out to be
758: similar for LIF and HH~neuronal models, we report here
759: results mainly for the HH~models of
760: Section~\ref{sec:hh-model-of-hvcra-neuron} and mention
761: briefly in the next section how the results differ for
762: a chain of~LIF neurons. Since the parameter spaces for
763: LIF and HH~neurons are high-dimensional (10~parameters
764: for~LIF, about 25~parameters for the HH~neurons, and
765: these do not include the choice of the functional
766: form~$I_s(t,v)$ for the~EPSC per spike), to establish
767: our key results we varied only one parameter
768: systematically, namely the maximum synaptic
769: conductance~$g_s$ of Eq.~(\ref{eq:hh-syn-current}) for
770: HH~neurons, or equivalently the number of synchronous
771: presynaptic spikes~$n$ of
772: Eq.~(\ref{eq:lif-syn-current}) for LIF neurons.
773: 
774: \begin{figure}[tb]
775:   \includegraphics[height=2.7in]{burstPropagate.eps}
776:   \caption{ Propagation of an initial 5-spike burst
777:     through an homogeneous one-dimensional excitatory
778:     chain of HH~neurons, for different maximum synaptic
779:     conductances~$g_s$ of
780:     Eq.~(\ref{eq:hh-syn-current}). Each column is
781:     corresponds to a fixed~$g_s$ value, and successive
782:     rows of a given column show the membrane
783:     potential~$v^k(t)$ as a function of time for
784:     successive neurons in the chain ($k=1, 2,
785:     \cdots$). {\bf (a)}~For a weak synaptic coupling
786:     strength~$g_s=0.1\,\rm nS$, the initial burst
787:     decays by the third neuron.  {\bf (b)}~For a
788:     stronger coupling~$g_s=0.2\,\rm nS$, the initial
789:     bursts decays to an invariant single spike.  {\bf
790:     (c)}~For~$g_s=0.3\,\rm nS$, the initial burst
791:     evolves to a stable state with three unevenly
792:     spaced spikes.  {\bf (d)}~For still stronger
793:     couplings $g_s \ge 0.45\,\rm nS$, the initial burst
794:     can be unstable for some initial states and the
795:     number of spikes increases steadily.}
796:   \label{fig:burst-propagation}
797: \end{figure}
798: 
799: Figure~\ref{fig:burst-propagation} shows how the same
800: initial burst of five spikes propagates through a
801: 1d~excitatory chain of HH~neurons for several different
802: values of the maximum synaptic conductance~$g_s$. The
803: initial burst was created by using an external
804: current~$I^1_e(t)$ of Eq.~(\ref{eq:HH-model}) for the
805: first neuron in the chain to inject a square pulse of
806: current that caused five spikes to appear in rapid
807: succession.  For a synaptic coupling strength~$g_s$
808: smaller than about~$0.1\,\rm nS$,
809: Fig.~\ref{fig:burst-propagation}a shows that the
810: initial burst rapidly dies out by the third neuron of
811: the chain and all spikes disappear. Over a range of
812: stronger couplings $0.2 < g_s < 0.4\,\rm nS$, the
813: initial bursts evolves into a stable invariant
814: propagating burst that can have one to five spikes
815: (columns (b) and~(c) of
816: Fig.~\ref{fig:burst-propagation}). For stronger
817: couplings $g_s > 0.4\,\rm nS$, the initial burst is
818: unstable and the number of spikes grows steadily
819: without limit. However, other initial conditions can
820: lead to stable bursts in this range, an example of
821: hysteresis.
822: 
823: \begin{figure}[tb]
824:   \centering
825:   \includegraphics[height=2.7in]{initCondProp2.eps}
826:   \caption{(Color onlin) The dependence of transient dynamics and
827:     resulting attractors of propagating bursts on the
828:     initial injecting current magnitude~$I^1_e(t)$ and
829:     on the synaptic conductance~$g_s$. For weak
830:     couplings with $g_s < 0.2\,\rm nS$, all initial
831:     states decay to zero (not shown here). {\bf (a)}
832:     For~$g_s=0.3\,\rm nS$, there are multiple basins of
833:     attraction corresponding to asymptotic propagating
834:     bursts with 1, 2, or~3 spikes. Transients decay
835:     rapidly by the third neuron of the chain. {\bf (b)}
836:     Increasing the synaptic strength to~$g_s =0.35\,\rm
837:     nS$ increases the number of attractors, with final
838:     bursts containing 1 to~5~spikes.  {\bf (c)} For a
839:     larger synaptic conductance~$g_s=0.45\,\rm nS$,
840:     initial conditions lead to stable or unstable
841:     states. An initial burst with 6~spikes will evolve
842:     to a slightly different invariant burst with also
843:     6~spikes (squares). }
844:   \label{fig:initCondProp}
845: \end{figure}
846: 
847: Fig.~\ref{fig:initCondProp} provides a more global
848: understanding of the transient dynamics and resulting
849: attractors. In each panel, the vertical axis indicates
850: the number of spikes observed in a burst while the
851: horizontal axis indicates the position~$k$ of a neuron
852: along the chain. For a weak synaptic coupling~$g_s<
853: 0.2\,\rm nS$, all initial states decay to zero spikes
854: (not shown in the figure).  For a synaptic coupling
855: of~$g_s=0.3\,\rm nS$ (Fig.~\ref{fig:initCondProp}a),
856: initial bursts with three or more spikes decay within
857: four neurons to a common final burst of three
858: spikes. An initial burst with two spikes evolves
859: slightly to an invariant burst with also two spikes,
860: and a similar conclusion holds for an initial burst
861: with a single spike. As the synaptic strength is
862: increased to~$g_s=0.35\,\rm nS$
863: (Fig.~\ref{fig:initCondProp}b), the number of
864: attractors increases so that stable bursts with one to
865: five spikes are observed depending on the initial state
866: of the first neuron. The transient time is still rather
867: short, with the final number of spikes stabilizing in
868: all cases within three neurons ($k \le 3$). Finally for
869: synaptic couplings stronger than about~$0.4\,\rm nS$
870: (Fig.~\ref{fig:initCondProp}c), stable and unstable
871: states are observed depending on the initial state of
872: the first neuron. The unstable bursts all grow at the
873: same rate, with an extra spike appearing for each next
874: neuron traversed.
875: 
876: \begin{figure}[tb]
877:   \centering
878:   \includegraphics[height=2.7in]{twoDMap2.eps}
879:   \caption{ (Color online) Different synaptic conductances~$g_s$ lead
880:     to different final propagating burst widths and
881:     spike numbers, which shows that the total burst
882:     duration is another important characteristic of the
883:     final attractor.  For $g_s = 0.3\,\rm nS$, the
884:     curve of block symbols evolves into the point~E1.
885:     For $g_s = 0.35 \,\rm nS$, the curve with diamond
886:     symbols ends in the point~E2.  For $g_s = 0.45\,\rm
887:     nS$, the propagation is unstable (upper triangle)
888:     except for the initial condition at point~B2.  }
889:   \label{fig:transient-states}
890: \end{figure}
891: 
892: The transient dynamics of Fig.~\ref{fig:initCondProp}
893: are presented in somewhat finer detail in
894: Fig.~\ref{fig:transient-states}, which shows that a
895: non-transient propagating bursts is characterized by at
896: least two parameters, their total width and the number
897: of spikes.  For example, the same beginning state~$B_1$
898: can evolve into three different end points depending on
899: the synaptic coupling strength~$g_s$: point~$E_1$ with
900: three spikes, point~$E_2$ with 4~spikes and with a
901: slightly longer width of~45~ms, and an unstable state
902: (triangles).
903: 
904: \begin{figure}[tb]
905:   \centering
906:   \includegraphics[height=2.7in]{steadyPlot.eps}
907:   \caption{ Plot of the number of spikes observed in a
908:   final non-transient propagating burst for different
909:   initial amplitudes of a square current pulse of fixed
910:   duration~20 ms and for different maximum synaptic
911:   conductances~$g_s$.  Different basins of attraction
912:   are observed, with 0-spike and unstable bursts having
913:   the largest basins, and roughly equal size basins for
914:   bursts with 1 to~5 spikes in the final burst.}
915:   \label{fig:steadyPlot}
916: \end{figure}
917: 
918: Finally, Fig.~\ref{fig:steadyPlot} indicates the
919: approximate basins of attraction that exist for
920: different initial current amplitudes (horizontal axis)
921: and different maximum synaptic conductances (vertical
922: axis). The plotted numbers indicate the number of
923: spikes observed in the asymptotic state for the
924: specified axis values.  This plot shows again how all
925: states decay away for sufficiently weak synaptic
926: coupling, all states are unstable for sufficiently
927: strong coupling, and that stable propagating bursts
928: with~1 to~6 spikes can be found for intermediate
929: coupling strengths. The basins of attraction turn out
930: to be of comparable size except for the tiny basin
931: corresponding to a 6-spike burst.
932: 
933: 
934: \subsection{Results for a 1d Homogeneous Chain of LIF
935: Neurons}
936: 
937: \label{sec:results-1d-homog-lif-chain}
938: 
939: Our results for homogeneous chains of~LIF neurons were
940: qualitatively similar to the above results for chains
941: of HH~neurons and we do not show the corresponding
942: results.  Instead of initiating the chain with a
943: current pulse, the first LIF~neuron of the chain was
944: initiated with a synaptic current corresponding to a
945: burst of equally-spaced spikes whose times of
946: occurrences~$t^0_i$ were specified as initial data (see
947: Eq.~(\ref{eq:total-synaptic-current})).  Similarly,
948: instead of a maximum conductance~$g_s$ being varied,
949: the synaptic coupling strength~$n$ of
950: Eq.~(\ref{eq:lif-syn-current}) was varied between~1
951: and~32 which we found to span the same kinds of states
952: discussed in Figs.~\ref{fig:burst-propagation}
953: and~~\ref{fig:steadyPlot}, from a 0-spike final state
954: to an initial burst that grew without bound in width
955: and in the number of spikes.  Because the LIF neurons
956: have no intracellular calcium dynamics, the number of
957: spikes in the final state was primarily determined by
958: the initial spike number. In contrast, the
959: intracellular calcium current of a HH neuron can boost
960: depolarization and maintain propagating bursts in
961: cohesion so that chains of HH~neurons have bigger
962: basins of attraction for stable bursts.
963: 
964: % When LIF synfire chains have no refractory period, the
965: % basins of attraction are narrower and the propagation 
966: % tends to amplify to become unstable.  The refractory period
967: % plays a role in shutting down excessive synaptic 
968: % current and in stabilizing a synfire chain.
969: 
970: 
971: \subsection{Other Hypotheses and Related Theory}
972: \label{sec:other-hypothese-and-related-theory}
973: 
974: We discuss briefly here some other hypotheses and
975: related theoretical work that might explain the data of
976: Hahnloser et al~\cite{Hahnloser02} and of
977: Mooney~\cite{Mooney00} but that do not involve the
978: propagation of bursts through a feedforward network.
979: 
980: While a feedforward network, especially a synfire
981: chain, is one of the simplest concepts that might
982: explain sparse precisely-timed bursting, there is much
983: theoretical work dating back to the~1980s which shows
984: that a recursive network with inhibitory and excitatory
985: connections is capable of learning and producing
986: different kinds of temporal
987: sequences~\cite{Kleinfeld89,Hertz91}. These networks
988: generalize the Hopfield attractor model of associative
989: memory~\cite{Hopfield82,Hertz91} by allowing
990: non-symmetric couplings between pairs of neurons, and
991: by using two kinds of synapses, ``fast'' synapses that
992: stabilize a given network state, and ``slow'' synapses
993: that cause successive transitions between the
994: quasistatic network states.  Thus there is no
995: difficulty in principle for a recursive network to
996: store, generate, or learn many temporal sequences of
997: the sort observed in~HVC. (We note that other classes
998: of recursive models are possible, e.g., Huerta and
999: Rabinovich~\cite{Huerta04} discuss a model with neurons
1000: that are strictly inhibitory or strictly excitatory,
1001: rather than allow any neuron to have inhibitory and
1002: excitatory connections with other neurons.)
1003:  
1004: Asymmetric Hopfield-like recursive networks have
1005: several attractive features for modeling
1006: HVC~dynamics. The HVC inhibitory
1007: neurons~\cite{Mooney00} can be incorporated in a
1008: natural way since temporal sequences are stored via
1009: neuronal connection strengths~$J_{ij}$ that typically
1010: have positive (excitatory) and negative (inhibitory)
1011: values. Recursive networks that generate sequences can
1012: evolve via a simple Hebbian learning
1013: rule~\cite{Kleinfeld89} from densely interconnected
1014: neurons (see Figs.~2 and~3 of Ref.~\cite{Mooney00}) and
1015: so are possibly easier to form during maturation of
1016: brain tissue than purely feedforward
1017: networks. Asymmetric Hopfield models are naturally
1018: fault-tolerant and error-correcting so that details of
1019: a stored sequence are weakly affected if neurons or
1020: synapses are modified, deleted, or added. Finally, a
1021: single network of this kind is capable of storing and
1022: generating many different temporal sequences, which is
1023: consistent with the ability of some songbirds to learn
1024: and sing many different songs~\cite{Catchpole95}.
1025: 
1026: However, further work is needed to determine whether
1027: existing recursive models~\cite{Kleinfeld89} are
1028: capable of taking into account specific experimental
1029: details of~HVC such as the fact that an \hvcra~neuron
1030: fires briefly just once during a one-second-long motif
1031: (a neuron is more likely to fire multiple times during
1032: a motif if there are recurrent pathways), that time
1033: interval between successive bursts is short (estimated
1034: to be about 10~ms~\cite{Fee04}, this is possibly too
1035: short for a separation of time scales to exist between
1036: slow and fast synapses~\cite{Kleinfeld89}), that there
1037: is a precise alignment of bursts with auditory features
1038: during singing~\cite{Hahnloser02} and during
1039: audition~\cite{Mooney00}, and especially that
1040: HVC~inhibitory neurons fire tonically while the
1041: HVC~projection neurons fire sparsely during
1042: singing~\cite{Hahnloser02} and during
1043: audition~\cite{Mooney00}. (There is no dynamical
1044: distinction between inhibitory and excitatory synapses
1045: in the recursive models~\cite{Kleinfeld89}.)
1046: 
1047: Another alternative explanation for the origin and
1048: precise alignment of \hvcra~bursts during singing and
1049: during audition is that these bursts are driven by
1050: precisely timed external inputs. This possibility is
1051: suggested by the fact that other nuclei such as~NIF
1052: and~Uva are known to provide distributed input to HVC,
1053: and to all three classes of
1054: HVC~neurons~\cite{Coleman04,Rosen06}. However, some
1055: experiments suggest that this is a less plausible
1056: explanation than intrinsic generation of bursts
1057: within~HVC. For example, preventing auditory input
1058: to~HVC by destroying a bird's cochlea (which deafens
1059: the bird) or by lesioning nucleus~NIF does not prevent
1060: a bird from singing, nor does it cause a bird's song to
1061: change substantially over time scales of days.
1062: %%% I need to get references for these points!
1063: Although extracellular recordings of an awake singing
1064: bird of the kind reported by Hahnloser et
1065: al~\cite{Hahnloser02} have not been carried out after
1066: deafening or lesioning, the fact that the song does not
1067: change immediately suggests that the same pattern of
1068: sparse bursts should still be observed
1069: in~\hvcra~neurons. Since the inputs to~HVC remain
1070: incompletely characterized, external driving, or some
1071: combination of external driving and intrinsic HVC
1072: circuitry, might be able to explain the observed bursts
1073: of \hvcra~neurons during singing.
1074: 
1075: A variation of the idea of precise external input
1076: driving the bursts in~HVC is the possibility that part
1077: of~HVC acts as a central pattern generator (CPG) that
1078: drives the \hvcra~neurons. For example, the
1079: time-ordering of bursts in \hvcra~neurons observed by
1080: Hahnloser et al~\cite{Hahnloser02} could arise from
1081: a~CPG composed of other HVC~neurons such that the~CPG
1082: connects to the \hvcra~neurons via axons of different
1083: lengths and so with different delay times. (The
1084: dynamics of the~CPG itself could be explained as a
1085: non-symmetric Hopfield model~\cite{Kleinfeld89} or a
1086: cyclic synfire chain.) A CPG-based model would differ
1087: from a synfire chain in that a given burst does not
1088: trigger causally the next burst like dominoes falling
1089: over along some path. For example, destroying all
1090: \hvcra~neurons that burst at a certain time in a
1091: CPG~model would not stop the firing of \hvcra~neurons
1092: that normally would burst a short time later. A recent
1093: \textit{in vitro} slice experiment~\cite{Solis05}
1094: suggests that HVC~has intrinsic oscillatory dynamics
1095: consistent with the existence of a~CPG, but it is not
1096: yet known how these slice data relate to the dynamics
1097: of \hvcra~neurons during song.
1098: 
1099: 
1100: \section{Conclusions}
1101: \label{sec:conclusions}
1102: 
1103: For an idealized one-dimensional homogeneous
1104: feedforward chain of excitatory non-bursting neurons,
1105: we have shown by numerical calculations that a brief
1106: high-frequency burst of two to six spikes can propagate
1107: in a stable way for various choices of parameter
1108: values. Previous studies of excitatory networks have
1109: not addressed the propagation of bursts and so have not
1110: been directly relevant for recent experiments
1111: concerning the properties of nucleus~HVC. Previous
1112: studies~\cite{Diesmann99} have also mainly involved
1113: networks of LIF~neurons (to reduce the computational
1114: effort) and so have not taken into account the more
1115: complex dynamics and realistic time scales of
1116: HH~neurons, especially the effects of a calcium current
1117: which experiments have shown to be present in
1118: HVC~neurons~\cite{Kubota91,Kubota98}. Our results show
1119: that stable bursts exist over a range of parameter
1120: values for both LIF and~HH neuronal models so that the
1121: propagation of stable bursts in a homogeneous chain is
1122: not sensitive to the choice of model or of model
1123: parameters.
1124: 
1125: These results make plausible the hypothesis that the
1126: sparse precisely-aligned high-frequency brief bursts
1127: observed by Hahnloser et al~\cite{Hahnloser02} in
1128: \hvcra~neurons during singing are intrinsic (do not
1129: need external input) and are causally connected in that
1130: one neuron initiates bursting in the next neuron and so
1131: on until the end of the chain. However, other
1132: explanations of the sparse precise bursts are possible
1133: as discussed in
1134: Section~\ref{sec:other-hypothese-and-related-theory},
1135: and further experiments, especially using methods that
1136: can analyze many neurons at once with good time
1137: resolution, would be valuable in helping to distinguish
1138: feedforward chains from external driving, recursive
1139: networks, or other mechanisms.
1140: 
1141: We have not addressed in this paper why \hvcra~neurons
1142: burst in the first place in, i.e., why information is
1143: transmitted as bursts rather than as a population of
1144: spikes. (Sparse firing can aid the learning of
1145: sequences as pointed out by Fiete et al~\cite{Fiete04}
1146: but this does not require bursts specifically.) There
1147: is somewhat of a paradox here in that the commonly
1148: posited purpose of a burst is to increase the
1149: likelihood of accurate transmission through a noisy
1150: network~\cite{Lisman97,Prida97}, while a similar goal
1151: is achieved without bursts by using a synfire
1152: chain~\cite{Abeles91}. Thus if a synfire chain is the
1153: correct architecture for \hvcra~neurons, it is less
1154: clear why bursts are needed, especially since HVC~has
1155: much less noise than cortical neurons and HVC~synapses
1156: are substantially stronger than cortical
1157: synapses~\cite{Mooney05}. Unpublished calculations by
1158: the authors~\cite{Li06II,Li06thesis} that extend the
1159: present paper to noisy heterogeneous synfire chains do
1160: suggest that bursts are important for achieving robust
1161: propagation even for a synfire architecture.
1162: 
1163: We finish this paper by summarizing some consequences
1164: of the chain hypothesis and of our calculations for
1165: recent and future songbird experiments.
1166: \begin{enumerate}
1167:   
1168: \item A simple consequence of the feedforward nature of
1169:   an excitatory chain is that an experimentalist in
1170:   principle should be able to initiate singing of an
1171:   arbitrary contiguous segment of a motif. For example,
1172:   if it is possible to stimulate an intermediate part
1173:   of a chain in such a way that a burst begins to
1174:   propagate toward the end of the chain, a songbird
1175:   might then sing a motif that starts somewhere in the
1176:   middle and that then continues to its end. Similarly,
1177:   physically terminating the chain at some intermediate
1178:   point (say with a local lesion) could cause a motif
1179:   to terminate before its usual endpoint, or some
1180:   combination of stimulating in the middle of the chain
1181:   and lesioning later in the chain could be implemented
1182:   in which case the motif could start somewhere in the
1183:   middle and terminate prematurely.
1184: 
1185:   Electrical stimulation of~HVC via a single
1186:   extracellular electrode in an awake behaving bird
1187:   does not initiate singing~\cite{Vu94,Vicario95,Vu98}
1188:   but instead resets a motif to its beginning (if the
1189:   stimulus occurs while a bird was singing). The
1190:   failure to initiate the singing of a motif, or part
1191:   of a motif, does not rule out the existence of an
1192:   excitatory chain but instead could imply that a
1193:   carefully-arranged pattern of input spikes might be
1194:   needed to initiate activity at some point in the
1195:   chain, especially if the network is a synfire chain
1196:   for which most neurons in a pool must fire in near
1197:   synchrony for neurons in the next pool to
1198:   fire. Identifying the location of \hvcra~neurons in a
1199:   chain by, say, optical imaging would be a valuable
1200:   prior step that could suggest how to work out an
1201:   electrical or photo-uncaging
1202:   protocol~\cite{Losonczy06} that could initiate a
1203:   chain at an arbitrary point along its length, or
1204:   could suggest how to lesion neurons that would
1205:   terminate propagation of a burst hence a motif
1206:   prematurely.
1207: 
1208: \item Our calculations suggest that a homogeneous chain
1209:   based on~LIF or HH~neuronal models should generally
1210:   have multiple basins of attraction.  Different kinds
1211:   of asymptotic stable bursts differing in the number
1212:   of spikes and in their total duration should be
1213:   observed in nucleus~HVC for different values of
1214:   synaptic strength and other neuronal parameters.
1215:   Further, hysteresis can be expected such that, for
1216:   identical experimental conditions, different initial
1217:   states will evolve to different kinds of asymptotic
1218:   bursts.
1219:   
1220: %   Verifying the existence of different basins of
1221: %   attraction is difficult and would require varying
1222: %   somehow the properties of HVC neurons or of their
1223: %   synapses and observing how a given initial burst
1224: %   evolves. Some properties can be altered by fatiguing
1225: %   synapses (say by repeated stimulation), by locally
1226: %   administering chemical agents (e.g., low doses of
1227: %   synaptic inhibitors to weaken synaptic transmission),
1228: %   by local photo-uncaging of neurotransmitters (which
1229: %   could strengthen synapses just before a burst
1230: %   arrives), or in the longer term, by genetic methods
1231: %   that, say, increase or decrease the amount of
1232: %   neurotransmitter stored in \hvcra\ synaptic vesicles
1233: %   or that alter the density of postsynaptic receptors.
1234: 
1235:   Hysteresis under fixed experimental conditions can be
1236:   explored by initiating different kinds of bursts.
1237:   This may be possible if the neurons at the beginning
1238:   of the chain can be identified and an appropriate
1239:   stimulation protocol worked out (see points~1 and~3).
1240: 
1241:   Most nonlinear network models, including asymmetric
1242:   Hopfield models and CPGs, also have multiple basins
1243:   of attraction so this is not a distinguishing
1244:   property of chain models. However, the number and
1245:   types of basins of attraction, and how these basins
1246:   vary under the influence of pharmaceuticals that alter
1247:   synaptic strengths, may lead to predictions that
1248:   distinguish feedforward from recursive models.
1249:   
1250: \item Chains of neurons, like the neuronal models that
1251:   make up the chains, are driven dissipative systems so
1252:   that initial states will generally evolve through a
1253:   transient before settling into an asymptotic
1254:   behavior~\cite{Berge84}.  Properties of a burst such
1255:   as its number of spikes and its set of interspike
1256:   intervals will evolve over time, with the transient
1257:   time depending on the choice of initial condition
1258:   (how far the initial state is from the attractor) and
1259:   on neuronal parameters.  Our results such as
1260:   Fig.~\ref{fig:initCondProp} for an homogeneous
1261:   HH~neuronal model suggest that transient times can in
1262:   fact be rapid, just four or fewer successive neurons.
1263:   Fig.~2 of Hahnloser et al~\cite{Hahnloser02} does not
1264:   show evidence of transient behavior since later
1265:   bursts do not seem to be statistically different from
1266:   earlier bursts during the same motif. However, only a
1267:   small number (about~20) of~\hvcra~neurons have been
1268:   sampled to date in singing birds, and neurons that
1269:   burst at the beginning of a motif have not been found
1270:   so further study involving more neurons would be
1271:   worthwhile. It may also be the case that the songbird
1272:   brain has evolved in such a way as to reduce or
1273:   eliminate the duration of transients, for example by
1274:   starting a chain with a state close to the asymptotic
1275:   burst, or by using inhibitory neurons to accelerate
1276:   convergence to synchronized firing as bursts
1277:   propagate from projection neuron to projection
1278:   neuron~\cite{Kopell04}.
1279:   
1280: \item A one dimensional chain is not robust since
1281:   propagation terminates if any neuron in the chain
1282:   dies.  A more realistic synfire chain must have at
1283:   least two neurons in a pool to be robust, and the
1284:   neurons in a pool must synchronize their spikes to
1285:   some extent to achieve reliable transmission from
1286:   pool to pool~\cite{Abeles91}. A synfire chain
1287:   hypothesis thus predicts that there must be more than
1288:   one \hvcra~neuron bursting at any give time (these
1289:   are the neurons that belong to the same pool) and
1290:   further that neurons that burst at about the same
1291:   time should have nearly synchronous bursts. Thus
1292:   Fig~2 of Ref.~\cite{Hahnloser02} may be incomplete
1293:   and a high-resolution optical study of~HVC (or
1294:   possibly a many-electrode study) during the singing
1295:   of a motif or during the audition of a bird's own
1296:   song~\cite{Mooney00} may reveal multiple
1297:   \hvcra~neurons firing bursts in synchrony.
1298:   
1299:   Abeles has pointed out that there is an inverse
1300:   relation between the number of neurons needed in a
1301:   synfire pool for reliable propagation and the average
1302:   strength of the synapses between
1303:   pools~\cite{Abeles91}. Using paired intracellular
1304:   recordings, Mooney and Prather~\cite{Mooney05} have
1305:   shown that synapses between \hvcra~neurons are
1306:   stronger than mammalian cortical neurons by about an
1307:   order of magnitude (the average~EPSP magnitude is
1308:   about 2~mV compared to about 0.2~mV in cortex) so
1309:   chains in~HVC would be expected to have about an
1310:   order of magnitude fewer neurons in a pool. Given
1311:   Abeles's estimate of about 50-100~neurons in a
1312:   cortical pool, perhaps ten or fewer neurons might be
1313:   needed in an HVC~synfire pool.  Assuming that the
1314:   HVC~bursts of average duration 6~ms are
1315:   non-overlapping, a zebra finch motif of duration
1316:   0.6~s would require about $0.6/0.006 \approx 100$
1317:   successive pools to span the motif, so a synfire
1318:   chain in~HVC might involve about $100 \times 10
1319:   \approx 1,000$ neurons. This is a much smaller number
1320:   than the 40,000 \hvcra~neurons estimated to exist
1321:   in~HVC. This suggests that a synfire chain might be
1322:   too simplistic an architecture to justify the large
1323:   number of \hvcra~neurons.
1324: 
1325: %  A more detailed understanding of the anatomy of
1326: %  HVC~synapses (numbers, kinds, spatial distributions,
1327: %  and strengths) between \hvcra~neurons would greatly
1328: %  help in our
1329: 
1330:   
1331: \item 
1332:   It is worth noting that a synfire chain architecture
1333:   can explain how an overall precise timing can be
1334:   maintained in~HVC even though there is a steady and
1335:   substantial turnover of neurons in~HVC throughout the
1336:   life of the songbird~\cite{NottebohmJN02,Wang02}.
1337:   Since perfect synchronization of a pool is not needed
1338:   to guarantee transmission of information through the
1339:   next pool, a fraction of neurons in a given pool can
1340:   alter their properties or fail without degrading the
1341:   transmission of information or its timing.
1342:   
1343: \end{enumerate}
1344: 
1345: In conclusion, our calculations support the hypothesis
1346: that the sparse bursts observed experimentally in
1347: \hvcra~neurons during singing can be understood as the
1348: propagation of bursts through an excitatory feedforward
1349: synfire chain. However, further experiments and
1350: computational studies are needed to confirm this
1351: hypothesis and to rule out the competing hypotheses
1352: discussed in
1353: Section~\ref{sec:other-hypothese-and-related-theory}.
1354: Especially interesting in the near term would be to
1355: understand some of the quantitative details of the
1356: experiments, for example the production of a brief 6~ms
1357: burst of four spikes corresponding to a frequency of
1358: about 600~Hz that is consistent with the known
1359: properties of HVC~neurons and of the HVC
1360: microcircuitry~\cite{Mooney05}. Although our results
1361: show that a brief burst can propagate in a stable
1362: manner, we and others~\cite{Jin05} have not been able
1363: to construct neuronal models, or simple networks of
1364: such models, that burst as rapidly and as briefly as
1365: actual \hvcra~neurons.
1366: 
1367: \vspace{.2in}
1368: 
1369: Note: Toward the completion of this paper, the authors
1370: received a preprint by Jin et al whose results overlap
1371: with the present paper (the preprint is related to a
1372: recent conference presentation~\cite{Jin05}). These
1373: authors also investigated whether brief high-frequency
1374: bursts analogous to those observed in \hvcra~neurons
1375: could propagate through a purely excitatory network,
1376: and they too found that this was possible for a range
1377: of parameters, thereby supporting the hypothesis that
1378: the bursting of \hvcra~neurons could be intrinsic
1379: to~HVC and causal.
1380: 
1381: The main difference of the present paper with the
1382: preprint is that we explored whether bursts could
1383: propagate under the simplest circumstances of a 1d,
1384: homogeneous, noiseless, purely-feedforward chain using
1385: simple neuronal models.  Jin et al studied a more
1386: complicated generalized (not purely feedforward),
1387: heterogeneous, noisy, purely excitatory synfire chain
1388: with many neurons per pool. Jin et al also used a more
1389: complicated two-compartment conductance-based neuronal
1390: model that included a low-threshold potassium channel
1391: that provided a strong spike-frequency adaptation
1392: similar to that observed experimentally. (This helped
1393: to create brief high-frequency bursts.) They also
1394: included calcium dynamics in the dendrite compartment
1395: that could lead to the firing of calcium spikes. Jin et
1396: al were able to show that adding the low-threshold
1397: potassium channel and calcium dynamics made the
1398: propagation of bursts through their generalized synfire
1399: chain more robust (one did not have to tune parameters
1400: so carefully to obtain a stable propagating
1401: burst). They also studied how jitter in the burst
1402: evolved along the length of the synfire chain, which we
1403: have also done in a forthcoming paper with a somewhat
1404: different model~\cite{Li06II}.
1405: 
1406: Given that our calculations and those of Jin et al have
1407: left out the role of inhibitory neurons and were based
1408: on a limited knowledge of~HVC neuronal properties and
1409: of the HVC microcircuitry (e.g., there is no evidence
1410: yet for a synfire chain in HVC, nor have the
1411: conductances of HVC~neurons been fully characterized),
1412: both calculations should be considered as modest but
1413: complementary and useful steps toward understanding
1414: possible mechanisms for the dynamics of~HVC during
1415: singing.
1416: 
1417: \appendix
1418: 
1419: \section{Hodgkin-Huxley Equations and Parameters}
1420: 
1421: \label{appendix:equations-parameters}
1422: 
1423: The single-compartment HH~model of
1424: Eq.~(\ref{eq:HH-model}) had the following five
1425: representative currents~$I_i(t)$:
1426: \begin{eqnarray}
1427:   I_{\rm leak} &=& \bar{g}_L (v-E_L), \nonumber \\
1428:   I_{\rm Na} &=& \bar{g}_{\rm Na} m^3 h (v-E_{\rm Na}), \nonumber\\
1429:   I_{\rm K} &=& \bar{g}_{\rm k} n^4 (v- E_{\rm K}),   \label{eq:hh-currents}
1430:  \\
1431:   I_{\rm Ca} &=& \bar{g}_{\rm Ca} m_1^2 h_1 (v-E_{\rm Ca}),\nonumber \\
1432:   I_{\rm KCa} &=& \bar{g}_{\rm KCa} n_1^3 (v-E_{\rm KCa}),\nonumber
1433: \end{eqnarray}
1434: with maximum conductances per unit area y (units of
1435: $\rm mS/mm^2$) given
1436: \begin{equation}
1437: \label{eq:hh-max-gbar-values}
1438:   \bar{g}_L = 0.003, \quad
1439:   \bar{g}_{\rm Na} = 1.0 , \quad
1440:   \bar{g}_{\rm K} = 0.22 , \qquad
1441:   \bar{g}_{\rm Ca} = 0.031 , \qquad
1442:   \bar{g}_{\rm KCa} = 0.007 , \qquad
1443: \end{equation}
1444: and corresponding reversal potentials (units of~mV):
1445: \begin{equation}
1446: \label{eq:hh-rev-potentials}
1447:   E_L = -70, \quad
1448:   E_{\rm Na} = 50 , \quad
1449:   E_{\rm K} = -80 , \qquad
1450:   E_{\rm Ca} = 120 , \qquad
1451:   E_{\rm KCa} = -80 . \qquad
1452: \end{equation}
1453: 
1454: \begin{table}[tbh]
1455:   \label{table:hh-parameters}
1456:   \begin{tabular}{c|c|c|c|c|c|c|}
1457:      \multicolumn{1}{c}{} 
1458:    & \multicolumn{1}{c}{\mbox{\ \ \ $m_{\ }$\ \ \ } }
1459:    & \multicolumn{1}{c}{\mbox{\ \ \ $n_{\ }$\ \ \  } }
1460:    & \multicolumn{1}{c}{\mbox{\ \ \  $h_{\ }$\ \ \  } }
1461:    & \multicolumn{1}{c}{\mbox{\ \ \  $m_1$\ \ \ }}
1462:    & \multicolumn{1}{c}{\mbox{\ \ \ $h_1$\ \ \  }}
1463:    & \multicolumn{1}{c}{\mbox{\ \ \ $n_1$\ \ \  }}\\
1464:     $v_0$ mV \, & -25.5 & -12.3  &  -60  & -50 & -80 &  \\
1465:     $v_1$ mV \, & 5.29 & 11.8 & -15 & 5 & -5 &  \\
1466:     $v_2$ mV \, & -120 & -28.3 & -62.9 & -50 & -80 & -35 \\
1467:     $v_3$ mV \, & 25 & 19.2 & 10 & 10 & 10 & 10 \\
1468:     $t_1$ ms \, & -1.26 & -6.4 & 1.1 & 20 & 60 & 30\\
1469:     $t_2$ ms \, & 1.32 & 7.2 & 0  & 10 & 10 & 10
1470:   \end{tabular}
1471:   \caption{Parameter values for the Hodgkin-Huxley model
1472:          Eq.~(\ref{eq:HH-model}) with currents
1473:   Eq.~(\ref{eq:hh-currents}) and gating equations
1474:   Eqs.~(\ref{eq:x-sigmoid-fn})-(\ref{eq:ca-tau-form}).
1475:   }
1476: \end{table}
1477: 
1478: The currents Eqs.~(\ref{eq:hh-currents}) depend on the
1479: membrane voltage~$v(t)$ and on gating variables $m$,
1480: $h$, $n$, $m_1$, $h_1$, and~$n_1$ that determine the
1481: probability as a function of time for certain channel
1482: subunits to be open. All the gating variables~$x(v)$
1483: except the KCa activation variable~$n_1$ obey the
1484: evolution equation
1485: \begin{equation}
1486:   \label{eq:gating-variable-eq}
1487:   \tau(v) \frac{dx}{dt} = x_{\infty}(v) - x  ,
1488: \end{equation}
1489: with the asymptotic value~$x_\infty(v)$ having the
1490: sigmoidal form
1491: \begin{equation}
1492:   x_{\infty}(v) = \left( 
1493:     1 + e^{\displaystyle -(v-v_0)/v_1} \right)^{-1} .
1494:   \label{eq:x-sigmoid-fn}
1495: \end{equation}
1496: The time constant~$\tau(v)$ has the form
1497: \begin{equation}
1498:   \tau(v) = t_2 + 
1499:   t_1 \left(
1500:     1 + e^{\displaystyle
1501:     -\left( (v-v_2)/v_3 \right) }
1502:    \right)^{-1} .
1503:    \label{eq:hh-tau-fn}
1504: \end{equation}
1505: for the standard HH channels~$m$, $h$, and~$n$, and the form
1506: \begin{equation}
1507:   \label{eq:ca-tau-form}
1508:   \tau(v) = t_2 + 
1509:     t_1 \exp\left[  
1510:       -\left( \frac{v-v_2}{v_3} \right)^2 
1511:     \right] ,
1512: \end{equation}
1513: for the Ca channels~$m_1$ and~$h_1$ and KCa channel $n_1$.
1514: Table~\ref{table:hh-parameters} gives the parameter
1515: values  that we used in these various expressions.
1516: 
1517: The KCa~activation variable~$n_1$ obeys the evolution
1518: equation
1519: \begin{equation}
1520:   \tau_{\rm KCa} \frac{dn_1}{dt} = f\left([{\rm
1521:   Ca}]_i\right)-n_1 ,
1522: \end{equation}
1523: where the expression~$[{\rm Ca}]_i$ denotes the average
1524: intracellular calcium concentration and the
1525: function~$f$ is given by
1526: \begin{equation}
1527:   f(\cai) = \frac{\cai}{(\cai+0.001)(1+\exp(-(v+40)/20))} .
1528: \end{equation}
1529: The intracellular calcium concentration accumulates
1530: when the low threshold calcium channel is open and
1531: calcium ions flow from extracellular solution into
1532: intracellular solution.  The phenomenological model
1533: for~$\cai$ was taken from Ref.~\cite{Turrigiano1995}
1534: \begin{equation}
1535:   \cai = ([{\rm Ca}]_{i0} - \cai - F*(I_{ca}))/\tau_{Ca} ,
1536: \end{equation}
1537: where the equilibrium Calcium concentration~$[{\rm
1538:   Ca}]_{i0}$ had the value~.00005~mM and the
1539: current-density factor~$F$ had the value 3~${\rm
1540:   mM-cm}^2/{\rm mA}$, and the time constant~$\tau_{Ca}$
1541: had the value~200 ms.  The KCa channel was chosen to
1542: imitate a low-threshold transient calcium current
1543: observed in HVC neurons by Kubota~\cite{Kubota98}.
1544: %%% MengRu: how was the KCa channel ``chosen to
1545: %%%  imitate'', i.e., what did you fit?
1546: 
1547: 
1548: \begin{acknowledgments}
1549:   The authors thank Richard Mooney, Jonathan Prather,
1550:   Melissa Coleman, Michale Fee, Stephen Shea, George
1551:   Augustine, and Paul Tiesinga for helpful discussions,
1552:   and Craig Henriquez and Duke University's Center for
1553:   Neural Engineering for providing support for one of
1554:   the authors (M.~Li).
1555: \end{acknowledgments}
1556: 
1557: 
1558: %\input{figure-captions}
1559: 
1560: \begin{thebibliography}{56}
1561: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1562: \expandafter\ifx\csname bibnamefont\endcsname\relax
1563:   \def\bibnamefont#1{#1}\fi
1564: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1565:   \def\bibfnamefont#1{#1}\fi
1566: \expandafter\ifx\csname citenamefont\endcsname\relax
1567:   \def\citenamefont#1{#1}\fi
1568: \expandafter\ifx\csname url\endcsname\relax
1569:   \def\url#1{\texttt{#1}}\fi
1570: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1571: \providecommand{\bibinfo}[2]{#2}
1572: \providecommand{\eprint}[2][]{\url{#2}}
1573: 
1574: \bibitem[{\citenamefont{Doupe and Kuhl}(1999)}]{Doupe99}
1575: \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Doupe}} \bibnamefont{and}
1576:   \bibinfo{author}{\bibfnamefont{P.~K.} \bibnamefont{Kuhl}},
1577:   \bibinfo{journal}{Annu. Rev. Neurosci.} \textbf{\bibinfo{volume}{22}},
1578:   \bibinfo{pages}{567} (\bibinfo{year}{1999}).
1579: 
1580: \bibitem[{\citenamefont{Catchpole and Slater}(1995)}]{Catchpole95}
1581: \bibinfo{author}{\bibfnamefont{C.~K.} \bibnamefont{Catchpole}}
1582:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{P.~J.~B.}
1583:   \bibnamefont{Slater}}, \emph{\bibinfo{title}{Bird Song: Biological themes and
1584:   variations}} (\bibinfo{publisher}{Cambridge University Press},
1585:   \bibinfo{address}{New York}, \bibinfo{year}{1995}).
1586: 
1587: \bibitem[{\citenamefont{Johnson et~al.}(2002)\citenamefont{Johnson, Soderstrom,
1588:   and Whitney}}]{Johnson02}
1589: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Johnson}},
1590:   \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Soderstrom}},
1591:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{O.}~\bibnamefont{Whitney}},
1592:   \bibinfo{journal}{Behavioural Brain Research} \textbf{\bibinfo{volume}{131}},
1593:   \bibinfo{pages}{57} (\bibinfo{year}{2002}).
1594: 
1595: \bibitem[{\citenamefont{Hahnloser et~al.}(2002)\citenamefont{Hahnloser,
1596:   Kozhevnikov, and Fee}}]{Hahnloser02}
1597: \bibinfo{author}{\bibfnamefont{R.~H.~R.} \bibnamefont{Hahnloser}},
1598:   \bibinfo{author}{\bibfnamefont{A.~A.} \bibnamefont{Kozhevnikov}},
1599:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Fee}},
1600:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{419}}, \bibinfo{pages}{65}
1601:   (\bibinfo{year}{2002}).
1602: 
1603: \bibitem[{\citenamefont{Reiner et~al.}(2004)\citenamefont{Reiner, Perkel,
1604:   Mello, and Jarvis}}]{Reiner04}
1605: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Reiner}},
1606:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Perkel}},
1607:   \bibinfo{author}{\bibfnamefont{C.~V.} \bibnamefont{Mello}}, \bibnamefont{and}
1608:   \bibinfo{author}{\bibfnamefont{E.~D.} \bibnamefont{Jarvis}},
1609:   \bibinfo{journal}{Ann. N. Y. Acad. Sci.} \textbf{\bibinfo{volume}{1016}},
1610:   \bibinfo{pages}{77} (\bibinfo{year}{2004}).
1611: 
1612: \bibitem[{\citenamefont{Fee et~al.}(2004)\citenamefont{Fee, Kozhevnikov, and
1613:   Hahnloser}}]{Fee04}
1614: \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Fee}},
1615:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kozhevnikov}},
1616:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{R.~H.~R.}
1617:   \bibnamefont{Hahnloser}}, \bibinfo{journal}{Ann. N. Y. Acad. Sci.}
1618:   \textbf{\bibinfo{volume}{1016}}, \bibinfo{pages}{153} (\bibinfo{year}{2004}).
1619: 
1620: \bibitem[{\citenamefont{Jin et~al.}(2005)\citenamefont{Jin, Ramazanoglu, and
1621:   Seung}}]{Jin05}
1622: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Jin}},
1623:   \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Ramazanoglu}},
1624:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{H.~S.} \bibnamefont{Seung}}
1625:   (\bibinfo{year}{2005}), \bibinfo{note}{society for Neuroscience Abstracts
1626:   No.~753.15}.
1627: 
1628: \bibitem[{\citenamefont{Abeles}(1991)}]{Abeles91}
1629: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Abeles}},
1630:   \emph{\bibinfo{title}{Corticonics: Neural Circuits of the Cerebral Cortex}}
1631:   (\bibinfo{publisher}{Cambridge University Press},
1632:   \bibinfo{address}{Cambridge, England}, \bibinfo{year}{1991}).
1633: 
1634: \bibitem[{\citenamefont{Diesmann et~al.}(1999)\citenamefont{Diesmann, Gewaltig,
1635:   and Aertsen}}]{Diesmann99}
1636: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Diesmann}},
1637:   \bibinfo{author}{\bibfnamefont{M.-O.} \bibnamefont{Gewaltig}},
1638:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Aertsen}},
1639:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{402}},
1640:   \bibinfo{pages}{529} (\bibinfo{year}{1999}).
1641: 
1642: \bibitem[{\citenamefont{O\c{s}an et~al.}(2004)\citenamefont{O\c{s}an, Curtu,
1643:   Rubin, and Ermentrout}}]{Osan04}
1644: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{O\c{s}an}},
1645:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Curtu}},
1646:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Rubin}}, \bibnamefont{and}
1647:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Ermentrout}},
1648:   \bibinfo{journal}{J. Math. Biol.} \textbf{\bibinfo{volume}{48}},
1649:   \bibinfo{pages}{243} (\bibinfo{year}{2004}).
1650: 
1651: \bibitem[{\citenamefont{Reyes}(2003)}]{Reyes03}
1652: \bibinfo{author}{\bibfnamefont{A.~D.} \bibnamefont{Reyes}},
1653:   \bibinfo{journal}{Nature Neuroscience} \textbf{\bibinfo{volume}{6}},
1654:   \bibinfo{pages}{593} (\bibinfo{year}{2003}).
1655: 
1656: \bibitem[{\citenamefont{Golomb and Amitai}(1997)}]{Golomb97}
1657: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Golomb}} \bibnamefont{and}
1658:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Amitai}},
1659:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{78}},
1660:   \bibinfo{pages}{1199} (\bibinfo{year}{1997}).
1661: 
1662: \bibitem[{\citenamefont{Litvak et~al.}(2003)\citenamefont{Litvak, Sompolinsky,
1663:   Segev, and Abeles}}]{Litvak03}
1664: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Litvak}},
1665:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Sompolinsky}},
1666:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Segev}}, \bibnamefont{and}
1667:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Abeles}},
1668:   \bibinfo{journal}{Journal of Neuroscience} \textbf{\bibinfo{volume}{23}},
1669:   \bibinfo{pages}{3096} (\bibinfo{year}{2003}).
1670: 
1671: \bibitem[{\citenamefont{Vogels and Abbott}(2095)}]{Vogels05}
1672: \bibinfo{author}{\bibfnamefont{T.~P.} \bibnamefont{Vogels}} \bibnamefont{and}
1673:   \bibinfo{author}{\bibfnamefont{L.~E.} \bibnamefont{Abbott}},
1674:   \bibinfo{journal}{J. of Neuroscience} \textbf{\bibinfo{volume}{25}},
1675:   \bibinfo{pages}{10786} (\bibinfo{year}{2095}).
1676: 
1677: \bibitem[{\citenamefont{Mooney}(2000)}]{Mooney00}
1678: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Mooney}}, \bibinfo{journal}{J.
1679:   Neurosci.} \textbf{\bibinfo{volume}{20}}, \bibinfo{pages}{5420}
1680:   (\bibinfo{year}{2000}).
1681: 
1682: \bibitem[{\citenamefont{Ohki et~al.}(2005)\citenamefont{Ohki, Chung, Ch'ng,
1683:   Kara, and Reid}}]{Ohki05}
1684: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Ohki}},
1685:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Chung}},
1686:   \bibinfo{author}{\bibfnamefont{Y.~H.} \bibnamefont{Ch'ng}},
1687:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Kara}}, \bibnamefont{and}
1688:   \bibinfo{author}{\bibfnamefont{R.~C.} \bibnamefont{Reid}},
1689:   \bibinfo{journal}{Nature} \textbf{\bibinfo{volume}{433}},
1690:   \bibinfo{pages}{597} (\bibinfo{year}{2005}).
1691: 
1692: \bibitem[{\citenamefont{Li and Greenside}(2006)}]{Li06II}
1693: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Li}} \bibnamefont{and}
1694:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Greenside}}
1695:   (\bibinfo{year}{2006}), \bibinfo{note}{in preparation.}
1696: 
1697: \bibitem[{\citenamefont{Li}(2006)}]{Li06thesis}
1698: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Li}}, Ph.D. thesis,
1699:   \bibinfo{school}{Duke University} (\bibinfo{year}{2006}).
1700: 
1701: \bibitem[{\citenamefont{Kubota and Taniguchi}(1998)}]{Kubota98}
1702: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kubota}} \bibnamefont{and}
1703:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Taniguchi}},
1704:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{80}},
1705:   \bibinfo{pages}{914} (\bibinfo{year}{1998}).
1706: 
1707: \bibitem[{\citenamefont{Mooney and Prather}(2005)}]{Mooney05}
1708: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Mooney}} \bibnamefont{and}
1709:   \bibinfo{author}{\bibfnamefont{J.~F.} \bibnamefont{Prather}},
1710:   \bibinfo{journal}{Journal of Neuroscience} \textbf{\bibinfo{volume}{25}},
1711:   \bibinfo{pages}{1952} (\bibinfo{year}{2005}).
1712: 
1713: \bibitem[{\citenamefont{Dutar et~al.}(1998)\citenamefont{Dutar, Vu, and
1714:   Perkel}}]{Dutar98}
1715: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Dutar}},
1716:   \bibinfo{author}{\bibfnamefont{H.~M.} \bibnamefont{Vu}}, \bibnamefont{and}
1717:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Perkel}},
1718:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{80}},
1719:   \bibinfo{pages}{1828} (\bibinfo{year}{1998}).
1720: 
1721: \bibitem[{\citenamefont{Wild et~al.}(2005)\citenamefont{Wild, Williams, Howie,
1722:   and Mooney}}]{Wild05}
1723: \bibinfo{author}{\bibfnamefont{J.~M.} \bibnamefont{Wild}},
1724:   \bibinfo{author}{\bibfnamefont{M.~N.} \bibnamefont{Williams}},
1725:   \bibinfo{author}{\bibfnamefont{G.~J.} \bibnamefont{Howie}}, \bibnamefont{and}
1726:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Mooney}},
1727:   \bibinfo{journal}{Journal of Comparative Neurology}
1728:   \textbf{\bibinfo{volume}{483}}, \bibinfo{pages}{76} (\bibinfo{year}{2005}).
1729: 
1730: \bibitem[{\citenamefont{Traub et~al.}(1999)\citenamefont{Traub, Jefferys, and
1731:   Whittington}}]{Traub99}
1732: \bibinfo{author}{\bibfnamefont{R.~D.} \bibnamefont{Traub}},
1733:   \bibinfo{author}{\bibfnamefont{J.~G.~R.} \bibnamefont{Jefferys}},
1734:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~A.}
1735:   \bibnamefont{Whittington}}, \emph{\bibinfo{title}{Fast Oscillations in
1736:   Cortical Circuits}} (\bibinfo{publisher}{Bradford Books},
1737:   \bibinfo{address}{New York}, \bibinfo{year}{1999}).
1738: 
1739: \bibitem[{\citenamefont{Nottebohm}(2002)}]{NottebohmJN02}
1740: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Nottebohm}},
1741:   \bibinfo{journal}{J Neuroscience} \textbf{\bibinfo{volume}{22}},
1742:   \bibinfo{pages}{624} (\bibinfo{year}{2002}).
1743: 
1744: \bibitem[{\citenamefont{Wang et~al.}(2002)\citenamefont{Wang, Hurley, Pytte,
1745:   and Kirn}}]{Wang02}
1746: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Wang}},
1747:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Hurley}},
1748:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Pytte}}, \bibnamefont{and}
1749:   \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Kirn}},
1750:   \bibinfo{journal}{J. Neurosci.} \textbf{\bibinfo{volume}{22}},
1751:   \bibinfo{pages}{10864} (\bibinfo{year}{2002}).
1752: 
1753: \bibitem[{\citenamefont{Ikegaya et~al.}(2004)\citenamefont{Ikegaya, Aaron,
1754:   Cossart, Aronov, Lampl, Ferster, and Yuste}}]{Ikegaya04}
1755: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Ikegaya}},
1756:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Aaron}},
1757:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cossart}},
1758:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Aronov}},
1759:   \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Lampl}},
1760:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Ferster}}, \bibnamefont{and}
1761:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Yuste}},
1762:   \bibinfo{journal}{Science} \textbf{\bibinfo{volume}{304}},
1763:   \bibinfo{pages}{559} (\bibinfo{year}{2004}).
1764: 
1765: \bibitem[{\citenamefont{Leonardo and Fee}(2005)}]{Leonardo05}
1766: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Leonardo}} \bibnamefont{and}
1767:   \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Fee}},
1768:   \bibinfo{journal}{Journal of Neuroscience} \textbf{\bibinfo{volume}{25}},
1769:   \bibinfo{pages}{652} (\bibinfo{year}{2005}).
1770: 
1771: \bibitem[{\citenamefont{Doya and Sejnowski}(1998)}]{Doya98}
1772: \bibinfo{author}{\bibfnamefont{K.}~\bibnamefont{Doya}} \bibnamefont{and}
1773:   \bibinfo{author}{\bibfnamefont{T.~J.} \bibnamefont{Sejnowski}}, in
1774:   \emph{\bibinfo{booktitle}{Central Auditory Processing and Neural Modeling}},
1775:   edited by \bibinfo{editor}{\bibnamefont{Poon}} \bibnamefont{and}
1776:   \bibinfo{editor}{\bibnamefont{Brugge}} (\bibinfo{publisher}{Plenum Press},
1777:   \bibinfo{address}{New York}, \bibinfo{year}{1998}).
1778: 
1779: \bibitem[{\citenamefont{Abarbanel et~al.}(2004)\citenamefont{Abarbanel, Gibb,
1780:   Mindlin, and Talathi}}]{Abarbanel04jn}
1781: \bibinfo{author}{\bibfnamefont{H.~D.~I.} \bibnamefont{Abarbanel}},
1782:   \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Gibb}},
1783:   \bibinfo{author}{\bibfnamefont{G.~B.} \bibnamefont{Mindlin}},
1784:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Talathi}},
1785:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{92}},
1786:   \bibinfo{pages}{96} (\bibinfo{year}{2004}).
1787: 
1788: \bibitem[{\citenamefont{Berger}(2004)}]{Berger04}
1789: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Berger}}, Master's thesis,
1790:   \bibinfo{school}{Institute of Neuroinformatics, University/ETH Z\"urich}
1791:   (\bibinfo{year}{2004}).
1792: 
1793: \bibitem[{\citenamefont{Drew and Abbott}(2003)}]{Drew03}
1794: \bibinfo{author}{\bibfnamefont{P.~J.} \bibnamefont{Drew}} \bibnamefont{and}
1795:   \bibinfo{author}{\bibfnamefont{L.~F.} \bibnamefont{Abbott}},
1796:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{89}},
1797:   \bibinfo{pages}{2697} (\bibinfo{year}{2003}).
1798: 
1799: \bibitem[{\citenamefont{Kleinfeld and Sompolinsky}(1989)}]{Kleinfeld89}
1800: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Kleinfeld}} \bibnamefont{and}
1801:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Sompolinsky}}, in
1802:   \emph{\bibinfo{booktitle}{In Methods in Neuronal Modeling: From Synapse to
1803:   Networks}}, edited by \bibinfo{editor}{\bibfnamefont{C.}~\bibnamefont{Koch}}
1804:   \bibnamefont{and} \bibinfo{editor}{\bibfnamefont{I.}~\bibnamefont{Segev}}
1805:   (\bibinfo{publisher}{MIT Press}, \bibinfo{address}{Cambridge, MA},
1806:   \bibinfo{year}{1989}), pp. \bibinfo{pages}{195--246}.
1807: 
1808: \bibitem[{\citenamefont{Huerta and Rabinovich}(2004)}]{Huerta04}
1809: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Huerta}} \bibnamefont{and}
1810:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Rabinovich}},
1811:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{93}},
1812:   \bibinfo{pages}{238104 (4 pages)} (\bibinfo{year}{2004}).
1813: 
1814: \bibitem[{\citenamefont{Dayan and Abbott}(2001)}]{Dayan01}
1815: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Dayan}} \bibnamefont{and}
1816:   \bibinfo{author}{\bibfnamefont{L.~F.} \bibnamefont{Abbott}},
1817:   \emph{\bibinfo{title}{Theoretical Neuroscience: Computational and
1818:   Mathematical Modeling of Neural Systems}} (\bibinfo{publisher}{MIT Press},
1819:   \bibinfo{year}{2001}).
1820: 
1821: \bibitem[{\citenamefont{Kincaid and Cheney}(1996)}]{Kincaid96}
1822: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Kincaid}} \bibnamefont{and}
1823:   \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Cheney}},
1824:   \emph{\bibinfo{title}{Numerical Analysis}} (\bibinfo{publisher}{Brooks/Cole},
1825:   \bibinfo{address}{Pacific Grove, California}, \bibinfo{year}{1996}),
1826:   \bibinfo{edition}{2nd} ed.
1827: 
1828: \bibitem[{\citenamefont{MathWorks}()}]{Matlab}
1829: \bibinfo{author}{\bibnamefont{MathWorks}}, \emph{\bibinfo{title}{Matlab}},
1830:   \bibinfo{howpublished}{http://www.mathworks.com/products/matlab/}.
1831: 
1832: \bibitem[{\citenamefont{Kubota and Saito}(1991)}]{Kubota91}
1833: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Kubota}} \bibnamefont{and}
1834:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Saito}},
1835:   \bibinfo{journal}{Journal of Physiology} \textbf{\bibinfo{volume}{440}},
1836:   \bibinfo{pages}{131} (\bibinfo{year}{1991}).
1837: 
1838: \bibitem[{\citenamefont{Prinz et~al.}(2003)\citenamefont{Prinz, Billimoria, and
1839:   Marder}}]{Prinz03}
1840: \bibinfo{author}{\bibfnamefont{A.~A.} \bibnamefont{Prinz}},
1841:   \bibinfo{author}{\bibfnamefont{C.~P.} \bibnamefont{Billimoria}},
1842:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Marder}},
1843:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{90}},
1844:   \bibinfo{pages}{3998} (\bibinfo{year}{2003}).
1845: 
1846: \bibitem[{\citenamefont{Huguenard and McCormick}(1992)}]{Huguenard92}
1847: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Huguenard}} \bibnamefont{and}
1848:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{McCormick}},
1849:   \bibinfo{journal}{J. Neurophysiology} \textbf{\bibinfo{volume}{68}},
1850:   \bibinfo{pages}{1373} (\bibinfo{year}{1992}).
1851: 
1852: \bibitem[{\citenamefont{Koch and Segev}(1998)}]{Yamada98}
1853: \bibinfo{editor}{\bibfnamefont{C.}~\bibnamefont{Koch}} \bibnamefont{and}
1854:   \bibinfo{editor}{\bibfnamefont{I.}~\bibnamefont{Segev}}, eds.,
1855:   \emph{\bibinfo{title}{Multiple Channels and Calcium Dynamics}}
1856:   (\bibinfo{publisher}{MIT}, \bibinfo{year}{1998}).
1857: 
1858: \bibitem[{\citenamefont{Hines et~al.}()\citenamefont{Hines, Moore, and
1859:   Carnevale}}]{NeuronCode}
1860: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Hines}},
1861:   \bibinfo{author}{\bibfnamefont{J.~W.} \bibnamefont{Moore}}, \bibnamefont{and}
1862:   \bibinfo{author}{\bibfnamefont{T.}~\bibnamefont{Carnevale}},
1863:   \emph{\bibinfo{title}{Neuron}},
1864:   \bibinfo{howpublished}{http://neuron.duke.edu/}.
1865: 
1866: \bibitem[{\citenamefont{Hertz et~al.}(1991)\citenamefont{Hertz, Krogh, and
1867:   Palmer}}]{Hertz91}
1868: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Hertz}},
1869:   \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Krogh}}, \bibnamefont{and}
1870:   \bibinfo{author}{\bibfnamefont{R.~G.} \bibnamefont{Palmer}},
1871:   \emph{\bibinfo{title}{Introduction to the Theory of Neural Computation}},
1872:   no.~\bibinfo{number}{I} in \bibinfo{series}{Santa Fe Institute Studies in the
1873:   Sciences of Complexity} (\bibinfo{publisher}{Westview},
1874:   \bibinfo{address}{Boulder, CO}, \bibinfo{year}{1991}).
1875: 
1876: \bibitem[{\citenamefont{Hopfield}(1982)}]{Hopfield82}
1877: \bibinfo{author}{\bibfnamefont{J.~J.} \bibnamefont{Hopfield}},
1878:   \bibinfo{journal}{Proc. Nat. Acad. Sci.} \textbf{\bibinfo{volume}{79}},
1879:   \bibinfo{pages}{2554} (\bibinfo{year}{1982}).
1880: 
1881: \bibitem[{\citenamefont{Coleman and Mooney}(2004)}]{Coleman04}
1882: \bibinfo{author}{\bibfnamefont{M.~J.} \bibnamefont{Coleman}} \bibnamefont{and}
1883:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Mooney}},
1884:   \bibinfo{journal}{Journal of Neuroscience} \textbf{\bibinfo{volume}{24}},
1885:   \bibinfo{pages}{7251} (\bibinfo{year}{2004}).
1886: 
1887: \bibitem[{\citenamefont{Rosen and Mooney}(2006)}]{Rosen06}
1888: \bibinfo{author}{\bibfnamefont{M.~J.} \bibnamefont{Rosen}} \bibnamefont{and}
1889:   \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Mooney}},
1890:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{95}},
1891:   \bibinfo{pages}{1158} (\bibinfo{year}{2006}).
1892: 
1893: \bibitem[{\citenamefont{Solis and Perkel}(2005)}]{Solis05}
1894: \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{Solis}} \bibnamefont{and}
1895:   \bibinfo{author}{\bibfnamefont{D.~J.} \bibnamefont{Perkel}},
1896:   \bibinfo{journal}{J. Neuroscience} \textbf{\bibinfo{volume}{25}},
1897:   \bibinfo{pages}{2811} (\bibinfo{year}{2005}).
1898: 
1899: \bibitem[{\citenamefont{Fiete et~al.}(2004)\citenamefont{Fiete, Hahnloser, Fee,
1900:   and Sueng}}]{Fiete04}
1901: \bibinfo{author}{\bibfnamefont{I.~R.} \bibnamefont{Fiete}},
1902:   \bibinfo{author}{\bibfnamefont{R.~H.~R.} \bibnamefont{Hahnloser}},
1903:   \bibinfo{author}{\bibfnamefont{M.~S.} \bibnamefont{Fee}}, \bibnamefont{and}
1904:   \bibinfo{author}{\bibfnamefont{H.~S.} \bibnamefont{Sueng}},
1905:   \bibinfo{journal}{J. Neurophysiol.} \textbf{\bibinfo{volume}{92}},
1906:   \bibinfo{pages}{2274} (\bibinfo{year}{2004}).
1907: 
1908: \bibitem[{\citenamefont{Lisman}(1997)}]{Lisman97}
1909: \bibinfo{author}{\bibfnamefont{J.~E.} \bibnamefont{Lisman}},
1910:   \bibinfo{journal}{Trends Neurosci.} \textbf{\bibinfo{volume}{20}},
1911:   \bibinfo{pages}{38} (\bibinfo{year}{1997}).
1912: 
1913: \bibitem[{\citenamefont{de~la Prida et~al.}(1997)\citenamefont{de~la Prida,
1914:   Stollenwerk, and Sanchez-Andres}}]{Prida97}
1915: \bibinfo{author}{\bibfnamefont{L.~M.} \bibnamefont{de~la Prida}},
1916:   \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Stollenwerk}},
1917:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{J.~V.}
1918:   \bibnamefont{Sanchez-Andres}}, \bibinfo{journal}{Physica D}
1919:   \textbf{\bibinfo{volume}{110}}, \bibinfo{pages}{323} (\bibinfo{year}{1997}).
1920: 
1921: \bibitem[{\citenamefont{Vu et~al.}(1994)\citenamefont{Vu, Mazurek, and
1922:   Kuo}}]{Vu94}
1923: \bibinfo{author}{\bibfnamefont{E.~T.} \bibnamefont{Vu}},
1924:   \bibinfo{author}{\bibfnamefont{M.~E.} \bibnamefont{Mazurek}},
1925:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{Y.-C.} \bibnamefont{Kuo}},
1926:   \bibinfo{journal}{J. Neuroscience} \textbf{\bibinfo{volume}{14}},
1927:   \bibinfo{pages}{6924} (\bibinfo{year}{1994}).
1928: 
1929: \bibitem[{\citenamefont{Vu et~al.}(1998)\citenamefont{Vu, Schmidt, and
1930:   Mazurek}}]{Vu98}
1931: \bibinfo{author}{\bibfnamefont{E.~T.} \bibnamefont{Vu}},
1932:   \bibinfo{author}{\bibfnamefont{M.~F.} \bibnamefont{Schmidt}},
1933:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{M.~E.}
1934:   \bibnamefont{Mazurek}}, \bibinfo{journal}{J. Neuroscience}
1935:   \textbf{\bibinfo{volume}{18}}, \bibinfo{pages}{9088} (\bibinfo{year}{1998}).
1936: 
1937: \bibitem[{\citenamefont{Vicario and Simpson}(1995)}]{Vicario95}
1938: \bibinfo{author}{\bibfnamefont{D.~S.} \bibnamefont{Vicario}} \bibnamefont{and}
1939:   \bibinfo{author}{\bibfnamefont{H.~B.} \bibnamefont{Simpson}},
1940:   \bibinfo{journal}{J. Neurophys.} \textbf{\bibinfo{volume}{73}},
1941:   \bibinfo{pages}{2602} (\bibinfo{year}{1995}).
1942: 
1943: \bibitem[{\citenamefont{Losonczy and Magee}(2006)}]{Losonczy06}
1944: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Losonczy}} \bibnamefont{and}
1945:   \bibinfo{author}{\bibfnamefont{J.~C.} \bibnamefont{Magee}},
1946:   \bibinfo{journal}{Neuron} \textbf{\bibinfo{volume}{50}}, \bibinfo{pages}{291}
1947:   (\bibinfo{year}{2006}).
1948: 
1949: \bibitem[{\citenamefont{Berg\'e et~al.}(1984)\citenamefont{Berg\'e, Pomeau, and
1950:   Vidal}}]{Berge84}
1951: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Berg\'e}},
1952:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Pomeau}}, \bibnamefont{and}
1953:   \bibinfo{author}{\bibfnamefont{C.}~\bibnamefont{Vidal}},
1954:   \emph{\bibinfo{title}{Order Within Chaos}} (\bibinfo{publisher}{John Wiley \&
1955:   Sons}, \bibinfo{address}{New York}, \bibinfo{year}{1984}).
1956: 
1957: \bibitem[{\citenamefont{Kopell and Ermentrout}(2004)}]{Kopell04}
1958: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Kopell}} \bibnamefont{and}
1959:   \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Ermentrout}},
1960:   \bibinfo{journal}{PNAS} \textbf{\bibinfo{volume}{101}},
1961:   \bibinfo{pages}{15482} (\bibinfo{year}{2004}).
1962: 
1963: \bibitem[{\citenamefont{Turrigiano et~al.}(1995)\citenamefont{Turrigiano,
1964:   Lemasson, and Marder}}]{Turrigiano1995}
1965: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Turrigiano}},
1966:   \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Lemasson}}, \bibnamefont{and}
1967:   \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Marder}}, \bibinfo{journal}{J
1968:   Neuroscience} \textbf{\bibinfo{volume}{15}}, \bibinfo{pages}{3640}
1969:   (\bibinfo{year}{1995}).
1970: 
1971: \end{thebibliography}
1972: 
1973: \end{document}