q-bio0703001/paper.tex
1: \documentclass[twocolumn,aps,showpacs,floatfix,superscriptaddress]{revtex4}
2: \usepackage{amsmath,amssymb,eucal,graphicx,subfigure}
3: \begin{document}
4: 
5: \title{Dynamics of an Idealized Model of Microtubule Growth and Catastrophe}
6: \author{T. Antal}
7: \affiliation{Program for Evolutionary Dynamics, Harvard University,
8:   Cambridge, MA 02138, USA}
9: \author{P. L. Krapivsky}
10: \affiliation{Center for Polymer Studies and Department of Physics, Boston University, Boston, MA 02215, USA}
11: \author{S. Redner}
12: \affiliation{Center for Polymer Studies and Department of Physics, Boston University, Boston, MA 02215, USA}
13: \author{M. Mailman}
14: \affiliation{Martin Fisher School of Physics, Brandeis University, Waltham, MA 02454, USA}
15: \author{B. Chakraborty}
16: \affiliation{Martin Fisher School of Physics, Brandeis University, Waltham, MA 02454, USA}
17: \begin{abstract}
18: 
19:   We investigate a simple dynamical model of a microtubule that evolves by
20:   attachment of guanosine triphosphate (GTP) tubulin to its end, irreversible
21:   conversion of GTP to guanosine diphosphate (GDP) tubulin by hydrolysis, and
22:   detachment of GDP at the end of a microtubule.  As a function of rates of
23:   these processes, the microtubule can grow steadily or its length can
24:   fluctuate wildly.  In the regime where detachment can be neglected, we find
25:   exact expressions for the tubule and GTP cap length distributions, as well
26:   as power-law length distributions of GTP and GDP islands.  In the opposite
27:   limit of instantaneous detachment, we find the time between catastrophes,
28:   where the microtubule shrinks to zero length, and determine the size
29:   distribution of avalanches (sequence of consecutive GDP detachment events).
30:   We obtain the phase diagram for general rates and verify our predictions by
31:   numerical simulations.
32: 
33: 
34: \end{abstract}
35: \pacs{87.16.Ka, 87.17.Aa, 02.50.Ey, 05.40.-a}
36: 
37: \maketitle
38: 
39: \section{INTRODUCTION AND MODEL}
40: 
41: Microtubules are polar linear polymers that perform major organizational
42: tasks in living cells \cite{FBL,VCJ}.  Through a unique feature of
43: microtubule assembly, termed dynamic instability \cite{MK}, they function as
44: molecular machines \cite{AH} that move cellular structures during processes
45: such as cell reproduction \cite{VCJ,WMD}.  A surprising feature of
46: microtubules is that they remain out of equilibrium under fixed external
47: conditions and can undergo alternating periods of rapid growth and even more
48: rapid shrinking (Fig.~\ref{l-vs-t}).
49: 
50: These sudden polymerization changes are driven by the interplay between
51: several fundamental processes.  Microtubules grow by the attachment of
52: guanosine triphosphate tubulin complexes (GTP) at one end \cite{MK,DM}.
53: Structural studies indicate that the end of a microtubule must consist of a
54: ``cap'' of consecutive GTP monomers \cite{dimer} for growth to continue
55: \cite{DM}.  Once polymerized, the GTP of this complex can irreversibly
56: hydrolyze into guanosine diphosphate (GDP).  If all the monomers in the cap
57: convert to GDP, the microtubule is destabilized and rapid shrinkage ensues by
58: the detachment of GDP tubulin units.  The competition between GTP attachment
59: and hydrolysis from GTP to GDP is believed to lead to the dynamic instability
60: in which the GTP cap hydrolyzes to GDP and then the microtubule rapidly
61: depolymerizes.  The stochastic attachment of GTP can, however, lead to a
62: rescue to the growing phase before the microtubule length shrinks to zero
63: \cite{FBL,DL}.
64: 
65: The origin of this dynamic instability has been actively investigated.  One
66: avenue of theoretical work on this dynamical instability is based on models
67: of mechanical stability \cite{JCF,vanburen05,mol}.  For example, a detailed
68: stochastic model of a microtubule that includes all the thirteen constituent
69: protofilaments has been investigated in Ref.~\cite{vanburen05}.  By using
70: model parameters that were inferred from equilibrium statistical physics,
71: VanBuren et al.\ \cite{vanburen05} found some characteristics of microtubule
72: evolution that agreed with experimental data \cite{mandelkow91}.  The
73: disadvantage of this detailed modeling, however, is its complexity, so that
74: it is generally not possible to develop an intuitive understanding of
75: microtubule evolution.
76: 
77: Another approach for modeling the dynamics of microtubules is based on
78: effective two-state models that describe the dynamics in terms of a switching
79: between a growing and a shrinking state \cite{DL,Bi,HZ,MKMC,MGGA,hill}.  The
80: essence of many of these models is that a microtubule exists either in a
81: growing phase (where a GTP cap exists at the end of the microtubule) or a
82: shrinking phase (without a GTP cap), and that there are stochastic
83: transitions between these two states.  By tuning parameters appropriately, it
84: is possible to reproduce the phase changes between the growing and shrinking
85: phases of microtubules that have been observed experimentally \cite{MK}.
86: While the two-state model has the advantage of having only a few parameters,
87: a constant rate of switching between a growing and shrinking microtubule is
88: built into the model.  Thus switching models cannot account for the
89: stochastic avalanches and catastrophes that occur in real microtubules.
90: 
91: On the other hand, a minimalist model of microtubule dynamics has been
92: proposed and investigated by Flyvbjerg et al.\ \cite{F}.  In their model,
93: they dispense with attempts to capture all of the myriad of experimental
94: parameters within a detailed model, but instead constructed an effective
95: continuous theory to describe microtubule dynamics.  Their goal was to
96: construct an effective theory that contained as few details as possible.  As
97: stated in Ref.~\cite{F}, they envision that their effective theory should be
98: derivable from a fundamental, microscopic theory and its parameters.
99: 
100: This minimalist modeling is the approach that we adopt in the present work.
101: We investigate a recently introduced \cite{ZLSW,CR} kinetic model that
102: accounts for many aspects of microtubule evolution.  Our main result is that
103: only a few essential parameters with simple physical interpretations are
104: needed to describe the rich features of microtubule growth, catastrophes, and
105: rescues \cite{short}.
106: 
107: We treat a microtubule as a linear polymer that consists of GTP or GDP
108: monomers that we denote as $+$ and $-$, respectively.  To emphasize this
109: connection between chemistry and the model, we will write the former as
110: GTP$^+$ and the latter as GDP$^-$.  The state of a microtubule evolves due to
111: the following three processes:
112: \begin{enumerate}
113: \item Attachment: A microtubule grows by attachment of a guanosine
114:   triphosphate (GTP$^+$) monomer.
115: \begin{eqnarray*}
116:     &|\cdots +\rangle \Longrightarrow |\cdots ++\rangle\qquad {\rm rate}~ \lambda\\
117:     &~~\,|\cdots -\rangle \Longrightarrow |\cdots -+\rangle\qquad {\rm  rate}~ p\lambda.
118: \end{eqnarray*}
119: 
120: \item Conversion: Once part of the microtubule, each GTP$^+$ can
121:   independently convert by hydrolysis to a guanosine diphosphate (GDP$^-$).
122: \begin{eqnarray*}
123:  ~~~~~~~~~~~|\cdots +\cdots\rangle \Longrightarrow |\cdots-\cdots \rangle \qquad {\rm rate}~ 1.
124: \end{eqnarray*}
125: 
126: \item Detachment: a microtubule shrinks due to detachment of a GDP$^-$
127:   monomer {\em only from the end of the microtubule}.
128: \begin{eqnarray*}
129: |\cdots-\,\rangle\Longrightarrow |\cdots\rangle\qquad {\rm rate}~ \mu.
130: \end{eqnarray*}
131: \end{enumerate}
132: Here the symbols $|$ and $\rangle$ denote the terminal and the active end of
133: the microtubule.  It is worth mentioning that these steps are similar to
134: those in a recently-introduced model of DNA sequence evolution \cite{MLA},
135: and that some of the results about the structure of DNA sequences seem to be
136: related to our results about island size distributions in microtubules.
137: 
138: \begin{figure}[ht]
139: \includegraphics*[width=0.39\textwidth]{mu5lam1.4.eps}
140: \includegraphics*[width=0.39\textwidth]{mu5lam1.5.eps}
141: \includegraphics*[width=0.39\textwidth]{mu5lam1.6.eps}
142: \caption{Numerical simulations of typical microtubule lengths versus time for
143:   detachment rate $\mu=5$ and attachment rates: (a) $\lambda=1.4,$ where the
144:   microtubule generally remains short, (b) $\lambda=1.5$, where the length
145:   fluctuates strongly, and (c) $\lambda= 1.6$, where the microtubule grows
146:   nearly steadily.}
147: \label{l-vs-t}
148: \end{figure}
149: 
150: Generically, the $(\lambda,\mu,p)$ phase space separates into a region where
151: the microtubule grows (on average) with a certain rate $V(\lambda,\mu,p)$,
152: and a compact phase where the average microtubule length is finite.  These
153: two phases are separated by a phase boundary $\mu=\mu_*(\lambda,p)$ along
154: which the growth rate $V(\lambda,\mu,p)$ vanishes.  While the behavior of a
155: microtubule for general parameter values is of interest, we will primarily
156: focus on extreme values of the governing parameters where we can obtain a
157: detailed statistical characterization of the microtubule structure.  For
158: certain properties, such as the shape of the phase diagram, we will also
159: present results of numerical simulations of the model.
160: 
161: In Sec.~\ref{growth}, we study the evolution of a microtubule under
162: unrestricted growth conditions---namely no detachment and an attachment rate
163: that does not depend on the identity of the last monomer.  Our results here
164: are relevant to understanding the distribution of cap length and the
165: diffusion coefficient of the tip of the microtubule in the growth phase.  The
166: predictions of the model in this limit could also be useful in understanding
167: the binding pattern of proteins to microtubules \cite{TGSM}.  Since proteins
168: are important regulatory factors in microtubule polymerization, these results
169: could prove useful in interpreting the effects of proteins on microtubule
170: growth.
171: 
172: By a master equation approach, we will determine both the number of GTP$^+$
173: monomers on a microtubule, as well as the length distributions of GTP$^+$
174: and GDP$^-$ islands (Fig.~\ref{cartoon}).  Many of these analytical
175: predictions are verified by numerical simulations.  In
176: Sec.~\ref{constrained}, we extend our approach to the case of constrained
177: growth, $p\ne 1$, in which microtubule growth depends on whether the last
178: monomer is a GTP$^+$ or a GDP$^-$.  In Sec.~\ref{instant}, we investigate the
179: phenomenon of ``catastrophe'' for infinite detachment rate $\mu$, in which a
180: microtubule shrinks to zero length when all of its constituent monomers
181: convert to GDP$^-$.  We derive the asymptotic behavior of the catastrophe
182: probability by expressing it as an infinite product and recognizing the
183: connection of this product with modular functions.  We also determine the
184: asymptotic behavior of the size distribution of avalanches, namely, sequences
185: of consecutive GDP$^-$ detachment events.  Finally, in Sec.~\ref{general}, we
186: discuss the behavior of a microtubule for general parameter values through a
187: combination of numerical and analytic results.  Here numerical simulations
188: are useful to extract quantitative results for parameter values that are note
189: amenable to theoretical analysis.  Several calculational details are given in
190: the appendixes.
191: 
192: \section{UNRESTRICTED GROWTH}
193: \label{growth}
194: 
195: We define unrestricted growth as the limit of detachment rate $\mu=0$, so
196: that a microtubule grows without bound.  Here we consider the special case
197: where the attachment rate does not depend on the identity of the last
198: monomer; that is, the limit of $p=1$, where the attachment is unconstrained.
199: Because of the latter condition, the number $N$ of GTP$^+$ monomers decouples
200: from the number of GDP$^-$, a greatly simplifying feature.
201: 
202: \begin{figure}[ht]
203: \includegraphics*[width=0.4\textwidth]{cartoon}
204: \caption{(Color online)~ Cartoon of a microtubule in unrestricted growth.
205:   Regions of GTP$^+$ are shown dark (blue) and regions of GDP$^-$ are light
206:   (yellow).  The GTP$^+$ regions get shorter further from the tip that
207:   advances as $\lambda t$, while the GDP$^-$ regions get longer.}
208: \label{cartoon}
209: \end{figure}
210: 
211: \subsection{Distribution of Positive Monomers}
212: 
213: The average number of GTP$^+$ monomers evolves as
214: \begin{equation}
215: \label{N-eq}
216: \frac{d}{d t}\,\langle N\rangle =\lambda -\langle N\rangle .
217: \end{equation}
218: The gain term accounts for the adsorption of a GTP$^+$ at rate $\lambda$,
219: while the loss term accounts for the conversion events GTP$^+$ $\to$ GDP$^-$,
220: each of which occurs with rate 1.  Thus $\langle N\rangle$ approaches its
221: stationary value of $\lambda$ exponentially quickly,
222: \begin{equation}
223: \label{N-av}
224: \langle N\rangle = \lambda (1-e^{-t}).
225: \end{equation}
226: 
227: More generally, consider the probability $\Pi_N(t)$ that there are $N$ GTP$^+$
228: monomers at time $t$.  This probability evolves according to
229: \begin{equation}
230: \label{PNt}
231: \frac{d\Pi_N}{dt} = -(N+\lambda)\Pi_N+\lambda\Pi_{N-1}+(N+1)\Pi_{N+1}.
232: \end{equation}
233: The loss term $(N+\lambda)\Pi_N$ accounts for conversion events GTP$^+$ $\to$
234: GDP$^-$ that occur with total rate $N$, and the attachment of a GTP$^+$ at
235: the end of the microtubule of length $N$ with rate $\lambda$.  The gain terms
236: can be explained similarly.
237: 
238: In terms of generating function $\Pi(z)\equiv\sum_{N=0}^\infty \Pi_N z^N$,
239: Eq.~\eqref{PNt} can be recast as the differential equation
240: \begin{equation}
241: \label{PNz}
242: \frac{\partial\Pi}{\partial t} = (1-z)\left(\frac{\partial \Pi}{\partial
243:       z}-\lambda \Pi\right).
244: \end{equation}
245: Introducing $\mathcal{Q}=\Pi e^{-\lambda z}$ and 
246: $y=\log (1-z)$, we
247: transform Eq.~\eqref{PNz} into the wave equation
248: \begin{equation}
249: \label{wave}
250: \frac{\partial \mathcal{Q}}{\partial t}+ \frac{\partial \mathcal{Q}}{\partial y}=0,
251: \end{equation}
252: whose solution is an arbitrary function of $t-y$ or, equivalently,
253: $(1-z)e^{-t}$.  If the system initially is a microtubule of zero length,
254: $\Pi_N(t=0)=\delta_{N,0}$, the initial generating function $\Pi(z,t=0)=1$, so
255: that $\mathcal{Q}=e^{-\lambda z}= e^{\lambda(1-z)} e^{-\lambda}$.  Thus for
256: $t>0$, $\mathcal{Q}= e^{\lambda(1-z)e^{-t}} e^{-\lambda}$, from which
257: \begin{equation}
258: \label{Poisson}
259: \Pi(z,t) = e^{-\lambda(1-z)(1-e^{-t})}.
260: \end{equation}
261: Expanding this expression in a power series in $z$, the probability for the
262: system to contain $N$ GTP$^+$ monomers is the time-dependent Poisson distribution
263: \begin{equation}
264: \label{Poisson-t}
265: \Pi_N(t) =  \frac{[\lambda(1-e^{-t})]^N}{N!}\,e^{-\lambda(1-e^{-t})}.
266: \end{equation}
267: From this result, the mean number of GTP$^+$ monomers and its variance are
268: \begin{equation}
269: \label{N-av-var}
270: \langle N\rangle = \langle N^2\rangle - \langle N\rangle^2 = \lambda(1-e^{-t}).
271: \end{equation}
272: 
273: \subsection{Tubule Length Distributions}
274: 
275: The length distribution $P(L,t)$ of the microtubule evolves according to the
276: master equation
277: \begin{equation}
278: \label{F}
279: \frac{d P(L,t)}{d t} = \lambda\left[P(L-1,t)-P(L,t)\right]
280: \end{equation}
281: For the initial condition $P(L,0)=\delta_{L,0}$, the solution is again the
282: Poisson distribution
283: \begin{equation}
284: \label{F-sol}
285: P(L,t) = \frac{(\lambda t)^L}{L!}\,e^{-\lambda t}
286: \end{equation}
287: {}from which the average and the variance are
288: \begin{equation}
289: \label{L-av-var}
290: \langle L\rangle=\lambda t,\quad \langle L^2\rangle-\langle L\rangle^2=\lambda t\,.
291: \end{equation}
292: Thus the growth rate of the microtubule and the diffusion coefficient of the
293: tip are
294: \begin{equation}
295: \label{VD-simple}
296: V=\lambda, \quad D=\lambda/2
297: \end{equation}
298: 
299: A more comprehensive description is provided by the joint distribution
300: $P(L,N,t)$ that a microtubule has length $L$ and contains $N$ GTP$^+$
301: monomers at time $t$.  This distribution evolves as
302: \begin{eqnarray}
303: \label{PLN}
304: \frac{d P(L,N)}{d t} &=& \lambda P(L-1,N-1)-(N+\lambda)P(L,N) \nonumber \\
305: &+&(N+1)P(L,N+1).
306: \end{eqnarray}
307: This joint distribution does {\em not} factorize, that is, $P(L,N,t)\ne
308: P(L,t)\,\Pi_N(t)$, because $\langle LN\rangle\ne \langle L\rangle \langle
309: N\rangle$.  To demonstrate this inequality, we compute $\langle LN\rangle$ by
310: multiplying Eq.~\eqref{PLN} by $LN$ and summing over all $L\geq N\geq 0$ to
311: give
312: \begin{equation}
313: \label{LN}
314: \frac{d}{dt}\,\langle LN\rangle=\lambda(\langle L\rangle+\langle N\rangle+1)
315: -\langle LN\rangle.
316: \end{equation}
317: Using Eqs.~\eqref{N-av-var} and \eqref{L-av-var} for $\langle N\rangle$ and
318: $\langle L\rangle$ and integrating we obtain
319: \begin{eqnarray}
320: \label{LN-av}
321: \langle LN\rangle &=& \lambda^2t\,(1-e^{-t})+\lambda(1-e^{-t})\nonumber\\
322: &=&\langle L\rangle \langle N\rangle +\langle N \rangle.
323: \end{eqnarray}
324: Using Eq.~\eqref{L-av-var}, we have $\langle
325: LN\rangle= \langle L\rangle \langle N\rangle(1+ \frac{1}{\lambda t})$, so
326: that the joint distribution is factorizable asymptotically.  For
327: completeness, we give the full solution for $P(L,N,t)$ in Appendix
328: \ref{app-P}.
329: 
330: \subsection{Cap Length Distribution}
331: \label{sec:cld}
332: 
333: Because of the conversion process GTP$^+$ $\to$ GDP$^-$, the tip of the
334: microtubule is comprised predominantly of GTP$^+$, while the tail exclusively
335: consists of GDP$^-$.  The region from the tip until the first GDP$^-$ monomer
336: is known as the {\em cap} (Fig.~\ref{cap-fig}) and it plays a fundamental
337: role in microtubule function.  We now use the master equation approach to
338: determine the cap length distribution.
339: 
340: \begin{figure}%[ht]
341: \includegraphics*[width=0.4\textwidth]{cap}
342: \caption{Representative configuration of a microtubule, with a GTP$^+$ cap of
343:   length 4, then three GTP$^+$ islands of lengths 1, 3, and 2, and three
344:   GDP$^-$ islands of lengths 3, 2, and a ``tail'' of length 5.  The rest of
345:   the microtubule consists of GDP$^-$.}
346: \label{cap-fig}
347: \end{figure}
348: 
349: \begin{figure}[ht]
350: \includegraphics*[width=0.8\columnwidth]{Cap_size}
351: \caption{Cap length distribution obtained from simulations at $\mu=0$,
352:   $\lambda=\ 100$, and $p=1$ compared to the theoretical prediction of
353:   Eq.~\eqref{nk-sol}.}
354: \label{capsize}
355: \end{figure}
356: 
357: Consider a cap of length $k$.  Its length increases by 1 due to the
358: attachment of a GTP$^+$ at rate $\lambda$.  The conversion of any GTP$^+$
359: into a GDP$^-$ at rate 1 reduces the cap length from $k$ to an arbitrary
360: value $s<k$.  These processes lead to the following master equation for the
361: probability $n_k$ that the cap length equals $k$:
362: \begin{equation}
363: \label{nk}
364: \dot n_k = \lambda(n_{k-1}-n_k)-kn_k+\sum_{s\geq k+1}n_s.
365: \end{equation}
366: Equation~\eqref{nk} is also valid for $k=0$ if we set $n_{-1}\equiv 0$. Note
367: that $n_0\equiv {\rm Prob}\{-\rangle\}$ is the probability for a cap of
368: length zero. We now solve for the stationary distribution by summing the
369: first $k-1$ of Eqs.~\eqref{nk} with $\dot n_k$ set to zero to obtain
370: \begin{equation}
371:  n_{k-1} = \frac{k}{\lambda} \sum_{s\geq k}n_s.
372: \end{equation}
373: The cumulative distribution, $ N_k=\sum_{s\geq k}n_s$, thus satisfies the
374: recursion
375: \begin{equation}
376: \label{Nk}
377:  N_k = \frac{\lambda}{k+\lambda} N_{k-1}.
378: \end{equation}
379: Using the normalization $N_0=1$ and iterating, we obtain the solution in
380: terms of the Gamma function \cite{AS}:
381: \begin{equation}
382: \label{Nk-soln}
383:  N_k = \frac{\lambda^k\, \Gamma(1+\lambda)}{\Gamma(k+1+\lambda)}.
384: \end{equation}
385: Hence the cap length distribution is
386: \begin{equation}
387: \label{nk-sol}
388:  n_k = \frac{\Gamma(1\!+\!\lambda)}{\Gamma(k\!+\!2\!+\!\lambda)}\, (k+1)\lambda^k
389: %\sim  \Gamma(1\!+\!\lambda)\,\frac{k^{-\lambda}\lambda^k}{k!},
390: \end{equation}
391: and the first few terms are
392: \begin{subequations}
393: \begin{align}
394: %\label{n0}
395:  n_0 &= \frac{1}{1+\lambda}\nonumber \\
396: %\label{n1}
397:  n_1 &= \frac{2\lambda}{(1+\lambda)(2+\lambda)\nonumber}\\
398: %\label{n2}
399:  n_2 &= \frac{3\lambda^2}{(1+\lambda)(2+\lambda)(3+\lambda)}.\nonumber
400: \end{align}
401: \end{subequations}
402: Results of direct simulation of the kinetic model are compared to the
403: predicted cap length distribution (Eq.~\eqref{nk-sol}) in Fig.~\ref{capsize}.
404: Because of the finite length of the simulated microtubule, there is a largest
405: cap length that is accessible numerically.  Aside from this limitation, the
406: simulations results are in agreement with theoretical predictions.
407: 
408: It is instructive to determine the dependence of the average cap length
409: $\langle k\rangle =\sum_{k\geq 0}kn_k$ on $\lambda$.  Using
410: $n_k=N_k-N_{k+1}$, we rearrange $\langle k\rangle$ into
411: \begin{equation}
412: \label{cap}
413: \langle k\rangle =\sum_{k\geq 1}N_k
414: \end{equation}
415: Using \eqref{Nk-soln}, the above sum may be written in terms of the confluent
416: hypergeometric series \cite{AS}:
417: % \begin{equation}
418: % \label{hyper}
419: % \sum_{k\geq 0}N_k=F(1;1+\lambda; \lambda)
420: % \end{equation}
421: \begin{equation}
422: \label{cap-av}
423: \langle k\rangle =-1+F(1;1+\lambda; \lambda).
424: \end{equation}
425: We now determine the asymptotic behavior of $\langle k\rangle$ by using the
426: integral representation
427: \begin{equation*}
428:  F(a;b; z)=\frac{\Gamma(b)}{\Gamma(b-a)\,\Gamma(a)}
429: \int_0^1 dt\,e^{zt}\,t^{a-1}(1-t)^{b-a-1}
430: \end{equation*}
431: to recast the average cap length \eqref{cap-av} as
432: \begin{equation}
433: \label{cap-average}
434: \langle k\rangle =-1+ \lambda \left(\frac{e}{\lambda}\right)^{\lambda} 
435: \gamma(\lambda,\lambda),
436: \end{equation}
437: where $\gamma(a,x)=\int_0^x dt\,t^{a-1} e^{-t}$ is the (lower) incomplete
438: gamma function.
439: 
440: In the realistic limit of $\lambda\gg 1$, we use the large $\lambda$
441: asymptotics
442: \begin{equation*}
443: \gamma(\lambda,\lambda)\to \frac{1}{2}\,\Gamma(\lambda), \quad 
444: \Gamma(\lambda)\sim \sqrt{\frac{2\pi}{\lambda}}\left(\frac{\lambda}{e}\right)^{\lambda},
445: \end{equation*}
446: to give
447: \begin{equation}
448: \label{kav-large}
449: \langle k\rangle\to \sqrt{\pi\lambda/2}\quad {\rm as}\quad \lambda\to\infty
450: \end{equation}
451: Thus even though the number of GTP$^+$ monomers equals $\lambda$, only
452: $\sqrt{\lambda}$ of them comprise the microtubule cap, as qualitatively
453: illustrated in Fig.~\ref{cartoon}.  Note that the average cap length is
454: proportional to the square-root of the velocity; essentially the same result
455: was obtained from the coarse-grained theory of Flyvbjerg et al.\ \cite{F}.
456: 
457: 
458: \subsection{Island Size Distributions}
459: 
460: At a finer level of resolution, we determine the distribution of island sizes
461: at the tip of a microtubule (Fig.~\ref{cap-fig}).  A simple characteristic of this
462: population is the average number $I$ of GTP$^+$ islands.  If all GTP$^+$ islands were
463: approximately as long as the cap, we would have $I\sim \langle
464: N\rangle/\langle k\rangle \sim \sqrt{\lambda}$.  As we shall see, however,
465: $I$ scales linearly with $\lambda$ because most islands are short.  A similar
466: dichotomy arises for negative islands.
467: 
468: To write the master equation for the average number of islands, note that the
469: conversion GTP$^+$ $\to$ GDP$^-$ eliminates islands of size 1.  Additionally, an
470: island of size $k\geq 3$ splits into two daughter islands, and hence the
471: number of islands increases by one, if conversion occurs at any one of the
472: $k-2$ in the interior of an island as illustrated below:
473: \begin{equation*}
474: -+\underbrace{+\cdots +}_{k-2} +-\,.
475: \end{equation*}
476: Conversely, if the cap has length 0, attachment creates a new cap of length 1
477: at rate $\lambda$.  The net result of these processes is encoded in the rate
478: equation
479: \begin{equation}
480: \label{island-eq}
481: \frac{d I}{dt} = \sum_{k\geq 1}(k-2)I_k +\lambda n_0,
482: \end{equation}
483: with $I_k$ the average number of GTP$^+$ islands of size $k$.
484: 
485: We now use the sum rules $I = \sum_{k\geq 1}I_k$ and $\langle N\rangle =
486: \sum_{k\geq 1}k I_k$ to recast \eqref{island-eq} as
487: \begin{equation}
488: \label{island-main}
489: \frac{d I}{dt} = \langle N\rangle - 2I +\lambda n_0
490: \end{equation}
491: from which the steady-state average number of islands is
492: \begin{equation}
493: \label{island-sol}
494: I = \frac{\langle N\rangle +\lambda n_0}{2}
495: = \frac{\lambda}{2}\,\frac{2+\lambda}{1+\lambda}.
496: \end{equation}
497: For large $\lambda$, the number of islands approaches $\lambda/2$, while the
498: number of GTP$^+$ monomers equals $\lambda$.  Thus the typical island size is 2.
499: Nevertheless, as we now show, the GTP$^+$ and GDP$^-$ island distributions actually have
500: power-law tails, with different exponents for each species.
501: 
502: The GTP$^+$ island size distribution evolves according to the master equation
503: \begin{equation}
504: \label{Ik-eq}
505: \dot I_k = -kI_k+2\sum_{s\geq k+1}I_s +\lambda(n_{k-1}-n_k)
506: \end{equation}
507: This equation is similar in spirit to Eq.~\eqref{nk} for the cap length
508: distribution.  As a useful self-consistency check, the sum of
509: Eqs.~\eqref{Ik-eq} gives \eqref{island-main}, while multiplying \eqref{Ik-eq}
510: by $k$ and summing over all $k\geq 1$ gives Eq.~\eqref{N-eq}.
511: 
512: \begin{figure}[ht]
513: \includegraphics*[width=0.9\columnwidth]{islands-p1-l100}
514: \caption{Simulation results at $\mu=0$, $\lambda=100$, and $p=1$ for the size
515:   distribution of positive islands, $I_{k}/\lambda$.  The solid line is the
516:   theoretical prediction of Eq.~\eqref{Ik-sol}.}
517: \label{islandsize}                                                                 
518: \end{figure}
519: 
520: The stationary distribution satisfies
521: \begin{equation}
522: \label{Ik}
523: kI_k=2\sum_{s\geq k+1}I_s +\lambda(n_{k-1}-n_k).
524: \end{equation}
525: Using $\sum_{s\geq 2}I_s=I-I_1$, we transform \eqref{Ik} at $k=1$ to
526: \begin{equation*}
527: \label{I1-eq}
528: 3I_1=2I+\lambda(n_0-n_1)
529: \end{equation*}
530: Similarly, using $\sum_{s\geq 3}I_s=I-I_1-I_2$ we transform \eqref{Ik} at
531: $k=2$ to
532: \begin{equation*}
533: \label{I2-eq}
534: 4I_2=2(I-I_1)+\lambda(n_1-n_2)
535: \end{equation*}
536: Thus using \eqref{nk-sol} and
537: \eqref{Ik} we obtain
538: \begin{subequations}
539: \begin{align}
540: \label{I1}
541:  I_1 &= \frac{\lambda}{3}+\frac{4\lambda}{3(1+\lambda)(2+\lambda)}\\
542: \label{I2}
543:  I_2 &= \frac{\lambda}{12}
544:  +\frac{25\lambda^2-6\lambda}{12(1+\lambda)(2+\lambda)(3+\lambda)}.
545: \end{align}
546: \end{subequations}
547: The same procedure gives $I_k$ for larger $k$.  
548: 
549: Since the $I_k$ represent the {\em average} number of islands of size $k$,
550: they become meaningful only for $\lambda\to\infty$ where an appreciable
551: number of such islands exist.  In this limit, we write $I_k$ more compactly
552: by first rearranging \eqref{Ik} into the equivalent form
553: \begin{equation}
554: \label{Ikk}
555: (k-1)I_{k-1}-(k+2)I_k=\lambda(n_{k-2}-2n_{k-1}+n_k).
556: \end{equation}
557: We  then use \eqref{nk-sol} and the asymptotic properties of the Gamma
558: function to find that the right-hand side of  Eq.~\eqref{Ikk} is
559: \begin{equation*}
560: \lambda(n_{k-2}-2n_{k-1}+n_k)=-\frac{3k+1}{\lambda}+O\left(\frac{1}{\lambda^2}\right),
561: \end{equation*}
562: and is therefore negligible in the large-$\lambda$ limit.  Thus \eqref{Ikk}
563: reduces to $(k-1)I_{k-1}=(k+2)I_k$, with solution $I_k=A/[k(k+1)(k+2)]$.  We
564: find the amplitude $A$ by matching with the exact result, Eq.~\eqref{I1}, to
565: give $I_1=\lambda/3$ for large $\lambda$.  The final result is
566: \begin{equation}
567: \label{Ik-sol}
568: I_k=\frac{2\lambda}{k(k+1)(k+2)}.
569: \end{equation}
570: In the large $\lambda$ limit, $I=\lambda/2$, and the above result can be re-written as
571: \begin{equation}
572: \label{Ik-I}
573: \frac{I_k}{I}=\frac{4}{k(k+1)(k+2)}.
574: \end{equation}
575: Remarkably, the size distribution of the positive islands is {\em
576:   identical\/} to the degree distribution of a growing network with strictly
577: linear preferential attachment \cite{network1,network2,network3}.
578: 
579: The results for the island size distribution in the large $\lambda$ limit are
580: compared to simulation results in Fig.~\ref{islandsize}.  These asymptotic
581: results are expected to apply to island sizes $k$ much smaller than the size
582: of the cap which scales as $\sqrt \lambda$.  The distributions obtained from
583: the numerical simulations should then obey the theoretical form but with a
584: finite-size cutoff.  The results in Fig.~\ref{islandsize} are consistent with
585: this picture but, interestingly, the numerical distribution rises above the
586: theoretical curve before falling sharply below it.  This anomaly occurs in
587: many heterogeneous growing network models, and it can be fully characterized
588: in terms of finite-size effects \cite{finite}.
589: 
590: \subsection{Continuum Limit, $\lambda\to\infty$}
591: \label{continuum}
592: 
593: When $\lambda\to\infty$, both the length of the cap and the length of the
594: region that contains GTP$^+$ become large.  In this limit, the results from
595: the discrete master equation can be expressed much more elegantly and
596: completely by a continuum approach.  The fundamental feature is that the
597: conversion process GTP$^+$ $\to$ GDP$^-$ occurs independently for each
598: monomer.  Since the residence time of each monomer increases linearly with
599: distance from the tip, the probability that a GTP$^+$ does not convert decays
600: exponentially with distance from the tip.  This fact alone is sufficient to
601: derive all the island distributions.
602: 
603: Consider first the length $\ell$ of the populated region
604: (Fig.~\ref{cap-fig}).  For a GTP$^+$ that is a distance $x$ from the tip, its
605: residence time is $\tau=x/\lambda$ in the limit of large $\lambda$.  Thus the
606: probability that this GTP$^+$ does not convert is $e^{-\tau}=e^{-x/\lambda}$.
607: We thus estimate $\ell$ from the extremal criterion \cite{extremal}
608: \begin{equation}
609: \label{ell}
610: 1=\sum_{x\geq \ell} e^{-x/\lambda}=(1-e^{-1/\lambda})^{-1}e^{-\ell/\lambda},
611: \end{equation}
612: that merely states that there is of the order of a single GTP$^+$ further
613: than a distance $\ell$ from the tip.  Since $(1-e^{-1/\lambda})^{-1}\to
614: \lambda$ when $\lambda$ is large, the length of the active region scales as
615: \begin{equation}
616: \label{ell-sol}
617: \ell=\lambda\ln\lambda
618: \end{equation}
619: 
620: The probability that the cap has length $k$ is given by
621: \begin{equation*}
622: (1-e^{-(k+1)/\lambda})\prod_{j=1}^k e^{-j/\lambda}.
623: \end{equation*}
624: The product ensures that all monomers between the tip and a distance $k$ from
625: the tip are GTP$^+$, while the prefactor gives the probability that a monomer
626: is a distance $k+1$ from the tip is a GDP$^-$.  Expanding the prefactor for
627: large $\lambda$ and rewriting the product as the sum in the exponent, we
628: obtain
629: \begin{equation}
630: \label{nk-cont}
631: n_k\sim \frac{k+1}{\lambda}\,\, e^{-k(k+1)/2\lambda},
632: \end{equation}
633: a result that also can be obtained by taking the large-$\lambda$ limit of
634: the exact result for $n_k$ given in Eq.~\eqref{nk-sol}.
635: 
636: Similarly, the probability to find a positive island of length $k$ that
637: occupies sites $x+1,x+2,\ldots,x+k$ is
638: \begin{equation}
639: \label{isl-x}
640: (1-e^{-x/\lambda})(1-e^{-(x+k+1)/\lambda})\prod_{j=1}^k e^{-(x+j)/\lambda}.
641: \end{equation}
642: The two prefactors ensure that sites $x$ and $x+k+1$ consist of GDP$^-$,
643: while the product ensures that all sites between $x+1$ and $x+k$ are GTP$^+$.
644: 
645: Most islands are far from the tip and they are relatively short, $k\ll x$, so
646: that \eqref{isl-x} simplifies to
647: \begin{equation}
648: \label{isl-x-simple}
649: (1-e^{-x/\lambda})^2 e^{-kx/\lambda}\, e^{-k^2/2\lambda}.
650: \end{equation}
651: The total number of islands of length $k$ is obtained by summing the island
652: density \eqref{isl-x-simple} over all $x$.  Since $\lambda\gg 1$, we replace
653: the summation by integration and obtain
654: \begin{eqnarray}
655: \label{isl-k}
656: I_k&=&\int_0^\infty \!\!dx\,(1-e^{-x/\lambda})^2\, e^{-kx/\lambda}\,
657: e^{-k^2/2\lambda}\nonumber \\
658: &=&\frac{2\lambda}{k(k\!+\!1)(k\!+\!2)}\, e^{-k^2/2\lambda}.
659: \end{eqnarray}
660: The power law tail agrees with Eq.~\eqref{Ik-sol}, whose derivation
661: explicitly invoked the $\lambda\to\infty$ limit.
662: 
663: We can also obtain the density of negative islands in this continuum
664: description, a result that seems impossible to derive by a microscopic master
665: equation description.  In parallel with \eqref{isl-x-simple}, the density of
666: negative islands of length $k\ll x$ with one end at $x$ is given by
667: \begin{equation}
668: \label{isl-negative}
669: e^{-2x/\lambda} (1-e^{-x/\lambda})^k, 
670: \end{equation}
671: and the total number of negative islands of length $k$ is
672: \begin{equation}
673: \label{isl-k-negative}
674: J_k=\int_0^\infty dx\,e^{-2x/\lambda} (1-e^{-x/\lambda})^k =\frac{\lambda}{(k+1)(k+2)}.
675: \end{equation}
676: Again, we find a power-law tail for the GDP$^-$ island size distribution, but
677: with exponent 2.  The total number of GDP$^-$ monomers within the populated
678: zone is then $\sum_{k\geq 1}kJ_k$.  While this sum formally diverges, we use
679: the upper size cutoff, $k_*\sim \lambda$ to obtain $\sum_{k\geq 1}kJ_k\simeq
680: \lambda\ln\lambda$.  Since the length of the populated zone $\ell\sim
681: \lambda\ln \lambda$,this zone therefore predominantly consists of GDP$^-$
682: islands.
683: 
684: In analogy with the cap, consider now the ``tail''---the last island of
685: GDP$^-$ within the populated zone (see Fig.~\ref{cap-fig}).  The probability
686: $m_k$ that it has length $k$ is
687: \begin{equation}
688: \label{last}
689: m_k = e^{-\ell/\lambda} (1-e^{-\ell/\lambda})^k.
690: \end{equation}
691: Using \eqref{ell-sol} we simplify the above expression to
692: \begin{equation*}
693: \label{last-simple}
694: m_k = \lambda^{-1} (1-\lambda^{-1})^k = \lambda^{-1} e^{-k/\lambda}.
695: \end{equation*}
696: Hence the average length of the tail is
697: \begin{equation}
698: \label{last-length}
699: \langle k\rangle = \sum_{k\geq 1} km_k = \lambda,
700: \end{equation}
701: which is much longer (on average) than the cap.
702: 
703: \section{CONSTRAINED GROWTH}
704: \label{constrained}
705: 
706: When $p\ne 1$, the rate of attachment depends on the state of the tip of the
707: microtubule---attachment to a GTP$^+$ occurs with rate $\lambda$ while
708: attachment to a GDP$^-$ occurs with rate $p\lambda$.  While this state
709: dependence makes the master equation description for the properties of the
710: tubule more complicated, qualitative features about the structure of the
711: populated zone are the same as those in the case $p=1$.  In this section, we
712: outline some of the basic features of the populated zone when $p\ne 1$, but
713: we still keep $\mu=0$.
714: 
715: \subsection{Distribution of GTP$^+$}
716: 
717: The average number of GTP$^+$ monomers now evolves according to the rate equation
718: \begin{equation}
719: \label{N-avt}
720:  \frac{d}{dt}\, \langle N\rangle = -\langle N\rangle+p\lambda n_0+\lambda (1-n_0),
721: \end{equation}
722: which should be compared to the rate equation Eq.~\eqref{N-eq} for the case
723: $p=1$.  The loss term on the right-hand side describes the conversion
724: GTP$^+\to $GDP$^-$, while the remaining terms represent gain due to
725: attachment to a GTP$^+$ with rate $\lambda$ and to a GDP$^-$ with rate
726: $p\lambda$.  Here $n_0$ is the probability for a cap of length zero, that is,
727: the last site is a GDP$^-$.  The stationary solution to \eqref{N-avt} is
728: \begin{equation}
729: \label{N-av-inf}
730:  \langle N\rangle = p\lambda n_0+\lambda (1-n_0),
731: \end{equation}
732: so we need to determine $n_0$.  By extending Eq.~\eqref{nk} to the case $p\ne
733: 1$, we then find that $n_0$ is governed by the rate equation
734: \begin{equation}
735: \label{n0-rate}
736: \dot n_0 = -p\lambda n_0+(1-n_0).
737: \end{equation}
738: Thus asymptotically $n_0 = \frac{1}{1+p\lambda}$ and substituting into
739: \eqref{N-av-inf}, the average number of GTP$^+$ monomers is
740: \begin{equation}
741: \label{N-av-sol}
742: \langle N\rangle=p\lambda\,\frac{1+\lambda}{1+p\lambda}
743: \end{equation}
744: 
745: More generally, we can determine the distribution of the number of GTP$^+$
746: monomers; the details of this calculation are presented in
747: Appendix~\ref{app}.
748: 
749: 
750: \subsection{Growth Rate and Diffusion Coefficient}
751: \label{XY-sub}
752: 
753: The growth rate of a microtubule equals $p\lambda$ when the cap length is
754: zero and to $\lambda$ otherwise.  Therefore
755: \begin{equation}
756: \label{Vn}
757: V(p,\lambda) = p\lambda n_0+\lambda(1-n_0) = p\lambda\,\frac{1+\lambda}{1+p\lambda}
758: \end{equation}
759: For the diffusion coefficient of the tip of a microtubule, we need its
760: mean-square position.  As in the case $p=1$, it is convenient to determine
761: the probability distribution for the tip position.  Thus we introduce
762: $X(L,t)$ and $Y(L,t)$, the probabilities that the microtubule length equals
763: $L$ and the last monomer is a GTP$^+$ or a GDP$^-$, respectively.  These
764: probabilities satisfy
765: \begin{subequations}
766: \begin{align}
767: \label{X}
768:    \frac{dX(L)}{d t}&= \lambda X(L\!-\!1)+p\lambda Y(L\!-\!1)-(1\!+\!\lambda)X(L)\\
769: \label{Y} 
770:    \frac{d Y(L)}{d t}&= X(L) - p\lambda Y(L),
771: \end{align} 
772: \end{subequations}
773: Summing these equations, the length distribution $P(L)=X(L)+Y(L)$ satisfies
774: \begin{equation}
775: \label{Fp}
776:    \frac{dP(L)}{d t} =\lambda X(L-1)+p\lambda Y(L-1)-\lambda X(L)-p\lambda Y(L).
777: \end{equation}
778: 
779: The state of the last monomer does not depend on the microtubule length $L$
780: for large $L$.  Thus asymptotically 
781: \begin{equation}
782: \label{XY}
783: X(L)=(1-n_0)P(L), \quad Y(L)=n_0 P(L).
784: \end{equation}
785: Substituting \eqref{XY} into \eqref{Fp} we obtain a master equation for the
786: tubule length distribution of the same form as Eq.~\eqref{F}, but with
787: prefactor $V$ given by \eqref{Vn} instead of $\lambda$.  As a result of this
788: correspondence, we infer that the diffusion coefficient is one-half of the
789: growth rate,
790: \begin{equation}
791: \label{D-sol}
792: D(p,\lambda)=\frac{1}{2}\,p\lambda\,\frac{1+\lambda}{1+p\lambda}.
793: \end{equation}
794: For large $\lambda$ both the growth rate of the tip of the microtubule and
795: its diffusion coefficient approach the corresponding expressions in
796: Eq.~\eqref{VD-simple} for the case $p=1$.
797: 
798: \subsection{Cap Length Distribution}
799: 
800: The master equations for the cap length distribution are the same as in the
801: $p=1$ case when $k\geq 2$.  The master equations for $k=0$ and $k=1$ are
802: slightly changed to account for the different rates at which attachment
803: occurs at a GDP$^-$ monomer:
804: \begin{eqnarray*}
805:  p\lambda n_0&=&N_{1}=1-n_0\\
806:  (1+\lambda)n_1-p\lambda n_0&=&N_2=1-n_0-n_1
807: \end{eqnarray*}
808: Solving iteratively we recover $n_0= \frac{1}{1+p\lambda}$ and also obtain
809: \begin{subequations}
810: \begin{align}
811: \label{n1-p}
812:  n_1 &= \frac{2p\lambda}{(1+p\lambda)(2+\lambda)}\\
813: \label{n2-p}
814:  n_2 &= \frac{3p\lambda^2}{(1+p\lambda)(2+\lambda)(3+\lambda)},
815:  \end{align}
816: \end{subequations} {\it etc.}  The general solution for the $n_k$ is found by
817: the same method as in Sec.~\ref{sec:cld} to be
818: \begin{equation}
819: \label{nkp}
820:  n_k = (k+1)\lambda^k\,\frac{p}{1+p\lambda}\, 
821:  \frac{\Gamma(2+\lambda)}{\Gamma(k+2+\lambda)},
822: \end{equation}
823: which are valid for $k\geq 1$.  With this length distribution, the average
824: cap length is then
825: \begin{equation}
826: \label{cap-p}
827: \langle k\rangle = \frac{p}{1+p\lambda}\sum_{k\geq 1} k(k+1)\lambda^k\,
828:  \frac{\Gamma(2+\lambda)}{\Gamma(k+2+\lambda)},
829: \end{equation}
830: and the sum can again be expressed in terms of hypergeometric series as in
831: Eq.~\eqref{cap-av}.  Rather than following this path, we focus on the most
832: interesting limit of large $\lambda$.  Then the cap length distribution
833: \eqref{nkp} approaches to the previous solution \eqref{nk-sol} for the case
834: $p=1$ and the mean length reduces to \eqref{kav-large}, independent of $p$.
835: 
836: \subsection{Island Size Distribution}
837: 
838: For the distribution of island sizes, the master equation remains the same as
839: in the $p=1$ case when $k\geq 2$.  However, when $k=1$, the master equation
840: becomes
841: \begin{equation}
842: \label{I1-eq-a}
843: I_1=2\sum_{s\geq 2}I_s +p\lambda n_0 - \lambda n_1
844: \end{equation}
845: in the stationary state. Then the average number of islands and the average
846: number of islands of size 1 are found from
847: \begin{eqnarray*}
848: 2I &=& \langle N\rangle +p\lambda n_0\\
849: 3I_1 &=& 2I +p\lambda n_0 - \lambda n_1
850: \end{eqnarray*}
851: Using $n_0 = \frac{1}{1+p\lambda}$ and Eqs.~\eqref{N-av-sol} and \eqref{n1-p} we obtain
852: \begin{subequations}
853: \begin{align}
854: \label{I-p}
855:  I &= \frac{p\lambda}{2}\,\frac{2+\lambda}{1+p\lambda}\\
856: \label{I1-p}
857:  I_1 &= \frac{p\lambda}{1+p\lambda}
858:  \left[\frac{\lambda}{3}+\frac{2+\lambda/3}{2+\lambda}\right]
859:  \end{align}
860: \end{subequations}
861: 
862: Again, in the limit of large $\lambda$, the average island size distribution
863: reduces to our previously-quoted results in \eqref{Ik-sol} or equivalently
864: \eqref{Ik-I}.  The leading behavior in the $\lambda\to\infty$ limit is again
865: independent of $p$.
866: 
867: \section{Instantaneous Detachment}
868: \label{instant}
869: 
870: For $\mu>0$, a microtubule can recede if its tip consists of GDP$^-$.  The
871: competition between this recession and growth by the attachment of GTP$^+$
872: leads to a rich dynamics in which the microtubule length can fluctuate wildly
873: under steady conditions.  In this section, we focus on the limiting case of
874: infinite detachment rate, $\mu=\infty$.  In this limit, any GDP$^-$
875: monomer(s) at the tip of a microtubule are immediately removed.  Thus the the
876: tip is always a GTP$^+$; this means that the parameter $p$ become immaterial.
877: Finally, for $\mu=\infty$, we also require the growth rate $\lambda\to\infty$
878: to have a microtubule with an appreciable length.  This is the limit
879: considered below.
880: 
881: As soon as the last monomer of the tubule changes from a GTP$^+$ to a
882: GDP$^-$, a string of $k$ contiguous GDP$^-$ monomers exist at the tip and
883: they detach immediately.  We term such an event an {\em avalanche of size
884:   $k$}.  We now investigate the statistical properties of these avalanches.
885: 
886: \subsection{Catastrophes}
887: The switches from a growing to a shrinking state of a microtubule are called
888: {\em catastrophes} \cite{DL}.  If a newly-attached GTP$^+$ at the tip
889: converts to a GDP$^-$ and the rest of the microtubule consists only of
890: GDP$^-$ at that moment, the microtubule instantaneously shrinks to zero
891: length, a phenomenon that can be termed a {\em global catastrophe}.  We now
892: determine the probability for such a catastrophe to occur.  Formally, the
893: probability of a global catastrophe is
894: \begin{equation}
895: \label{sweep}
896: \mathcal{C}(\lambda)=\frac{1}{1+\lambda}\prod_{n=1}^\infty (1-e^{-n/\lambda}).
897: \end{equation}
898: The factor $(1+\lambda)^{-1}$ gives the probability that the monomer at the
899: tip converts to a GDP$^-$ before the next attachment event, while the product
900: gives the probability that the rest of the microtubule consists of GDP$^-$.
901: In principle, the upper limit in the product is set by the microtubule
902: length.  However, for $n>\lambda$, each factor in the product is close to 1
903: and the error made in extending the product to infinity is small.
904: The expression within the product is obtained under the assumption that the
905: tubule grows steadily between these complete catastrophes and the smaller,
906: local catastrophes, are therefore ignored in this calculation.
907: 
908: The leading asymptotic behavior of the infinite product in \eqref{sweep} is
909: found by expressing it in terms of the Dedekind $\eta$ function
910: \cite{apostol}
911: \begin{equation}
912: \label{eta}
913: \eta(z)=e^{i\pi z/12}\prod_{n=1}^\infty (1-e^{2\pi i nz}),
914: \end{equation}
915: and using a remarkable identity satisfied by this function,
916: \begin{equation*}
917: \label{ident}
918: \eta(-1/z)=\sqrt{-i z}\,\eta(z).
919: \end{equation*}
920: For our purposes, we define $a=-i\pi z$ to rewrite this identity as
921: \begin{equation*}
922: \label{identity}
923: \prod_{n=1}^\infty (1-e^{-2an})=\sqrt{\frac{\pi}{a}}\,e^{(a-b)/12}
924: \prod_{n=1}^\infty (1-e^{-2bn})
925: \end{equation*}
926: where $b=\pi^2/a$.  Specializing to the case $a=(2\lambda)^{-1}$ yields
927: \begin{eqnarray}
928: \label{S}
929: \mathcal{C}(\lambda)&=&\frac{\sqrt{2\pi\lambda}}{1+\lambda} \,\,e^{-\pi^2\lambda/6}\, e^{1/24\lambda}\,
930: \prod_{n\geq 1} (1-e^{-4\pi^2\lambda  n})\nonumber \\
931: &\sim& \sqrt{\frac{2\pi}{\lambda}}\,\,e^{-\pi^2\lambda/6}.
932: \end{eqnarray}
933: Since the time between catastrophes scales as the inverse of the occurrence
934: probability, this inter-event time becomes very long for large $\lambda$.
935: 
936: \subsection{Avalanche Size Distribution}
937: 
938: In the instantaneous detachment limit, $\mu=\infty$, the catastrophes are
939: avalanches whose size is determined by the number of GDP$^-$s between the tip
940: and the first GTP$^+$ island.  A global catastrophe is an avalanche of size
941: equal to the length of the tubule, whose occurrence probability was
942: calculated in the preceding section.  Similar arguments can be used to
943: calculate the size distribution of the smaller avalanches.
944: 
945: Since the cap is large when $\lambda$ is large, an avalanche of size 1 arises
946: only through the reaction scheme
947: \begin{equation*}
948:   |\cdots ++\rangle \Longrightarrow |\cdots +-\rangle
949:   \Longrightarrow |\cdots +\rangle,
950: \end{equation*}
951: where the first step occurs at rate 1 and the second step is instantaneous.
952: Since attachment proceeds with rate $\lambda$, the probability that
953: conversion occurs before attachment is $A_1=(1+\lambda)^{-1}$; this expression
954: gives the relative frequency of avalanches of sizes $\geq 1$.
955: Analogously, an avalanche of size two is formed by the events
956: \begin{equation*}
957: |\cdots +++\rangle \Longrightarrow |\cdots+-+\rangle
958:  \Longrightarrow |\cdots+--\rangle \Longrightarrow |\cdots+\rangle,
959: \end{equation*}
960: and the probability that the first two steps occur before an attachment event
961: is $A_2\sim \lambda^{-2}$ to lowest order.  At this level of approximation, the
962: relative frequency of avalanches of size equal to 1 is $A_1-A_2\sim
963: \lambda^{-1}$.  Since we are interested in the regime $\lambda\gg 1$, we
964: shall only consider the leading term in the avalanche size distribution.
965: 
966: 
967: Generally an avalanche of size $k$ is formed if the system starts in the
968: configuration
969: \begin{equation*}
970: |+\underbrace{+\cdots +}_{k-1} +\rangle,
971: \end{equation*}
972: then $k-1$ contiguous GTP$^+$ monomers next to the tip convert, and finally
973: the GTP$^+$ at the tip converts to GDP$^-$ before the next attachment event.
974: The probability for the first $k-1$ conversion events is $\lambda^{-(k-1)}
975: (k-1)!$, where the factorial arises because the order of these steps is
976: irrelevant.  The probability of the last step is $\lambda^{-1}$.  Thus the
977: relative frequency of avalanches of size $k$ is
978: \begin{equation}
979: \label{Ak}
980: A_k\sim \lambda^{-k}\Gamma(k).
981: \end{equation}
982: 
983: The result can also be derived by the approach of Sec.~\ref{continuum}.  We
984: use the fact that the configuration
985: \begin{equation*}
986: |+\underbrace{-\cdots -}_{k-1} +\rangle
987: \end{equation*}
988: occurs with probability $\prod_{1\leq n\leq k-1} (1-e^{-n/\lambda})$.
989: Multiplying by the probability that the monomer at the tip converts before
990: the next attachment event then gives the probability for an avalanche of size
991: $k$:
992: \begin{equation}
993: \label{Ak-gen}
994: A_k = (1+\lambda)^{-1}\prod_{n=1}^{k-1} (1-e^{-n/\lambda})
995: \end{equation}
996: Using $1-e^{-n/\lambda}=n/\lambda$ we recover \eqref{Ak}.  If we expand the
997: exponent to the next order, $1-e^{-n/\lambda}\approx n/\lambda-n^2/(2\lambda^2)$,
998: Eq.~\eqref{Ak-gen} becomes
999: \begin{equation*}
1000: A_k = \lambda^{-k}\,\Gamma(k)\prod_{n=1}^{k-1} \left(1-\frac{n}{2\lambda}\right)\sim
1001: \lambda^{-k}\,\Gamma(k)\,e^{-k^2/4\lambda}.
1002: \end{equation*}
1003: 
1004: 
1005: \section{GENERAL GROWTH CONDITIONS}
1006: \label{general}
1007: 
1008: The general situation where the attachment and detachment rates, $\lambda$
1009: and $\mu$, have arbitrary values, and where the parameter $p\ne 1$ seems
1010: analytically intractable because the master equations for basic observables
1011: are coupled to an infinite hierarchy of equations to higher-order variables.
1012: For example, the quantity $n_0\equiv {\rm Prob}\{-\rangle\}$, the probability
1013: for a cap of length zero, satisfies the exact equation
1014: \begin{equation}
1015: \label{n0-mu}
1016: \dot n_0 = -p\lambda n_0+(1-n_0)-\mu\mathcal{N}_0 ~,
1017: \end{equation}
1018: and the speed of the tip is
1019: \begin{equation}
1020: \label{speed-gen}
1021: V(\lambda, \nu, p) = p\lambda n_0 + \lambda(1-n_0)- \mu n_0 ~.
1022: \end{equation}
1023: Here $\mathcal{N}_0\equiv {\rm Prob}\{+-\rangle\}$ is the probability that
1024: there is a GDP$^-$ at the front position with a GTP$^+$ on its left.  Thus to
1025: determine $n_0$ we must find $\mathcal{N}_0$, which then requires
1026: higher-order correlation functions, {\it etc}.  This hierarchical nature
1027: prevents an exact analysis and we turn to approximate approaches to map out
1028: the behavior in different regions of the parameter space.
1029: 
1030: \subsection{Limiting Cases}
1031: 
1032: For $\lambda, \mu \ll 1$, the conversion GTP$^+\to$ GDP$^-$ at rate one
1033: greatly exceeds the rates $\lambda, p\lambda, \mu$ of the other three basic
1034: processes that govern microtubule dynamics.  Hence we can assume that
1035: conversion is instantaneous.  Consequently, the end of a microtubule consists
1036: of a string of GDP$^-$, $|\cdots ---\rangle$, in which the tip advances at
1037: rate $p\lambda$ and retreats at rate $\mu$.  Thus from \eqref{speed-gen} the
1038: speed of the tip is
1039: \begin{equation}
1040: \label{V-small}
1041:   V(p,\lambda, \mu) = p\lambda-\mu
1042: \end{equation}
1043: when $p\lambda>\mu$. 
1044:  
1045: On the other hand, for $\lambda\gg 1$, $n_0\equiv {\rm Prob}\{-\rangle\}$ is small
1046: and ${\rm Prob}\{--\rangle\}$ is exceedingly small.  Hence $n_0={\rm
1047:   Prob}\{--\rangle\}+\mathcal{N}_0\approx\mathcal{N}_0$.  Substituting this result
1048: into \eqref{n0-mu} and solving for $n_0$ we find
1049: \begin{equation}
1050: \label{n0-solution}
1051: n_0 =  \frac{1}{1+p\lambda+\mu}.
1052: \end{equation}
1053: Note that indeed $n_0\ll 1$ when $\lambda\gg 1$.  Using \eqref{n0-solution}
1054: in \eqref{speed-gen} we obtain the general result for the growth velocity
1055: \begin{equation}
1056: \label{V-large}
1057: V=\lambda-\frac{(1-p)\lambda+\mu}{1+p\lambda+\mu} 
1058: \quad{\rm when}\quad \lambda\gg 1.
1059: \end{equation}
1060: 
1061: \subsection{The Phase Boundary}
1062: 
1063: A basic characteristic of microtubule dynamics is the phase boundary in the
1064: parameter space that separates the region where the microtubule grows without
1065: bound and a region where the mean microtubule length remains finite.  For
1066: small $\lambda$, this boundary is found from setting $V=0$ in
1067: Eq.~\eqref{V-small} to give $\mu_* = p\lambda$ for $\lambda\ll 1$.  The phase
1068: boundary is a straight line in this limit, but for larger $\lambda$ the
1069: boundary is a convex function of $\lambda$ (see Fig.~\ref{phase-diag}).  We
1070: can compute the velocity to second order in $\mu$ and $\lambda$ by assuming
1071: $\mathcal{N}_0=0$ and then using \eqref{n0-mu} and \eqref{speed-gen}. This
1072: leads to the  phase boundary
1073: \begin{equation}
1074: \label{crit-small}
1075: \mu_*=p\lambda+p\lambda^2,
1076: \end{equation}
1077: that is both convex and more precise.  On this phase boundary, the average
1078: tubule length grows as $\sqrt{t}$.
1079: 
1080: \begin{figure}[ht]
1081: \includegraphics*[width=0.315\textwidth]{mu-vs-lambda-1.eps}
1082: \includegraphics*[width=0.315\textwidth]{mu-vs-lambda-0.1.eps}
1083: \caption{Phase diagrams of a microtubule from simulations for (a) $p=1$ and
1084:   (b) $p=0.1$.  The dashed line represents the prediction \eqref{crit-small}
1085:   that is appropriate for small $\mu$.  The extremes of the error bars are
1086:   the points for which the tubule velocities are 0.005 and
1087:   0.015, and their average defines the data point. }
1088: \label{phase-diag}
1089: \end{figure}
1090: 
1091: When $\lambda$ is large, Eq.~\eqref{V-large} implies that $V$ is positive and
1092: it reduces to $V=\lambda-1$ for $\mu\gg \lambda$.  This simple result follows
1093: from the fact that recession of the microtubule is controlled by the unit
1094: conversion rate.  As soon as conversion occurs, detachment occurs immediately
1095: for $\mu\gg\lambda$ and the microtubule recedes by one step.  Since advancement
1096: occurs at rate $\lambda$, the speed of the tip is simply $\lambda-1$.
1097: However, for extremely large $\mu$ the prediction $V=\lambda-1$ breaks
1098: down and the microtubule becomes compact.  To determine the phase boundary in
1099: this limit, consider first the case $\mu=\infty$.  As shown in the previous
1100: section, the probability of a catastrophe roughly scales as
1101: $e^{-\pi^2\lambda/6}$ so that the typical time between catastrophes is
1102: $e^{\pi^2\lambda/6}$.  Since $V=\lambda-1$, the typical length of a
1103: microtubule just before a catastrophe is $(\lambda-1) e^{\pi^2\lambda/6}$.
1104: Suppose now the detachment rate $\mu$ is very large but finite.  The
1105: microtubule is compact if the time to shrink a microtubule of length $\lambda
1106: e^{\pi^2\lambda/6}/\mu$ is smaller than the time $(p\lambda)^{-1}$ required
1107: to generate a GTP$^+$ by the attachment $|\cdots -\rangle \Longrightarrow
1108: |\cdots -+\rangle$ and thereby stop the shrinking.  We estimate the location
1109: of the phase boundary by equating the two times to give
1110: \begin{equation}
1111: \label{crit-large}
1112: \mu_*\sim p \lambda^2 e^{\pi^2\lambda/6}\quad{\rm when}\quad \lambda\gg 1.
1113: \end{equation}
1114: 
1115: We checked the theoretical predictions \eqref{crit-small} and
1116: \eqref{crit-large} for the phase boundary numerically
1117: (Fig.~\ref{phase-diag}).  For small $\lambda$, the agreement between theory
1118: and the simulation is excellent.  For larger $\lambda$, the tubule growth is
1119: more intermittent and it becomes increasingly difficult to determine the
1120: phase boundary with precision.  Nevertheless, the qualitative expectations of
1121: our theory remain valid.
1122: 
1123: \subsection{Fluctuations of the Tip}
1124: 
1125: Finally, we study the fluctuations of the tip in the small and large
1126: $\lambda$ limits.  In the former case but also on the growth phase
1127: $p\lambda>\mu$, the tip undergoes a biased random walk with diffusion
1128: coefficient
1129: \begin{equation}
1130: \label{D-small}
1131:   D(p,\lambda, \mu) = \frac{p\lambda+\mu}{2} 
1132:   \quad{\rm when}\quad 1\gg p\lambda>\mu
1133: \end{equation}
1134: 
1135: For large $\lambda$, the analysis is simplified by the principle that can be
1136: summarized by: ``The leading behaviors in the $\lambda\to\infty$ limit are
1137: universal, that is, independent of $p$ and $\mu$.'' This is not true if $p$
1138: is particularly small [like $\lambda^{-1}$] and/or if $\mu$ is particularly
1139: large [like $\mu_*$ given by \eqref{crit-large}]. But when $p\lambda\gg 1$
1140: and $\mu\ll \mu_*$, the above principle is true, and
1141: \begin{equation}
1142: \label{VD-sol}
1143: V=\lambda, \quad D=\frac{\lambda}{2}
1144: \end{equation}
1145: in the leading order. 
1146: 
1147: The computation of sub-leading behaviors is more challenging. We merely state
1148: here two asymptotic results. When $\mu\ll\lambda$, we again have the relation
1149: $D=V/2$, with $V=\lambda+1-p^{-1}$. If $\mu_*\gg \mu\gg\lambda\gg 1$, we
1150: find
1151: \begin{equation}
1152: \label{VD-large}
1153: V=\lambda-1, \quad D=\frac{\lambda+1}{2}
1154: \end{equation}
1155: The derivation of the latter uses the probabilities $X(L,t)$ and $Y(L,t)$ and
1156: follows similar steps as in Sect.~\ref{XY-sub}.
1157: 
1158: \section{SUMMARY}
1159: 
1160: We investigated a simple dynamical model of a microtubule that grows by
1161: attachment of a GTP$^+$ to its end at rate $\lambda$, irreversible conversion
1162: of any GTP$^+$ to GDP$^-$ at rate 1, and detachment of a GDP$^-$ from the end
1163: of a microtubule at rate $\mu$.  Remarkably, these simple update rules for a
1164: one-dimensional system lead to steady growth, wild fluctuations, or a steady
1165: state.  Our model has a minimalist formulation and therefore is not meant to
1166: account for all of the microscopic details of microtubule dynamics.  Rather,
1167: our main goal has been to solve for the structural and dynamical properties
1168: of this idealized microtubule model.  Some of the quantities that we
1169: determined, such as island size distributions, have not been studied
1170: previously.  Thus our predictions about the cap and island size distributions
1171: may help motivate experimental studies of these features of microtubules.
1172: 
1173: A rich phenomenology was found as a function of the three fundamental rates
1174: in our model.  When GTP$^+$ attachment is dominant ($\lambda\gg 1$) and the
1175: attachment is independent of the identity of the last monomer on the free end
1176: of the microtubule ($p=1$), the GTP$^+$ and GDP$^-$ organize into alternating
1177: domains, with gradually longer GTP$^+$ domains and gradually shorter GDP$^-$
1178: domains toward the tip of the microtubule (Fig.~\ref{cartoon}).  Here, the
1179: parameter $\lambda$ could naturally be varied experimentally by either
1180: changing the temperature or the concentration of tubulins (free GTP$^+$) in
1181: the solution.  
1182: 
1183: The basic geometrical features in this growing phase of a microtubule can be
1184: summarized as:
1185: 
1186: \medskip
1187: \begin{tabular}{|c|l|c|}
1188: 	\hline
1189: symbol &  meaning   &   scaling behavior  \\
1190: 	\hline
1191: $N$  & \# GTP$^+$ monomers & $\lambda$  \\
1192: $L$  & tubule length & $\lambda t$  \\
1193: $\langle k\rangle$ & average cap length &  $\sqrt{\pi\lambda/2}$\\
1194: $I$ & \# islands  & $ \lambda/2$\\
1195: $I_k$ & \# GTP$^+$ $k$-islands  & $2\lambda/k^3$\\
1196: $J_k$ & \# GDP$^-$ $k$-islands  & $\lambda/k^2$\\
1197: \hline
1198: \end{tabular}
1199: 
1200: \smallskip
1201: \noindent We emphasize that the island size distributions of GTP$^+$ and
1202: GDP$^-$ are robust power laws with respective exponents of $3$ and $2$.  In
1203: the limit of $p\ll 1$, in which attachment is suppressed when a GDP$^-$ is at
1204: the free end of the microtubule, the average number of GTP$^+$ monomers on
1205: the microtubule asymptotically is $p\lambda$, while the rest of the results
1206: in the above table remain robust in the long-time limit.
1207: 
1208: Conversely, when detachment of GDP$^-$ from the end of the tubule is dominant
1209: (detachment rate $\mu\to\infty$, a rate that also could be controlled by the
1210: temperature), the microtubule length remains bounded but its length can
1211: fluctuate strongly.  When the attachment rate is also large, the strong
1212: competition between attachment and detachment leads to wild fluctuations in
1213: the microtubule length even with steady external conditions.  We developed a
1214: probabilistic approach that shows that the time between catastrophes, where
1215: the microtubule shrinks to zero length, scales exponentially with the
1216: attachment rate $\lambda$.  Thus a microtubule can grow essentially freely
1217: for a very long time before undergoing a catastrophe.
1218: 
1219: For the more biologically relevant case of intermediate parameter values, we
1220: extended our theoretical approaches to determine basic properties of the
1221: tubule, such as its rate of growth, fluctuations of the tip around this mean
1222: growth behavior, and the phase diagram in the ($\lambda,\mu$) parameter
1223: space.  In this intermediate regime, numerical simulations provide more
1224: detailed picture of the geometrical structure and time history of a
1225: microtubule.
1226: 
1227: \acknowledgments{Paul Weinger generated the early numerical results that led
1228:   to this work.  We also thank Rajesh Ravindran, Allison Ferguson and Daniel
1229:   Needleman for many helpful conversations.  We acknowledge financial support
1230:   to the Program for Evolutionary Dynamics at Harvard University by Jeffrey
1231:   Epstein and NIH grant R01GM078986 (TA), NSF grant DMR0403997 (BC and MM),
1232:   and NSF grants CHE0532969 (PLK) and DMR0535503 (SR).}
1233: 
1234: 
1235: \appendix
1236: 
1237: \section{JOINT DISTRIBUTION FOR $p=1$}
1238: \label{app-P}
1239: 
1240: Introducing the two-variable generating function 
1241: \begin{equation}
1242: \label{gen-P}
1243: \mathcal{P}(x,y,t)=\sum_{L\geq N\geq 0}x^L y^N P(L,N,t)
1244: \end{equation}
1245: we recast \eqref{PLN} into a partial differential equation 
1246: \begin{equation}
1247: \label{P-diff}
1248: \frac{\partial \mathcal{P}}{\partial t}=(1-y)\frac{\partial \mathcal{P}}{\partial y}
1249: -\lambda(1-xy)\mathcal{P}
1250: \end{equation}
1251: Writing
1252: \begin{equation}
1253: \label{PQ}
1254: \mathcal{P}(x,y,t)=e^{\lambda[xy-(1-x)\ln(1-y)]}\,\,\mathcal{Q}(x,y,t)
1255: \end{equation}
1256: we transform \eqref{P-diff} into a wave equation for an auxiliary function 
1257: $\mathcal{Q}(x,y,t)$
1258: \begin{equation}
1259: \label{Q-diff}
1260: \frac{\partial \mathcal{Q}}{\partial t}=(1-y)\frac{\partial \mathcal{Q}}{\partial y}
1261: \end{equation}
1262: whose general solution is 
1263: \begin{equation}
1264: \label{Q-sol}
1265: \mathcal{Q}(x,y,t)=\Phi(x,\ln(1-y)-t)
1266: \end{equation}
1267: %with $\Phi$ being an arbitrary differentiable function of two variables. 
1268: The initial condition $P(L,N,0)=\delta_{L,0}\,\delta_{N,0}$ implies
1269: $\mathcal{P}(x,y,0)=1$ and therefore
1270: \begin{equation*}
1271: e^{\lambda[xy-(1-x)\ln(1-y)]}\,\,\Phi(x,\ln(1-y))=1
1272: \end{equation*}
1273: from which
1274: \begin{equation}
1275: \label{Phi-sol}
1276: \Phi(a,b)=e^{\lambda[(1-a)b+a(e^b-1)]}
1277: \end{equation}
1278: Combining Eqs.~\eqref{PQ}, \eqref{Q-sol}--\eqref{Phi-sol} we arrive at
1279: \begin{equation}
1280: \label{P-sol}
1281: \lambda^{-1}\ln\mathcal{P}=xy(1-e^{-t})-t-x(1-e^{-t}-t)
1282: \end{equation}
1283: 
1284: 
1285: \section{JOINT DISTRIBUTION FOR $p\ne1$}
1286: \label{app}
1287: 
1288: For $p\ne 1$, we consider the distributions $X_N$ and $Y_N$, defined as the
1289: probabilities to have $N$ GTP$^+$ monomers with the tip being either GTP$^+$
1290: or GDP$^-$, respectively.  These probabilities satisfy a closed set of
1291: coupled equations.  In the stationary state these equations become
1292: \begin{subequations}
1293: \begin{align}
1294: \label{XN}
1295:    &(N+\lambda) X_N=\lambda X_{N-1}+p\lambda Y_{N-1}+NX_{N+1}\\
1296: \label{YN} 
1297:    &(N+p\lambda) Y_N=X_{N+1}+(N+1)Y_{N+1}.
1298: \end{align} 
1299: \end{subequations}
1300: 
1301: Since $X_0\equiv 0$, it is convenient to define the generating functions
1302: corresponding to $X_N$ and $Y_N$ as follows:
1303: \begin{subequations}
1304: \begin{align}
1305: \label{X-gen-def}
1306: \mathcal{X}(z)&= \sum_{N\geq 1}z^{N-1} X_N\\
1307: \label{Y-gen-def}
1308: \mathcal{Y}(z)&=\sum_{N\geq 0}z^{N} Y_N.
1309: \end{align} 
1310: \end{subequations}
1311: Now multiply Eq.~\eqref{XN} by $z^{N}$ and Eq.~\eqref{YN} by $z^{N-1}$ and
1312: sum over all $N\geq 1$ or $N\geq 0$, respectively, to obtain
1313: \begin{subequations}
1314: \begin{align}
1315: \label{XY1}
1316:    p\lambda \mathcal{Y}&=\mathcal{X}-\zeta(\mathcal{X}-\mathcal{X}')\\
1317: \label{XY2} 
1318:    p\lambda \mathcal{Y}&=\mathcal{X}-\zeta\mathcal{Y}'.
1319: \end{align} 
1320: \end{subequations}
1321: where $\zeta=\lambda(z-1)$ and prime denotes a derivative in $\zeta$.  We can
1322: reduce these two coupled first-order differential equations to uncoupled
1323: second-order equations:
1324: \begin{subequations}
1325: \begin{align}
1326: \label{X-gen}
1327:    &\zeta\mathcal{X}'' + (2+p\lambda-\zeta)\mathcal{X}'-(1+p\lambda)\mathcal{X}=0\\
1328: \label{Y-gen} 
1329:    &\zeta\mathcal{Y}'' + (2+p\lambda-\zeta)\mathcal{Y}'-p\lambda\mathcal{Y}=0.
1330: \end{align} 
1331: \end{subequations}
1332: The solutions are the confluent hypergeometric functions
1333: \begin{subequations}
1334: \begin{align}
1335: \label{X-gen-sol}
1336:    \mathcal{X}(z)&=\frac{p\lambda}{1+p\lambda}\,F(1+p\lambda; 2+p\lambda; \zeta)\\
1337: \label{Y-gen-c} 
1338:     \mathcal{Y}(z)&=\frac{1}{1+p\lambda}\,F(p\lambda; 2+p\lambda; \zeta).
1339: \end{align} 
1340: \end{subequations}
1341: 
1342: These generating functions have seemingly compact expressions but one has to
1343: keep in mind that the $X$ and $Y$ probabilities are actually infinite sums.
1344: For instance, $Y_0=\mathcal{Y}(z=0)=\mathcal{Y}(\zeta=-\lambda)$. Recalling
1345: the definition of the confluent hypergeometric function we obtain
1346: \begin{eqnarray*}
1347:  Y_0 &=&\frac{1}{1+p\lambda}\,F(p\lambda; 2+p\lambda; -\lambda)\\
1348:  &=& \frac{1}{1+p\lambda}\sum_{n\geq 0}
1349:  \frac{(p\lambda)_n}{(2+p\lambda)_n}\,\frac{(-\lambda)^n}{n!}
1350: \end{eqnarray*}
1351: where $(a)_n=a(a+1)\ldots (a+n-1)=\Gamma(a+n)/\Gamma(a)$ is the rising factorial. 
1352: Note that $\Pi_0=Y_0$. Computing 
1353: \begin{eqnarray*}
1354: X_1 &=&\frac{p\lambda}{1+p\lambda}\,F(1+p\lambda; 2+p\lambda; -\lambda)\\
1355: Y_1 &=&\lambda\,\frac{p\lambda}{(1+p\lambda)(2+p\lambda)}\,
1356: F(1+p\lambda; 3+p\lambda; -\lambda)
1357: \end{eqnarray*}
1358: one can determine $\Pi_1=X_1+Y_1$.  Some of these formulas can be simplified
1359: using the Kummer relation
1360: \begin{equation*}
1361: F(a;b;\zeta)=e^\zeta F(b-a;b;-\zeta)
1362: \end{equation*}
1363: For instance,
1364: \begin{eqnarray*}
1365:   Y_0 &=&\sum_{n\geq 0}
1366:   \frac{(n+1)\lambda^n e^{-\lambda}}{(1+p\lambda)_{n+1}}\\
1367:   X_1 &=&p\lambda\sum_{n\geq 0}
1368:   \frac{\lambda^n e^{-\lambda}}{(1+p\lambda)_{n+1}}\\
1369:   Y_1 &=&p\lambda^2
1370:   \sum_{n\geq 0}\frac{(n+1)\lambda^n e^{-\lambda}}{(1+p\lambda)_{n+2}}.
1371: \end{eqnarray*}
1372: 
1373: \begin{thebibliography}{19}
1374: 
1375: \bibitem{FBL} D.~K. Fygenson, E.~Braun, and A.~Libchaber, Phys.\ Rev.\ E {\bf
1376:     50}, 1579 (1994).
1377: 
1378: \bibitem{VCJ} O.~Valiron, N. Caudron, and D. Job, Cell.\ Mol.\ Life Sci.\
1379:   {\bf 58}, 2069 (2001).
1380: 
1381: \bibitem{MK} T. Mitchison and M.~Kirschner, Nature \textbf{312}, 237
1382:   (1984).
1383: 
1384: \bibitem{AH} J. Howard and A.~A. Hyman, Nature \textbf{422}, 753 (2003).
1385: 
1386: \bibitem{WMD} C.~E. Walczak, T.~J. Mitchison, and A. Desai, Cell
1387:   \textbf{84}, 37 (1996).
1388: 
1389: \bibitem{DM} A. Desai and T. J. Mitchison, Annu.\ Rev.\ Cell Dev.\ Biol.
1390:   \textbf{13}, 83 (1997).
1391: 
1392: \bibitem{dimer} We term a heterodimer that consists of $\alpha$- and
1393:   $\beta$-tubulin as a ``monomer'' throughout this paper.
1394: 
1395: \bibitem{DL} M. Dogterom and S. Leibler, Phys.\ Rev.\ Lett. \textbf{70},
1396:   1347 (1993).
1397: 
1398: \bibitem{JCF} I.~M. J$\acute{\rm a}$nosi, D. Chr\'etien, and H.
1399:   Flyvbjerg, Eur.\ Biophys.\ J. {\bf 27}, 501 (1998); Biophys.\ J.
1400:   \textbf{83}, 1317 (2002).
1401: 
1402: \bibitem{vanburen05} V. VanBuren, D. J. Odde, and L. Cassimeris, Proc.\ Nat.\
1403:   Acad.\ Sci.\ USA {\bf 99}, 6035 (2002); V. VanBuren, L. Cassimeris, and
1404:   D. J. Odde, Biophys.\ J. {\bf 89}, 2911 (2005).
1405: 
1406: \bibitem{mol} M. I. Molodtsov, E. A. Ermakova, E. E. Shnol, E. L. Grishchuk,
1407:   J. R. McIntosh, and F. I. Ataullakhanov, Biophys.\ J. {\bf 88} 3167 (2005).
1408: 
1409: \bibitem{mandelkow91} E. M. Mandelkow, E. Mandelkow, R. A. Milligan, J. Cell
1410:   Biol.\ {\bf 114}, 997 (1991); D.~Chr\'etien, S.~D.~Fuller, and E.~Karsenti,
1411:   J.  Cell Biol.\ {\bf 129}, 1311 (1995).
1412: 
1413: \bibitem{Bi} D.~J. Bicout, Phys.\ Rev.\ E \textbf{56}, 6656 (1997);
1414:  D.~J. Bicout and R. J. Rubin, Phys.\ Rev.\ E \textbf{59}, 913 (1999).
1415: 
1416: \bibitem{HZ} M. Hammele and W. Zimmermann, Phys.\ Rev.\ E {\bf 67}, 021903
1417:   (2003).
1418: 
1419: \bibitem{MKMC} P. K. Mishra, A. Kunwar, S. Mukherji, and D. Chowdhury, Phys.\
1420:   Rev.\ E {\bf 72}, 051914 (2005).
1421: 
1422: \bibitem{MGGA} G. Margolin, I. V. Gregoretti, H. V. Goodson, and M.~S.~Alber,
1423:   Phys.\ Rev.\ E {\bf 74}, 041920 (2006).
1424: 
1425: \bibitem{hill} T. L. Hill and Y. Chen, Proc.\ Nat.\ Acad.\ Sci.\ USA {\bf 81},
1426:   5772 (1984).
1427: 
1428: \bibitem{F} H. Flyvbjerg, T.~E. Holy, and S. Leibler,Phys.\ Rev.\ Lett.
1429:   \textbf{73}, 2372 (1994); Phys.\ Rev.\ E \textbf{54}, 5538 (1996).
1430: 
1431: \bibitem{ZLSW} C. Zong, T. Lu, T. Shen, and P.~G. Wolynes, Phys.\ Biol.\
1432:   \textbf{3}, 83 (2006).
1433: 
1434: \bibitem{CR} B. Chakraborty and R. Rajesh, unpublished.
1435: 
1436: \bibitem{short} An abbreviated account of this work is given in T.~Antal, P.
1437:   L. Krapivsky, and S. Redner, J. Stat.\ Mech.\ L05004 (2007).
1438: 
1439: \bibitem{MLA} P. W. Messer, M. L\"assig, and P. F. Arndt, J. Stat.\ Mech.\
1440:   P10004 (2005).
1441: 
1442: \bibitem{TGSM} J.~S. Tirnauer, S. Grego, E. Slamon, and T.~J.~Mitchison,
1443:   Mol.\ Biol.\ Cell \textbf{13}, 3614 (2002).
1444: 
1445: \bibitem{AS} M. Abramowitz and I.~A. Stegun, {\it Handbook of Mathematical
1446:     Functions} (Dover, New York, 1972).
1447: 
1448: \bibitem{network1} H.~A. Simon, Biometrica \textbf{42}, 425 (1955); A.~L.
1449:   Bar\'abasi and R. Albert, Science \textbf{286}, 509 (1999).
1450: 
1451: \bibitem{network2} P.~L. Krapivsky, S. Redner, and F. Leyvraz, Phys.\ Rev.\
1452:   Lett.\ \textbf{85}, 4629 (2000);  P.~L. Krapivsky and S. Redner, Phys.\
1453:   Rev.\ E {\bf 63}, 066123  (2001).
1454: 
1455: \bibitem{network3} S.~N. Dorogovtsev, J.~F.~F.~Mendes, and A.~N. Samukhin,
1456:   Phys.\ Rev.\ Lett.\ \textbf{85}, 4633 (2000).
1457: 
1458: \bibitem{finite} P.~L. Krapivsky and S. Redner, J. Phys.\ A \textbf{35}, 9517
1459:   (2002).
1460: 
1461: \bibitem{extremal} J. Galambos, {\it The Asymptotic Theory of Extreme Order
1462:     Statistics} (Krieger Publishing Co., Florida, 1987).
1463: 
1464: \bibitem{apostol} T.~M. Apostol, {\it Modular Functions and Dirichlet Series
1465:     in Number Theory, $2^{\rm nd}$ {\rm ed.}} (Springer-Verlag, New York,
1466:   1990).
1467: 
1468: 
1469: 
1470: 
1471: \end{thebibliography}
1472: \end{document}
1473: 
1474: 
1475: