1: %%%%%%%%%%%%%%%%%% Start of LaTeX file %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%% LaTeX2e
3: \documentclass[
4: a4paper,
5: 12pt, onecolumn,
6: % 10pt, twocolumn,
7: % draft
8: ]{article}
9:
10: \title{\bf Bethe Ansatz calculation of
11: the spectral gap of the asymmetric exclusion process
12: }
13: \author{ O. Golinelli, K. Mallick
14: \bigskip
15: \\ \ad email: \{golinelli, mallick\}@cea.fr
16: \\ \ad Service de Physique Th\'eorique,
17: \\ \ad Cea Saclay, 91191 Gif-sur-Yvette, France
18: % fax: (+33) 1 69 08 81 20
19: }
20: \date{\normalsize December 15, 2003
21: \\ Preprint T03/196 ; arXiv:cond-mat/0312371
22: }
23: %_______________________________________________________________________
24:
25: %%% page style
26: \newcommand {\ad}{\normalsize\em} % style for address
27: \pagestyle{myheadings}
28: \markright{O. Golinelli, K. Mallick --- ASEP spectral gap}
29:
30: %%% for two-columns
31: %\textwidth 18cm \oddsidemargin -1cm \evensidemargin -1cm
32: %\textheight 23.5cm \topmargin -0.5cm
33:
34: %%% encapsulated figures
35: \usepackage[final]{graphicx}
36: \newcommand{\figwidth}{\columnwidth}
37: %\newcommand{\figwidth}{8.8truecm}
38:
39:
40: \newcommand{\Li}{\mathrm{Li}} % Notation for polylog functions
41: \newcommand{\Arg}{\mathrm{Arg}}
42:
43: \begin{document}
44: %\twocolumn[\hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname
45: \maketitle
46: %_______________________________________________________________________
47:
48: \begin{abstract}
49: %===============
50: \normalsize
51:
52: We present a new derivation of the spectral gap of the totally
53: asymmetric exclusion process on a half-filled ring of size $L$ by using
54: the Bethe Ansatz. We show that, in the large $L$ limit, the Bethe
55: equations reduce to a simple transcendental equation involving the
56: polylogarithm, a classical special function. By solving that equation,
57: the gap and the dynamical exponent are readily obtained. Our method can
58: be extended to a system with an arbitrary density of particles.
59:
60: \medskip \noindent Keywords: ASEP, Bethe Ansatz, Dynamical Exponent,
61: Spectral Gap.
62:
63: \medskip \noindent Pacs numbers: 05.40.-a; 05.60.-k.
64:
65: \end{abstract}
66: %\vskip2pc ] % for twocolumn
67:
68:
69: %=====================
70: \section{Introduction}
71: %=====================
72:
73: The asymmetric simple exclusion process (ASEP) is a model of driven
74: diffusive particles on a lattice with hard-core exclusion (for general
75: review see Spohn 1991). The ASEP appears as a minimal building block in a
76: large variety of models for hopping conductivity (Richards 1977), polymer
77: reptation (Widom {\it et al.} 1991), traffic flow (Schreckenberg and Wolf
78: 1998) or surface growth (Krug 1997). In particular, the ASEP in one
79: dimension is a discrete version of the Kardar-Parisi-Zhang (KPZ) equation
80: (Halpin-Healy and Zhang 1995). In biophysics, the ASEP has been used to
81: describe the diffusion of macromolecules through narrow vessels (Levitt
82: 1973) and the kinetics of protein synthesis on RNA (MacDonald and Gibbs
83: 1969); more recently, a mapping between sequence-alignment and the
84: exclusion process has been proposed (Bundschuh 2002). From a theoretical
85: point of view, the ASEP plays the role of a paradigm in non-equilibrium
86: statistical mechanics: it displays a variety of features such as boundary
87: induced phase transitions (Krug 1991), spontaneous symmetry breaking in
88: one dimension (Evans {\it et al.} 1995) and dynamical phase separation
89: (Evans {\it et al.} 1998).
90:
91: Exact results for the ASEP in one dimension have been derived using two
92: complementary approaches (for a review see Derrida 1998). The Matrix
93: Ansatz (Derrida {\it et al.} 1993) allows to calculate steady state
94: properties such as invariant measures (Speer 1993), current fluctuations in
95: the stationary state and large deviation functionals (Derrida {\it et al.}
96: 2003). The Bethe Ansatz (Dhar 1987) provides spectral information about the
97: evolution (Markov) operator (Gwa and Spohn 1992; Sch\"utz 1993; Kim 1995)
98: which can be used to derive large deviation functions (Derrida and Lebowitz
99: 1998; Derrida and Appert 1999; Derrida and Evans 1999). The exact relation
100: between these two techniques is still a matter of investigation (Alcaraz
101: {\it et al.} 1994; Stinchcombe and Sch\"utz 1995; Alcaraz and Lazo 2003).
102:
103: The relaxation time to the stationary state for the ASEP on a lattice of
104: size $L$ scales typically as a power law, $L^z$, $z$ being the ASEP
105: dynamical exponent. The calculation of $z$ for a one dimensional system by
106: Bethe Ansatz is an important exact result that was first announced by Dhar,
107: who found $z = 3/2$ (Dhar 1987). The spectral gap (and thus $z$) was
108: subsequently calculated for the half-filling case by Gwa and Spohn (1992)
109: and for an arbitrary density by Kim (1995) who mapped the ASEP into a
110: non-Hermitian XXZ Heisenberg spin chain. The one-dimensional ASEP belongs
111: to the KPZ universality class and therefore the KPZ dynamical exponent in
112: one dimension is equal to 3/2; this result was previously deduced from
113: Galilean invariance and renormalization group arguments (for a review see
114: Krug 1997).
115:
116: In this work, we present a new method of calculating the spectral gap of
117: the totally asymmetric exclusion process (TASEP) starting from the Bethe
118: Ansatz equations. Our method, based on an analytic continuation formula,
119: circumvents the technical difficulties involved in the derivation of Gwa
120: and Spohn (1992) and renders the calculation of the ASEP dynamical exponent
121: much more concise and transparent. Besides, our technique can be readily
122: extended to the arbitrary density case and allows us to calculate the
123: spectral gap for the asymmetric exclusion process with a tagged particle.
124:
125: The outline of this paper is as follows. In section 2, we recall the
126: definition of the TASEP, present the Bethe Ansatz equations without
127: deriving them and summarize their analysis. In section 3, we present our
128: original calculation of the spectral gap in the half-filling case.
129: Concluding remarks and generalizations of our method are given in section~4.
130:
131:
132: %===============================================
133: \section{Bethe Ansatz for the TASEP}
134: %===============================================
135:
136: \subsection{The TASEP model}
137:
138: We consider the totally asymmetric simple exclusion process on a periodic
139: one dimensional lattice with $L$ sites (sites $i$ and $L + i$ are
140: identical). In this model, the total number $n$ of particles is conserved.
141: Each lattice site $i$ ($1 \le i \le L$) is either occupied by one particle
142: or is empty ({\em exclusion rule}). Stochastic dynamical rules govern the
143: evolution of the system: a particle on a site $i$ at time $t$ jumps, in
144: the interval between times $t$ and $t+dt$, with probability $dt$ to the
145: neighbouring site $i+1$ if this site is empty. The total number of
146: configurations for $n$ particles on a ring with $L$ sites is given by
147: $\Omega = L! / [ n! (L-n)!]$. In the stationary state, all configurations
148: have the same probability $1/\Omega$ (Derrida 1998).
149:
150: A configuration can be characterized by the positions of the $n$ particles
151: on the ring, $(x_1, x_2, \dots, x_n)$ with $1 \le x_1 < x_2 < \dots < x_n
152: \le L$. We call $\psi_t(x_1,\dots, x_n)$ the probability of this
153: configuration at time $t$; the probability distribution $\psi_t$ of the
154: system at time $t$ is thus a $\Omega$-dimensional vector. As the ASEP is
155: a continuous-time Markov ({\it i.e.}, memoryless) process, the time
156: evolution of $\psi_t$ is determined by the master equation
157: \begin{equation}
158: \frac{d\psi_t}{dt} = M \psi_t \, ,
159: \end{equation}
160: where the transition rate $\Omega\times\Omega$ matrix $M$ is the Markov
161: matrix. A right eigenvector $\psi$ is associated with the eigenvalue $E$
162: of $M$ if
163: \begin{equation}
164: M \psi = E\psi \, .
165: \label{eq:mpsi=epsi}
166: \end{equation}
167: The Markov matrix $M$ is a real non-symmetric matrix and, therefore, its
168: eigenvalues (and eigenvectors) are either real numbers or complex
169: conjugate pairs. The spectrum of $M$ contains the eigenvalue $E = 0$ and
170: the associated right eigenvector is the stationary state $\psi(x_1,\dots,
171: x_n) = 1/\Omega$. Because the dynamics is ergodic ({\it i.e.}, $M$ is an
172: irreducible and aperiodic Markov matrix), the Perron-Frobenius theorem
173: implies that all eigenvalues $E$ except $0$ have a strictly negative real
174: part; the relaxation time of the corresponding eigenmode is $\tau =
175: -1/\mathrm{Re}(E)$. (The imaginary part of $E$ gives rise to an oscillatory
176: behaviour).
177:
178: In this paper, we shall calculate the gap $E_1$, {\it i.e.}, the non-zero
179: eigenvalue of $M$ with largest real part. The eigenmode associated with
180: $E_1$ has thus the longest relaxation time that scales as $L^z$, $z$
181: being the dynamical exponent.
182:
183:
184: %===============================
185: \subsection{The Bethe equations}
186: %===============================
187:
188: Writing $M$ explicitly, the eigenvalue equation~(\ref{eq:mpsi=epsi})
189: becomes
190: \begin{equation}
191: E \psi(x_1,\dots, x_n) = \sum_i \left[
192: \psi(x_1, \dots, x_{i-1},\ x_i-1,\ x_{i+1}, \dots, x_n)
193: - \psi(x_1,\dots, x_n) \right] \, ,
194: \end{equation}
195: where the sum runs over the indexes $i$ such that $ x_{i-1} < x_i-1$, {\it
196: i.e.}, such that the corresponding jump is allowed. The {\em Bethe Ansatz}
197: assumes that the eigenvectors $\psi$ can be written in the form
198: \begin{equation}
199: \psi(x_1,\dots,x_n) = \sum_{\sigma \in \Sigma_n} A_{\sigma} \,
200: z_{\sigma(1)}^{x_1} \, z_{\sigma(2)}^{x_2} \dots z_{\sigma(n)}^{x_n}
201: \label{eq:ba} \, ,
202: \end{equation}
203: where $\Sigma_n$ is the group of the $n!$ permutations of $n$ indexes. The
204: coefficients $\{A_{\sigma}\}$ and the fugacities $\{z_1, \dots, z_n\}$ are
205: complex numbers to be determined. The eigenvalue $E$ associated with an
206: eigenvector of the form~(\ref{eq:ba}) is given by
207: \begin{equation}
208: E = -n + \sum_{i=1}^n 1/z_i \, .
209: \label{eq:e}
210: \end{equation}
211: Using matching conditions at the boundary surfaces $ x_{i-1} = x_i-1$ and
212: the periodicity of the lattice, it can be shown that a vector $\psi$ of
213: the type~(\ref{eq:ba}) is an eigenvector of $M$ if the fugacities
214: $\{z_1,\dots, z_n\}$ satisfy the {\em Bethe equations}
215: \begin{equation}
216: (z_i-1)^n z_i^{-L} = - \prod_{j=1}^n(1-z_j) \ \ \
217: \hbox{for} \ \ \ i=1,\dots,n \, .
218: \label{eq:be}
219: \end{equation}
220: The procedure for deriving these equations has been thoroughly explained in
221: Halpin-Healy and Zhang (1995) and Derrida (1998).
222:
223: The obvious solution $z_1 = \dots = z_n =1$ provides the stationary
224: distribution with eigenvalue $0$. More generally, given a solution
225: $\{z_1,\dots, z_n\}$ of Eq.~(\ref{eq:be}), the corresponding eigenvalue
226: $E$ is obtained from Eq.~(\ref{eq:e}); moreover, the coefficients
227: $\{A_{\sigma}\}$ and thus the eigenvector $\psi$ are uniquely determined.
228: In order to have a complete basis of eigenvectors, $\Omega$ independent
229: solutions of the Bethe equations~(\ref{eq:be}) are needed.
230:
231: Following Gwa and Spohn (1992), we introduce $Z_i = 2/z_i -1$. The
232: equations~(\ref{eq:e}) and (\ref{eq:be}) then become, respectively,
233: \begin{equation}
234: 2E = -n + \sum_{j=1}^n Z_j
235: \label{eq:eZ} \, ,
236: \end{equation}
237: and
238: \begin{equation}
239: (1-Z_i)^n \ (1+Z_i)^{L-n}
240: = - 2^L \prod_{j=1}^n \frac{Z_j - 1}{Z_j + 1} \ \ \
241: \hbox{for} \ \ \ i=1,\dots,n \, .
242: \label{eq:bez}
243: \end{equation}
244: We notice that the left hand side of Eq.~(\ref{eq:bez}) is a polynomial in
245: $Z_i$ whereas the right hand side (r.h.s.) is independent of the index
246: $i$.
247:
248: The analysis of the Bethe equations is simplified if only half-filled
249: models are considered, that is, if $L = 2n$ (Gwa and Spohn 1992).
250: Equation~(\ref{eq:bez}) then reduces to
251: \begin{equation}
252: (1-Z_i^2)^n = - 4^n \prod_{j=1}^n \frac{Z_j - 1}{Z_j + 1} \ \ \
253: \hbox{for} \ \ \ i=1,\dots,n \, .
254: \label{eq:behf}
255: \end{equation}
256: The half-filling restriction does not affect the physical behaviour of the
257: ASEP: in the large $L$ limit, models with arbitrary density, $\rho = n/L
258: \in \ ]0,1[$, belong to the same universality class. However, systems with
259: vanishingly small density of particles ($\rho \to 0$) or holes ($\rho \to
260: 1$) exhibit a different behaviour and will not be discussed here.
261:
262:
263: %=====================================
264: \subsection{Analysis of Bethe equations}
265: %=====================================
266:
267: Taking advantage of the fact that the r.h.s. of Eq.~(\ref{eq:behf}) is
268: independent of the index $i$, the Bethe equations can be reformulated as
269: follows. Consider the polynomial equation
270: \begin{equation}
271: (1-Z^2)^n = Y \, ,
272: \label{eq:z}
273: \end{equation}
274: where $Y$ is a given complex number. Writing
275: \begin{equation}
276: Y = -e^{u\pi} \, ,
277: \label{eq:u}
278: \end{equation}
279: $u$ being a complex number with $ -1 \le \mathrm{Im}(u) < 1$, we obtain the
280: $n$-th roots of $Y$
281: \begin{equation}
282: y_m = e^{(u+i)\pi/n} e^{(m-1)2i\pi/n} \ \ \, \hbox{ for } m = 1, \ldots, n \, .
283: \label{eq:ym}
284: \end{equation}
285: The $y_m$'s are evenly spaced on a circle of center 0 and radius
286: $|Y|^{1/n}$ and are labeled counter-clockwise $0 \le \Arg(y_1) < \Arg(y_2)
287: < \dots < \Arg(y_n) < 2 \pi$. Thus, the $2n$ solutions $(Z_1, \dots,
288: Z_{2n})$ of Eq.~(\ref{eq:z}) are
289: \begin{equation}
290: Z_m = (1 - y_m)^{1/2} \,; \ \ \ Z_{m+n} = - Z_m \ \
291: \hbox{ with } \,\, m = 1, \ldots, n \, .
292: \label{eq:zm}
293: \end{equation}
294: The branch cut of the function $z^{1/2}$ is, as usual, the real semi-axis
295: $(-\infty,0]$, {\it i.e.}, for $m = 1, \ldots, n,$ the argument of $Z_m$
296: belongs to $[-\pi/2,\pi/2[$. We explain in Appendix~\ref{ap:cassini} that
297: each $Z_m$ is an analytic function of $Y$ in the complex plane with a
298: branch cut along $[0,+\infty)$ and that the locus of the $Z_m$'s is a
299: remarkable curve called a Cassini oval.
300:
301: We now choose $n$ different roots $(Z_{c(j)})_{j=1,n}$ among $(Z_1, \dots,
302: Z_{2n})$, such that the choice function $c: \{1, \dots, n\} \rightarrow
303: \{1, \dots, 2n\} $ satisfies
304: \begin{equation}
305: 1 \le c(1) < \dots < c(n) \le 2n \, .
306: \end{equation}
307: There are precisely $\Omega$ such choice functions, $\Omega = (2n)! / n!^2$
308: being the size of the Markov matrix. Finally, we define
309: \begin{equation}
310: A_c(Y) = -4^n \prod_{j=1}^n \frac{Z_{c(j)} - 1}{Z_{c(j)} + 1} \, ,
311: \label{eq:ac}
312: \end{equation}
313: where $A_c$ is a function of $Y$ and of the choice function $c$. The
314: Bethe equations~(\ref{eq:behf}) are equivalent to the {\it
315: self-consistency} equation
316: \begin{equation}
317: A_c(Y) = Y \, .
318: \label{eq:ay=y}
319: \end{equation}
320: Given the choice function $c$ and a root $Y$ of this equation, the
321: $Z_{c(j)}$'s are determined from Eq.~(\ref{eq:z}) and the corresponding
322: eigenvalue $E_c$ is obtained from Eq.~(\ref{eq:eZ}).
323:
324: For small values of $n$, the above described procedure allows us to compute
325: numerical solutions of the Bethe equations. From our numerical
326: observations, we conjecture that for each choice function $c$ (among the
327: $\Omega$ possible choice functions), the self-consistency
328: Eq.~(\ref{eq:ay=y}) has a unique solution $Y$ that yields one eigenvector
329: $\psi_c$ and one eigenvalue $E_c$. This suggests that the Bethe equations
330: yield a complete basis of eigenvectors for the ASEP.
331:
332: Let us first consider the choice function $c(j) = j$, {\it i.e.}, the
333: selected $Z_j$'s are the $n$ solutions of Eq.~(\ref{eq:z}) with largest real
334: parts. As this choice plays an important role in the following analysis,
335: we define
336: \begin{eqnarray}
337: A_0(Y) &=& -4^n \prod_{j=1}^n \frac{Z_j - 1}{Z_j + 1} \, , \label{eq:a0} \\
338: 2 E_0 &=& -n + \sum_{j=1}^n Z_j \, .\label{eq:e0}
339: \end{eqnarray}
340: We emphasize that $E_0$ is an implicit function of $Y$. The equation
341: $A_0(Y) = Y$ has the solution $Y=0$ that yields $Z_j = 1$ for all $j$ and
342: provides the stationary distribution (or ground state) with eigenvalue $0$.
343:
344: Numerical observations (Gwa and Spohn 1992) indicate that the first excited
345: eigenvalue $E_1$ corresponds to the choice $c(j) = j$ for $j = 1, \ldots,
346: n-1$ and $c(n) = n+1$; {\it i.e.}, the first excited state is obtained from
347: the ground state by the {\it excitation} $(n \to n+1)$ . Writing $A_1(Y)$
348: and $E_1$ for the functions $A_c$ and $E_c$ corresponding to this choice
349: function, we have, from Eqs.~(\ref{eq:ac}, \ref{eq:a0} and \ref{eq:e0}),
350: \begin{eqnarray}
351: A_1(Y) &=& A_0(Y) \left( \frac{Z_1-1}{Z_1+1} \ \frac{Z_n-1}{Z_n+1}
352: \right)^{-1} \, ,
353: \label{eq:a1} \\
354: 2E_1 &=& 2E_0 - (Z_1 + Z_n) \, ,
355: \label{eq:e1e0}
356: \end{eqnarray}
357: where we have used $Z_{n+1} = -Z_1$. The excitation $(1 \to 2n)$, that is,
358: $c(j) = j+1$ for $j = 1, \ldots, n-1$ and $c(n) = 2n$, also leads to
359: Eqs.~(\ref{eq:a1})~and~(\ref{eq:e1e0}) and thus to the same eigenvalue
360: $E_1$. The first excited state has therefore a degeneracy of order 2.
361:
362: Consequently, in order to find the expression for the gap $E_1$, we must
363: solve the self-consistency equation
364: \begin{equation}
365: A_1(Y) = Y \, ,
366: \label{eq:scY}
367: \end{equation}
368: calculate the $Z_j$'s for $j = 1, \ldots , n, $ and finally deduce $E_1$
369: from Eq.~(\ref{eq:e1e0}).
370:
371: In the above discussion, we closely followed Gwa and Spohn (1992) to
372: present the Bethe Ansatz equations for the TASEP. We shall now solve
373: these equations and calculate the gap by a radically different and simpler
374: method.
375:
376:
377: %===============================
378: \section{Calculation of the gap}
379: %===============================
380:
381: Let us define $F(Y)$ as
382: \begin{equation}
383: A_0(Y) = Y \exp(F(Y)).
384: \label{eq:a0f}
385: \end{equation}
386: In Appendix \ref{ap:deriv}, we have derived the following identities,
387: valid for $|Y| \le 1$,
388: \begin{eqnarray}
389: F(Y) &=& \sum_{k=1}^{\infty} \frac{w_{kn}}{k} \ Y^k \, ,
390: \label{eq:fn} \\
391: -4E_0 &=& \sum_{k=1}^{\infty} \frac{w_{kn-1}}{k} \ Y^k \, ,
392: \label{eq:e0w}
393: \end{eqnarray}
394: where the $w_k$'s are given by
395: \begin{equation}
396: w_k = \frac{(2k-1)!!}{(2k)!!} = \frac{(2k)!}{(k! \ 2^k)^2} \, .
397: \label{eq:wk}
398: \end{equation}
399: From the Stirling formula, the leading order behaviour of $ w_k $ for $k
400: \to\infty$, is given by
401: \begin{equation}
402: w_k \sim \frac{1}{\sqrt{\pi k}} \, .
403: \label{eq:awk}
404: \end{equation}
405: From the power series~(\ref{eq:fn} and \ref{eq:e0w}) we deduce that
406: $A_0(Y)$ and $E_0$ are analytic functions of the complex variable $Y$
407: inside the unit circle. This property is not obvious {\it a priori}: the
408: functions $A_0(Y)$ and $E_0$, defined in Eqs.~(\ref{eq:a0}) and
409: (\ref{eq:e0}), respectively, depend implicitly on $Y$ via the $Z_m$'s
410: that involve a branch cut along $[0,+\infty)$. Indeed, for a generic
411: choice function $c(j)$, $A_c(Y)$ and $E_c$ are analytic only in the
412: complex $Y$ plane with a cut along $[0,+\infty)$ and, therefore, are not
413: analytic in the neighbourhood of $Y=0$. This special property of
414: $A_0(Y)$ and $E_0$ is obtained by an explicit calculation in Appendix
415: \ref{ap:deriv} and from geometrical considerations in Appendix
416: \ref{ap:cassini}.
417:
418: Using Eqs.~(\ref{eq:a1}),~(\ref{eq:a0f})~and~(\ref{eq:fn}), the
419: self-consistency equation~(\ref{eq:scY}), that determines the gap, reduces
420: to
421: \begin{equation}
422: \sum_{k=1}^{\infty} \frac{w_{kn}}{k} \ Y^k
423: = \ln \frac{1-Z_1}{1+Z_1} + \ln \frac{1-Z_n}{1+Z_n} \, .
424: \label{eq:sll}
425: \end{equation}
426: From Eq.~(\ref{eq:e1e0}) and Eq.~(\ref{eq:e0w}), we have
427: \begin{equation}
428: -4E_1 = \sum_{k=1}^{\infty} \frac{w_{kn-1}}{k} \ Y^k + 2Z_1 + 2Z_n \, .
429: \label{eq:e1}
430: \end{equation}
431: Combining Eq.~(\ref{eq:sll}) and Eq.~(\ref{eq:e1}), we obtain
432: \begin{equation}
433: -4E_1 = \sum_{k=1}^{\infty} \frac{w_{kn}}{k(2kn-1)} \ Y^k
434: + \left( 2Z_1 + \ln \frac{1-Z_1}{1+Z_1} \right)
435: + \left( 2Z_n + \ln \frac{1-Z_n}{1+Z_n} \right) \, ,
436: \label{eq:e1s}
437: \end{equation}
438: where we have used $(2kn)w_{kn} = (2kn-1) w_{kn-1}$.
439:
440: Thus, to find the gap for a half-filled system with $n$ particles, we
441: must solve Eq.~(\ref{eq:sll}) for $Y$ and substitute the result in
442: Eq.~(\ref{eq:e1s}). We emphasize that the power series in these
443: equations represent, inside the unit disk, analytic functions of $Y$ that
444: are defined in the whole complex plane with a cut along $[1,+\infty)$.
445:
446: We now consider the thermodynamic limit, $n \to \infty$. We obtain, at
447: leading order, from Eq.~(\ref{eq:awk})
448: \begin{eqnarray}
449: F(Y) = \sum_{k=1}^{\infty} \frac{w_{kn}}{k} \ Y^k
450: & \to & \frac{1}{\sqrt{\pi n}} \ \Li_{3/2}(Y), \\
451: \sum_{k=1}^{\infty} \frac{w_{kn}}{k(2kn-1)} \ Y^k
452: & \to & \frac{1}{2\sqrt{\pi n^3}} \ \Li_{5/2}(Y),
453: \end{eqnarray}
454: where we have used the {\em polylogarithm} function of index $s$
455: \begin{equation}
456: \Li_s(z) = \sum_{k=1}^{\infty} \frac{z^k}{k^s} \, .
457: \end{equation}
458: By virtue of the integral representation
459: \begin{equation}
460: \Li_s(z) = \frac{z}{\Gamma(s)} \ \int_0^{\infty} \frac{t^{s-1} \
461: dt}{e^t -z} \, ,
462: \label{eq:integr}
463: \end{equation}
464: the function $\Li_s$ can be extended by analytic continuation to the whole
465: complex plane with a branch cut along the real semi-axis $[1,+\infty)$.
466: In the large $n$ limit, we deduce from Eqs.~(\ref{eq:u}, \ref{eq:ym} and
467: \ref{eq:zm}) that
468: \begin{eqnarray}
469: Z_1 = (1-y_1)^{1/2} &=& \sqrt{\frac{\pi}{n}} \ (-u-i)^{1/2}
470: + O \left( \frac{1}{n^{3/2}} \right) \, , \\
471: Z_n = (1-y_n)^{1/2} &=& \sqrt{\frac{\pi}{n}} \ (-u+i)^{1/2}
472: + O \left( \frac{1}{n^{3/2}} \right) \, ,
473: \end{eqnarray}
474: where we have supposed that $ Y = -e^{u\pi}$ remains finite
475: when $n \to \infty$. Using these expressions and the expansion
476: $ \, \ln \frac{1-Z}{1+Z} = -2Z - \frac{2}{3} Z^3 + O(Z^5), $
477: Eq.~(\ref{eq:sll}) reduces to
478: \begin{equation}
479: \Li_{3/2}(-e^{u\pi}) = -2\pi \left[ (-u+i)^{1/2} + (-u-i)^{1/2} \right] ,
480: \label{eq:li32}
481: \end{equation}
482: and the gap~(\ref{eq:e1s}), at the leading order, is given by
483: \begin{equation}
484: E_1 = \frac{1}{n^{3/2}} \left\{
485: \frac{-1}{8\sqrt{\pi}} \ \Li_{5/2}(-e^{u\pi})
486: + \frac{\pi^{3/2}}{6} \ \left[ (-u+i)^{3/2} + (-u-i)^{3/2}
487: \right] \right\}.
488: \label{eq:e1li52}
489: \end{equation}
490: [Notice that the r.h.s. of Eq.~(\ref{eq:li32}) and of Eq.~(\ref{eq:e1li52})
491: are real when $u$ is real.] With the help of the Maple software, we find a
492: unique solution of Eq.~(\ref{eq:li32}) in the strip $ -1 \le \mathrm{Im}(u)
493: < 1$ that is real and is given by
494: \begin{equation}
495: u = 1.119 \, 068 \, 802 \, 804 \, 474 \dots
496: \label{eq:solu}
497: \end{equation}
498: Inserting this value of $u$ in Eq.~(\ref{eq:e1li52}) yields the large $n$
499: (or large $L$) behaviour of the gap
500: \begin{equation}
501: E_1 = - \frac{2.301 \, 345 \, 960 \, 455 \, 050 \dots}{n^{3/2}}
502: = - \frac{6.509 \, 189 \, 337 \, 976 \, 136 \dots}{L^{3/2}}.
503: %-6.509189337976136493323714
504: \label{eq:solgap}
505: \end{equation}
506: This is precisely the result obtained by Gwa and Spohn (1992). This gap
507: scales as $L^{-3/2}$ and is real for the TASEP in the half-filling case.
508:
509:
510: %===============================
511: \section{Summary and discussion}
512: %===============================
513:
514: In this work, we have calculated the gap of the TASEP in the limit of a
515: large size system by using the Bethe Ansatz. We first take the large~$n$
516: limit of the Bethe equations inside the unit circle, then perform the
517: analytic continuation of these equations in the whole complex plane with
518: a branch cut along $[1,+\infty)$ and finally solve them. Gwa and Spohn
519: (1992), on the contrary, first represent the analytic continuation of the
520: Bethe equations for a fixed value of $n$ as a $n$-dependent complex
521: integral (thanks to the Euler-Maclaurin formula) and then extract the
522: gap from the large $n$ limit of this integral representation which is
523: rather a delicate operation. We have shown here that the derivation of the
524: TASEP gap is greatly simplified by performing the calculations in the
525: reverse order, that is taking the large~$n$ limit first and the analytic
526: continuation afterwards.
527:
528: We do not claim that it is always true that large $n$ limit and
529: analytic continuation are commuting operations. If the solution $Y$ of
530: Eq.~(\ref{eq:scY}), in the large $n$ limit, diverges to $\infty$ or
531: approaches asymptotically the branch cut, reversing the order of
532: operations may not be possible. Fortunately, in our problem, $Y$ remains
533: a bounded negative real number when $n \to \infty$.
534:
535: We have applied our method to several other problems but, for sake of
536: conciseness, we simply list these additional calculations without giving
537: the details. Our technique provides the subleading corrections to the
538: gap and allows us to calculate the eigenvalue of the highest excited
539: state, of a finite excitation above the ground state, or below the
540: highest excited state. The gap for the totally asymmetric exclusion
541: process with an arbitrary density $\rho$ can also be obtained by
542: elementary means, and we find in agreement with (Kim 1995)
543: \begin{equation}
544: E_1\left( \rho \right) =
545: 2 \sqrt{ {\rho( 1 -\rho)} } \,\, E_1 \left(\rho = {1}/{2}\right)
546: \pm \frac{2i\pi}{L}(2\rho -1)
547: \, . \nonumber
548: \end{equation}
549: where $ E_1 (\rho = {1}/{2})$ is the gap for the half-filling case given
550: in Eq.~(\ref{eq:solgap}); we notice that for a density other than
551: one-half the gap has a non-zero imaginary part. We have also studied
552: generalizations of the ASEP by introducing a tagged particle that has the
553: same dynamics as the other particles: the gap then scales as $L^{-5/2}$.
554:
555:
556:
557: Finally, we emphasize that the formula~(\ref{eq:integr}) already appeared
558: in the work of Derrida and Appert (1999): indeed, Eqs.~(\ref{eq:sll}) and
559: (\ref{eq:e1}) are similar to those used in their calculation of large
560: deviation functions of the ASEP by Bethe Ansatz.
561:
562:
563: \subsection*{Acknowledgments}
564: %============================
565:
566: It is a pleasure to thank C. Appert, B. Derrida, C. Godr\`eche, J.-M. Luck
567: and S. Mallick for inspiring discussions and remarks about the manuscript.
568:
569: \appendix
570:
571: \section{Derivation of Eq.~(\ref{eq:fn}) and Eq.~(\ref{eq:e0w})}
572: %===============================================================
573: \label{ap:deriv}
574:
575: The numbers $w_k$ defined in Eq.~(\ref{eq:wk}) are the coefficients of the
576: Taylor series
577: \begin{equation}
578: \frac{1}{\sqrt{1-x}} = \sum_{k=0}^{\infty} w_k x^k \, .
579: \end{equation}
580: By integration, we find
581: \begin{equation}
582: \sqrt{1-x} = 1 - \frac{1}{2} \ \sum_{k=1}^{\infty} \frac{w_{k-1}}{k} x^k
583: \, .
584: \end{equation}
585: Recalling that $Z_m = \sqrt{1-y_m}$ [see
586: Eqs.~(\ref{eq:ym})~and~(\ref{eq:zm})], we obtain
587: \begin{equation}
588: \sum_{m=1}^n Z_m = n - \frac{1}{2} \ \sum_{k=1}^{\infty}
589: \frac{w_{k-1}}{k} \sum_{m=1}^n y_m^k \, .
590: \label{eq:sumZ}
591: \end{equation}
592: The fact that the $y_m$'s are the $n$-th roots of $Y$ leads to the
593: following relation
594: \begin{equation}
595: \sum_{m=1}^n y_m^k = \left\{ \begin{array}{ll}
596: nY^{k/n} & \mbox{if $k$ is a multiple of $n$} \\
597: 0 & \mbox{otherwise.}
598: \end{array} \right.
599: \label{eq:ymk}
600: \end{equation}
601: Inserting this relation in Eq.~(\ref{eq:sumZ}), we obtain
602: \begin{equation}
603: \sum_{m=1}^n Z_m = n - \frac{1}{2} \ \sum_{k=1}^{\infty}
604: \frac{w_{nk-1}}{k} Y^k \, .
605: \label{eq:sumZinY}
606: \end{equation}
607: We thus find, thanks to the crucial identity~(\ref{eq:ymk}), that
608: $\sum_{m=1}^n Z_m$ is analytic in $Y$ inside the unit circle. Finally,
609: substituting Eq.~(\ref{eq:sumZinY}) in Eq.~(\ref{eq:e0}), we obtain
610: Eq.~(\ref{eq:e0w}).
611:
612: The derivation of Eq.~(\ref{eq:fn}) follows similar steps. We first notice
613: that the Taylor expansion of the function
614: \begin{equation}
615: f(x) = \ln\left( \frac{4}{x}\ \frac{1-\sqrt{1-x}}{1+\sqrt{1-x}} \right)
616: \, \label{eq:deff}
617: \end{equation}
618: is given by
619: \begin{equation}
620: f(x) = \sum_{k=1}^{\infty} \frac{w_k}{k} x^k \, . \label{eq:taylorf}
621: \end{equation}
622: [This follows from $f(0) = 0$ and
623: $ f'(x) = \frac{1}{x} \left(\frac{1}{\sqrt{1-x}} - 1 \right)\,].$
624: Using Eq.~(\ref{eq:a0}) and the identity $\prod_{m=1}^n (-y_m) = -Y, $
625: we deduce that
626: \begin{equation}
627: A_0(Y) = Y \prod_{m=1}^n \frac{4}{y_m}\ \frac{1-Z_m}{1+Z_m} \, .
628: \label{eq:prodA}
629: \end{equation}
630: From Eq.~(\ref{eq:a0f}) and Eqs.~(\ref{eq:deff}),~(\ref{eq:taylorf}) and
631: ~(\ref{eq:prodA}), we obtain
632: \begin{equation}
633: F(Y) = \sum_{m=1}^n f(y_m) = \sum_{k=1}^{\infty} \frac{w_k}{k} \sum_{m=1}^n
634: y_m^k \, .
635: \end{equation}
636: This equation reduces to Eq.~(\ref{eq:fn}) by virtue of the
637: identity~(\ref{eq:ymk}).
638:
639: \section{Roots of the Bethe Equations and Cassini ovals}
640: %======================================================
641:
642: \label{ap:cassini}
643:
644: The polynomial equation
645: \begin{equation}
646: (1-Z^2)^n = Y \,
647: \end{equation}
648: where $Y$ is a fixed complex number, has $2n$ solutions, $(Z_1, \dots,
649: Z_{2n})$. The purpose of this appendix is to explain how these solutions
650: can be labeled in a coherent way so that each root $Z_m(Y)$ is an analytic
651: function of $Y$.
652:
653: We first notice that $y_m$, defined in Eq.~(\ref{eq:ym}), is an analytic
654: function of $Y$ in the complex plane with a branch cut along the real
655: semi-axis $[0,+\infty)$. Nevertheless when $Y$ crosses $[0,+\infty)$, the
656: functions $y_1, y_2 \dots, y_n$ are the analytic continuations (above the
657: axis) of respectively $y_n, y_1, \dots, y_{n-1}$: thus the existence
658: of the branch cut along $[0,+\infty)$ is due to the labeling of the roots.
659:
660:
661: The complex numbers $(Z_1, \dots, Z_{2n})$ belong to the curve defined by
662: \begin{equation}
663: \left|Z-1 \right| \, \left| Z+1\right| = r \,\,\,
664: \hbox{ with } \,\,\, r = |Y|^{1/n} \,
665: \label{eq:cassini}
666: \end{equation}
667: and called a {\em Cassini oval}. A Cassini oval is the conformal
668: transformation of the circle of center 1 and radius $r$ by the function
669: $z \rightarrow z^{1/2}$. Its shape depends on whether the point 0 is
670: inside or outside the circle, {\it i.e.}, whether $r<1$ or not. When
671: $r<1$, the curve of equation~(\ref{eq:cassini}) consists of two ovals
672: around the points $Z=\pm 1$. The numbers $(Z_1,\dots,Z_n)$ lie on the
673: right oval and $(Z_{n+1},\dots, Z_{2n})$ on the left oval. For the
674: marginal case, $r=1$, the curve is the {\em lemniscate of Bernoulli},
675: with a multiple point at $Z=0$. When $r>1$, the Cassini oval is a single
676: loop with a peanut shape (when $r \in ]1,2[$) or an oval shape ($r \ge
677: 2$). See Fig.~\ref{fig} where the cases $r = 0.9, 1$ and $1.1$ are
678: drawn. In the large-$r$ limit, the oval tends to the circle of radius
679: $\sqrt{r}$.
680:
681: \begin{figure}
682: \centering
683: \includegraphics[width=\figwidth, keepaspectratio]{fig.eps}
684: \caption{\em
685: Labeling the roots $Z_m$ of the equation $(1-Z^2)^n=Y$.
686: Here $Y = e^{i\phi} r^n$ with $n=10$, $\phi = \pi/2$, and $r \in
687: \{0.9, 1, 1.1\}$. The continuous curves are the corresponding Cassini
688: ovals (see text). When $r$ is fixed and $\phi$ goes from 0 to $2\pi$,
689: each $Z_m$ slips counter-clockwise along the Cassini ovals. Then, the
690: jump $\phi = 2\pi \to 0$, {\rm i.e.}, $Y$ crosses $[0,+\infty)$,
691: consists of a global shift of the labels $m$ around each continuous
692: curve.
693: }
694: \label{fig}
695: \end{figure}
696:
697: We now discuss the analyticity properties of $Z_m(Y)$: $Z_m$ is an
698: analytic function of $y_m$ with a branch cut along $(-\infty,0]$; this
699: branch cut is compatible with that of $y_m(Y)$. Consequently $Z_m$ is
700: an analytic function of $Y$ with a branch cut along $[0,+\infty)$.
701: Moreover when $Y$ crosses the real segment $[0,1[$, the functions $Z_1(Y),
702: Z_2(Y) \dots, Z_n(Y)$ are the analytic continuations (above the axis) of
703: the functions $Z_n(Y), Z_1(Y), \dots, Z_{n-1}(Y)$ respectively. See
704: Fig.~\ref{fig}. (A similar property is true for the functions
705: $Z_{n+1}(Y), Z_{n+2}(Y) \dots, Z_{2n}(Y)$ that belong to the left oval).
706: But when $Y$ crosses $]1,+\infty)$, the functions $Z_1(Y), Z_2(Y), \dots,
707: Z_{2n}(Y)$ are the analytic continuations (above the axis) of the
708: functions $Z_{2n}(Y), Z_1(Y), \dots, Z_{2n-1}(Y)$ respectively.
709: Consequently the branch cut of the function $A_0(Y)$ defined in
710: Eq.~(\ref{eq:a0}) is $[1, +\infty)$ and not $[0, +\infty)$.
711:
712:
713: \section*{References}
714: %=========================
715: \begin{itemize}
716:
717: \item
718: F.~C.~Alcaraz, M.~Droz, M.~Henkel, V. Rittenberg, 1994,
719: {\em Reaction-diffusion processes, critical dynamics and quantum
720: chains},
721: Ann. Phys. {\bf 230}, 250.
722:
723: \item
724: F.~C.~Alcaraz and M.~J.~Lazo, 2003,
725: {\em The Bethe Ansatz as a matrix product Ansatz},
726: preprint cond-mat/0304170.
727:
728: \item
729: R. Bundschuh, 2002,
730: {\em Asymmetric exclusion process and extremal statistics of
731: random sequences},
732: Phys. Rev. E {\bf 65}, 031911.
733:
734: \item
735: B. Derrida, 1998,
736: {\em An exactly soluble non-equilibrium system: the asymmetric simple
737: exclusion process},
738: Phys. Rep. {\bf 301}, 65.
739:
740: \item
741: B. Derrida, C. Appert, 1999,
742: {\em Universal large-deviation function of the Kardar-Parisi-Zhang
743: equation in one dimension},
744: J. Stat. Phys. {\bf 94}, 1.
745:
746: \item
747: B. Derrida, M. R. Evans, 1999,
748: {\em Bethe Ansatz solution for a defect particle in
749: the asymmetric exclusion process},
750: J. Phys. A: Math. Gen. {\bf 32}, 4833.
751:
752: \item
753: B. Derrida, M.~R.~Evans, V. Hakim, V. Pasquier, 1993,
754: {\em Exact solution of a 1D asymmetric exclusion
755: model using a matrix formulation},
756: J. Phys. A: Math. Gen. {\bf 26}, 1493.
757:
758: \item
759: B. Derrida, J. L. Lebowitz, 1998,
760: {\em Exact large deviation function in the asymmetric exclusion process},
761: Phys. Rev. Lett. {\bf 80}, 209.
762:
763: \item
764: B.~Derrida, J.~L.~Lebowitz, E.~R.~Speer, 2003,
765: {\em Exact large deviation functional of a
766: stationary open driven diffusive system: the asymmetric exclusion process},
767: J. Stat. Phys. {\bf 110}, 775.
768:
769: \item
770: D. Dhar, 1987,
771: {\em An exactly solved model for interfacial growth},
772: Phase Transitions {\bf 9}, 51.
773:
774: \item
775: M.~R.~Evans, D.~P.~Foster, C.~Godr\`eche, D.~Mukamel, 1995,
776: {\em Asymmetric exclusion model with two species: spontaneous
777: symmetry breaking},
778: J. Stat. Phys. {\bf 80}, 69.
779:
780: \item
781: M.~R.~Evans, Y.~Kafri, H.~M.~Koduvely, D.~Mukamel, 1998,
782: {\em Phase separation in one-dimensional driven diffusive systems},
783: Phys. Rev. Lett. {\bf 80}, 425.
784:
785: \item
786: L.-H. Gwa, H. Spohn, 1992,
787: {\em Bethe solution for the dynamical-scaling exponent of the noisy
788: Burgers equation},
789: Phys. Rev. A {\bf 46}, 844.
790:
791: \item
792: T. Halpin-Healy, Y.-C.~Zhang, 1995,
793: {\em Kinetic roughening phenomena, stochastic growth, directed polymers and
794: all that},
795: Phys. Rep. {\bf 254}, 215.
796:
797: \item
798: D. Kim, 1995,
799: {\em Bethe Ansatz solution for crossover scaling functions
800: of the asymmetric XXZ chain and the Kardar-Parisi-Zhang-type
801: growth model},
802: Phys. Rev. E {\bf 52}, 3512.
803:
804: \item
805: J. Krug, 1991,
806: {\em Boundary-induced phase transitions in driven diffusive systems},
807: Phys. Rev. Lett. {\bf 67}, 1882.
808:
809: \item
810: J. Krug, 1997,
811: {\em Origins of scale invariance in growth processes},
812: Adv. Phys. {\bf 46}, 139.
813:
814: \item
815: D. G. Levitt, 1973,
816: {\em Dynamics of a single-file pore: Non-Fickian behavior},
817: Phys. Rev. A {\bf 8}, 3050.
818:
819: \item
820: C.~T.~MacDonald, J.~H.~Gibbs, 1969,
821: {\em Concerning the kinetics of poly\-peptide synthesis
822: on polyribosomes},
823: Biopolymers {\bf 7}, 707.
824:
825: \item
826: P. M. Richards, 1977,
827: {\em Theory of one-dimensional hopping conductivity and diffusion},
828: Phys. Rev. B {\bf 16}, 1393.
829:
830: \item
831: M.~Schreckenberg, D.~E.~Wolf (ed.), 1998,
832: {\em Traffic and granular flow '97}
833: (Berlin: Springer-Verlag).
834:
835: \item
836: G.~Sch\"utz, 1993,
837: {\em Generalized Bethe Ansatz solution of a one-dimen\-sional
838: asymmetric exclusion process on a ring with blockage},
839: J. Stat. Phys. {\bf 71}, 471.
840:
841: \item
842: E.~R. ~Speer, 1993,
843: {\em The two species totally asymmetric exclusion process},
844: in Micro, Meso and Macroscopic approaches in Physics, M.~Fannes
845: C. Maes and A. Verbeure Ed. NATO Workshop 'On three levels',
846: Leuven, July 1993.
847:
848: \item
849: H. Spohn, 1991,
850: {\em Large scale dynamics of interacting particles},
851: (New-York: Springer-Verlag).
852:
853: \item
854: R.~B.~Stinchcombe and G.~M.~Sch\"utz, 1995,
855: {\em Application of operator algebras to stochastic dynamics
856: and the Heisenberg chain},
857: Phys. Rev. Lett. {\bf 75}, 140.
858:
859: \item
860: B.~Widom, J.~L.~Viovy, A.~D.~Defontaines, 1991,
861: {\em Repton model of gel electrophoresis and diffusion},
862: J. Phys. I France {\bf 1}, 1759.
863:
864: \end{itemize}
865:
866: \end{document}
867: %%%%%%%%%%%%%%%%%%%%%%% End of LaTeX file %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
868: