cond-mat0404358/dlg.tex
1: \documentclass[12pt,epsf,amstex]{article}
2: \usepackage [dvips]{graphicx}
3: \usepackage{amsmath}
4: \usepackage{times}
5: \usepackage{psfig}
6: \usepackage{epsfig}
7: \usepackage{amssymb}
8: %\usepackage{theorem}
9: %\renewcommand{\baselinestretch}{2}
10: \newcommand{\bE}{\text{\bf E}}
11: \newcommand{\bA}{\text{\bf A}}
12: \newcommand{\bB}{\text{\bf B}}
13: \newcommand{\bP}{\text{\bf P}}
14: \newcommand{\bj}{\text{\bf j}}
15: \newcommand{\dd}{\text{d}}
16: \newcommand{\sR}{\text{\tiny R}}
17: \newcommand{\ee}{\text{e}}
18: \newcommand{\ii}{\text{i}}
19: \newcommand{\thy}{\text{th}}
20: \newcommand{\ji}{\text{\small i}}
21: \newcommand{\p}{\partial}
22: \newcommand{\bx}{\text{\bf x}}
23: \newcommand{\br}{\text{\bf r}}
24: \newcommand{\by}{\text{\bf y}}
25: \newcommand{\be}{\text{\bf e}}
26: \newcommand{\eps}{\varepsilon}
27: \newcommand{\bk}{\text{\bf k}}
28: \newcommand{\bo}{\text{\bf 0}}
29: \newcommand{\bq}{\text{\bf q}}
30: \newcommand{\bp}{\text{\bf p}}
31: \newcommand{\bu}{\text{\bf u}}
32: \newcommand{\bGamma}{\boldmath$\Gamma$\unboldmath}
33: \newcommand{\bF}{\text{\bf F}}
34: \begin{document}
35: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
36: \begin{center}
37: {\Large{\bf Large deviations in weakly interacting\\
38: \vspace{8pt}
39: boundary driven lattice gases}}
40: \end{center}
41: 
42: \vspace{2cm}
43: \begin{center}
44: {Fr\'ed\'eric van Wijland$^{1,2}$ and Zolt\'an R\'acz$^{1,3}$}
45: \end{center}
46: 
47: \noindent ${}^1$Laboratoire de Physique Th\'eorique, B\^at. 210,
48: Universit\'e de Paris-Sud,\\ 91405 Orsay Cedex, France.\\
49: 
50: \noindent ${}^2${P\^ole Mati\`ere et Syst\`emes Complexes, Universit\'e de Paris VII, France.\\
51: 
52: \noindent ${}^3$Institute for Theoretical Physics, HAS Research Group,
53: E\"otv\"os University \\1117 Budapest, P\'azm\'any s\'et\'any 1/a, Hungary.\\
54: }\\\\
55: 
56: \begin{center}{\bf Abstract}\\
57: \end{center}
58: One-dimensional, boundary-driven lattice gases with local interactions
59: are studied in the weakly interacting limit. The density profiles
60: and the correlation functions are calculated to first order in the interaction
61: strength for zero-range and short-range processes differing only
62: in the specifics of the detailed-balance dynamics. Furthermore,
63: the effective free-energy (large-deviation function) and the integrated
64: current distribution are also found to this order. From the former, we find
65: that the boundary drive generates long-range correlations only for
66: the short-range dynamics while the latter provides support to an
67: additivity principle recently proposed by Bodineau and Derrida.
68: 
69: \vskip 2cm
70: %Pr\'epublication L.P.T. Orsay 04/??.
71: \newpage
72: 
73: \section{Introduction}
74: 
75: One-dimensional boundary driven lattice gases are simple model systems which
76: allow detailed studies of nonequilibrium steady-states.
77: They are relevant in the sense that steady states in experiments
78: are often produced by choosing appropriate boundary conditions
79: (heating a horizontal layer of liquid from below being the most widely quoted
80: example) and, furthermore, these models appear to capture some
81: important consequences of the nonequilibrium drive such as e.g. the
82: generation of long-range correlations.
83: 
84: 
85: Analytical approaches to driven lattice gases proceed along two distinct paths.
86: On one hand, exploiting the matrix product method, Derrida and
87: coworkers \cite{{derridalebowitzspeer1},{derridalebowitz},{derridalebowitzspeer2},
88: {derridalebowitzspeer3},{derridaenaud},{derridadoucotroche}} have come up with a
89: series of exact results for the exclusion process
90: (hard-core particles undergoing symmetric or
91: biased diffusive motion) in one dimension .
92: On the other hand, in a parallel series of papers,
93: Bertini and coworkers
94: \cite{{bertinisolegabriellijonalasinioladim1},{bertinisolegabriellijonalasinioladim2},
95: {bertinisolegabriellijonalasinioladim3}}
96: solved some of the same problems employing
97: the formalism of fluctuating hydrodynamics.
98: The results obtained for the probability functional of a given
99: density profile or for the integrated current distribution are
100: important since they provide
101: new insights into the properties of nonequilibrium steady states.
102: Thus it would be highly desirable to establish how general or universal
103: these results are.
104: In order to investigate this generality issue, in this paper,
105: we go beyond the specificity of the exclusion process, and study the effects of
106: varying the interactions between the particles as well as varying the microscopic
107: dynamics in boundary driven lattice gases.
108: 
109: Our main "technical" results are the derivation of
110: the weak interaction limits of both the nonequilibrium free energy and
111: the distribution of the integrated particle current. From the
112: explicit form of the free energy, we can then deduce the density profile and the
113: correlations (effective interactions) for various interactions (pair or triplet)
114: and various dynamics [zero-range~\cite{zrp-origin} and short-range
115: (misanthropic~\cite{Cocozza}) processes].
116: 
117: Our main "physical"  findings are that the density profiles are nonlinear and
118: they depend on both the interactions and the dynamics. Furthermore, the details
119: of the dynamics are found to change the
120: the correlations from short- to long-range, and they are shown to be able to
121: change the sign of the effective interactions. As to the
122: distributions of the integrated particle current, we verify that, for
123: all the interaction- and dynamics combinations we studied, they
124: are in agreement with the recently proposed addititvity principle which
125: allows to construct this distribution function from the knowledge of
126: its first two moments \cite{bodineauderrida}.
127: 
128: We start by introducing the models of interacting lattice gases and
129: deriving dynamical rules satisfying detailed balance in the absence of
130: boundary drive (Sec.\ref{Model}). Next we describe on the example of
131: free particles how the models can be formulated in terms of field
132: theories (Sec.\ref{Free}). Zero-range and short-range (misanthropic) processes
133: are worked out in Sections \ref{ZRP} and \ref{SRP}, respectively.
134: There we provide explicit expressions for the effective free energies.
135: Finally, Sec.\ref{JJJ} is devoted to a separate treatment
136: of the integrated current distribution.
137: 
138: 
139: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
140: \section{The model}\label{Model}
141: \subsection{Lattice gas with onsite interaction and boundary drive}
142: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
143: A $d=1$ dimensional lattice gas is considered with hopping dynamics
144: in the bulk and with particle injection and removal at the ends of the chain.
145: The state of the system $\vec n\equiv \{n_0,n_1,...,n_L\}$ is
146: specified by the number of particles $n_i=0,1,...,\infty$
147: at lattice sites $i$ ($i=0,1,...,L$), and
148: the interactions are assumed to be local
149: \begin{equation}
150: E(\vec n)=\sum_{i=0}^{L}h_i
151: \end{equation}
152: where $h_i=h(n_i)$ is the energy of interactions among particles at the same site.
153: We shall treat the simplest case
154: of pair interactions, $h(n)=\eps n (n-1)$. However triplet interactions,
155: $h(n)=\eps n (n-1)(n-2)$,
156: will also be considered occasionally either for demonstrating the effects of competing interactions
157: \begin{equation}\label{compete}
158: h_i=h(n_i)=\eps \left(\pm n_i(n_i-1)+
159:  \lambda n_i(n_i-1)(n_i-2) \right)\, .
160: \end{equation}
161: or for probing the universality of our results with respect to varying the type
162: of interaction. For $\lambda=0$, we restrict our study to the + sign in the
163: r.h.s. of (\ref{compete}) in order to have stability against
164: collapse of all particles on a single site, while for $\lambda > 0$ both signs
165: will be considered.
166: 
167: The dynamics of the system is described in terms
168: of a master equation for the
169: time evolution of the probability $P(\vec n,t)$ of a given particle configuration
170: \begin{equation}
171: \partial_t P(\vec n,t)={\cal L}_{D}P(\vec n,t)+{\cal L}_{BC}P(\vec n,t) \,
172: \end{equation}
173: where the first and second terms on the right-hand side represent the
174: nearest-neighbor hopping in the bulk and the injection-removal terms at the
175: boundaries. The bulk term is given through the rate of hopping $w_{i\rightarrow i\pm1}(\vec n)$
176: from site $i$ to $i\pm1$ in a state $\vec n$ as
177: \begin{eqnarray}
178: &&\hspace{-1.2cm}{\cal L}_{D}P(\vec n,t)=-\sum_{i=0}^{L-1}w_{i\rightarrow i+1}(\vec n)P(\vec n,t)-
179: 			\sum_{i=1}^{L}w_{i\rightarrow i-1}(\vec n)P(\vec n,t)	       \\
180: &&+\sum_{i=0}^{L-1}w_{i\rightarrow i+1}(\vec n_{i+1\rightarrow i})P(\vec n_{i+1\rightarrow i},t)
181: +\sum_{i=1}^{L}w_{i\rightarrow i-1}(\vec n_{i-1\rightarrow i})P(\vec n_{i-1\rightarrow i},t)
182: \nonumber
183: \end{eqnarray}
184: where state $\vec n_{i\rightarrow j}$ is obtained from $\vec n$ by moving a particle from $i$ to $j$.
185: The injection and removal of particles at the boundaries $i=0$ and $i=L$
186: are described by the terms
187: \begin{eqnarray}
188: {\cal L}_{BC}P(\vec n,t)=&&-\left [\,w_0^+(\vec n)+w_L^+(\vec n)+
189: 		       w_0^-(\vec n)+w_L^-(\vec n)\,\right ]\,P(\vec n,t)
190: 			\\
191: &&+\,w_0^+(\vec {n}_{0^-})\,P(\vec n_{0^-},t)+w_0^-(\vec n_{0^+})\,P(\vec n_{0^+},t)
192: 			\nonumber \\
193: &&+\,w_L^+(\vec n_{L^-})\,P(\vec n_{L^-},t)+w_L^-(\vec n_{L^+})\,P(\vec n_{L^+},t)
194: \nonumber
195: \end{eqnarray}
196: where $w_0^+(\vec n)$, $w_0^-(\vec n)$, $w_L^+(\vec n)$ and $w_L^-(\vec n)$ are the rates
197: of adding ($+$) or removing ($-$) a particle at site $0$ or $L$ in the state $\vec n$
198: and, furthermore, the state $\vec n_{i+(-)}$ is obtained from $\vec n$ by adding (removing)
199: a particle at site $i$.
200: 
201: \subsection{Choice of dynamics}
202: \label{choiceofdyn}
203: \subsubsection{Equilibrium distribution}
204: 
205: In order to specify the dynamics, we shall assume that if the
206: injection and removal rates are such that there is not
207: net flux (particle or energy current) through the system then
208: equilibrium is reached. Due to the onsite nature of the interaction, the
209: equilibrium distribution factorizes, it becomes the grand-canonical distribution
210: \begin{equation}
211: P_{\text{eq}}(\vec n)=\prod_i p_{\text{eq}}(n_i),\quad
212: p_{\text{eq}}(n_i)=\frac{1}{\Xi}\frac{\zeta^{n_i}\ee^{-h_i/T}}{n_i!}
213: \label{Peq}
214: \end{equation}
215: where $T$ and $\zeta$ are characterizing the temperature and fugacity
216: of the outside heat and particle bath, and $\Xi$ is a normalization factor. In the following,
217: we shall absorb $T$ in the interaction strength $\eps$ and the
218: high temperature limit studied below will simply mean that  $\eps$ is small.
219: 
220: \subsubsection{Hopping rates}
221: 
222: The bulk hopping rates may now be defined by making the natural (though not
223: necessary) assumption that they satisfy
224: detailed balance with $P_{\text{eq}}$, and that the detailed balance remains valid
225: in the presence of the boundary drive, as well. This assumption yields the following condition
226: for the rate of hopping to the right
227: \begin{eqnarray}
228: \frac{w_{i\rightarrow i+1}(\vec n)}{w_{i+1\rightarrow i}(\vec n_{i\rightarrow i+1})}&=&
229: \frac{n_i\ee^{-[E(\vec n_{i\rightarrow i+1})-E{(\vec n)]/2}}}
230: {(n_{i+1}+1)\ee^{-[E{(\vec n)}-E{(\vec n_{i\rightarrow i+1})]/2}}}\label{sr-rate}\\
231: &=&\frac{n_i\ee^{h(n_i)-h(n_i-1)}}{(n_{i+1}+1)\ee^{h(n_{i+1}+1)-h(n_{i+1})}}\label{zr-rate}
232: \, ,
233: \end{eqnarray}
234: and similar condition with $i+1$ replaced by $i-1$ applies for the rate of hopping to the left.
235: Clearly, the hopping rates are not uniquely determined by detailed balance.
236: The remaining arbitrariness is not relevant for equilibrium but,
237: in the presence of a drive,
238: the steady state properties do depend on details of the
239: dynamics. In order to
240: have some idea about the importance of various choices, we shall consider two
241: significantly different rates.
242: 
243: First, the rates will be chosen so that they
244: depend only on the energy change at the site from which the hopping originates. As can be
245: seen from (\ref{zr-rate}), a choice satisfying this condition is as follows
246: \begin{equation}
247: w_{i\rightarrow i+1}^{\rm (zr)}(\vec n)=Dn_i\ee^{h(n_i)-h(n_i-1)}\,
248: \label{zr-diff-rate}
249: \end{equation}
250: where $D$ is the "diffusion coefficient" setting the timescale.
251: Dynamics defined by the above rate is an example of the so called
252: {\em zero range processes} \cite{{zrp-origin},{zrp-review}}
253: where, in general, the rate is an arbitrary function of $n_i$. Thus, the first half
254: of our study can be viewed as an investigation of
255: zero range processes in the presence of a boundary drive.
256: 
257: For a second choice, a more standard hopping dynamics will be used with rates
258: depending not only on the energy change at the starting position of the hopping particle
259: but on the total energy change due to the hopping.
260: This choice follows from the eq.(\ref{sr-rate}) version of detailed balance
261: \begin{equation}
262: w_{i\rightarrow i+1}^{\rm (sr)}(\vec n)=Dn_i\ee^{-[h(n_i-1)+h(n_{i+1})-h(n_i)-h(n_{i+1})]/2}\,
263: \label{sr-diff-rate}
264: \end{equation}
265: where a subscript in $w^{\rm (sr)}$ signifies the name
266: {\it short range process} \cite{Cocozza}
267: we shall use in order to distinguish the resulting dynamics
268: from the zero range process. An important feature of the short range process is that
269: tuning the interactions $(0\le \eps < \infty, \lambda=0)$ so that one extrapolates
270: between the exactly solvable noninteracting limit $(\eps =0)$ and the much investigated
271: and also exactly solvable \cite{derridalebowitzspeer1,bertinisolegabriellijonalasinioladim3}
272: boundary driven symmetric exclusion process $(\eps \to \infty )$.
273: 
274: \subsubsection{Boundary drive (injection and removal rates)}
275: Nonequilibrium drive is introduced into the model through injection and removal of particles
276: at the ends ($j=0$ and $L$) of the chain. The rates can be set again by referring to
277: detailed balance condition with the baths attached to the ends
278: of the chain. The baths are assumed to contain noninteracting particles with the chemical
279: potentials set to produce the appropriate injection and removal rates.
280: 
281: For zero range
282: process, the removal can be imagined as diffusion into the baths. The rates are
283: independent of the properties of the baths and given by (\ref{zr-diff-rate}) with
284: $D$ replaced by new (externally controlled) parameters $\beta$ and $\gamma$, yielding
285: \begin{equation}
286: w^{(\rm zr)-}_0(\vec n)=\gamma n_0\ee^{h(n_0)-h(n_0-1)} \quad ; \quad
287: w^{(\rm zr)-}_L(\vec n)=\beta n_L\ee^{h(n_L)-h(n_L-1)} \, .
288: \label{zr-removal-rate}
289: \end{equation}
290: Within the picture of zero range process, the injections
291: (moving a particle from the bath to the end sites)
292: are independent of state of the end sites, and thus they are constants $\alpha$ and $\delta$
293: determined by the properties of the baths
294: \begin{equation}
295: w^{(\rm zr)+}_0(\vec n)=\alpha \quad ; \quad
296: w^{(\rm zr)+}_L(\vec n)=\delta \, .
297: \label{zr-injection-rate}
298: \end{equation}
299: 
300: The boundary rates for the short range process are determined along similar lines.
301: The rate of removal follows from (\ref{sr-rate}) and, since the particles in the
302: baths are not interacting, they are given by
303: \begin{equation}
304: w^{(\rm sr)-}_0(\vec n)=\gamma n_0\ee^{[h(n_0)-h(n_0-1)]/2} \quad ; \quad
305: w^{(\rm sr)-}_L(\vec n)=\beta n_L\ee^{[h(n_L)-h(n_L-1)]/2} \, .
306: \label{sr-removal-rate}
307: \end{equation}
308: In order to satisfy the detailed balance (\ref{sr-rate}), the injection rates must also depend
309: on the energy change due to the adding a particle and, using again that the particles
310: in the bath do not interact, we find
311: \begin{equation}
312: w^{(\rm sr)+}_0(\vec n)=\alpha \ee^{[h(n_0)-h(n_0+1)]/2} \quad ; \quad
313: w^{(\rm sr)+}_L(\vec n)=\delta \ee^{[h(n_L)-h(n_L+1)]/2} \, .
314: \label{sr-injection-rate}
315: \end{equation}
316: 
317: We have now finished the description of the two models we shall be
318: concerned with in this paper. Eqs.
319: (\ref{zr-diff-rate},\ref{zr-removal-rate},\ref{zr-injection-rate})
320: define the boundary driven zero range process while the rates given by eqs.
321: (\ref{sr-diff-rate},\ref{sr-removal-rate},\ref{sr-injection-rate})
322: define the corresponding short range process. As one can easily verify,
323: the case of
324: \begin{equation}
325: \frac{\alpha}{\gamma}=\frac{\delta}{\beta}
326: \end{equation}
327: corresponds to equilibrium with the baths (no drive) and the steady state
328: distribution is the equilibrium distribution given by eq.(\ref{Peq}).
329: Nonequilibrium drive is present when the above equality is violated.
330: 
331: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
332: \section{Preparing for the perturbation expansion: Noninteracting particles}
333: \label{Free}
334: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
335: \subsection{From the master equation to path integrals}
336: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
337: We shall now perform the mapping of the master equation for the probability
338: that the system is in
339: the occupation numbers configuration $\vec{n}$ at time $t$, namely,
340: $P(\vec{n};t)$, to a
341: field theory, where approximation techniques inherited from our experience
342: with equilibrium systems can be transferred without any difficulty.
343: 
344: We only briefly sketch the procedure (see \cite{mattisglasser} for a recent
345: review). We build up the state vector
346: $|\Psi(t)\rangle=\sum_{\vec{n}}P(\vec{n},t)|\vec{n}\rangle$ and rewrite the
347: master equation for $P$
348: in terms of and evolution equation for $|\Psi(t)\rangle$.
349: This equation can be cast in the form
350: \begin{equation}
351: \frac{\dd|\Psi(t)\rangle}{\dd t}=\hat{\cal L}_0|\Psi(t)\rangle
352: \end{equation}
353: where $\hat{\cal L}_0$ is an evolution operator acting on the
354: microstates $|\vec{n}\rangle$ indexed by the local particle
355: numbers. Not suprisingly, the evolution operator $\hat{\cal L}_0$ is
356: conveniently expressed in terms of
357: the creation and annihilation operators $a^{\dagger}_i, a_i$ related
358: to each local occupation
359: number $n_i$. For the rates specified in the previous subsection,
360: but at $\eps=0$, it takes the following form
361: \begin{equation}\begin{split}
362: -\hat{\cal L}_0=&D\sum_i\sum_{j\text{ nn of }i}
363: (a_i^\dagger-a_j^\dagger)a_i\\&+\alpha(1-a_0^\dagger)+\delta(1-a_L^\dagger)\\
364: &+\gamma(a_0^\dagger-1)a_0+\beta
365: (a_L^\dagger-1)a_L
366: \end{split}\end{equation}
367: One can of course recognize that at $\eps=0$ one is formally dealing with
368: a free boson hamiltonian.
369: The next step consists in converting the computation of expectation values
370: of physical observables
371: into the evaluation of a path-integral over two fields $\hat{a}_i(t)$ and $a_i(t)$.
372: This step leads
373: to averaging a physical observable $A(\{n_i\})$ in the following way:
374: \begin{equation}
375: \langle A(\vec{n})\rangle=\int{\cal D}\hat{a}_i{\cal D}a_i\;
376: \tilde{A}(\vec{a}(t))\;\ee^{-S_\eps[\hat{a},a]}
377: \end{equation}
378: where
379: \begin{equation}\begin{split}
380: S_0[\hat{a},a]=-\sum_i
381: a_i(t)+\int_0^t\dd\tau\Big[&\sum_i\Big(\hat{a}_i\p_\tau
382: a_i+D\sum_j(\hat{a}_i-\hat{a}_j)a_i\Big)\\&+\alpha(1-\hat{a}_0)+\gamma
383: (\hat{a}_0-1)a_0\\&+
384: \delta(1-\hat{a}_L)+\beta
385: (\hat{a}_L-1)a_L\Big]
386: \end{split}\end{equation}
387: Note that we have moved the initial condition to $t=-\infty$, in order to sit
388: directly in the
389: steady-state (initial condition independent). The prescription to obtain $\tilde{A}$
390: is as follows: normal order $A(\{a_i^\dagger a_i\})$ and then
391: replace the operators $a^\dagger_i$ and $a_i$ by the fields one and $a_i(t)$
392: respectively. This yields $\tilde{A}(\vec{a}(t))$.
393: We have adopted a path-integral formulation because of the large toolbox that
394: goes with it. Besides, it formally provides us with a sort of dynamical
395: partition function on which series expansion are straightforward.
396: The efficiency of the mapping lies in the following observation: at $\eps=0$,
397: that is for independent
398: particles, the Gaussian field theory exactly encodes Poissonian statistics.
399: Hence the high-temperature ($\eps\to 0$) expansion that will be performed next
400: consists in expanding around the
401: Poissonian distribution. For free particles, as expected, the stationary
402: measure factorizes as a
403: product of local Poissonian distributions,
404: \begin{equation}
405: P(\vec{n})=\prod_i\ee^{-\zeta_i}\frac{\zeta_i^{n_i}}{n_i!}
406: \label{Prod-state}
407: \end{equation}
408: Hence, for free particles, there is local equilibrium, and the density field
409: $\rho_i=\langle n_i\rangle$ is identical to the fugacity field $\zeta_i$. This means that the
410: properties of the system are the same as those that one would obtain in
411: equilibrium if one imposed a space dependent fugacity $\zeta_i$. Here the explicit
412: expression of the fugacity is given by, in the continuum limit, setting
413: $x=i/L\in[0,1]$,
414: \begin{equation}\label{freeprofile}
415: \zeta(x)=\zeta_0+(\zeta_1-\zeta_0)x+\frac{\zeta_0-\zeta_1}{\beta\gamma L}\left(\gamma
416: x-\beta(1-x)\right)+{\cal O}(L^{-2})
417: \end{equation}
418: where we have set $\zeta_0=\frac{\alpha}{\gamma}$ and
419: $\zeta_1=\frac{\delta}{\beta}$ as the fugacities of the reservoirs. The field theory described by $S_0$ is free, which results in $a(x,t)$ being a
420: nonfluctuating field set to its average expression $\rho(x)=\zeta(x)$. This means in particular that, for $\eps=0$,
421: \begin{equation}
422: \int{\cal D}\hat{a}_i{\cal D}a_i\tilde{A}(\{a_i(t)\})\ee^{-S_0}=\tilde{A}(\{\rho_i\})
423: \end{equation}
424: whatever the observable $A$.\\
425: 
426: 
427: Another central quantity is the response function of
428: the system to particle injection: the probability $G(x,y;t-t')$ that there is a particle at $x$ at time $t$ given
429: that there was one at $y$ at time $t'$ is given by
430: \begin{equation}
431: G(x,y;\tau)=\frac{2}{L}\sum_{n\in\mathbb{Z}}\sin(n\pi x)\sin(n\pi y)\ee^{-\pi^2 n^2 \tau/L^2}
432: \end{equation}
433: The time integrated response function $\hat{G}(x,y)=\int_0^\infty\dd\tau
434: G(x,y;\tau)$ will also be needed
435: \begin{equation}\begin{split}
436: \hat{G}(x,y)=&L\left[x(1-y)\Theta(y-x)+y(1-x)\Theta(x-y)\right]+\\
437: &\frac{1}{\gamma}(1-x)(1-y)-\frac{1}{\beta}x y\\&-\frac{1}{L\beta^2\gamma^2}\left(\gamma
438: x-\beta(1-x)\right)\left(\gamma
439: y-\beta(1-y)\right)+{\cal
440: O}(L^{-2})
441: \end{split}\end{equation}
442: In technical terms $G$ is simply the free propagator of the theory.\\
443: \subsection{Effective free energy in the steady-state}
444: The probability $P[\{n_i\}]$ to observe a given occupation number configuration
445: $\{n_i\}$ (or alternatively, a given proifle $n(x)$) is given by
446: \begin{equation}
447: P[\{n_i\}]=\langle\prod_i\delta(n_i-a_i^\dagger a_i)\rangle\end{equation}
448: We further define the effective free energy ${\cal F}[n]$ of the profile $n(x)$ by
449: \begin{equation}
450: {\cal F}[n]=-\lim_{L\to\infty}\frac{\ln P[n]}{L}
451: \end{equation}
452: To access ${\cal F}$ we first pass to the generating
453: function
454: \begin{equation}
455: \hat{P}[\{z_i\}]=\sum_{\{n_i\}}\prod_jz_j^{n_j}P[\{n_i\}]
456: \end{equation}
457: then work directly on
458: \begin{equation}
459: \Omega[\{z_i\}]=-\lim_{L\to\infty}\frac{\ln\hat{P}[\{z_i\}]}{L}
460: \end{equation}
461: which plays the r\^ole of an effective grand-potential. Our task is to compute
462: \begin{equation}
463: \hat{P}[\{z_i\}]=\langle\ee^{\sum_i(z_i-1) a_i(t)}\rangle
464: \end{equation}
465: where the brackets denote the weighted path-integral defined by the action $S_0$. Using again a
466: continuum limit notation, we find, as expected, that
467: \begin{equation}
468: \hat{P}[\{z(x)\}]\Big|_{\eps=0}=\ee^{L\int_0^1(z(x)-1) \rho(x)}
469: \end{equation}
470: from which we recover that the a steady-state probability distribution function is a product
471: of local Poissonian distributions:
472: \begin{equation}
473: P[\{n_i\}]=\prod_i\ee^{-\rho_i}\frac{\rho_i^{n_i}}{n_i!}
474: \end{equation}
475: Going to a continuum notation we find
476: \begin{equation}
477: {\cal F}[n]=\int_0^1\dd x\left[\rho(x)-n(x)+n(x)\ln\frac{n(x)}{\rho(x)}\right]
478: \end{equation}
479: in agreement with the mathematically precise construction of
480: the continuum limit \cite{derridalebowitzspeer2}.
481: At this stage, the present paragraph looks like a very technical rephrasing of simple
482: properties. We are now ready, however, to attack the case of interacting systems.
483: 
484: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
485: \section{Zero Range Process}
486: \label{ZRP}
487: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
488: \subsection{Evolution operator and field-theory}
489: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
490: We will restrict our analysis to the case of pair repulsion, namely
491: \begin{equation}
492: h(n)=\eps n (n-1)
493: \end{equation}
494: The master equation equation can be cast in the form
495: \begin{equation}
496: \frac{\dd|\Psi(t)\rangle}{\dd t}=\hat{\cal L}_\eps|\Psi(t)\rangle
497: \end{equation}
498: where for the rates specified in (\ref{zr-diff-rate},\ref{zr-removal-rate},\ref{zr-injection-rate}), the evolution operator $\hat{\cal L}_\eps$ takes the following form
499: \begin{equation}\begin{split}
500: -\hat{\cal L}_\eps=&D\sum_i\sum_{j\text{ nn of }i}
501: (a_i^\dagger-a_j^\dagger)\ee^{2\eps a_i^\dagger
502: a_i}a_i\\&+\alpha(1-a_0^\dagger)+\delta(1-a_L^\dagger)\\&+\gamma(a_0^\dagger-1)\ee^{2\eps a_0^\dagger a_0}a_0+\beta
503: (a_L^\dagger-1)\ee^{2\eps a_L^\dagger a_L}a_L
504: \end{split}\end{equation}
505: One can of course recognize that at $\eps=0$ one recovers the free boson hamiltonian of the
506: previous section. The corresponding action reads
507: \begin{equation}\begin{split}
508: S_\eps[\hat{a},a]=-\sum_i
509: a_i(t)+\int_0^t\dd\tau\Big[&\sum_i\Big(\hat{a}_i\p_\tau
510: a_i+D\sum_j(\hat{a}_i-\hat{a}_j)\ee^{(\ee^{2\eps}-1)\hat{a}_i
511: a_i}a_i\Big)\\&+\alpha(1-\hat{a}_0)+\gamma
512: (\hat{a}_0-1)\ee^{(\ee^{2\eps}-1)\hat{a}_0a_0}a_0\\&+
513: \delta(1-\hat{a}_L)+\beta
514: (\hat{a}_L-1)\ee^{(\ee^{2\eps}-1)\hat{a}_L a_L}a_L\Big]
515: \end{split}\end{equation}
516: Our calculations will be based on an expansion of $S_\eps$ to first order in $\eps$:
517: \begin{equation}\label{actionZRP}
518: S_\eps=S_0+2\eps \int\dd
519: t\Bigg[\sum_{i,j}(\hat{a}_j-\hat{a}_i)\hat{a}_ia_i^2+\gamma(\hat{a}_0-1)\hat{a}_0a_0^2+\beta(\hat{a}_L-1)\hat{a}_La_L^2\Bigg]
520: +{\cal O}(\eps^2)\end{equation}
521: \subsection{Effective free energy, profile and correlations}
522: In order to evaluate
523: \begin{equation}
524: \hat{P}[z(x)]=\langle\ee^{L\int_0^1\dd x(z(x)-1)a(x,t)}\rangle
525: \end{equation}
526: we rely on a cumulant expansion. This form is particularly well-suited for a cumulant expansion,
527: which we write
528: as
529: \begin{equation}
530: \hat{P}[z(x)]=\langle\exp\left[L\sum_{n\geq 1}\frac{1}{n!}\int_0^1\prod_{j=1}^n (z(x_j)-1)
531: W^{(n)}(x_1,...,x_n)+\right]
532: \end{equation}
533: where $W^{(n)}$ is the $n$-point connected correlation function of the field
534: $a$ at equal times, in the steady state.
535: In practice it is convenient to introduce the auxiliary fields
536: $\bar{\phi}=\hat{a}-1$ and $\phi=a-{\zeta}$. Note that
537: \begin{equation}
538: {\rho}(x)=\langle n(x)\rangle=\langle a(x)\rangle=W^{(1)}(x)
539: \end{equation}
540: is so far undetermined. It will be useful to split $W^{(2)}$
541: into two pieces, denoted by $W^{(2)}_{\text{loc}}$ and $W^{(2)}_{\text{NE}}$,
542: corresponding to the local delta correlated piece in $W^{(2)}$ (namely the
543: diagonal part) and the genuinely nonequilibrium contribution, respectively. In order to obtain $W^{(2)}$ one first directly computes the field connected
544: correlation function to leading order in $\eps$:
545: \begin{equation}\begin{split}
546: \langle\phi(x_1,t_1)\phi(x_2,t_2)\rangle_c=&
547: \frac{2\eps}{L}\int_0^\infty \dd t
548: \int_0^1\dd y\Big( \zeta(y)^2(\p_y^2G(x_1,y;t_1-t)G(x_2,y;t_2-t)\\&+G(x_1,y;t_1-t)\p_y^2G(x_2,y;t_2-t) )\Big)
549: \end{split}\end{equation}
550: where the usual contractions between a bared and a non-bared field were carried out. This
551: yields,
552: \begin{equation}\label{corrfieldZRP}
553: W^{(2)}(x,y)=W_{\text{loc}}^{(2)}(x,y)=-2\eps\zeta^2(x)\delta(L(x-y))
554: \end{equation}
555: as if local equilibrium would hold (to this order in $\eps$ at least).
556: Note that the time depependent correlations are obtained by the same token.
557: The profile for the ZRP reads
558: \begin{equation}\label{profZRP}\begin{split}
559: {\rho}(x)=\zeta(x)-2\eps\zeta(x)^2
560: \end{split}\end{equation}
561: Let us compare with what we would obtain in equilibrium for the distribution
562: (\ref{Peq}). The density-fugacity relationship and the local particle
563: number fluctuations would read
564: \begin{equation}
565: \rho=\langle n \rangle=\zeta-2\eps \zeta^2,\;\;\langle n^2\rangle_c=\zeta-4\eps
566: \zeta^2
567: \end{equation}
568: A quick glance at (\ref{profZRP},\ref{corrfieldZRP}) shows that the zero range
569: process, to first order in $\eps$ and to leading order in $L$ at least, appears
570: to be consistent with local equilibrium. The existence of local equilibrium
571: can actually be proved in general, independently of the explicit
572: form of $h(n)$ provided the transition rates are given by the expressions
573: (\ref{zr-diff-rate},\ref{zr-removal-rate},\ref{zr-injection-rate}).
574: At the field theory level, it follows by inspection of the
575: corresponding Feynman diagrams which indicates that
576: local equilibrium holds to all orders in $\eps$
577: due to the fact that the interaction arising from diffusion
578: in the bulk is proportional to the Laplacian of the response field.
579: Hence, by Wick's theorem, the interactions are proportional to the
580: Laplacian of the propagator $G$ which is a delta function in space.
581: Thus spatial correlations cannot be built up and the steady state
582: measure remains a product measure just as is the case for
583: noninteracting particles (\ref{Prod-state}). It should be noted,
584: that the existence of local equilibrium can be more simply proved \cite{Gunter}
585: by just substituting product measure into the steady state
586: master equation and deriving an equation for the local fugacity
587: (\ref{freeprofile}).\\
588: 
589: For zero range processes the effective free energy can thus directly be
590: obtained from the equilibrium distribution (\ref{Peq}) in which one has
591: substituted the local fugacity by its expression in terms of the local average
592: density.
593: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
594: 
595: 
596: \section{Short range process}
597: \label{SRP}
598: \subsection{Evolution operator and action}
599: As announced in \ref{choiceofdyn}, we now wish to study a different set of dynamical rules, those of the
600: short-range process. For definiteness, we confine the present paragraph to the case of
601: on-site pair repulsion, namely, $h(n)=\eps n(n-1)$. One of the reasons for this
602: choice is that, part from the $\eps=0$ limit which reduces to free particles,
603: the $\eps\to\infty$ limit exactly corresponds to the Symmetric Exclusion
604: Process (SEP). Given that the scaling variable that appear in our expansions is
605: $\eps\rho$, we may have the hope to connect our results with the low density
606: behavior of the results obtained for the SEP. Again, for this dynamics, it is possible to
607: write the master equation in the form of an imaginary time Schr\"odinger equation with an
608: evolution operator
609: \begin{equation}\begin{split}
610: -\hat{\cal L}_\eps=&D\sum_i\sum_{j\text{ nn of }i}
611: (a_i^\dagger-a_j^\dagger)\ee^{\eps\hat{n}_i-\eps\hat{n}_j}a_i\\&+\alpha(1-a_0^\dagger)\ee^{-\eps\hat{n}_0}+\delta(1-a_L^\dagger)\ee^{-\eps\hat{n}_L}\\
612: &+\gamma(a_0^\dagger-1)\ee^{\eps\hat{n}_0}a_0+\beta
613: (a_L^\dagger-1)\ee^{\eps\hat{n}_L}a_L
614: \end{split}\end{equation}
615: In practice our analysis will be limited to the first nontrivial order in
616: $\eps$, and the corresponding action reads
617: \begin{equation}\label{actionSRP}\begin{split}
618: S_\eps=&S_0+\eps \int\dd
619: t\Bigg[\sum_{i,j}(\hat{a}_j-\hat{a}_i)(\hat{a}_ia_i-\hat{a}_j a_j)\\&-\alpha
620: (1-\hat{a}_0)\hat{a}_0
621: a_0+\gamma(\hat{a}_0-1)\hat{a}_0a_0^2\\&-\delta(1-\hat{a}_L)\hat{a}_L
622: a_L+\beta(\hat{a}_L-1)\hat{a}_La_L^2\Bigg]
623: \end{split}\end{equation}
624: The expanded action (\ref{actionSRP}) will be the starting point of our computations.
625: 
626: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
627: \subsection{Effective free energy for onsite pair repulsion}
628: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
629: As explained earlier, we shall focus on the generating function of the
630: probability distribution,
631: \begin{equation}
632: \hat{P}[z(x)]=\langle\ee^{L\int_0^1\dd x (z(x)-1) a(x,t)}\rangle
633: \end{equation}
634: Note that if one would start from a hamiltonian of the
635: form
636: \begin{equation}
637: h(n)=\eps\sum_{p\geq 2} g_p n(n-1)...(n-p+1)
638: \end{equation}
639: where the $g_p$'s are order one constants, then the cumulant expansion would
640: require to go as far as $W^{(p)}$, to leading order in $\eps$. Of course, at
641: finite $\eps$, all cumulants are needed. In the present case of onsite pair
642: repulsion, namely with
643: $
644: h(n)=\eps n(n-1)
645: $
646: we are thus left with the computation of the first and second cumulant of
647: $a(x,t)$. We find that ${\rho}(x)=\langle a(x,t)\rangle$ has the following
648: expression:
649: \begin{equation}\begin{split}
650: {\rho}(x)=&\zeta(x)-2\eps\gamma
651: (\zeta_0+\zeta(0))\zeta(0)\hat{G}(x,0)
652: -2\eps\beta(\zeta_1+\zeta(1))\zeta(1)\hat{G}(x,1)\\
653: &-\frac{\eps}{L}\int_0^1\dd y\;\zeta(y)\p_y\zeta(y)\p_y \hat{G}(x,y)
654: \\=&(\zeta_0-2\eps\zeta_0^2)(1-x)+(\zeta_1-2\eps \zeta_1^2)x
655: +\eps(\zeta_1-\zeta_0)^2x(1-x)\\&+\frac{\zeta_0-\zeta_1}{\beta\gamma L}\Bigg(\gamma
656: x\Big(1-\eps[\zeta_0(3-2x)+\zeta_1(2+2x)]\Big)\\&\qquad-\beta(1-x)\Big(1-\eps[\zeta_1(1+2x)+\zeta_0(4-2x)]\Big)\Bigg)+{\cal
657: O}(L^{-2})
658: \end{split}\end{equation}
659: It may be seen that
660: \begin{equation}
661: {\rho}(x)=\zeta(x)-2\eps\zeta(x)^2-\eps(\zeta_0-\zeta_1)^2 x(1-x)+{\cal O}(L^{-1})
662: \end{equation}
663: thus showing that there is no local equilibrium in the short range process.
664: The density-density correlation function $C^{(2)}(x,y)=\langle n(x)n(y)\rangle_c$ takes the simple form
665: \begin{equation}\begin{split}
666: C^{(2)}(x,y)=(\zeta(x)-2\eps\zeta(x)^2)\delta(L(x-y))+W^{(2)}(x,y)
667: \end{split}\end{equation}
668: with
669: \begin{equation}\label{W2pairs}\begin{split}
670: W^{(2)}(x,y)=&-2\eps\zeta(x)^2\delta(L(x-y))
671: \\&-\frac{\eps}{L}(\zeta_0-\zeta_1)^2\left[x(1-y)\Theta(y-x)+y(1-x)\Theta(x-y)\right]
672: \end{split}\end{equation}
673: which now features a nonzero long range $W^{(2)}_{\text{NE}}$ piece:
674: \begin{equation}
675: \begin{split}
676: W_{\text{NE}}^{(2)}(x,y)=&-\frac{\eps}{L}(\zeta_0-\zeta_1)^2\left[x(1-y)\Theta(y-x)+y(1-x)\Theta(x-y)\right]
677: \end{split}\end{equation}
678: It is possible to return to the effective free energy
679: \begin{equation}\label{freeenergypairs}\begin{split}
680: {\cal F}[n(x)]=&\int_0^1\dd x
681: \left[{\rho}(x)+\eps\rho(x)^2-n(x)+n(x)\ln\frac{n(x)}{{\rho}(x)+2\eps\rho(x)^2}+\eps
682: n(x)(n(x)-1)\right]\\&-\frac 12\int_0^1\dd x\dd
683: y\left(\frac{n(x)}{{\rho}(x)}-1\right)W_{\text{NE}}^{(2)}(x,y)\left(\frac{n(y)}{{\rho}(y)}-1\right)
684: \end{split}\end{equation}
685: The first brackets on the rhs of (\ref{freeenergypairs}) corresponds to a system
686: in local equilibrium with respect to an effective fugacity $\rho+2\eps\rho^2$;
687: it already appeared for the zero range process. The integral in the second line illustrates the nonlocal
688: long-range nature
689: of the effective interactions in a nonequilibrium steady-state
690: ($W^{(2)}_{\text{NE}}$ being negative, the corresponding contribution is
691: positive, which expresses the repulsive nature of the effective interactions). The fluctuations
692: of the total number of particles read
693: \begin{equation}\label{fluctSRP}
694: \Delta N^2=\Delta N^2_{\text{loc. eq.}}-L\frac{\eps}{12}(\rho_0-\rho_1)^2
695: \end{equation}
696: where $\Delta N^2_{\text{loc. eq.}}$ denotes the fluctuations of a
697: systems in local thermal equilibrium with the same fugacity profile (such as the
698: ZRP). The decrease of the global fluctuations is yet another consequence of the
699: development of long-range anticorrelations. As a coincidence, note that setting
700: in (\ref{fluctSRP}) $\eps=1$ yields exactly the symmetric exclusion process
701: expression for this quantity~\cite{derridalebowitzspeer2}.
702: 
703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
704: \subsection{Effective free energy with onsite triplet repulsion}
705: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
706: Now we wish to explore the effect of varying the type of interaction by
707: studying the case
708: \begin{equation}
709: h(n)=\eps n (n-1)(n-2)
710: \end{equation}
711: which assigns an energy cost to the site proportional to the number
712: of triplets of particles are
713: present at the site. We find that the profile is given by
714: \begin{equation}
715: \rho(x)=\underbrace{\zeta(x)-3\eps\zeta(x)^3}_{\text{local
716: eq.}}-\eps x(1-x)(\zeta_0-\zeta_1)^2(\zeta_0+\zeta_1+\zeta_0(1-x)+\zeta_1 x)
717: \end{equation}
718: while the correlation function of the field reads
719: \begin{equation}\label{W2triplets}\begin{split}
720: W^{(2)}(x,y)=&-6\eps\zeta(x)^3\delta(L(x-y))
721: \\&-\frac{12\eps}{L}(\zeta_0-\zeta_1)^2(\zeta_0+\zeta_1)
722: \left[x(1-y)\Theta(y-x)+y(1-x)\Theta(x-y)\right]\\&
723: +\frac{12\eps}{L}(\zeta_0-\zeta_1)^3 w(x,y)
724: \end{split}\end{equation}
725: In (\ref{W2triplets}) the extra function $w(x,y)$ has the explicit expression
726: \begin{equation}
727: w(x,y)=\frac{16}{\pi^4}\sum^\prime_{n,m\geq 1}\frac{1}{m^2+n^2}
728: \left[\frac{1}{(m+n)^2}-\frac{1}{(m-n)^2}\right]\sin(n\pi x)\sin (m\pi y)
729: \label{wxy}
730: \end{equation}
731: where $\displaystyle{\sum^\prime_{n,m\geq 1}}$ means a summation over $n$ and $m$ of opposite parities. We have not been
732: able to come up with a closed expression for $w$ (though one might exist). The function $w$
733: vanishes at $(1/2,1/2)$ and does not contribute to the number of particles
734: fluctuations. Figure (\ref{fig-w})
735: shows a two-dimensional plot of $w$ which reveals that $w(x,y)$ conveys
736: anticorrelations for $x,y$ in the vicinity of the reservoir with the highest
737: density, while positive correlations develop close to the reservoir imposing
738: the lowest density.
739: %\begin{figure}[htb]
740: %  \centerline{ \epsfxsize=8cm \epsfbox{w-function-dlg.eps} }
741: %   \vspace{0.cm}
742: %\caption{Plot of $w(x,y)$ [see eq.(\ref{wxy})] as a function of $x,y\in[0,1]$.}
743: %\label{fig-w}
744: %\end{figure}
745: \begin{figure}
746: $$\input{w-function-dlg2.pstex_t}$$
747: \caption{Plot of $w(x,y)$ [see eq.(\ref{wxy})] as a function of $x,y\in[0,1]$.}\label{fig-w}
748: \end{figure}
749: The function $w$ stands for the deviation of the long-range component of the
750: density correlation function with respect to the pair interaction case (whether
751: at small $\eps$ as in (\ref{W2pairs}), or at $\eps\to\infty$ as computed by
752: Spohn~\cite{spohn}). It is worth stressing that, to our knowledge, no
753: microscopic model had, up to now, shown such strong deviations.\\ 
754: 
755: The fluctuations
756: of the total number of particles now read
757: \begin{equation}
758: \Delta N^2=\Delta N^2_{\text{loc. eq.}}-2L\eps(\zeta_0-\zeta_1)^2(\zeta_0+\zeta_1) 
759: \end{equation}
760: where $\Delta N^2_{\text{loc. eq.}}$ denotes the fluctuations of a
761: systems in local thermal equilibrium with the same fugacity profile (such as the
762: ZRP). In the present case the third cumulant is of order $\eps$ as well, and it
763: has the formal expression
764: and that 
765: \begin{equation}\label{W3triplet}\begin{split}
766: &W^{(3)}(x_1,x_2,x_3)=-6\eps\zeta(x_1)^3\delta(L(x_1-x_2))\delta(L(x_2-x_3))
767: \\&-\frac{6\eps}{L}\int\dd\tau\dd z(\zeta_0(1-z)+\zeta_1 z)G(x_1,z;\tau)G(x_2,z;\tau)G(x_3,z;\tau)
768: \end{split}\end{equation}
769: Denoting by $W^{(3)}_{\text{NE}}$ the function appearing in
770: the second line of the r.h.s. of Eq.~(\ref{W3triplet}) allows to write the
771: effective free energy in the form
772: \begin{equation}\label{freeenergytriplets}\begin{split}
773: {\cal F}[n(x)]&=\int_0^1\dd x
774: \left[{\rho}+\eps\rho^3-n+n\ln\frac{n}{{\rho}+3\eps\rho^3}+\eps
775: n(n-1)(n-2)\right]\\&-\frac 12\int_0^1\dd x\dd
776: y\left[\frac{n(x)}{{\rho}(x)}-1\right]
777: W_{\text{NE}}^{(2)}(x,y)\left[\frac{n(y)}{{\rho}(y)}-1\right]\\&
778: -\frac{1}{3!}\int_0^1\dd x\dd y\dd z
779: \left(\frac{n(x)}{{\rho}(x)}-1\right)
780: \left(\frac{n(y)}{{\rho}(y)}-1\right)
781: \left(\frac{n(z)}{{\rho}(z)}-1\right)
782: W_{\text{NE}}^{(3)}(x,y,z)
783: \end{split}\end{equation}
784: As one can see, the triplet repulsion generates not only two-body
785: long ranged repulsive interactions, but also three body interactions, repulsive
786: as well. This establishes a correspondence between the microscopics (the precise
787: form of $h(n)$) and the effective interactions generated by the boundary drive.
788: 
789: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
790: \subsection{Competing interactions}
791: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
792: Suppose we now focus on a local interaction energy of the form
793: \begin{equation}
794: h(n)=\eps\left[- n (n-1)+\lambda n (n-1) (n-2)\right],\;\lambda >0
795: \end{equation}
796: where the first term on the r.h.s., which favors pair condensation, is competing
797: with the second term on the r.h.s. which assigns a relative energy cost
798: $\lambda$ to the piling up of three or more particles. To first order in
799: $\eps$ the density correlation function is merely a linear combination of that
800: obtained for the pair and triplet interactions, namely
801: \begin{equation}\begin{split}
802: C^{(2)}(x<y)=\frac{\eps}{L}(\zeta_0-\zeta_1)^2&\Big[
803: (1-12\lambda(\zeta_0+\zeta_1))x(1-y)+12\lambda(\zeta_0-\zeta_1) w(x,y)\Big]
804: \end{split}\end{equation}
805: We can clearly see that, at sufficiently low densities, positive correlations
806: develop, thus providing a microscopic counterexample to the general belief
807: that  the driving of a current builds up anticorrelations (note, however, that
808: Spohn~\cite{spohn} had left this possibility {\it a priori} open).
809: 
810: 
811: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
812: \section{Integrated current distribution}
813: \label{JJJ}
814: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
815: Let $Q(t)$ be the net number of particles which have jumped from the left
816: reservoir into the system over the time interval $[0,t]$ (counted positively for
817: an actual jump from the reservoir into the system, and negatively for a jump out
818: of the system to the reservoir). The physical motivations for studying the
819: properties of $Q$ are detailed by Lebowitz and
820: Spohn~\cite{lebowitzspohn} who showed that this quantity, defined for a Markov
821: process (our boundary driven lattice gas with stochastic dynamics) plays a role analogous to
822: the phase space contraction rate for dynamical systems (and for which
823: Gallavotti and Cohen~\cite{gallavotticohen} proved their fluctuation theorem).
824: Below, we shall concerned with the calculation of the distribution of $Q(t)$.
825: 
826: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
827: \subsection{Free particles}
828: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
829: Following the procedure described in \cite{derridalebowitz,derridadoucotroche},
830: we construct a master equation for
831: $P(\{n_i\},Q,t)$, the probability that the system is in state $\{n_i\}$ and with
832: $Q(t)=Q$ by explicitly separating those moves in phase space that increase $Q$
833: by one, decrease it by one, or leave it unchanged. Introducing a state vector
834: $|\psi(Q,t)\rangle=\sum_{\{n_i\}}P(\{n_i\},Q,t)|\{n_i\}\rangle$, we are left with an evolution
835: equation of the form
836: \begin{equation}
837: \frac{\dd|\psi(Q,t)\rangle}{\dd t}=-(\hat{H}_{1}+\hat{H}_{-1}+\hat{H}_0)|\psi(Q,t)\rangle
838: \end{equation}
839: where the operators $\hat{H}_\pm 1$ increase/decrease $Q$ by one, and where
840: $\hat{H}_0$ leave $Q$ unchanged. We are interested in the distribution function
841: of $Q(t)$ in the steady state, over large time intervals,
842: \begin{equation}
843: p(Q,t)=\langle \bp|\psi(Q,t)\rangle
844: \end{equation}
845: or, more conveniently, its generating function, namely,
846: \begin{equation}
847: \hat{p}(z,t)=\sum_{Q=-\infty}^{+\infty}z^Qp(Q,t)
848: \end{equation}
849: It turns out that the generating function can conveniently be expressed in terms of a path-integral
850: \begin{equation}
851: \hat{p}(z,t)=\int{\cal D}\hat{a}_i{\cal D}a_i\; \ee^{-S_{0,z}[\hat{a},a]}
852: \end{equation} 
853: where, for noninteracting particles
854: \begin{equation}\begin{split}
855: S_{0,z}[\hat{a},a]=-\sum_i
856: a_i(t)+\int_0^t\dd\tau\Big[&\sum_i\Big(\hat{a}_i\p_\tau
857: a_i+D\sum_j(\hat{a}_i-\hat{a}_j)a_i\Big)\\&+\alpha(1-z\hat{a}_0)+\gamma
858: (\hat{a}_0-z^{-1})a_0\\&+
859: \delta(1-\hat{a}_L)+\beta
860: (\hat{a}_L-1)a_L\Big]
861: \end{split}\end{equation} 
862: The action $S_{0,z}$ does not describe a stochastic process (unless $z=1$ for
863: which it reduces to $S_0$). However, it may be
864: readily seen that
865: \begin{equation}
866: \hat{p}(z,t)=\langle 
867: \ee^{\int_0^t\dd t (\alpha (z-1)-\gamma(1-z^{-1})a_0(t)) }\rangle_z
868: \end{equation}
869: where the brackets $\langle..\rangle_z$ now denote an average with respect to the 
870: process governed by $S_0$ in which $\alpha$ is formally replaced with $\alpha z$. Using that $a_0(t)$, for free
871: particles, is a nonfluctuating field with expression taken from (\ref{freeprofile}) by changing
872: $\alpha$ into $\alpha z$,
873: \begin{equation}
874: a_0(t)=\zeta_0 z-\frac{z\zeta_0-\zeta_1}{\gamma L}\left[1-\eps(4\zeta_0 z+\zeta_1)\right]
875: \end{equation}
876: we find that, in the long time limit, 
877: \begin{equation}
878: \lim_{t\to\infty}\frac{\ln\hat{p}(z,t)}{t}=\mu(z)=\frac{1}{L}\frac{z-1}{z}(\zeta_0 z-\zeta_1)
879: \end{equation}
880: It is {\it a posteriori} clear why subleading finite size corrections had to be
881: kept all along. The function $\mu(z)$ is the generating function for the cumulants of $Q$.
882: It is worth commenting on two
883: remarkable, yet expected, properties of $\mu(\zeta_0,\zeta_1,z)$. Namely, if we
884: exchange the roles of the
885: reservoirs, the distribution of $Q$ is turned into that of $-Q$, hence
886: \begin{equation}\label{P}
887: \mu(\zeta_0,\zeta_1,z)=\mu(\zeta_1,\zeta_0,z^{-1})
888: \end{equation}
889: But it also
890: satisfies the Gallavotti--Cohen property (see \cite{lebowitzspohn,derridadoucotroche} for a readable proof), namely
891: \begin{equation}\label{GC}
892: \mu(\zeta_0,\zeta_1,z)=\mu(\zeta_0,\zeta_1,\frac{\zeta_1}{\zeta_0 z})
893: \end{equation}
894: which is best known when rephrased as follows. Let $\pi(Q)$ be the large
895: deviation function related to $p(Q,t)$:
896: \begin{equation}
897: \pi(Q)=\lim_{t\to\infty}\frac{\ln p(Q,t)}{t}
898: \end{equation}
899: Besides, $\pi(Q)$ appears as the Legendre transform of $\mu(z)$ with respect to
900:  $\ln z$:
901: \begin{equation}
902: \pi(Q)=\text{max}_z\{\mu(\zeta_0,\zeta_1,z)-\ln z Q\}
903: \end{equation}
904: hence
905: \begin{equation}
906: \lim_{t\to\infty}\frac{1}{t}\ln\frac{p(Q,t)}{p(-Q,t)}=\pi(Q)-\pi(-Q)=\ln\frac{\zeta_0}{\zeta_1}Q
907: \end{equation}
908: Of course this can {\it a posteriori} be verified on the explicit expression of $\pi(Q)$:
909: \begin{equation}\begin{split}
910: \pi(Q)=&\frac{Q}{2}+\sqrt{\frac{Q^2}{4}+\zeta_0\zeta_1}
911: +\frac{\zeta_0\zeta_1}{\frac{Q}{2}+\sqrt{\frac{Q^2}{4}+\zeta_0\zeta_1}}\\&-(\zeta_0+\zeta_1)
912: -Q\ln\left[\frac{\sqrt{\frac{Q^2}{4}+\zeta_0\zeta_1}+\frac{Q}{2}}{\zeta_0}\right]
913: \end{split}
914: \label{piQeq}
915: \end{equation}
916: which is plotted on Fig.~(\ref{fig-piQ}).
917: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
918: \begin{figure}
919: $$\input{piQfree2.pstex_t}$$
920: \caption{Plot of $\pi(Q)$ in eq.(\ref{piQeq})
921: for free particles with $\rho_0=1\gg\rho_1$. Note the
922: strong asymmetry of $\pi$.}
923: \label{fig-piQ}
924: \end{figure}
925: %\begin{figure}[htb]
926: %  \centerline{ \epsfxsize=8cm \epsfbox{piQfree.eps} }
927: %   \vspace{0.cm}
928: %\caption{Plot of $\pi(Q)$ in eq.(\ref{piQeq})
929: %for free particles with $\rho_0=1\gg\rho_1$. Note the
930: %strong asymmetry of $\pi$.}
931: %\label{fig-piQ}
932: %\end{figure}
933: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
934: The Gallavotti-Cohen theorem was shown~\cite{kurchan,lebowitzspohn} to hold under quite general conditions
935: for nonequilibrium steady-states resulting from Markovian dynamics and we shall
936: further comment upon it in subsection \ref{discaddprinc}.
937: \subsection{Short Range Process with pair repulsion}
938: We now repeat the strategy outlined in the previous paragraph, but for the
939: interacting case, with the short range process dynamics. We split the evolution operator $\hat{H}_\eps$ into three
940: pieces, each describing moves that increase, decrease of leave $Q$ unchanged.
941: Then we pass to a path integral in terms of which we find
942: \begin{equation}
943: \hat{p}(z,t)=\int{\cal D}\hat{a}_i{\cal D}a_i\; \ee^{-S_{\eps,z}[\hat{a},a]}
944: \end{equation} 
945: 
946: Having found $S_{\eps,z}$ we bring the calculation of $\hat{p}$ to that of a given observable with
947: respect to the original stochastic process in which $\alpha$ is replaced with
948: $\alpha z$:
949: \begin{equation}\begin{split}
950: \hat{p}(z,t)=\langle\exp\Big(-\int_0^t\dd
951: t\Big[\alpha(1-z)(1-\eps\hat{a}_0a_0)+\gamma(1-z^{-1})a_0(1+\eps\hat{a}_0a_0)\Big]\Big)\rangle_z
952: \end{split}\end{equation}
953: In order to evaluate the latter expectation value to leading order in $\eps$ we rely on a cumulant
954: expansion. 
955: Introducing the variable 
956: \begin{equation}
957: \omega(\zeta_0,\zeta_1,z)=\frac{z-1}{z}\left[(\zeta_0-\eps\zeta_0^2)z-(\zeta_1-\eps \zeta_1^2)-\eps (z-1)\zeta_0\zeta_1\right]
958: \end{equation}
959: we find that
960: \begin{equation}
961: L\mu=\omega-\frac{\eps}{3}\omega^2
962: \end{equation}
963: It is remarkable that the intermediate variable $\omega$ itself satisfies the two
964: invariance properties that $\mu$ is known to fulfill. Namely, the function
965: $\omega(\rho_0,\rho_1,z)$ verifies the left-right symmetry
966: \begin{equation}
967: \omega(\zeta_0,\zeta_1,z)=\omega(\zeta_1,\zeta_0,z^{-1})
968: \end{equation}
969: and the Gallavotti--Cohen property
970: \begin{equation}
971: \omega(\zeta_0,\zeta_1,z)=\omega(\zeta_0,\zeta_1,\frac{\zeta_1}{\zeta_0 z})
972: \end{equation}
973: That $\omega$ verifies (\ref{P},\ref{GC}) implies directly that $\mu$ possesses
974: the same invariance properties, as already identified in the SEP~\cite{derridadoucotroche}. One invariance property that our result does not possess, in
975: constrast to \cite{derridadoucotroche}, is the particle--hole symmetry.\\
976: 
977: The effective small parameter of the expansion is $\eps\rho$, where $\rho$ is the typical
978: density, hence our small $\eps$ expansion is seen to coincide with a low density expansion of the exact
979: result obtained at $\eps\to\infty$ by \cite{derridadoucotroche},
980: \begin{equation}
981: L\mu=\omega-\frac13
982: \omega^2+\frac{8}{45}\omega^3+{\cal O}(\omega^4)
983: \end{equation}
984: with
985: \begin{equation}
986: \omega=\frac{z-1}{z}\left[\frac{\zeta_0}{1+\zeta_0}z-\frac{\zeta_1}{1+\zeta_1}-
987: (z-1)\frac{\zeta_0}{1+\zeta_0}\frac{\zeta_1}{1+\zeta_1}\right]
988: \end{equation}
989: A natural question that arises next is how universal the result  obtained in the
990: exact calculation \cite{derridadoucotroche} is? How sensitive is it to
991: varying the microscopic interactions. In order to sort out this issue we have
992: performed a similar analysis for repulsive triplet interactions, with short
993: range dynamics.
994: 
995: \subsection{Short Range Process with triplet repulsion}
996: In the  same spirit as in the previous paragraph, we express the
997: generating function $\hat{p}(z)$
998: as the expectation value of an exponential observable with
999: respect to the lattice gas measure in
1000: which $\alpha$ is replaced with $\alpha z$. 
1001: We rely again on a cumulant expansion, and after a rather tedious
1002: calculation, we arrive at the
1003: following results. We now define the auxiliary variable $\omega$ by
1004: \begin{equation}\begin{split}
1005: \omega(\zeta_0,\zeta_1,z)=&\frac{z-1}{z}\left[(\zeta_0-\eps\zeta_0^3)z-(\zeta_1-\eps \zeta_1^3)-\eps
1006: (z-1)\zeta_0\zeta_1(\zeta_0+\zeta_1)\right]\\&\times\left[1-\frac{\eps}{2}\frac{z-1}{z}(\zeta_0 z-\zeta_1)(\zeta_0+\zeta_1)\right]
1007: \end{split}\end{equation}
1008: which also obeys (\ref{P},\ref{GC}). We find that
1009: \begin{equation}
1010: L\mu=\omega-\frac{\eps}{10}\omega^3+{\cal O}(\eps^2)
1011: \end{equation}
1012: This expression unambiguously points at a different distribution function for the
1013: integrated current. It further allows to connect the microscopics --the triplet
1014: repulsion-- with the final form of $\mu$. A $p$-body interaction would yield a 
1015: first correction to $\mu$ of the form $\omega^p$. Unfortunately we have not been
1016: able to come by a physical interpretation for the intermediate quantity $\omega$.
1017: %\subsection{Dimension $d$}
1018: \subsection{Zero Range Process with pair repulsion}
1019: Finally we examine the case of the zero range process with pair repulsion, for
1020: which we know that the steady state distribution follows local equilibrium.
1021: It is then sufficient to start from the equilibrium expression for the free process
1022: in which one has subsituted the current with its appropriate expression. It is
1023: not hard to see, through a direct evaluation, that
1024: \begin{equation}\begin{split}
1025: \frac{\langle Q\rangle}{t}=\alpha-\gamma\langle
1026: n_0\ee^{2\eps(n_0-1)}\rangle=\frac{(\zeta_0-2\eps\zeta_0^2)
1027: -(\zeta_1-2\eps\zeta_1^2)}{L}\simeq\frac{\rho_0-\rho_1}{L}
1028: \end{split}\end{equation}
1029: which leads us to
1030: \begin{equation}\label{muZRP}\begin{split}
1031: L\mu(\zeta_0,\zeta_1,z)=
1032: \frac{z-1}{z}\left[(\zeta_0-2\eps\zeta_0^2) z-(\zeta_1-2\eps\zeta_1^2)\right]
1033: \end{split}\end{equation}
1034: This form is identical to that predicted by Bodineau and
1035: Derrida~\cite{bodineauderrida}) for zero range processes.
1036: The Gallavotti-Cohen relation, in our particular case, now takes the form
1037: \begin{equation}\begin{split}
1038: \mu(\zeta_0,\zeta_1,z)=\mu(\zeta_0,\zeta_1,\frac{\zeta_0-2\eps\zeta_0^2}{\zeta_1-2\eps\zeta_1^2}\frac{1}{z})
1039: \end{split}\end{equation}
1040: It is remarkable that at equilibrium, that is for $\rho_0=\rho_1$, the current fluctuations $\langle Q^2\rangle$ is the
1041: same for the ZRP and the SRP, while higher order cumulants (that is the whole distribution) are
1042: different.
1043: 
1044: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1045: \subsection{Illustration of the additivity principle}\label{discaddprinc}
1046: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1047: To conlude this section, we would like to illustrate how our explicit results
1048: fit into the general framework provided by Bodineau and
1049: Derrida~\cite{bodineauderrida}. In particular, they have deduced, from a postulated additivity principle, a
1050: general expression for the integrated current distribution, from the sole
1051: knowledge of $\langle Q\rangle(\rho_0,\rho_1)$ and $\langle
1052: Q^2\rangle_c(\rho_0,\rho_1)$. One of the consequences of their findings is a
1053: direct computation of the analog, in the Gallavotti-Cohen theorem, of the
1054: "entropy production rate" (as formally defined by Lebowitz and
1055: Spohn~\cite{lebowitzspohn}) in these
1056: driven stochastic lattice gases. Defining, as in \cite{bodineauderrida},
1057: \begin{equation}
1058: D(\rho)=\frac{1}{t}\frac{\p}{\p \rho_0}\langle
1059: Q\rangle(\rho_0,\rho)\Big|_{\rho_0=\rho},\;\sigma(\rho)=\frac{\langle
1060: Q^2\rangle_c(\rho,\rho)}{t}
1061: \end{equation}
1062: they have shown that
1063: \begin{equation}\label{BDGC}\begin{split}
1064: \mu(z)=\mu\left(2\int^{\rho_0}_{\rho_1}\dd\rho\frac{D(\rho)}{\sigma(\rho)}\frac{1}{z}\right)
1065: \end{split}\end{equation}
1066: To leading order in $\eps$, it is straighforward to see from (\ref{muZRP}) that, for the Zero Range
1067: Process,
1068: \begin{equation}
1069: D(\rho)=1,\;\sigma(\rho)=2\rho
1070: \end{equation}
1071: For the short range process,
1072: \begin{equation}\begin{split}
1073: \text{for pair interaction: }D(\rho)=1+2\eps \rho,\;\sigma(\rho)=2\rho\\
1074: \text{for triplet interaction: }D(\rho)=1+6\eps \rho^2,\;\sigma(\rho)=2\rho\\
1075: \end{split}\end{equation}
1076: thus leading, in both cases, to 
1077: \begin{equation}\begin{split}
1078: \mu(z)=\mu(\frac{\zeta_0}{\zeta_1 z})
1079: \end{split}\end{equation}
1080: Hence formula (\ref{BDGC}) is in perfect agreement with our explicit
1081: computations, not only for zero range processes, but also for the non trivial
1082: short range processes with pair or triplet repulsion.
1083: 
1084: \section{Final remarks}
1085: 
1086: We have shown on specific examples that driving a system out of equilibrium
1087: magnifies the differences in the underlying microscopic dynamics. Not only the
1088: density profiles are different, but some dynamical rules lead the
1089: system to a state of local equilibrium, while others let it develop long range
1090: effective interactions. In all cases, however,
1091: our explicit results for the integrated current distribution provide
1092: support for the postulated additivity principle of Bodineau
1093: and Derrida \cite{bodineauderrida}.\\
1094: 
1095: We should emphasize that the above results have been worked out
1096: by setting up a formalism, based on path integrals, which provides an alternative
1097: both to exact solutions and to the
1098: fluctuating hydrodynamics approach. The path integral formalism
1099: fills a gap in the sense that it allows us to go directly from a
1100: microscopic formulation to macroscopic properties and, furthermore,
1101: it allows formulating approximate approaches in nonequilibrium
1102: settings. Most of our results rely on a
1103: high temperature or virial like expansion, thus providing a intuitive parallel
1104: to the standard techniques of equilibrium statistical mechanics.\\
1105: 
1106: We believe that several lines could now be explored. First, the present
1107: toolbox allows us to investigate the effect of additional space dimensions at
1108: little extra formal cost (even though calculations will undoubtedly
1109: be more involved). This would help to clearly isolate those feature which are
1110: characteristic of one-dimensional systems from those that generalize to
1111: more realistic ones. An interesting problem arising in higher
1112: dimension is the interplay between
1113: longitudinal and transverse current fluctuations. Second, most of the quantities
1114: studied in the present work are
1115: time-independent, but the present toolbox makes possible
1116: the sudies of time-dependent quantities such as the time-dependent
1117: profile or effective free energy. Third, a natural extension of our formalism
1118: would consist of establishing nonperturbative results in $\eps$. Much in the
1119: same way as in liquid state theory, infinite families of appropriately
1120: chosen Mayer diagrams can be
1121: summed up, and one could investigate which types of Feynman diagrams will
1122: contribute in building up the strongly nonlocal nature of the free energy
1123: functional found in \cite{derridalebowitzspeer1}.
1124: Finally, extending our results for
1125: more complex geometries with more than two particle reservoirs, as suggested in
1126: \cite{bodineauderrida} or done in \cite{blanterbuttiker},
1127: should allow us to identify the universal emerging features and to bring
1128: the theory closer to experiments.\\
1129: 
1130: \noindent {\bf Ackownledgments.} This research has been partially supported
1131: by the Hungarian Academy
1132: of Sciences (Grant No. OTKA T043734). The authors would like to thank Gunter
1133: Sch\"utz, Henk Hilhorst, Emmanuel Trizac and Alain Barrat for useful
1134: discussions, and further acknowledge Bernard Derrida for numerous
1135: useful suggestions.
1136: \newpage
1137: \begin{thebibliography}{0}
1138: 
1139: \bibitem{derridalebowitzspeer1}
1140: B. Derrida, J. L. Lebowitz and E. R. Speer,
1141: Phys. Rev. Lett. {\bf 87}, 150601 (2001)
1142: [{\tt cond-mat/0105110}], {\it
1143: Free Energy Functional for Nonequilibrium Systems: An Exactly Solvable Case}.
1144: 
1145: \bibitem{derridalebowitz}
1146: B. Derrida and J. L. Lebowitz, Phys. Rev. Lett. {\bf 80}, 209 (1998) [{\tt cond-mat/9809044}], {\it
1147: Exact Large Deviation Function in the Asymmetric Exclusion Process}.
1148: 
1149: \bibitem{derridalebowitzspeer2}
1150: B. Derrida, J. L. Lebowitz and E. R. Speer
1151: J. Stat Phys. {\bf 107}, 599 (2002) [{\tt cond-mat/0109346}], {\it
1152: Large Deviation of the Density Profile in the Steady State of the Open Symmetric 
1153: Simple  Exclusion Process}.
1154: 
1155: \bibitem{derridalebowitzspeer3}
1156: B. Derrida, J. L. Lebowitz and E. R. Speer, 
1157: J. Stat. Phys. {\bf 110}, 775 (2003) [{\tt cond-mat/0205353}],
1158: {\it  Exact Large Deviation Functional of a Stationary Open Driven Diffusive System:
1159: The Asymmetric Exclusion Process}.
1160: 
1161: \bibitem{derridaenaud}
1162: B. Derrida and C. Enaud, J. Stat. Phys. {\bf 114}, 537 (2004),
1163: [{\tt cond-mat/0307023}], {\it  Large deviation functional of the weakly asymmetric
1164: exclusion process}.
1165: 
1166: \bibitem{derridadoucotroche}
1167: B. Derrida, B, Dou\c{c}ot and P.-E. Roche, J. Stat. Phys. {\bf 115}, 713 (2004),
1168: [{\tt cond-mat/0310453}], {\it  Current fluctuations in the one dimensional
1169: symmetric exclusion process with open boundaries}.
1170: 
1171: \bibitem{bertinisolegabriellijonalasinioladim1}
1172: L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio and C. Landim,
1173: Phys. Rev. Lett. {\bf 87}, 040601 (2001)
1174: [{\tt cond-mat/0104153}], {\it
1175: Fluctuations in Stationary non Equilibrium States}.
1176: 
1177: \bibitem{bertinisolegabriellijonalasinioladim2}
1178: L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio, C. Landim,
1179: J. Stat. Phys. {\bf 107}, 635 (2002)
1180: [{\tt cond-mat/0108040}], {\it
1181: Macroscopic fluctuation theory for stationary non equilibrium states}.
1182: 
1183: \bibitem{bertinisolegabriellijonalasinioladim3}
1184: L. Bertini, A. De Sole, D. Gabrielli, G. Jona-Lasinio and C. Landim, Math. Phys.
1185: Anal. Geom. {\bf 6}, 231 (2003) [{\tt cond-mat/0307280}], {\it
1186: Large deviations for the boundary driven symmetric simple exclusion process}.
1187: 
1188: \bibitem{zrp-origin} F. Spitzer, Adv. Math. {\bf 5}, 246 (1970),
1189: 		 {\it Interaction of Markov processes}.
1190: 		 
1191: \bibitem{Cocozza} C. Cocozza-Thivent, Z. Warhsch. Verw. Gebiete, {\bf 70}, 509 (1995),
1192: {\it Processus des misanthropes}.
1193: 
1194: \bibitem{bodineauderrida}
1195: T. Bodineau and B. Derrida,  Phys. Rev. Lett. in press [{\tt cond-mat/0402305}],
1196: {\it Current fluctuations
1197: in nonequilibrium diffusive systems: an additivity principle}.
1198: 
1199: \bibitem{zrp-review} For references and a review of recent developments
1200: in zero range processes see G. M. Sch\"utz, J. Phys. A {\bf 36}, R339 (2003)
1201: [{\tt cond-mat/0308450}],
1202: {\it Critical phenomena and universal dynamics in one-dimensional driven
1203: diffusive systems with two species of particles}.
1204: 
1205: \bibitem{mattisglasser}
1206: D.C. Mattis and M.L. Glasser, { Rev. Mod. Phys.} {\bf 70}, 979 (1998), {\it
1207: The uses of quantum field theory in diffusion-limited reactions}.
1208: 
1209: \bibitem{Gunter} This was shown to us by G. Sch\"utz, private communication.
1210: 
1211: \bibitem{spohn}
1212: H. Spohn, J. Phys. A {\bf 16}, 4275 (1983), {\it 
1213: Long range correlations for stochastic lattice gases in a nonequilibrium
1214: steady-state}.
1215: 
1216: \bibitem{lebowitzspohn}
1217: J.L. Lebowitz and H. Spohn, J. Stat. Phys. {\bf 95}, 333 (1999)  [{\tt cond-mat/9811220}], {\it A Gallavotti-Cohen
1218: Type Symmetry in the Large Deviation Functional for Stochastic Dynamics}.
1219: 
1220: \bibitem{gallavotticohen}
1221: G. Gallavotti and E.G.D. Cohen, Phys. Rev. Lett. {\bf 74}, 2694 (1995) [{\tt
1222: chao-dyn/9410007}], {\it Dynamical ensembles in nonequilibrium statistical mechanics}. 
1223: 
1224: \bibitem{kurchan}
1225: J. Kurchan, { J. Phys. A} {\bf 31}, 3719 (1998) [{\tt cond-mat/9709304}], {\it Fluctuation theorem for stochastic
1226: dynamics}.
1227: 
1228: %\bibitem{janssenschmittmann}
1229: %H.K. Janssen and B. Schmittmann, {Z. Phys. B} {\bf 63}, 517 (1986), {\it
1230: %Field theory of long time behaviour in driven diffusive systems}.
1231: 
1232: \bibitem{blanterbuttiker}
1233: Ya.M. Blanter and M. B\"uttiker,  {Phys. Rev. B}
1234: {\bf 56}, 2127 (1997) [{\tt cond-mat/9702047}], {\it Shot noise current-current correlations in
1235: multi-terminal diffusive conductors}.
1236: 
1237: 
1238: \end{thebibliography}
1239: 
1240: \end{document}
1241: 
1242: 
1243: 
1244: 
1245: 
1246: 
1247: