%Paper accepted for publication in Physics Letters A. \magnification=1200 \baselineskip=2\baselineskip \centerline{GENERAL LINEAR RESPONSE FORMULA IN STATISTICAL MECHANICS,} \centerline{AND THE FLUCTUATION-DISSIPATION THEOREM FAR FROM EQUILIBRIUM.} \bigskip \centerline{by David Ruelle} \medskip \centerline{IHES, 91440 Bures sur Yvette, France. $<$ruelle@ihes.fr$>$.} \bigskip Abstract. {\it Given a nonequilibrium steady state $\rho$ we derive formally the linear response formula given by equation (6) in the text for the variation of an expectation value at time $t$ under a time-dependent infinitesimal perturbation $\delta_\tau F$ of the acting forces. This leads to a form of the fluctuation-dissipation theorem valid far from equilibrium: the complex singularities of the susceptibility are in part those of the spectral density, and in part of a different nature to be discussed.} \bigskip PACS: 05.45.+b, 05.70.Ln, 47.70.-n. \medskip Keywords: {\sl fluctuation-dissipation, linear response, nonequilibrium, statistical mechanics}\bigskip. \bigskip Serious attention is now starting to be paid to the chaotic dynamics underlying nonequilibrium statistical mechanics. (See for instance Chernov et al. [1]). A statistical mechanical system is kept far away from equilibrium by nonhamiltonian forces, and ``cooled'' by a ``Gaussian thermostat'' ( Hoover [2], Evans and Morriss [3]). What this amounts to is that the phase space $M$ of the system is taken to be compact, and its time evolution given by $$ {dx\over dt}=F(x,t)\eqno{(1)} $$ where no particular assumption is made on $F$ (except smoothness, and often $t$-indepen- dence). \medskip It would at first appear that the framework provided by $$ {dx\over dt}=F(x)\eqno{(2)} $$ is much too general to provide results of interest for nonequilibrium statistical mechanics. It is reasonable however to assume that (2) defines a chaotic time evolution, and that we may exclude a set of Lebesgue measure 0 of initial conditions in the distant past. These assumptions have surprisingly strong consequences. If we translate {\it chaos} mathematically by {\it uniform hyperbolicity}, then time averages are uniquely determined and given by a so-called SRB measure (see below). Physical time evolutions are often hyperbolic in the weaker sense that most Lyapunov exponents are different from 0. A natural idea is thus to proceed as if physical systems were uniformly hyperbolic (and then compare the results with experiments). This has been called the {\it chaotic hypothesis} by Gallavotti and Cohen (who also assume microscopic reversibility). The approach just outlined has been vindicated in the case of the Gallavotti-Cohen fluctuation theorem [4], which agrees with numerical experiments far from equilibrium [5], [6]. From the chaotic principle one also recovers near equilibrium the Onsager reciprocity relations [7], [8]. The present letter follows the same philosophy for the study of linear response far from equilibrium. Our calculations will be formal and easy. They can be made rigorous if uniform hyperbolicity holds [9], but this is harder. A rigorous analysis in a more general setup seems difficult and one may have to untroduce new ideas like the limit of a large system (thermodynamic limit). Such a limit will be used anyway to compute transport coefficients like the viscosity. What may be said now is that if linear response for physical systems has a simple expression, it must be the one given by our formal calculation. The formulae that we obtain appear thus unavoidable, and should be fundamental for nonequilibrium statistical mechanics far from equilibrium. \medskip Integrating (1) with initial condition $x$ at time $s$ gives at time $t$ a point $x(s,t)$. Suppose now that we are in the time independent situation of (2). We may thus write $$ x(s,t)=x(t-s)=f^{t-s}x\eqno{(3)} $$ If $m$ is a probability measure absolutely continuous with respect to the volume element $dx$ on the phase space $M$, and if $m$ under the time evolution $f^t$ tends weakly to a limit $\rho$, then $\rho$ is a good candidate to describe a nonequilibrium steady state. Using the notation $$ m(\Phi)=\int m(dx)\Phi(x)\qquad,\qquad\rho(\Phi)=\int\rho(dx)\Phi(x) $$ we have by assumption, for every continuous $\Phi$, $$ \lim_{s\to-\infty}\int m(dx)\Phi(f^{-s}x)=\int\rho(dx)\Phi(x)\eqno{(4)} $$ We call $\rho=\rho_F$ an SRB measure, it is usually not absolutely continuous with respect to $dx$. Such measures were introduced by Sinai, Ruelle, and Bowen [10], [11], [12] for uniformly hyperbolic dynamical systems, where it was shown that there is a unique SRB measure on each mixing attractor. A much more general study of SRB measures was then made by Ledrappier, Strelcyn, and Young [13], [14] (see Young [15] for recent results). The use of SRB measures to describe nonequilibrium steady states in statistical mechanics was advocated early by Ruelle [16], but only recently did it lead to useful results with the {\it Fluctuation Theorem} of Gallavotti and Cohen [4]. \medskip One problem with SRB measures is that their characterizations [10], [11], [12], [13], [14] are difficult to use. It is however relatively easy to expand $\rho_{F+\delta F}$ with respect to $\delta F$ [17]. This has been done rigorously in a special case [9]. Here we proceed instead formally, and to first order with respect to a {\it time dependent} perturbation $\delta_tF$ (we keep $F$ {\it time independent}). We shall thus recover in a new manner some classical results of nonequilibrium statistical mechanics, for which see for instance [18], and also obtain new results. \medskip Using (4), we have formally $$ \delta_t\rho(\Phi)=\delta\lim_{s\to-\infty}\int m(dx)\,\Phi(x(s,t)) =\lim_{s\to-\infty}\int m(dx)\,\delta x(s,t)\cdot\nabla_{x(s,t)}\Phi $$ and we may assume $s0$. >From this one can deduce the Kramers-Kronig {\it dispersion relations} [21]. The dispersion relations therefore also hold far from equilibrium. We discuss now the singularities of $\hat\kappa_\omega$ for ${\rm Im}\,\omega<0$. For the sake of clarity, we consider the situation first at equilibrium, then away from equilibrium. Our discussion will not make use of microscopic reversibility. \medskip {\it At equilibrium}, the SRB measure $\rho_F$ is absolutely continuous with respect to the volume element of $M$, and we simply write $\rho_F(dx)=dx$. From (7) and (8), it follows that the singularities of $(\hat\kappa_\omega^{s,u}X)\Phi$ for ${\rm Im}\,\omega<0$ are the same as those of $$ \int_{-\infty}^\infty e^{i\omega\sigma}d\sigma \rho_F((\Phi\circ f^\sigma).\Psi)=S_\omega(\Phi,\Psi)\eqno{(9)} $$ for $\Psi={\rm div}^s X^s$, or $F\cdot\nabla\phi+{\rm div}^u X^u$. The right-hand side of (9) is the Fourier transform of the correlation function $\sigma\mapsto\rho_F((\Phi\circ f^\sigma).\Psi)$, and $S_\omega$ is called the {\it spectral density}. For sufficiently regular $\Phi$, $\Psi$, one can extend $S_\omega(\Phi,\Psi)$ to complex $\omega$. The singularities of $\hat\kappa_\omega^s$ and $\hat\kappa_\omega^u$ (and $\hat\kappa_\omega$) are thus expected to be the same as those of $S_\omega$ for ${\rm Im}\,\omega<0$. This connection between $\hat\kappa_\omega$ and $S_\omega$ is basically the fluctation-dissipation theorem. \medskip {\it Far away from equilibrium}, the singularities of $\hat\kappa_\omega^u$ are (in view of (8)) again the same as the singularities with ${\rm Im}\,\omega<0$ of the spectral density $S_\omega$. For instance a simple pole of $S_\omega$ at $\omega_0$ (with ${\rm Im}\,\omega_0<0$) corresponds to a simple pole $a(\omega-\omega_0)^{-1}$ of $\hat\kappa_\omega^u$ (even if it is not clear how to determine the residue $a$ in practice). This is what remains here of the fluctuation-dissipation theorem. The singularities of $\hat\kappa_\omega^s$, however, become different from those of $\hat\kappa_\omega^u$. Define ${\cal T}^\sigma$ on the vector fields $X^s$ (in the stable direction) by $$ ({\cal T}^\sigma X^s)(x) =(T_{f^{-\sigma}x}f^\sigma)X^s(f^{-\sigma}x)\eqno{(10)} $$ Then, $({\cal T}^\sigma)_{\sigma\ge0}$ is a contraction semigroup and, if $-H$ is its infinitesimal generator, we have by the Hille-Yosida theorem [22] $$ \int_0^\infty e^{i\omega\sigma}d\sigma((Tf^\sigma)X^s)\circ f^{-\sigma} =(H-i\omega)^{-1} $$ so that $(\hat\kappa_\omega^sX)(\Phi) =\rho_F[(H-i\omega)^{-1}X^s\cdot\nabla\Phi]$. The singularities of $\hat\kappa_\omega^s$ are thus related to the spectrum of $H$. \medskip To summarize, we see that at equilibrium the singularities of the susceptibility $\hat\kappa_\omega$ are the singularities of the spectral density $S_\omega$ with ${\rm Im}\,\omega<0$. Outside of equilibrium, the singularities of the susceptibility $\hat\kappa_\omega$ bifurcate into those of $\hat\kappa_\omega^u$ which are again the singularities of the spectral density $S_\omega$ with ${\rm Im}\,\omega<0$, and those of $\hat\kappa_\omega^s$ which are different (and related to the spectrum of the infinitesimal generator of the semigroup $({\cal T}^\sigma)_{\sigma\ge0}$ defined by (10)). It is thus in principle possible to distinguish the singularities of $\hat\kappa_\omega^u$ and of $\hat\kappa_\omega^s$, and it would be interesting to see them in an experimental study bifurcating from each other as one moves away from equilibrium [23]. \medskip {\it Acknowledgements.} \medskip Part of this work was performed in the stimulating atmosphere of the Santa Fe Institute under a grant from the Sloan Foundation. \medskip {\it References.} \medskip [1] N.I. Chernov et al. Phys. Rev. Letters {\bf 70}, 2209-2212(1993); Commun. Math. Phys. {\bf154},569-601(1993). [2] W.G. Hoover. {\it Molecular dynamics}. Lecture Notes in Physics {\bf 258}. Springer, Heidelberg, 1986. [3] D.J. Evans and G.P.Morriss. {\it Statistical mechanics of nonequilibrium fluids}. Academic Press, NY, 1990. [4] G. Gallavotti and E.G.D. Cohen. Phys. Rev. Letters {\bf 74}, 2694-2697(1995); J. Statist. Phys. {\bf 80},931-970(1995). [5] D.J. Evans, E.G.D. Cohen, and G.P.Morriss. Phys. Rev. Letters {\bf 71},2401-2404 (1993). [6] F. Bonetto et al. Physica D, {\bf 105},226-252(1997). [7] G. Gallavotti. Phys. Rev. Letters, {\bf 77},4334-4337(1996). [8] G. Gallavotti and D. Ruelle. Commun. Math. Phys. {\bf 190}, 279-285(1997). [9] D. Ruelle. Commun. Math. Phys. {\bf 187},227-241(1997). [10] Ya.G. Sinai. Russian Math. Surveys {\bf 27},No 4,21-69(1972). [11] D. Ruelle. Am. J. Math. {\bf 98},619-654(1976). [12] R. Bowen and D. Ruelle. Invent. Math. {\bf 29},181-202(1975). [13] F. Ledrappier and J.-M. Strelcyn. Ergod. Th. and Dynam. Syst. {\bf 2},203-219(1982). [14] F. Ledrappier and L.S. Young. Ann. of Math. {\bf 122},509-539, 540-574(1985). [15] L.-S. Young. ``Ergodic theory of chaotic dynamical systems.'' Lecture at the International Congress of Mathematical Physics 1997. [16] G. Gallavotti mentions a seminar talk in Zurich in 1973. See also D. Ruelle. Ann. N.Y. Acad. Sci. {\bf 316},408-416(1978). [17] A first version of this idea, near equilibrium, appears in G. Gallavotti. J.Stat. Phys. {\bf 84},899-926(1996). [18] S.R. de Groot and P. Mazur. {\it Nonequilibrium thermodynamics.} Dover, New York, 1984. [19] D. Ruelle. "Nonequilibrium statistical mechanics near equilibrium: computing higher order terms." Nonlinearity, to appear. [20] The names of H. Nyquist, M.S. Green, H.B. Callen, and R. Kubo are attached to this result. For a general discussion see for instance [18] Ch. VIII. [21] See [18] Ch. VIII.3. [22] See for example Section X.8 of M. Reed and B. Simon. {\it Methods of Mathematical Physics II.} Academic Press, New York, 1975. A concise presentation is also given in Section {\bf 378B} of the {\it Encyclopedic Dictionary of Mathematics (2nd English ed.)} MIT Press, Cambridge MA, 1987. [23] Eric Siggia pointed out to the author that a change in the nature of singularities of the susceptibility when one moves away from equilibrium is implicit in P.C. Martin et al. Phys. Rev. A.{\bf 8},423-437(1973). \end