This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Ergodic observables in non-ergodic systems: the example of the harmonic chain

Marco Baldovin Institute for Complex Systems - CNR, P.le Aldo Moro 2, 00185, Rome, Italy Université Paris-Saclay, CNRS, LPTMS,530 Rue André Rivière, 91405, Orsay, France    Raffaele Marino Dipartimento di Fisica e Astronomia, Universitá degli Studi di Firenze, Via Giovanni Sansone 1, 50019, Sesto Fiorentino, Italy    Angelo Vulpiani Dipartimento di Fisica, Sapienza Universitá di Roma, P.le Aldo Moro 5, 00185, Rome, Italy
(December 20, 2024)
Abstract

In the framework of statistical mechanics the properties of macroscopic systems are deduced starting from the laws of their microscopic dynamics. One of the key assumptions in this procedure is the ergodic property, namely the equivalence between time averages and ensemble averages. This property can be proved only for a limited number of systems; however, as proved by Khinchin [1], weak forms of it hold even in systems that are not ergodic at the microscopic scale, provided that extensive observables are considered.

Here we show in a pedagogical way the validity of the ergodic hypothesis, at a practical level, in the paradigmatic case of a chain of harmonic oscillators. By using analytical results and numerical computations, we provide evidence that this non-chaotic integrable system shows ergodic behavior in the limit of many degrees of freedom. In particular, the Maxwell-Boltzmann distribution turns out to fairly describe the statistics of the single particle velocity. A study of the typical time-scales for relaxation is also provided.

I Introduction

Since the seminal works by Maxwell, Boltzmann and Gibbs, statistical mechanics has been conceived as a link between the microscopic world of atoms and molecules and the macroscopic one where everyday phenomena are observed [2]. The same physical system can be described, in the former, by an enormous number of degrees of freedom NN (of the same order of the Avogadro number) or, in the latter, in terms of just a few thermodynamics quantities. Statistical mechanics is able to describe in a precise way the behavior of these macroscopic observables, by exploiting the knowledge of the laws for the microscopic dynamics and classical results from probability theory. Paradigmatic examples of this success are, for instance, the possibility to describe the probability distribution of the single-particle velocity in an ideal gas [2, 3], as well as the detailed behavior of phase transitions [4, 5] and critical phenomena [6, 7]. In some cases (Bose-Einstein condensation [8], absolute negative temperature systems [9, 10, 11]) the results of statistical mechanics were able to predict states of the matter that were never been observed before.

In spite of the above achievements, a complete consensus about the actual reasons for such a success has not been yet reached within the statistical mechanics community. The main source of disagreement is the so-called “ergodic hypothesis”, stating that time averages (the ones actually measured in physics experiments) can be computed as ensemble averages (the ones appearing in statistical mechanics calculations). Specifically, a system is called ergodic when the value of the time average of any observable is the same as the one obtained by taking the average over the energy surface, using the microcanonical distribution [12]. It is worth mentioning that, from a mathematical point of view, ergodicity holds only for a small amount of physical systems: the KAM theorem [13, 14, 15] establishes that, strictly speaking, non-trivial dynamics cannot be ergodic. Nonetheless, the ergodic hypothesis happens to work extremely well also for non-ergodic systems. It provides results in perfect agreement with the numerical and experimental observations, as seen in a wealth of physical situations [16, 17, 18].

Different explanations for this behavior have been provided. Without going into the details of the controversy, three main points of view can be identified: (i) the “classical” school based on the seminal works by Boltzmann and the important contribution of Khinchin, where the main role is played by the presence of many degrees of freedom in the considered systems  [1, 19, 20, 21, 22, 23]; (ii) those, like the Prigogine school, who recognize in the chaotic nature of the microscopic evolution the dominant ingredient [24, 25, 26, 27]; (iii) the maximum entropy point of view, which does not consider statistical mechanics as a physical theory but as an inference methodology based on incomplete information [28, 29, 30, 31].

The main aim of the present contribution is to clarify, at a pedagogical level, how ergodicity manifests itself for some relevant degrees of freedom, in non-ergodic systems. We say that ergodicity occurs “at a practical level”. To this end, a classical chain of NN coupled harmonic oscillators turns out to be an excellent case study: being an integrable system, it cannot be suspected of being chaotic; still, “practical” ergodicity is recovered for relevant observables, in the limit of N1N\gg 1. We believe that this kind of analysis supports the traditional point of view of Boltzmann, which identifies the large number of degrees of freedom as the reason for the occurrence of ergodic behavior for physically relevant observables. Of course, these conclusions are not new. In the works of Khinchin (and then Mazur and van der Lynden) [1, 32, 33, 34, 35] it is rigorously shown that the ergodic hypothesis holds for observables that are computed as an average over a finite fraction of the degrees of freedom, in the limit of N1N\gg 1. Specifically, if we limit our interest to this particular (but non-trivial) class of observables, the ergodic hypothesis holds for almost all initial conditions (but for a set whose probability goes to zero for NN\to\infty), within arbitrary accuracy. In addition, several numerical results for weakly non-linear systems  [36, 37, 38], as well as integrable systems [14, 39], present strong indications of the poor role of chaotic behaviour, implying the dominant relevance of the many degrees of freedom. Still, we think it may be useful, at least from a pedagogical point of view, to analyze an explicit example where analytical calculations can be made (to some extent), without losing physical intuition about the model.

The rest of this paper is organized as follows. In Section II we briefly recall basic facts about the chosen model, to fix the notation and introduce some formulae that will be useful in the following. Section III contains the main result of the paper. We present an explicit calculation of the empirical distribution of the single-particle momentum, given a system starting from out-of-equilibrium initial conditions. We show that in this case the Maxwell-Boltzmann distribution is an excellent approximation in the NN\to\infty limit. Section IV is devoted to an analysis of the typical times at which the described ergodic behavior is expected to be observed; a comparison with a noisy version of the model (which is ergodic by definition) is also provided. In Section V we draw our final considerations.

II Model

We are interested in the dynamics of a one-dimensional chain of NN classical harmonic oscillators of mass mm. The state of the system is described by the canonical coordinates {qj(t),pj(t)}\{q_{j}(t),p_{j}(t)\} with j=1,..,Nj=1,..,N; here pj(t)p_{j}(t) identifies the momentum of the jj-th oscillator at time tt, while qj(t)q_{j}(t) represents its position. The jj-th and the (j+1)(j+1)-th particles of the chain interact through a linear force of intensity κ|qj+1qj|\kappa|q_{j+1}-q_{j}|, where κ\kappa is the elastic constant. We will assume that the first and the last oscillator of the chain are coupled to virtual particles at rest, with infinite inertia (the walls), i.e. q0qN+10q_{0}\equiv q_{N+1}\equiv 0.

The Hamiltonian of the model reads therefore

(𝐪,𝐩)=j=0Npj22m+j=0Nmω022(qj+1qj)2,\mathcal{H}(\mathbf{q},\mathbf{p})=\sum_{j=0}^{N}\frac{p_{j}^{2}}{2m}+\sum_{j=0}^{N}\frac{m\omega_{0}^{2}}{2}(q_{j+1}-q_{j})^{2}, (1)

where ω0=κm\omega_{0}=\sqrt{\frac{\kappa}{m}}.

Such a system is integrable and, therefore, trivially non-ergodic. This can be easily seen by considering the normal modes of the chain, i.e. the set of canonical coordinates

Qk=2N+1j=1Nqjsin(jkπN+1)Q_{k}=\sqrt{\frac{2}{N+1}}\sum_{j=1}^{N}\,q_{j}\sin\left(\frac{jk\pi}{N+1}\right) (2a)
Pk=2N+1j=1Npjsin(jkπN+1),P_{k}=\sqrt{\frac{2}{N+1}}\sum_{j=1}^{N}\,p_{j}\sin\left(\frac{jk\pi}{N+1}\right)\,, (2b)

with k=1,,Nk=1,...,N. Indeed, by rewriting the Hamiltonian in terms of these new canonical coordinates one gets

(𝐐,𝐏)=12k=1N(Pk2m+ωk2Qk2),\mathcal{H}(\mathbf{Q},\mathbf{P})=\frac{1}{2}\sum_{k=1}^{N}\left(\frac{P_{k}^{2}}{m}+\omega_{k}^{2}Q_{k}^{2}\right), (3)

where the frequencies of the normal modes are given by

ωk=2ω0sin(πk2N+2).\omega_{k}=2\omega_{0}\sin\left(\frac{\pi k}{2N+2}\right)\,. (4)

In other words, the system can be mapped into a collection of independent harmonic oscillators with characteristic frequencies {ωk}\{\omega_{k}\}. This system is clearly non-ergodic, as it admits NN integrals of motion, namely the energies

Ek=12(Pk2m+ωk2Qk2)E_{k}=\frac{1}{2}\left(\frac{P_{k}^{2}}{m}+\omega_{k}^{2}Q_{k}^{2}\right)

associated to the normal modes.

In spite of its apparent simplicity, the above system allows the investigation of some nontrivial aspects of the ergodic hypothesis, and helps clarifying the physical meaning of this assumption.

III Ergodic behavior of the momenta

In this section we analyze the statistics of the single-particle momenta of the chain. We aim to show that they approximately follow a Maxwell-Boltzmann distribution

𝒫MB(p)=β2πmeβp2/2m\mathcal{P}_{MB}(p)=\sqrt{\frac{\beta}{2\pi m}}e^{-\beta p^{2}/2m} (5)

in the limit of large NN, where β\beta is the inverse temperature of the system. With the chosen initial conditions, β=N/Etot\beta=N/E_{tot}. Firstly, extending some classical results by Kac [40, 41], we focus on the empirical distribution of the momentum of one particle, computed from a unique long trajectory, namely

𝒫e(j)(p)=1T0T𝑑tδ(ppj(t)).\mathcal{P}_{e}^{(j)}\left(p\right)={1\over T}\int_{0}^{T}\,dt\,\delta\left(p-p_{j}(t)\right)\,. (6)

Then we consider the marginal probability distribution 𝒫e(p,t)\mathcal{P}_{e}\left(p,t\right) computed from the momenta {pj}\{p_{j}\} of all the particles at a specific time tt, i.e.

𝒫e(p,t)=1Nj=1Nδ(ppj(t)).\mathcal{P}_{e}\left(p,t\right)={1\over N}\sum_{j=1}^{N}\delta\left(p-p_{j}(t)\right)\,. (7)

In both cases we assume that the system is prepared in an atypical initial condition. More precisely, we consider the case in which Qj(0)=0Q_{j}(0)=0, for all jj, and the total energy EtotE_{tot}, at time t=0t=0, is equally distributed among the momenta of the first NN^{\star} normal modes, with 1NN1\ll N^{\star}\ll N:

Pj(0)={2mEtot/Nfor1jN0forN<jN.P_{j}(0)=\begin{cases}\sqrt{2mE_{tot}/N^{\star}}\quad&\text{for}\quad 1\leq j\leq N^{\star}\\ 0\quad&\text{for}\quad N^{\star}<j\leq N\,.\end{cases} (8)

In this case, the dynamics of the first NN^{\star} normal modes is given by

Q(t)\displaystyle Q(t) =2Etotωk2Nsin(ωkt)\displaystyle=\sqrt{\frac{2E_{tot}}{\omega_{k}^{2}N^{\star}}}\sin\left(\omega_{k}t\right) (9)
P(t)\displaystyle P(t) =2mEtotNcos(ωkt).\displaystyle=\sqrt{\frac{2mE_{tot}}{N^{\star}}}\cos\left(\omega_{k}t\right)\,.

III.1 Empirical distribution of single-particle momentum

Our aim is to compute the empirical distribution of the momentum of a given particle pjp_{j}, i.e., the distribution of its values measured in time. This analytical calculation was carried out rigorously by Mazur and Montroll in Ref. [42]. Here, we provide an alternative argument that has the advantage of being more concise and intuitive, in contrast to the mathematical rigour of [42]. Our approach exploits the computation of the moments of the distribution; by showing that they are the same, in the limit of infinite measurement time, as those of a Gaussian, it is possible to conclude that the considered momentum follows the equilibrium Maxwell-Boltzmann distribution. The assumption N1N\gg 1 will enter explicitly the calculation.

The momentum of the jj-th particle can be written as a linear combination of the momenta of the normal modes by inverting Eq. (2b):

pj(t)\displaystyle p_{j}(t) =2N+1k=1Nsin(jkπN+1)Pk(t)\displaystyle=\sqrt{\frac{2}{N+1}}\sum_{k=1}^{N}\,\sin\left(\frac{jk\pi}{N+1}\right)P_{k}(t) (10)
=2mEtot(N+1)Nk=1Nsin(kjπN+1)cos(ωkt)\displaystyle=2\sqrt{\frac{mE_{tot}}{(N+1)N^{\star}}}\sum_{k=1}^{N^{\star}}\sin\left(\frac{kj\pi}{N+1}\right)\cos\left(\omega_{k}t\right)

where the ωk\omega_{k}’s are defined by Eq. (4), and the dynamics (9) has been taken into account. The nn-th empirical moment of the distribution is defined as the average pjn¯\overline{p_{j}^{n}} of the nn-th powerof pjp_{j} over a measurement time TT:

pjn¯\displaystyle\overline{p_{j}^{n}} =1T0T𝑑tpjn(t)\displaystyle=\frac{1}{T}\int_{0}^{T}dtp_{j}^{n}(t) (11)
=1T0T𝑑t(CN)nl=1n[kl=1Nsin(kljπN+1)cos(ωklt)]\displaystyle=\frac{1}{T}\int_{0}^{T}dt\,(C_{N^{\star}})^{n}\prod_{l=1}^{n}\left[\sum_{k_{l}=1}^{N^{\star}}\sin\left(\frac{k_{l}j\pi}{N+1}\right)\cos\left(\omega_{k_{l}}t\right)\right]
=(CN)nk1=1Nkn=1Nsin(k1jπN+1)sin(knjπN+1)1T0T𝑑tcos(ωk1t)cos(ωknt)\displaystyle=(C_{N^{\star}})^{n}\sum_{k_{1}=1}^{N^{\star}}\dots\sum_{k_{n}=1}^{N^{\star}}\sin\left(\frac{k_{1}j\pi}{N+1}\right)\dots\sin\left(\frac{k_{n}j\pi}{N+1}\right)\,\frac{1}{T}\int_{0}^{T}dt\,\cos\left(\omega_{k_{1}}t\right)\dots\cos\left(\omega_{k_{n}}t\right)

with

CN=2mEtot(N+1)N.C_{N^{\star}}=2\sqrt{\frac{mE_{tot}}{(N+1)N^{\star}}}\,. (12)

We want to study the integral appearing in the last term of the above equation. To this end it is useful to recall that

12π02π𝑑θcosn(θ)={(n1)!!n!!for n even0for n odd.\frac{1}{2\pi}\int_{0}^{2\pi}d\theta\cos^{n}(\theta)=\begin{cases}\frac{(n-1)!!}{n!!}\quad&\text{for $n$ even}\\ 0\quad&\text{for $n$ odd}\,.\end{cases} (13)

As a consequence, one has

1T0T𝑑tcosn(ωt){(n1)!!n!!for n even0for n odd.\frac{1}{T}\int_{0}^{T}dt\cos^{n}(\omega t)\simeq\begin{cases}\frac{(n-1)!!}{n!!}\quad&\text{for $n$ even}\\ 0\quad&\text{for $n$ odd}\,.\end{cases} (14)

Indeed, we are just averaging over ωT/2π\simeq\omega T/2\pi periods of the integrated function, obtaining the same result we get for a single period, with a correction of the order O((ωT)1)O\left((\omega T)^{-1}\right). This correction comes from the fact that TT is not, in general, an exact multiple of 2π/ω2\pi/\omega. If ω1\omega_{1}, ω2\omega_{2}, …, ωq\omega_{q} are incommensurable (i.e., their ratios cannot be expressed as rational numbers), provided that TT is much larger than (ωjωk)1(\omega_{j}-\omega_{k})^{-1} for each choice of 1k<jq1\leq k<j\leq q, a well known result [40] assures that

1T0T𝑑tcosn1(ω1t)cosnq(ωqt)\displaystyle\frac{1}{T}\int_{0}^{T}dt\cos^{n_{1}}(\omega_{1}t)\cdot...\cdot\cos^{n_{q}}(\omega_{q}t)\simeq (1T0T𝑑tcosn1(ω1t))(1T0T𝑑tcosnq(ω1t))\displaystyle\left(\frac{1}{T}\int_{0}^{T}dt\cos^{n_{1}}(\omega_{1}t)\right)\cdot...\cdot\left(\frac{1}{T}\int_{0}^{T}dt\cos^{n_{q}}(\omega_{1}t)\right) (15)
\displaystyle\simeq (n11)!!n1!!(nq1)!!nq!!if all n’s are even ,\displaystyle\frac{(n_{1}-1)!!}{n_{1}!!}\cdot...\cdot\frac{(n_{q}-1)!!}{n_{q}!!}\,\quad\text{if all $n$'s are even\,,}

where the last step is a consequence of Eq. (14). Instead, if at least one of the nn’s is odd, the above quantity vanishes, again with corrections due to the finite time TT. Since the smallest sfrequency is ω1\omega_{1}, one has that the error is at most of the order O(q(ω1T)1)O(qN/ω0T)O\left(q(\omega_{1}T)^{-1}\right)\simeq O(qN/\omega_{0}T).

Refer to caption
Figure 1: Convergence of the empirical distribution of the single particle momentum. The four panels show histograms for the momentum of the jj-th particle, computed by considering the value of pjp_{j} at integral times 0,1,2,,t0,1,2,...,t. The system is prepared at time 0 in the atypical initial condition described in the text, where only NN^{\star} normal modes are excited. Different values of tt are considered: when the total time is large enough, the distribution approcahes the Maxwell-Boltzmann equilibrium (red curve). Here N=103N=10^{3}, N=0.1NN^{\star}=0.1N, Etot=NE_{tot}=N, ω0=1\omega_{0}=1, j=123j=123.

Let us consider again the integral in the last term of Eq. (11). The ωk\omega_{k}’s are, in general, incommensurable. Therefore, the integral vanishes when nn is odd, since in that case at least one of the {nl}\{n_{l}\}, l=1,,ql=1,...,q, will be odd. When nn is even, the considered quantity is different from zero as soon as the kk’s are pairwise equal, so that n1==nq=2n_{1}=...=n_{q}=2. In the following we will neglect the contribution of terms containing groups of four or more equal kk’s: if nNn\ll N^{\star}, the number of these terms is indeed O(N)\sim O(N^{\star}) times less numerous than the pairings, and it can be neglected if N1N^{\star}\gg 1 (which is one of our assumptions on the initial condition). Calling Ωn\Omega_{n} the set of possible pairings for the vector 𝐤=(k1,,kl)\mathbf{k}=(k_{1},...,k_{l}), we have then

pjn¯(CN2)n𝐤Ωnl=1nsin(kljπN+1),\overline{p_{j}^{n}}\simeq\left(\frac{C_{N^{\star}}}{\sqrt{2}}\right)^{n}\,\sum_{\mathbf{k}\in\Omega_{n}}\prod_{l=1}^{n}\sin\left(\frac{k_{l}j\pi}{N+1}\right)\,, (16)

with an error of O(1/N)O(1/N^{\star}) due to neglecting groups of 4, 6 and so on, and an error O(nN/ω0T)O(nN/\omega_{0}T) due to the finite averaging time TT, as discussed before. Factor 2n/22^{-n/2} comes from the explicit evaluation of Eq. (15) .

At fixed jj, we need now to estimate the sums appearing in the above equation, recalling that the kk’s are pairwise equal. If j>NNj>\frac{N}{N^{\star}}, the arguments of the periodic functions can be thought as if independently extracted from a uniform distribution 𝒫(k)=1/N\mathcal{P}(k)=1/N^{\star}. One has:

sin2(kjπN+1)k=1N1Nsin2(kjπN+1)12πππ𝑑θsin2(θ)=12,\left\langle\sin^{2}\left(\frac{kj\pi}{N+1}\right)\right\rangle\simeq\sum_{k=1}^{N^{\star}}\frac{1}{N^{\star}}\sin^{2}\left(\frac{kj\pi}{N+1}\right)\simeq\frac{1}{2\pi}\int_{-\pi}^{\pi}d\theta\,\sin^{2}(\theta)=\frac{1}{2}\,, (17)

and

l=1nsin(kljπN+1)2n/2,\left\langle\prod_{l=1}^{n}\sin\left(\frac{k_{l}j\pi}{N+1}\right)\right\rangle\simeq 2^{-n/2}\,, (18)

if 𝐤Ωn\mathbf{k}\in\Omega_{n}. As a consequence

pjn¯\displaystyle\overline{p_{j}^{n}} (CN2)n(N)n/2𝒩(Ωn)(mEtotN+1)n/2𝒩(Ωn),\displaystyle\simeq\left(\frac{C_{N^{\star}}}{2}\right)^{n}(N^{\star})^{n/2}\,\mathcal{N}(\Omega_{n})\simeq\left(\frac{mE_{tot}}{N+1}\right)^{n/2}\mathcal{N}(\Omega_{n})\,, (19)

where 𝒩(Ωn)\mathcal{N}(\Omega_{n}) is the number of ways in which we can choose the pairings. These are the moments of a Gaussian distribution with zero average and mEtotN+1\frac{mE_{tot}}{N+1} variance.

Summarising, it is possible to show that, if nNNn\ll N^{\star}\ll N, the first nn moments of the distribution are those of a Maxwell-Boltzmann distribution. In the limit of N1N\gg 1 with N/NN^{\star}/N fixed, the Gaussian distribution is thus recovered up to an arbitrary number of moments. Let us note that the assumption Qj(0)=0Q_{j}(0)=0, while allowing to make the calculations clearer, is not really relevant. Indeed, if Qj(0)0Q_{j}(0)\neq 0 we can repeat the above computation while replacing ωkt\omega_{k}t by ωkt+ϕk\omega_{k}t+\phi_{k}, where the phases ϕk\phi_{k} take into account the initial conditions.

Fig. 1 shows the standardized histogram of the relative frequencies of single-particle velocities of the considered system, in the N1N\gg 1 limit, with the initial conditions discussed before. As expected, the shape of the distribution tends to a Gaussian in the large-time limit.

III.2 Distribution of momenta at a given time

A similar strategy can be used to show that, at any given time tt large enough, the histogram of the momenta is well approximated by a Gaussian distribution. Again, the large number of degrees of freedom plays an important role. We want to compute the empirical moments

pn(t)=1Nj=1Npjn(t),\left\langle p^{n}\right\rangle(t)=\frac{1}{N}\sum_{j=1}^{N}p_{j}^{n}(t)\,, (20)

defined according to the distribution 𝒫e(j)(p)\mathcal{P}_{e}^{(j)}\left(p\right) introduced by Eq. (6). Using again Eq. (10) we get

pn(t)=\displaystyle\left\langle p^{n}\right\rangle(t)= 1Nj=1N(CN)n[k=1Nsin(kjπN+1)cos(ωkt)]n\displaystyle\frac{1}{N}\sum_{j=1}^{N}(C_{N^{\star}})^{n}\left[\sum_{k=1}^{N^{\star}}\sin\left(\frac{kj\pi}{N+1}\right)\cos\left(\omega_{k}t\right)\right]^{n} (21)
=\displaystyle= 1N(CN)nk1Nkn=1N[l=1Ncos(ωklt)]j=1Nsin(k1jπN+1)sin(knjπN+1).\displaystyle\frac{1}{N}(C_{N^{\star}})^{n}\sum_{k_{1}}^{N^{\star}}\dots\sum_{k_{n}=1}^{N^{\star}}\left[\prod_{l=1}^{N}\cos\left(\omega_{k_{l}}t\right)\right]\sum_{j=1}^{N}\sin\left(\frac{k_{1}j\pi}{N+1}\right)\dots\sin\left(\frac{k_{n}j\pi}{N+1}\right)\,.

Reasoning as before, we see that the sum over jj vanishes in the large NN limit unless the kk’s are pairwise equal. Again, we neglect the terms including groups of 4 or more equal kk’s, assuming that nNn\ll N^{\star}, so that their relative contribution is O(1/N)O(1/N^{\star}). That sum selects paired values of kk for the product inside the square brackets, and we end with

pn(t)1N(CN)n𝐤Ωn[l=1Ncos(ωklt)].\left\langle p^{n}\right\rangle(t)\simeq\frac{1}{N}(C_{N^{\star}})^{n}\sum_{\mathbf{k}\in\Omega_{n}}\left[\prod_{l=1}^{N}\cos\left(\omega_{k_{l}}t\right)\right]\,. (22)

If tt is “large enough” (we will come back to this point in the following section), different values of ωkl\omega_{k_{l}} lead to completely uncorrelated values of cos(ωklt)\cos(\omega_{k_{l}}t). Hence, as before, we can consider the arguments of the cosines as extracted from a uniform distribution, obtaining

pn(t)(CN2)n(N)n/2𝒩(Ωn)(mEtotN+1)n/2𝒩(Ωn).\left\langle p^{n}\right\rangle(t)\simeq\left(\frac{C_{N^{\star}}}{2}\right)^{n}(N^{\star})^{n/2}\,\mathcal{N}(\Omega_{n})\simeq\left(\frac{mE_{tot}}{N+1}\right)^{n/2}\mathcal{N}(\Omega_{n})\,. (23)

These are again the moments of the equilibrium Maxwell-Boltzmann distribution. We had to assume nNn\ll N^{\star}, meaning that a Gaussian distribution is recovered only in the limit of large number of degrees of freedom.

The empirical distribution can be compared with the Maxwell-Boltzmann by looking at the Kullback-Leibler divergence K(𝒫e(p,t),𝒫MB(p))K(\mathcal{P}_{e}(p,t),\mathcal{P}_{MB}(p)) which provides a sort of distance between the empirical 𝒫e(p,t)\mathcal{P}_{e}(p,t) and the Maxwell-Boltzmann:

K[𝒫e(p,t),𝒫MB(p)]=𝒫e(p,t)ln𝒫MB(p)𝒫e(p,t)dp.K[\mathcal{P}_{e}(p,t),\mathcal{P}_{MB}(p)]=-\int\mathcal{P}_{e}(p,t)\ln\frac{\mathcal{P}_{MB}(p)}{\mathcal{P}_{e}(p,t)}dp. (24)

Figure 2 shows how the Kullback-Leibler divergences approach their equilibrium limit, for different values of NN. As expected, the transition happens on a time scale that depends linearly on NN.

A comment is in order: even if this behaviour may look similar to the H-Theorem for diluited gases, such a resemblance is only superficial. Indeed, while in the cases of diluited gases the approach to the Maxwell-Boltzmann is due to the collisions among different particles that actually exchange energy and momentum, in the considered case the “thermalization” is due to a dephasing mechanism.

Refer to caption
Figure 2: Kullback-Leibler divergence between the empirical distribution of the momenta (7) and the Maxwell-Boltzmann (5), as a function of time. Panel (a) shows the behaviour of K[𝒫e(p,t),𝒫MB(p)]K[\mathcal{P}_{e}(p,t),\mathcal{P}_{MB}(p)], defined by Eq. (24), for different values of NN. As before, the system is initialized at time t=0t=0 in a far-from-equilibrium state where only a small fraction of the modes is excited. The transition to equilibrium takes place on a time scale that is proportional to NN, as expected from the argument in the main text. Panels (b)-(d) show the detail of the empirical distribution 𝒫e(p,t)\mathcal{P}_{e}(p,t) at different times, for the N=104N=10^{4} case. Unlike in Fig. 1, here the histogram is built including the momenta of all particles in the system, at the considered time. Parameters: N=0.1NN^{\star}=0.1N, Etot=NE_{tot}=N, ω0=1\omega_{0}=1.

IV Analysis of the time scales

In the previous section, when considering the distribution of the momenta at a given time, we had to assume that tt was “large enough” in order for our approximations to hold. In particular we required cos(ωk1t)\cos(\omega_{k_{1}}t) and cos(ωk2t)\cos(\omega_{k_{2}}t) to be uncorrelated as soon as k1k2k_{1}\neq k_{2}. Such a dephasing hypothesis amounts to asking that

|ωk1tωk2t|>2πc,|\omega_{k_{1}}t-\omega_{k_{2}}t|>2\pi c\,, (25)

where cc is the number of phases by which the two oscillator have to differ before they can be considered uncorrelated. The constant cc may be much larger than 1, but it is not expected to depend strongly on the size NN of the system. In other words, we require

t>c|ωk1ωk2|t>\frac{c}{|\omega_{k_{1}}-\omega_{k_{2}}|} (26)

for each choice of k1k_{1} and k2k_{2}. To estimate this typical relaxation time, we need to pick the minimum value of |ωk1ωk2||\omega_{k_{1}}-\omega_{k_{2}}| among the possible pairs (k1,k2)(k_{1},k_{2}). This term is minimized when k1=k~k_{1}=\tilde{k} and k2=k~1k_{2}=\tilde{k}-1 (or vice-versa), with k~\tilde{k} chosen such that ωk~ωk~1\omega_{\tilde{k}}-\omega_{\tilde{k}-1} is minimum. In the large-NN limit this quantity is approximated by

ωk~ωk~1=ω0sin(k~π2N+2)ω0sin(k~ππ2N+2)ω0cos(k~π2N+2)π2N+2,\omega_{\tilde{k}}-\omega_{\tilde{k}-1}=\omega_{0}\sin\left(\frac{\tilde{k}\pi}{2N+2}\right)-\omega_{0}\sin\left(\frac{\tilde{k}\pi-\pi}{2N+2}\right)\simeq\omega_{0}\cos\left(\frac{\tilde{k}\pi}{2N+2}\right)\frac{\pi}{2N+2}\,, (27)

which is minimum when k~\tilde{k} is maximum, i.e. for k~=N\tilde{k}=N^{\star}.

Dephasing is thus expected to occur at

t>4cNω0cos(Nπ2N),t>\frac{4cN}{\omega_{0}\cos\left(\frac{N^{\star}\pi}{2N}\right)}\,, (28)

i.e. t>4cN/ω0t>4cN/\omega_{0} in the N/N1N^{\star}/N\ll 1 limit.

It is instructive to compare this characteristic time with the typical relaxation time of the “damped” version of the considered system. For doing so, we assume that our chain of oscillators is now in contact with a viscous medium which acts at the same time as a thermal bath and as a source of viscous friction. By considering the (stochastic) effect of the medium, one gets the Klein-Kramers stochastic process [43, 44]

qjt=pjmpjt=ω02(qj+12qj+qj1)γpj+2γTξj\begin{split}&\frac{\partial q_{j}}{\partial t}=\frac{p_{j}}{m}\\ &\frac{\partial p_{j}}{\partial t}=\omega_{0}^{2}(q_{j+1}-2q_{j}+q_{j-1})-\gamma p_{j}+\sqrt{2\gamma T}\xi_{j}\,\end{split} (29)

where γ\gamma is the damping coefficient and TT is the temperature of the thermal bath (we are taking the Boltzmann constant kBk_{B} equal to 1). Here the {ξj}\{\xi_{j}\} are time-dependent, delta-correlated Gaussian noises such that ξj(t)ξk(t)=δjkδ(tt)\left\langle\xi_{j}(t)\xi_{k}(t^{\prime})\right\rangle=\delta_{jk}\delta(t-t^{\prime}). Such a system is surely ergodic and the stationary probability distribution is the familiar equilibrium one

𝒫s(𝐪,𝐩)eH(𝐪,𝐩)T.\mathcal{P}_{s}(\mathbf{q},\mathbf{p})\propto e^{-\frac{H(\mathbf{q},\mathbf{p})}{T}}. (30)

Also in this case we can consider the evolution of the normal modes. By taking into account Eqs. (2) and (29) one gets

Qk˙\displaystyle\dot{Q_{k}} =1mPk\displaystyle=\frac{1}{m}P_{k} (31)
Pk˙\displaystyle\dot{P_{k}} =ωk2QkγmP+2γTζk\displaystyle=-\omega_{k}^{2}Q_{k}-\frac{\gamma}{m}P+\sqrt{2\gamma T}\zeta_{k}

where the {ζk}\{\zeta_{k}\} are again delta-correlated Gaussian noises. It is important to notice that also in this case the motion of the modes is independent (i.e. the friction does not couple normal modes with different kk); nonetheless, the system is ergodic, because the presence of the noise allows it to explore, in principle, any point of the phase-space.

The Fokker-Planck equation for the evolution of the probability density function 𝒫(Qk,Pk,t)\mathcal{P}\left(Q_{k},P_{k},t\right) of the kk-th normal mode can be derived using standard methods [43]:

t𝒫=Qk(Pk𝒫)+Pk(ωk2Qk𝒫+γmPk𝒫)+γTPk2𝒫.\partial_{t}\mathcal{P}=-\partial_{Q_{k}}\left(P_{k}\mathcal{P}\right)+\partial_{P_{k}}\left(\omega_{k}^{2}Q_{k}\mathcal{P}+\frac{\gamma}{m}P_{k}\mathcal{P}\right)+\gamma T\partial_{P_{k}}^{2}\mathcal{P}\,. (32)

The above equation allows to compute also the time dependence of the correlation functions of the system in the stationary state. In particular one gets

ddtQk(t)Qk(0)=1mPk(t)Qk(0)\frac{d}{dt}\left\langle Q_{k}(t)Q_{k}(0)\right\rangle=\frac{1}{m}\left\langle P_{k}(t)Q_{k}(0)\right\rangle (33)

and

ddtPk(t)Qk(0)ωk2mQk(t)Qk(0)γmPk(t)Qk(0),\frac{d}{dt}\left\langle P_{k}(t)Q_{k}(0)\right\rangle-\omega_{k}^{2}m\left\langle Q_{k}(t)Q_{k}(0)\right\rangle-\frac{\gamma}{m}\left\langle P_{k}(t)Q_{k}(0)\right\rangle\,, (34)

which, once combined together, lead to

d2dt2Qk(t)Qk(0)+γmddtQk(t)Qk(0)+ωk2Qk(t)Qk(0)=0.\frac{d^{2}}{dt^{2}}\left\langle Q_{k}(t)Q_{k}(0)\right\rangle+\frac{\gamma}{m}\frac{d}{dt}\left\langle Q_{k}(t)Q_{k}(0)\right\rangle+\omega_{k}^{2}\left\langle Q_{k}(t)Q_{k}(0)\right\rangle=0\,. (35)

For ωk<γ/m\omega_{k}<\gamma/m the solution of this equation admits two characteristic frequencies ω~±\tilde{\omega}_{\pm}, namely

ω~±=γ2m(1±1m2ωk2γ2).\tilde{\omega}_{\pm}=\frac{\gamma}{2m}\left(1\pm\sqrt{1-\frac{m^{2}\omega_{k}^{2}}{\gamma^{2}}}\right). (36)

In the limit ωkγ/m\omega_{k}\ll\gamma/m one has therefore

ω~m4γωk2mω02π2k2γN2.\tilde{\omega}_{-}\simeq\frac{m}{4\gamma}\omega_{k}^{2}\simeq\frac{m\omega_{0}^{2}\pi^{2}k^{2}}{\gamma N^{2}}\,. (37)

Therefore, as a matter of fact, even in the damped case the system needs a time that scales as N2N^{2} in order to get complete relaxation for the modes. As we discussed before, the dephasing mechanism that guarantees for “practical” ergodicity in the deterministic version is instead expected to occur on time scales of order O(N)O(N).

V Conclusions

The main aim of this paper was to expose, at a pedagogical level, some aspects of the foundation of statistical mechanics, namely the role of ergodicity for the validity of the statistical approach to the study of complex systems.

We analyzed a chain of classical harmonic oscillators (i.e. a paradigmatic example of integrable system, which cannot be suspected to show chaotic behaviour). By extending some well-known results by Kac [40], we showed that the Maxwell-Bolzmann distribution approximates with arbitrary precision (in the limit of large number of degrees of freedom) the empirical distribution of the momenta of the system, after a dephasing time which scales with the size of the chain. This is true also for quite pathological initial conditions, where only a small fraction of the normal modes is excited at time t=0t=0. The scaling of the typical dephasing time with the number of oscillators NN may appear as a limit of our argument, since this time will diverge in the thermodynamic limit; on the other hand one should consider, as explicitely shown before, that the damped version of this model (which is ergodic by definition) needs times of the order O(N2)O(N^{2}) to reach thermalization for each normal mode.

This comparison clearly shows that the effective thermalization observed in large systems has little to do with the mathematical concept of ergodicity, and it is instead related to the large number of components concurring to define the global observales that are usually taken into account (in our case, the large number of normal modes that define the momentum of a single particle). When these components cease to be in phase, the predictions of statistical mechanics start to be effective; this can be observed even in integrable systems, without need for the mathematical notion of ergodicity to hold.

In other words, we believe that the present work give further evidence of the idea (which had been substantiated mathematically by Khinchin, Mazur and van der Linden) that the most relevant ingredient of statistical mechanics is the large number of degrees of freedom, and the global nature of the observables that are typically taken into account.

Acknowledgements

RM is supported by #NEXTGENERATIONEU (NGEU) and funded by the Ministry of University and Research (MUR), National Recovery and Resilience Plan (NRRP), project MNESYS (PE0000006) ”A Multiscale integrated approach to the study of the nervous system in health and disease” (DN. 1553 11.10.2022).

References

  • Khinchin [1949] A. I. Khinchin, Mathematical foundations of statistical mechanics (Courier Corporation, 1949).
  • Huang [2009] K. Huang, Introduction to statistical physics (CRC press, 2009).
  • Peliti and Epstein [2003] L. Peliti and M. Epstein, Statistical Mechanics in a Nutshell (Princeton University Press, 2003).
  • Ehrenfest and Ehrenfest [1990] P. Ehrenfest and T. Ehrenfest, The conceptual foundations of the statistical approach in mechanics (Courier Corporation, 1990).
  • Di Castro and Raimondi [2015] C. Di Castro and R. Raimondi, Statistical mechanics and applications in condensed matter (Cambridge University Press, 2015).
  • Ma [1985] S.-K. Ma, Statistical mechanics (World Scientific Publishing Company, 1985).
  • Caracciolo et al. [2021] S. Caracciolo, R. Fabbricatore, M. Gherardi, R. Marino, G. Parisi, and G. Sicuro, Criticality and conformality in the random dimer model, Phys. Rev. E 103, 042127 (2021).
  • Anderson et al. [1995] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell, Observation of Bose-Einstein condensation in a dilute atomic vapor, Science 269, 198 (1995).
  • Ramsey [1956] N. F. Ramsey, Thermodynamics and statistical mechanics at negative absolute temperatures, Physical Review 103, 20 (1956).
  • Braun et al. [2013] S. Braun, J. P. Ronzheimer, M. Schreiber, S. S. Hodgman, T. Rom, I. Bloch, and U. Schneider, Negative absolute temperature for motional degrees of freedom, Science 339, 52 (2013).
  • Baldovin et al. [2021a] M. Baldovin, S. Iubini, R. Livi, and A. Vulpiani, Statistical mechanics of systems with negative temperature, Phys. Rep. 923, 1 (2021a).
  • Lebowitz and Penrose [1973] J. L. Lebowitz and O. Penrose, Modern ergodic theory, Physics Today 26, 23 (1973).
  • Arnold [2009] V. I. Arnold, Proof of a theorem of A N Kolmogorov on the invariance of quasi-periodic motions under small perturbations of the Hamiltonian, Collected Works: Representations of Functions, Celestial Mechanics and KAM Theory, 1957–1965 , 267 (2009).
  • Arnold [2013] V. I. Arnold, Mathematical methods of classical mechanics, Vol. 60 (Springer Science & Business Media, 2013).
  • Arnold et al. [1988] V. Arnold, V. Kozlov, and A. Neishtadt, Dynamical Systems III, Encyclopaedia of Mathematical Sciences, Vol. 3 (1988).
  • Livi et al. [1987] R. Livi, M. Pettini, S. Ruffo, and A. Vulpiani, Chaotic behavior in nonlinear Hamiltonian systems and equilibrium statistical mechanics, Journal of statistical physics 48, 539 (1987).
  • Baldovin et al. [2021b] M. Baldovin, A. Vulpiani, and G. Gradenigo, Statistical mechanics of an integrable system, Journal of Statistical Physics 183, 41 (2021b).
  • Cocciaglia et al. [2022] N. Cocciaglia, A. Vulpiani, and G. Gradenigo, Thermalization without chaos in harmonic systems, Physica A: Statistical Mechanics and its Applications 601, 127581 (2022).
  • Schrödinger [1944] E. Schrödinger, The statistical law in nature, Nature 153, 704 (1944).
  • Darrigol [2018] O. Darrigol, Atoms, Mechanics, and Probability: Ludwig Boltzmann’s Statistico-Mechanical Writings - An Exegesis (Oxford University Press, 2018).
  • Boltzmann [2022] L. Boltzmann, Lectures on gas theory (Univ of California Press, 2022).
  • Boltzmann [2012] L. Boltzmann, Theoretical physics and philosophical problems: Selected writings, Vol. 5 (Springer Science & Business Media, 2012).
  • Castiglione et al. [2008] P. Castiglione, M. Falcioni, A. Lesne, and A. Vulpiani, Chaos and Coarse Graining in Statistical Mechanics (Cambridge University Press, 2008).
  • Prigogine [1978] I. Prigogine, Time, structure, and fluctuations, Science 201, 777 (1978).
  • Prigogine [2017] I. Prigogine, Non-equilibrium statistical mechanics (Courier Dover Publications, 2017).
  • Prigogine [1996] I. Prigogine, The laws of chaos, Review (Fernand Braudel Center) , 1 (1996).
  • Lando and Flach [2023] G. M. Lando and S. Flach, Thermalization slowing-down in multidimensional Josephson junction networks (2023), arXiv:2306.14398 [cond-mat.stat-mech] .
  • Jaynes [1982] E. T. Jaynes, On the rationale of maximum-entropy methods, Proceedings of the IEEE 70, 939 (1982).
  • Uffink [1995] J. Uffink, Can the maximum entropy principle be explained as a consistency requirement?, Studies in History and Philosophy of Science Part B: Studies in History and Philosophy of Modern Physics 26, 223 (1995).
  • Jaynes [1957] E. T. Jaynes, Information theory and statistical mechanics, Physical review 106, 620 (1957).
  • Jaynes [2003] E. T. Jaynes, Probability theory: The logic of science (Cambridge university press, 2003).
  • Halmos [2017] P. R. Halmos, Lectures on ergodic theory (Courier Dover Publications, 2017).
  • Van der Linden and Mazur [1967] J. Van der Linden and P. Mazur, On the asymptotic problem of statistical thermodynamics for a real system ii. equivalence of canonical and pressure ensemble, Physica 36, 491 (1967).
  • Mazur and Van der Linden [1963] P. Mazur and J. Van der Linden, Asymptotic form of the structure function for real systems, Journal of Mathematical Physics 4, 271 (1963).
  • Casetti et al. [1997] L. Casetti, M. Cerruti-Sola, M. Pettini, and E. Cohen, The fermi-pasta-ulam problem revisited: Stochasticity thresholds in nonlinear hamiltonian systems, Physical Review E 55, 6566 (1997).
  • Chakraborty and Mallik [2001] G. Chakraborty and A. Mallik, Dynamics of a weakly non-linear periodic chain, International Journal of Non-Linear Mechanics 36, 375 (2001).
  • Arnold et al. [1978] L. Arnold, W. Horsthemke, and R. Lefever, White and coloured external noise and transition phenomena in nonlinear systems, Zeitschrift für Physik B Condensed Matter 29, 367 (1978).
  • Arnold and Lefever [2012] L. Arnold and R. Lefever, Stochastic Nonlinear Systems in Physics, Chemistry, and Biology: Proceedings of the Workshop Bielefeld, Fed. Rep. of Germany, October 5–11, 1980, Vol. 8 (Springer Science & Business Media, 2012).
  • Goldstein et al. [2002] H. Goldstein, C. Poole, and J. Safko, Classical mechanics (2002).
  • Kac [1959] M. Kac, Statistical Independence in Probability, Analysis, and Number Theory, Carus Mathematical Monographs (Mathematical Association of America, 1959).
  • Ford et al. [1965] G. Ford, M. Kac, and P. Mazur, Statistical mechanics of assemblies of coupled oscillators, Journal of Mathematical Physics 6, 504 (1965).
  • Mazur and Montroll [1960] P. Mazur and E. Montroll, Poincaré cycles, ergodicity, and irreversibility in assemblies of coupled harmonic oscillators, Journal of mathematical physics 1, 70 (1960).
  • Gardiner [1985] C. W. Gardiner, Handbook of stochastic methods (Springer, 1985).
  • Marino and Aurell [2016] R. Marino and E. Aurell, Advective-diffusive motion on large scales from small-scale dynamics with an internal symmetry, Phys. Rev. E 93, 062147 (2016).