Processing math: 100%
Research article Special Issues

Large deviations for a binary collision model: energy evaporation

  • We analyze the large deviations for a discrete energy Kac-like walk. In particular, we exhibit a path, with probability exponentially small in the number of particles, that looses energy.

    Citation: Giada Basile, Dario Benedetto, Emanuele Caglioti, Lorenzo Bertini. Large deviations for a binary collision model: energy evaporation[J]. Mathematics in Engineering, 2023, 5(1): 1-12. doi: 10.3934/mine.2023001

    Related Papers:

    [1] Kentaro Fujie, Jie Jiang . A note on construction of nonnegative initial data inducing unbounded solutions to some two-dimensional Keller–Segel systems. Mathematics in Engineering, 2022, 4(6): 1-12. doi: 10.3934/mine.2022045
    [2] Lina Zhao, Eun-Jae Park . A locally conservative staggered least squares method on polygonal meshes. Mathematics in Engineering, 2024, 6(2): 339-362. doi: 10.3934/mine.2024014
    [3] Ansgar Jüngel, Ulisse Stefanelli, Lara Trussardi . A minimizing-movements approach to GENERIC systems. Mathematics in Engineering, 2022, 4(1): 1-18. doi: 10.3934/mine.2022005
    [4] José Antonio Vélez-Pérez, Panayotis Panayotaros . Wannier functions and discrete NLS equations for nematicons. Mathematics in Engineering, 2019, 1(2): 309-326. doi: 10.3934/mine.2019.2.309
    [5] Massimiliano Giona, Luigi Pucci . Hyperbolic heat/mass transport and stochastic modelling - Three simple problems. Mathematics in Engineering, 2019, 1(2): 224-251. doi: 10.3934/mine.2019.2.224
    [6] Mario Pulvirenti . On the particle approximation to stationary solutions of the Boltzmann equation. Mathematics in Engineering, 2019, 1(4): 699-714. doi: 10.3934/mine.2019.4.699
    [7] Alessia Nota, Juan J. L. Velázquez . Homoenergetic solutions of the Boltzmann equation: the case of simple-shear deformations. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023019
    [8] Lorenzo Pistone, Sergio Chibbaro, Miguel D. Bustamante, Yuri V. Lvov, Miguel Onorato . Universal route to thermalization in weakly-nonlinear one-dimensional chains. Mathematics in Engineering, 2019, 1(4): 672-698. doi: 10.3934/mine.2019.4.672
    [9] Giacomo Canevari, Arghir Zarnescu . Polydispersity and surface energy strength in nematic colloids. Mathematics in Engineering, 2020, 2(2): 290-312. doi: 10.3934/mine.2020015
    [10] Miyuki Koiso . Stable anisotropic capillary hypersurfaces in a wedge. Mathematics in Engineering, 2023, 5(2): 1-22. doi: 10.3934/mine.2023029
  • We analyze the large deviations for a discrete energy Kac-like walk. In particular, we exhibit a path, with probability exponentially small in the number of particles, that looses energy.



    Large deviations associated to Boltzmann-type equations have been the object of recent investigations. The most challenging case of Newtonian dynamics of hard spheres in the Boltzmann-Grad limit has been heuristically discussed in [5] and rigorously, by means of cluster expansion, in [4]. The case of microscopic stochastic dynamics has been originally analyzed in [8], where a large deviation upper bound is derived. In [15] a large deviation principle is obtained for a space inhomogeneous model with a finite set of velocities. More recently, in [2] it is considered a homogeneous model which conserves momentum but not energy. The large deviation upper bound is achieved, while the lower bound is obtained for a restricted class of paths. In [6] the analogous results are provided for the Kac walk, which conserves also the energy.

    We emphasize that, except in the case of bounded velocities, a proof of a large deviation principle with matching upper and lower bound is still missing, even in the homogeneous case. A key issue is the possible occurrence of macroscopic paths with finite rate function that violate the conservation of the energy. A class of examples has been constructed in [6] by exploiting the solutions of homogeneous Boltzmann equations provided by Lu and Wennberg in [10], for which the energy is increasing.

    Here we consider a Kac-like microscopic dynamics with discrete energy, that is inspired by the so-called KMP model [3,7], and described as follows: N particles, with equally spaced energy levels, evolve via random binary collisions such that in each collision the total energy is preserved. At the kinetic level the one-particle energy distribution evolves according to a discrete homogeneous Boltzmann equation, that is an infinite system of coupled ordinary differential equations. We focus on the large deviation properties of the pair empirical measure and flux and propose a candidate rate function I when the initial distribution of the energies satisfies the micro-canonical constraint, i.e., the total energy is fixed. By the arguments in [1,2,6], it can be shown that the large deviation upper bound holds with rate function I, and a matching lower bound can be proven for a restricted class of paths which conserve the energy. Our main novel point is the construction of a path (ˉf,ˉQ) which looses the energy and whose probability is exponentially small with rate I(ˉf,ˉQ). This result quantifies the probability to violate the conservation of the energy, which is exponentially small in the number of particles N. It can be compared with the result in [14] where, in the contest of the derivation of incompressible Navier-Stokes equations from stochastic lattice gas, it is shown that the probability of violating the incompressibility condition is of the order eN2.

    Referring to [1] for a general discussion, we emphasize that, due to the micro-canonical constraint, at the kinetic level the energy cannot increase and the candidate rate function I is different from the one in [6,8].

    Given N2, a configuration is defined by N energies in N. The configuration space is therefore given by ΣN=NN. Elements of ΣN are denoted by ε:=(ε1,..,εN) and we denote by ΣN,E the configuration space with total energy EN, i.e.,

    ΣN,E:={εNN:Ni=1εi=E}.

    The microscopic dynamics is defining by choosing at random a pair {i,j} and redistributing uniformly the corresponding energies. Therefore we consider the Markov processes on ΣN whose generator acts on bounded functions f:ΣNR as

    LNf=1N{i,j}Lijf,

    where the sum is carried over the unordered pairs {i,j}{1,..,N}, ij, and

    Lijf(ε)=1εi+εj+1εi+εj=0[f(Tijε)f(ε)], (1.1)

    in which

    (Tijε)k:={ifk=iεi+εjifk=jεkotherwise.

    Observe that, for each E>0, the dynamics preserves ΣN,E. Moreover, it is ergodic when restricted to ΣN,E and reversible with respect to the uniform measure on ΣN,E.

    We denote by (ε(t))t0 the continuous time Markov chain generated by LN. In particular, the path ε() is piecewise constant, and the transition probability of its jumps is given in (1.1). Fix hereafter T>0. Given a probability ν on ΣN,E we denote by PNν the law of this chain on the time interval [0,T], when the initial datum is sampled according to ν. Observe that PNν is a probability on the Skorokhod space D([0,T];ΣN,E). As usual if ν=δε for some εΣN,E, the corresponding law is simply denoted by PNε. We refer to [9] for a gentle introduction to continuous time Markov chains.

    Given e(0,+), we denote by Pe(N) the set of probability measures π on N with mean bounded by e, i.e., such that εεπ(ε)e. We consider Pe(N) as a closed subset of the space of probability measure on N equipped with the weak topology. Then Pe(N) endowed with the relative topology is a compact Polish space. Indeed, Pe(N) is the weak closure of the set of probabilities on N with mean e. The empirical measure records the energy of the particles forgetting their labels, for E=Ne it is defined as the map πN:ΣN,EPe(N) given by

    πN(ε):=1NNi=1δεi. (1.2)

    Let D([0,T];Pe(N)) the set of Pe(N)-valued c{á}dl{á}g paths endowed with the Skorokhod topology and the corresponding Borel σ-algebra. With a slight abuse of notation we denote also by πN the map from D([0,T];ΣN,E) to D([0,T];Pe(N)) defined by πNt(ε):=πN(ε(t)), t[0,T].

    We also introduce the empirical flow that records the collisions of the particles forgetting their labels. To this end, we denote by M the subset of the finite measures Q on [0,T]×N2×N2 that satisfy Q(dt;ε,ε,ε,ε)=Q(dt;ε,ε,ε,ε)=Q(dt;ε,ε,ε,ε). We endow M with the weak* topology and the associated Borel σ-algebra. The empirical flow is the map QN:D([0,T];ΣN,E)M defined by

    QN(ε)(F):=1N{i,j}k1F(τi,jk;εi(τi,jk),εj(τi,jk),εi(τi,jk),εj(τi,jk)) (1.3)

    where F:[0,T]×N2×N2R is continuous, bounded, and F(t;ε,ε,ε,ε)=F(t;ε,ε,ε,ε)=F(t;ε,ε,ε,ε), while (τi,jk)k1 are the jump times of the pair (εi,εj). Here, εi(t)=limstεi(s). In view of the conservation of the energy, the measure QN(dt;) is supported on E:={ε+ε=ε+ε}N2×N2.

    For each εΣN,E, with PNε probability one, the pair (πN,QN) satisfies the following balance equation that express the conservation of probability. For each ϕ:[0,T]×NR bounded and continuously differentiable with respect to time

    πNT(ϕT)πN0(ϕ0)T0dtπNt(tϕt)+T0ε,ε,ε,εQN(dt;ε,ε,ε,ε)[ϕt(ε)+ϕt(ε)ϕt(ε)ϕt(ε)]=0. (1.4)

    Fix mP(N) and assume one of the following condition: m is a point mass or the support of m does not generate a proper sub-lattice of Z. Note that in the second case m satisfies the condition for the local central limit theorem for i.i.d. lattice random variables, see [13,§ VII.1]. For γR we set

    Zγ=Zγ(m):=εm(ε)eγε. (1.5)

    We assume that there exists γ(0,+] such that Zγ<+ for γ(0,γ) and Zγ+ for γγ. For e(0,+), we then define the probability μN,e on ΣN,Ne by considering i.i.d. m-distributed energies and conditioning to the total energy, i.e.,

    μN,e:=mN(|Ni=1εi=Ne), (1.6)

    that will be chosen as the initial distribution of the microscopic dynamics. In the case of point mass, we require that {e} is exactly the support of m. Observe that, by the equivalence of the ensembles, as N+ the one-marginal of μN,e converges to the probability me given by

    me(ε):=eγeεm(ε)Zγe,whereγe<γissuchthatεεme(ε)=e. (1.7)

    Denoting by B the collision kernel in (1.1), i.e.,

    B(ε,ε,ε,ε)=1ε+ε+11I{ε+ε=ε+ε}1I{{ε,ε}{ε,ε}}, (1.8)

    the law of the large numbers for the empirical measure is described by the following discrete homogeneous Boltzmann equation

    tft(ε)=ε,ε,εB(ε,ε,ε,ε)[ft(ε)ft(ε)ft(ε)ft(ε)]. (1.9)

    More precisely, in probability with respect to PNμN,e, the empirical path (πNt)t[0,T] converges to (ft)t[0,T] where ft is the unique solution of the Cauchy problem associated to (1.9) with initial datum f0=me. As the proof of this statement can be achieved by adapting the chaos propagation arguments in [16], we omit the details. The law of the large numbers of the empirical flow QN reads

    QN(dt;ε,ε,ε,ε)12dtft(ε)ft(ε)B(ε,ε,ε,ε), (1.10)

    where the convergence is in probability with respect to PNμN,e. We refer to Lemma 3.3 below for the proof.

    In the general contest of homogeneous Boltzmann equations, uniqueness of the Cauchy problem associated to (1.9) holds for paths ft that conserves the energy, see e.g., [12]. However in the present case, since the supε,εε,εB(ε,ε,ε,ε) is bounded, Gronwall's inequality implies the uniqueness without assuming the energy conservation, see e.g., Lemma 4.1 in [2]. In particular, by uniqueness, for this model Lu and Wennberg like solutions do not exist. We finally observe that (1.9) admits a one-parameter family of stationary solutions given by fstat(ε)=p(1p)ε, p(0,1].

    For e(0,+), let Se be the (closed) subset of D([0,T];Pe(N))×M given by elements (π,Q) that satisfies the balance equation

    πT(ϕT)π0(ϕ0)T0dtπt(tϕt)+T0ε,ε,ε,εQ(dt;ε,ε,ε,ε)[ϕt(ε)+ϕt(ε)ϕt(ε)ϕt(ε)]=0 (1.11)

    for each ϕ:[0,T]×NR bounded and continuously differentiable in t. We consider Se endowed with the relative topology and the corresponding Borel σ-algebra.

    For πD([0,T];Pe(N)) let Qπ be the measure defined by

    Qπ(dt;ε,ε,ε,ε):=12dtπt(ε)πt(ε)B(ε,ε,ε,ε), (1.12)

    where B is the collision kernel in (1.8). Observe that Qπ(dt,) is supported on E. Let Sace be the subset of Se given by the elements (π,Q) such that πC([0,T];Pe(N)) and QQπ. The dynamical rate function J:Se[0,+] is defined by

    J(π,Q):={T0ε,ε,ε,εdQπ[dQπdQπlogdQπdQπdQπdQπ+1]if(π,Q)Seac+otherwise (1.13)

    Given two probabilities μ1,μ2, the relative entropy Ent(μ2|μ1) is defined as Ent(μ2|μ1)=dμ1ρlogρ, where dμ2=ρdμ1, understanding that Ent(μ2|μ1)=+ if μ2 is not absolutely continuous with respect to μ1. Let He:Pe(N)[0,+] be defined by

    He(π)=Ent(π|me)+(γγe)[eεεπ(ε)], (1.14)

    where me and γe are as in (1.7). When m is the point mass on e, He(π) is zero when π=δe and + otherwise. The candidate large deviation rate function is given by

    I(π,Q):=He(π0)+J(π,Q). (1.15)

    As discussed in [1], the (static) large deviations of the empirical measure with respect to the probability μN,e are described by the rate function He, where the second term on the r.h.s. of (1.14) is the cost of having energy less than e. Note that if γ=+, then He(π) is finite only if the energy of π is e. A key ingredient in the proof is the local central limit theorem for the sum of independent me distributed random variables.

    Denote by ˆSe the subset of S given by the pair (π,Q) such that

    T0ε,ε,ε,εQ(dt,ε,ε,ε,ε)(ε+ε)<+.

    If (π,Q)ˆSe, the balance equation (1.11) implies that the path πt conserves the energy.

    As already mentioned, a proof of a large deviations principle for the pair (πN,QN) with matching upper and lower has not been yet achieved. The analysis in [2,6] implies however the large deviation upper bound with rate I with a matching lower bound on the set ˆSe. The precise statement is the following.

    Theorem 1.1. Fix e(0,+) and let μN,e be the family of probabilities on ΣN,Ne defined in (1.6). The family {PNμN,e(πN,QN)1} satisfies a large deviations upper bound with good rate function I:Se[0,+], namely I has compact level sets and for each closed CS

    ¯limN+1NlogPNμN,e((πN,QN)C)inf(π,Q)CI(π,Q). (1.16)

    Moreover, for each open OSe

    lim_N+1NlogPNμN,e((πN,QN)O)inf(π,Q)OˆSeI(π,Q). (1.17)

    Referring to [2,6] for comments on the technicalities involved in the lower bound, we now turn to the novel point of the present analysis, that is the construction of paths (π,Q) – with π not energy conserving – whose probability is precisely of the order exp{NI(π,Q)}. Since these paths do not belong to ˆSe, this result provides insights on the large deviations properties of Kac's walk not covered by (1.17). As the large deviations upper bound is already covered by (1.16), we focus on the matching lower bound.

    Theorem 1.2. Fix e(0,+) and let μN,e be the family of probabilities on ΣN,Ne defined in (1.6). For each t(0,T) there exists a path (ˉf,ˉQ), satisfying εˉft(ε)ε=e for t[0,t) and εˉft(ε)ε<e for t[t,T], such that I(ˉf,ˉQ)<+ and

    lim_N1NlogPNμN,e((πN,QN)O)I(ˉf,ˉQ), (1.18)

    for any open neighborhood O(ˉf,ˉQ).

    We will provide a self-contained proof of this statement that do not rely on Theorem 1.1. In the argument we take advantage of the fact that the energies are in N. However we expect that the strategy can be extended also to the continuous case. In Section 2 we construct a path (ˉf,ˉQ) satisfying the above requirements, i.e., with evaporating energy for t>t and such that I(ˉf,ˉQ)<+. The lower bound (1.18) is then proven in Section 3. For the sake of concreteness, the proposed path (ˉf,ˉQ) has zero energy for t(t,T].

    Fix t(0,T). In order to construct a path (ˉf,ˉQ), satisfying εˉft(ε)ε=e for t[0,t) and εˉft(ε)ε<e for t(t,T], we start by considering a solution to a perturbed Boltzmann equation, namely a Boltzmann equation with a suitable modified collision kernel.

    Consider the collision kernel ˜B given by

    ˜B(ε,ε,ε,ε)=12δε,εδε+ε,ε+ε[δε,ε+ε+δε,ε+ε]1I{{ε,ε}{ε,ε}}, (2.1)

    that describes a scenario in which only particles with the same energy collide, and in each collision the whole energy is transferred to a single particle. The Cauchy problem for the corresponding modified homogeneous Boltzmann equation reads

    {tft(ε)=ε,ε,ε[˜B(ε,ε,ε,ε)ft(ε)ft(ε)˜B(ε,ε,ε,ε)ft(ε)ft(ε)],f0(ε)=m(ε). (2.2)

    Proposition 2.1. Assume that m has energy e, and let f be the unique solution to (2.2). Then its energy is conserved, i.e., for any t[0,+) ε0ft(ε)ε=e, while ft weakly converges to δ0, as t+. Moreover, for every t0,

    i)  ft(ε)21+t,  forε1,ii)  ε1ft(ε)c1+t,iii)  ε1ft(ε)logεc11+t(1+log(1+t)) (2.3)

    where c=c(e) does not depend on t and the initial datum m.

    Proof. We prove Eq (2.3), from which the convergence of ft follows. The modified Boltzmann equation reads as

    ˙ft(0)=12ε1ft(ε)2,˙ft(ε)=ft(ε)2  forε1odd,˙ft(ε)=12ft(ε/2)2ft(ε)2    forε2even. (2.4)

    Note that the equation for ft(ε) involves only ft(ε) with εε, then the system has global and unique solution. If ε is odd,

    ft(ε)=f0(ε)1+tf0(ε)11+t.

    If ε is even, set ξt(ε)=(1+t)ft(ε). Let T0 be the first time t such that ξt(ε)=2 for some εε. The time T0 is strictly positive, since ξ0(ε)1 for any ε. For t<T0 it holds

    ˙ξt(ε)=11+t(ξt(ε)ξ2t(ε)+12ξ2t(ε/2))<ξt(ε)ξ2t(ε)+23(2ξt(ε)),

    then T0=+, and this concludes the proof of i) in (2.3).

    In order to prove ii) and iii) we first note that

    2n0εft(ε)=2n0εf0(ε)t0ds2n2n1+1εfs(ε)2, (2.5)

    and then +0εft(ε)e. Inequality ii) and iii) follows by using this fact and the Chebyshev's inequality. We conclude the proof by noticing that the energy is in fact conserved, since ft(ε)e/ε and then for any h1,

    +hεft(ε)2emaxεhft(ε)e2h,

    which assures that the right-hand-side of Eq (2.5), is vanishing as n+.

    For the modified Boltzmann equation with rate in (2.1) the energy vanishes for dispersion to infinity as t+. We reparametrize the time so that this happens at a finite time. Fixed t(0,T), let α:[0,t)[0,+) given by α(t)=t1t/t. Letting f the solution to (2.2), set

    ˉft(ε)={fα(t)(ε)t[0,t)δε,0t[t,T], (2.6)

    which satisfies the homogeneous Boltzmann equation with time dependent collision kernel

    ˉBt(ε,ε,ε,ε)={˙α(t)˜B(ε,ε,ε,ε)t[0,t)0t[t,T]. (2.7)

    We define the corresponding flux dˉQ=dtˉqt, where

    ˉqt(ε,ε,ε,ε)=12ˉft(ε)ˉft(ε)˜Bt(ε,ε,ε,ε), (2.8)

    so that the pair (ˉf,ˉQ) satisfies the balance equation (1.11). Observe that, by construction, εˉft(ε)ε=e for t[0,t) and εˉft(ε)ε=0 for t[t,T]. We now show that the pair (ˉf,ˉQ) is such that I(ˉf,ˉQ)<+. Since ˉf0=m, it is enough to show J(ˉf,ˉQ)<+. This is stated in the next Proposition.

    Proposition 2.2. For (ˉf,ˉQ) defined above the dynamical rate function J(ˉf,ˉQ) is finite.

    Proof. Since (ˉf,ˉQ)Seac, the dynamical rate function defined in Eq (1.13) is given by

    12t0dtε,ε,ε,εˉft(ε)ˉft(ε)ˉBt(logˉBtB1)+12T0ˉft(ε)ˉft(ε)B. (2.9)

    For t<t we have that

    ˉBt(logˉBtB1)=˙α˜B(log˙α+log1+2ε21).

    Since log˙α=2log(1+α/t), the first integral is

    12+0dαε1f2α(ε)(2log(1+αt)+log1+2ε21).

    Using (2.3) we bound this term by

    c+01(1+α)3/2(1+log(1+α))<+,

    where c depend only on e and t.

    The second integral in Eq (2.9) is

    12t0dt(ε1fα(t))2+12(Tt)T2,

    which completes the proof.

    In order to explain the strategy to prove (1.18), we recall some basic facts on the large deviations lower bound. Let {Pn} be a sequence of probabilities on a topological space X. Fix xX and a open neighborhood Ox. To obtain a lower bound for Pn(O) we modify the probability Pn so that x becomes the typical behavior. If we are able to do so by paying – as measured by the relative entropy with respect to Pn – not too much then we obtain a good lower bound. The precise statement is summarized in the next lemma, see e.g., [11] for its proof.

    Lemma 3.1. Let {Pn}nN be a sequence of probabilities on a completely regular topological space X and fix xX. Assume that there exists a sequence {Pxn} weakly convergent to δx and such that

    ¯limn1nEnt(Pxn|Pn)I(x) (3.1)

    for some I:X[0,+]. Then for any open neighborhood Ox

    lim_n1nlogPn(O)I(x).

    In most of the applications, and indeed also in our case, the strategy suggested by the above lemma is implemented together with a density argument. The family of perturbed probabilities Pxn is not constructed for the point x itself but rather for an approximation xk; if the function I is continuous along the sequence xk then this will do as well. We emphasize that in typical infinite dimensional applications – as in the present case – the rate function I is only lower semicontinuous so the sequence xk has to be properly chosen. We summarize the argument in the next statement which is deduced from Lemma 3.1 by a straightforward diagonal argument.

    Lemma 3.2. Let {Pn}nN be a sequence of probabilities on a completely regular topological space X, fix xX, and a sequence xkx. Assume that there exists I:X[0,+] meeting the following conditions:

    (i) for each kN there exists a family {Pxkn}nN satisfying the conditions in Lemma 3.1;

    (ii) ¯limkI(xk)I(x).

    Then, for any open neighborhoods Ox

    lim_n1nlogPn(O)I(x).

    To implement condition (i) and (ii) in the previous lemma, for 0<δ<t, define the pair (ˉfδ,ˉQδ) by

    ˉfδt(ε)={ˉft(ε)t[0,tδ)ˉftδ(ε)t[tδ,T], (3.2)

    and dˉQδ=dtˉqδt with ˉqδt=ˉqt1I[0,tδ)(t). Let μN,e as in the statement of Theorem 1.2, and denote by ˉPN,δμN,e the law of the microscopic dynamics with the perturbed collision kernel ˉBδ=ˉB1I[0,tδ)(t), ˉB in (2.7). Then the following two Lemmata imply the large deviation lower bound (1.18).

    Lemma 3.3. For each δ(0,t) as N+ the pair (πN,QN) converges in ˉPN,δμN,e probability to (ˉfδ,ˉQδ). Furthermore,

    limN+1NEnt(ˉPN,δμN,e|PNμN,e)=I(ˉfδ,ˉQδ). (3.3)

    Lemma 3.4. As δ0 we have (ˉfδ,ˉQδ)(ˉf,ˉQ) and I(ˉfδ,ˉQδ)I(ˉf,ˉQ).

    Proof of Lemma 3.3. By definition of ˉBδ, suptsupε,εε,εˉBδt(ε,ε,ε)cδ. Therefore, by classical chaos propagation argument, πN converges in ˉPN,δμN,e probability to ˉfδ. To deduce the convergence of the empirical flow, it is enough to observe that for each bounded Ft(ε,ε,ε,ε)

    MFt:=t0ε,ε,ε,εQN(ds,ε,ε,ε,ε)Fs(ε,ε,ε,ε)12t0dsε,ε,ε,επNs(ε)πNs(ε)ˉBδs(ε,ε,ε,ε)Fs(ε,ε,ε,ε)+121Nt0dsε,ε,επNs(ε)ˉBδs(ε,ε,ε,ε)Fs(ε,ε,ε,ε) (3.4)

    is a ˉPN,δμN,e martingale with predictable quadratic variation

    MFt=121Nt0dsε,ε,ε,επNs(ε)πNs(ε)ˉBδs(ε,ε,ε,ε)F2s(ε,ε,ε,ε)121N2t0dsε,ε,επNs(ε)ˉBδs(ε,ε,ε,ε)F2s(ε,ε,ε,ε).

    Set Fδt(ε,ε,ε,ε)=log(ˉBδt/B). By standard Markov chain computation, the relative entropy of ˉPN,δμN,e with respect to PNμN,e is given by

    1NEnt(ˉPN,δμN,e|PNμN,e)=ˉEN,δμN,e(QN(Fδ)12T0dtε,ε,ε,επNt(ε)πNt(ε)[ˉBδt(ε,ε,ε,ε)B(ε,ε,ε,ε)]+1N12T0dtε,ε,επNt(ε)[ˉBδt(ε,ε,ε,ε)B(ε,ε,ε,ε)]). (3.5)

    Since suptsupε,εε,εˉBδt(ε,ε,ε)cδ, by the law of large numbers

    limNˉEN,δμN,e(12T0dtε,ε,ε,επNt(ε)πNt(ε)[ˉBδt(ε,ε,ε,ε)B(ε,ε,ε,ε)])=12T0dtε,ε,ε,εˉfδt(ε)ˉfδt(ε)[ˉBδt(ε,ε,ε,ε)B(ε,ε,ε,ε)],

    while the last term on the right hand side of (3.5) vanishes as N diverges. Again, by the law of large numbers, in order to prove ˉEN,δμN,e(QN(Fδ))ˉQδ(Fδ) it is enough to show the uniform integrability of QN(Fδ) with respect to ˉPN,δμN,e. By exploiting the martingale decomposition (3.4), since |Fδ|cδ(1+log(1+ε+ε)), a direct computation yields ˉEN,δμN,e(QN(Fδ)2)cδ, which implies the requested uniform integrability. Observing that He(ˉfδ0)=0, and recalling (1.13) and (1.15), the proof is concluded.

    Proof of Lemma 3.4. The convergence of (ˉfδ,ˉQδ) to (ˉf,ˉQ) follows from (3.2), the continuity of tˉft, and the integrability of ˉqt. The convergence of the rate function is achieved by the arguments in the proof of Proposition 2.2 and dominated convergence.

    The authors declare no conflict of interest.



    [1] G. Basile, D. Benedetto, L. Bertini, E. Caglioti, On the probability of observing energy increasing solutions to homegeneous Boltzmann equation, unpublished work.
    [2] G. Basile, D. Benedetto, L. Bertini, C. Orrieri, Large deviations for Kac-Like Walks, J. Stat. Phys., 184 (2021), 10. http://dx.doi.org/10.1007/s10955-021-02794-2 doi: 10.1007/s10955-021-02794-2
    [3] L. Bertini, D. Gabrielli, J. L. Lebowitz, Large deviations for a stochastic model of heat flow, J. Stat. Phys., 121 (2005), 843–885. http://dx.doi.org/10.1007/s10955-005-5527-2 doi: 10.1007/s10955-005-5527-2
    [4] T. Bodineau, I. Gallagher, L. Saint-Raymond, S. Simonella, Statistical dynamics of a hard sphere gas: fluctuating Boltzmann equation and large deviations, 2021, arXiv: 2008.10403.
    [5] F. Bouchet, Is the Boltzmann equation reversible? A large deviation perspective on the irreversibility paradox, J. Stat. Phys., 181 (2020), 515–550. http://dx.doi.org/10.1007/s10955-020-02588-y doi: 10.1007/s10955-020-02588-y
    [6] D. Heydecker, Large deviations of Kac's conservative particle system and energy non-conserving solutions to the Boltzmann equation: a counterexample to the predicted rate function, 2021, arXiv: 2103.14550.
    [7] C. Kipnis, C. Marchioro, E. Presutti, Heat flow in an exactly solvable model, J. Stat. Phys., 27 (1982), 65–74. http://dx.doi.org/10.1007/BF01011740 doi: 10.1007/BF01011740
    [8] C. Léonard, On large deviations for particle systems associated with spatially homogeneous Boltzmann type equations, Probab. Th. Rel. Fields, 101 (1995), 1–44. http://dx.doi.org/10.1007/BF01192194 doi: 10.1007/BF01192194
    [9] T. M. Liggett, Continuous time Markov processes: an introduction, Providence: American Mathematical Society, 2010.
    [10] X. Lu, B. Wennberg, Solutions with increasing energy for the spatially homogeneous Boltzmann equation, Nonlinear Anal. Real, 3 (2002), 243–258. http://dx.doi.org/10.1016/S1468-1218(01)00026-8 doi: 10.1016/S1468-1218(01)00026-8
    [11] M. Mariani, A Γ-convergence approach to large deviations, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), XVIII (2018), 951–976. http://dx.doi.org/10.2422/2036-2145.201301_010 doi: 10.2422/2036-2145.201301_010
    [12] S. Mischler, B. Wennberg, On the spatially homogeneous Boltzmann equation, Ann. Inst. Henri Poincaré, Analyse non linéaire, 16 (1999), 467–501. http://dx.doi.org/10.1016/S1468-1218(01)00026-8 doi: 10.1016/S1468-1218(01)00026-8
    [13] V. V. Petrov, Sums of independent random variables, Berlin, Heidelberg: Springer, 1975. http://dx.doi.org/10.1007/978-3-642-65809-9
    [14] J. Quastel, H.-T. Yau, Lattice gases, large deviations, and the incompressible Navier-Stokes equations, Ann. Math., 148 (1998), 51–108.
    [15] F. Rezakhanlou, Large deviations from a kinetic limit, Ann. Probab., 26 (1998), 1259–1340. http://dx.doi.org/10.1214/aop/1022855753 doi: 10.1214/aop/1022855753
    [16] A. S. Sznitman, Topics in propagation of chaos, In: Ecole d'Eté de Probabilités de Saint-Flour XIX–1989, Berlin, Heidelberg: Springer, 1991,165–251. http://dx.doi.org/10.1007/BFb0085169
  • This article has been cited by:

    1. Roberta Bianchini, Chiara Saffirio, Fluid instabilities, waves and non-equilibrium dynamics of interacting particles: a short overview, 2022, 5, 2640-3501, 1, 10.3934/mine.2023033
    2. Pietro Caputo, Daniel Parisi, Nonlinear recombinations and generalized random transpositions, 2024, 7, 2644-9463, 1245, 10.5802/ahl.219
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2292) PDF downloads(250) Cited by(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog