Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

A minimizing-movements approach to GENERIC systems

  • We present a new time discretization scheme adapted to the structure of GENERIC systems. The scheme is based on incremental minimization and is therefore variational in nature. The GENERIC structure of the scheme provides stability and conditional convergence. We show that the scheme can be rigorously implemented in the classical case of the damped harmonic oscillator. Numerical evidence is collected, illustrating the performance of the method and, in particular, the conservation of the energy at the discrete level.

    Citation: Ansgar Jüngel, Ulisse Stefanelli, Lara Trussardi. A minimizing-movements approach to GENERIC systems[J]. Mathematics in Engineering, 2022, 4(1): 1-18. doi: 10.3934/mine.2022005

    Related Papers:

    [1] Jana Kopfová, Petra Nábělková . Thermoelastoplastic oscillator with Prandtl-Ishlinskii operator. Mathematics in Engineering, 2025, 7(3): 264-280. doi: 10.3934/mine.2025012
    [2] Jérôme Droniou, Jia Jia Qian . Two arbitrary-order constraint-preserving schemes for the Yang–Mills equations on polyhedral meshes. Mathematics in Engineering, 2024, 6(3): 468-493. doi: 10.3934/mine.2024019
    [3] François Alouges, Aline Lefebvre-Lepot, Jessie Levillain . A limiting model for a low Reynolds number swimmer with N passive elastic arms. Mathematics in Engineering, 2023, 5(5): 1-20. doi: 10.3934/mine.2023087
    [4] Kyungkeun Kang, Dongkwang Kim . Existence of generalized solutions for Keller-Segel-Navier-Stokes equations with degradation in dimension three. Mathematics in Engineering, 2022, 4(5): 1-25. doi: 10.3934/mine.2022041
    [5] Jayme Vaz Jr., Edmundo Capelas de Oliveira . On the fractional Kelvin-Voigt oscillator. Mathematics in Engineering, 2022, 4(1): 1-23. doi: 10.3934/mine.2022006
    [6] Christopher Chong, Andre Foehr, Efstathios G. Charalampidis, Panayotis G. Kevrekidis, Chiara Daraio . Breathers and other time-periodic solutions in an array of cantilevers decorated with magnets. Mathematics in Engineering, 2019, 1(3): 489-507. doi: 10.3934/mine.2019.3.489
    [7] Simon Lemaire, Julien Moatti . Structure preservation in high-order hybrid discretisations of potential-driven advection-diffusion: linear and nonlinear approaches. Mathematics in Engineering, 2024, 6(1): 100-136. doi: 10.3934/mine.2024005
    [8] Carlo Danieli, Bertin Many Manda, Thudiyangal Mithun, Charalampos Skokos . Computational efficiency of numerical integration methods for the tangent dynamics of many-body Hamiltonian systems in one and two spatial dimensions. Mathematics in Engineering, 2019, 1(3): 447-488. doi: 10.3934/mine.2019.3.447
    [9] Takeyuki Nagasawa, Kohei Nakamura . Asymptotic analysis for non-local curvature flows for plane curves with a general rotation number. Mathematics in Engineering, 2021, 3(6): 1-26. doi: 10.3934/mine.2021047
    [10] James McCoy, Jahne Meyer . Semi-discrete linear hyperbolic polyharmonic flows of closed polygons. Mathematics in Engineering, 2025, 7(3): 281-315. doi: 10.3934/mine.2025013
  • We present a new time discretization scheme adapted to the structure of GENERIC systems. The scheme is based on incremental minimization and is therefore variational in nature. The GENERIC structure of the scheme provides stability and conditional convergence. We show that the scheme can be rigorously implemented in the classical case of the damped harmonic oscillator. Numerical evidence is collected, illustrating the performance of the method and, in particular, the conservation of the energy at the discrete level.



    The aim of this note is to discuss a new variational time-discretization scheme adapted to the structure of General Equations for Non-Equilibrium Reversible-Irreversible Coupling (GENERIC). Introduced by Grmela & Öttinger, this formulation provides a unified frame for describing the time evolution of physical systems out of equilibrium in presence of reversible and irreversible dynamics [37].

    Let y denote the state of a closed, nonequilibrium physical system and let E(y) and S(y) be the corresponding total energy and total entropy, respectively. The GENERIC formulation of the time evolution of the system reads

    y=L(y)DE(y)+K(y)DS(y). (1.1)

    Here, y denotes the time derivative, L is the antisymmetric Poisson operator, and K is the symmetric and positive definite Onsager operator (details in Section 2).

    The gist of the GENERIC system (1.1) is that conservative and dissipative dynamics are clearly separated. Under a specific compatibility condition, see (2.4) below, this entails that a solution ty(t) to (1.1) conserves energy and accumulates entropy in a quantifiable way, namely

    ddtE(y)=0andddtS(y)=DS(y),K(y)DS0.

    The conservation of energy and the quantified dissipative character are the distinguishing traits of the GENERIC system (1.1). To replicate these properties at the discrete level leads to so-called structure-preserving approximations. These have drawn interest in the last years, giving rise to different numerical solutions adapted to diverse applicative contexts.

    In the last years, GENERIC has attracted increasing attention and has been applied to a number of situations ranging from complex fluids [18], to dissipative quantum mechanics [32], to thermomechanics [5,21,22,23,31], to electromagnetism [24], to shape-memory alloys [3], to the Vlasov-Fokker-Planck [13,14,20] and the Boltzmann equation [35], and to large-deviation limits of reversible stochastic processes [26,27].

    Numerical schemes conserving energy can be found for instance in [16,41,43,44,45,48]; see [8] for a review and [50] for a contribution explicitly focusing on GENERIC formulations. These schemes are often discrete-gradient methods, where gradients of functionals are specifically modified in order to fulfill a discrete chain rule and exactly replicate conservation [16,17,19,30]. In the thermomechanical context, structure-preserving discretizations either in terms of the absolute temperature [9,10,15,39] or of the internal energy or the entropy [5,6] have been obtained. The reader is referred to [28] for an approach to open systems.

    GENERIC integrators able to conserve energy and accurately describe entropy accumulation have been proposed in [38], where however some limitations are also mentioned. In particular, energy and Onsager operators have to be suitably modified in a time-step dependent manner in order energy conservation to hold. Integrators are actually constructed in case of a single dissipation mechanism only (were K be a matrix, it would have to have rank one) and no convergence theory is provided. The discretization of the damped harmonic oscillator is addressed in [38], where nonetheless the temperature is given by a prescribed heat bath. In this case, explicit solutions have to be used in order to specify the GENERIC integrator.

    To the best of our knowledge, all available numerical schemes directly target GENERIC systems in their differential form (1.1). Our focus is here on a time-discretization of variational nature instead, fitting into the general scheme of so-called minimizing movements [1,2]. In particular, at all discrete steps we aim at solving a specific minimization problem. We start by defining the entropy-production potential Ψ(y,ξ)=ξ,K(y)ξ/2 and denoting by Ψ(y,) its conjugate in the second variable. Then, the GENERIC relation (1.1) in [0,T] can be equivalently rewritten in terms of the scalar equation

    S(y(t))+t0Ψ(y,yL(y)DE(y))dr+t0Ψ(y,DS(y))dr+S(y(0))=0 (1.2)

    for all t[0,T]. The reformulation of dissipative systems in terms of scalar equations as (1.2) is usually referred to as De Giorgi's Energy-Dissipation Principle [12,29,33]. This principle has already been applied in a variety of different contexts, including generalized gradient flows [4,47], curves of maximal slope in metric spaces [2,11], rate-independent systems [34,42], and optimal control [40]. For GENERIC systems, the reformulation of (1.1) in terms of such variational principle has been already presented in [14].

    Our new minimizing-movements approach to (1.1) consists in tackling a discrete version of relation (1.2). Assume to be given a time partition 0=t0<t1<<tN=T with steps τi=titi1, and an initial state y0. We define the time-discrete trajectory {yi}Ni=0 by letting y0=y0 and subsequently perform the minimization

    miny{S(y)+τiΨ(yϕ,yyi1τiL(y)DE(yϕ))+τiΨ(yϕ,DS(y))+S(yi1)}

    for i=1,,N where yϕ:=ϕy+(1ϕ)yi1 for some fixed ϕ[0,1]. The latter is nothing but a localized and discretized version of the De Giorgi's Energy-Dissipation Principle (1.2).

    The minimizing-movements scheme above has been introduced in [25]. The theory in [25] is however restricted to ϕ=1 and to the case of state-independent operators L and K, which severely limits the applicability to real GENERIC systems. In addition, the convergence analysis in [25] relies on a suitable set of a priori assumptions, leaving open the discussion whether these can be met in practice.

    The aim of this note is then threefold. First, we extend the reach of the numerical method to include the case of state dependent operators L(y) and K(y), hence covering the full extent of the GENERIC theory (Section 2). In addition, we investigate the general case ϕ[0,1]. Our main result is the conditional convergence of Theorem 2.1. Second, we provide the detailed analysis in the classical case of the damped harmonic oscillator for the choice ϕ=1/2. In this case, the above-mentioned convergence assumptions can actually be proved to hold (Section 3). Eventually, we present numerical experiments assessing the performance of the minimizing-movements scheme (Subsection 3.4). In particular, we show that the scheme conserves the energy.

    In this section, we recall the structure of a GENERIC system [37] by specifying the assumptions on functionals and operators that will be used throughout. Moreover, we formulate our minimizing-movements scheme and present a conditional convergence result, namely Theorem 2.1.

    The GENERIC system

    y=L(y)DE(y)+K(y)S(y)a.e. in [0,T] (2.1)

    is defined by specifying the quintuple (Y,E,S,L,K). In the following, the state space Y is assumed to be a reflexive Banach space. The functionals E and S represent the total energy and the total entropy of the system, respectively. We assume E to be Fréchet differentiable, with strongly-weakly continuous differential DE, and S:Y(,] to be proper and lower semicontinuous with single-valued and strongly-weakly continuous Fréchet subdifferential (S) [36]. Recall that one has ξ(S)(y) iff S(y)> and

    lim infxyS(y)S(x)ξ,xyxyY0.

    In the following, we also make use of the obvious notation S=(S).

    The operators L and K define a Poisson and an Onsager geometric structure on Y, respectively. In particular, for all states yY we assume that L(y) and K(y) are linear and continuous from Y to Y. Moreover, L(y) is required to be antisymmetric, L(y)=L(y), and to fulfill the Jacobi identity {{g1,g2},g3}+{{g2,g3},g1}+{{g3,g1},g2}=0. Here, gi denotes any differentiable function on Y and the Poisson bracket is defined as {g,˜g}(y)=Dg(y),L(y)D˜g(y), where , denotes the duality pairing between Y (dual) and Y. Moreover, we assume the strong-weak continuity

    ynYy  and  ξnYξ    L(yn)ξnYL(y)ξ. (2.2)

    The mapping K(y) is asked to be symmetric and positive definite, namely K(y)=K(y)0. We associate with K the so-called entropy-production potential Ψ:Y×Y[0,) given by Ψ(y,ξ)=ξ,K(y)ξ/2 and let Ψ be its conjugate in the second variable, namely Ψ(y,η)=supξ(ξ,ηΨ(y,ξ)). We also assume the lower semicontinuity of the sum of the entropy-production potential and its dual, that is

    ynYy,  ηnYη,  and  ξnYξ  lim infn(Ψ(yn,ηn)+Ψ(yn,ξn))Ψ(y,η)+Ψ(y,ξ). (2.3)

    In addition, functionals and operators are asked to satisfy the crucial noninteraction condition

    L(y)S(y)=K(y)DE(y)=0yY. (2.4)

    This condition ensures that the system of dissipative actions K(y)S(y) does not spoil energy conservation and the system of conservative actions L(y)DE(y) does not contribute to dissipation. Indeed, by assuming sufficient smoothness one checks that

    ddtE(y)=DE(y),y(2.1)=DE(y),L(y)DE(y)+DE(y),K(y)S(y)=0+S(y),K(y)DE(y)(2.4)=0,ddtS(y)=S(y),y(2.1)=S(y),L(y)DE(y)+S(y),K(y)S(y)=DE(y),L(y)S(y)+S(y),K(y)S(y)(2.4)=S(y),K(y)S(y)0. (2.5)

    The noninteraction condition (2.4) hence implies that trajectories y solving (2.1) have constant total energy, and that the entropy rate is S(y),K(y)S(y). In particular, the entropy is nondecreasing and entropy production results solely from irreversible processes. In computing (2.5) we have used the chain rule (d/dt)S(y)=S(y),y almost everywhere. This is classical in case S is (a regular perturbation of) a convex functional [7,Prop. 3.3,p. 73]. The reader is referred to [46] for a general discussion out of the convex case.

    Before moving on, let us remark that the structure of GENERIC is geometric in nature. Indeed, it is invariant by coordinate changes. Let y=ϕ(˜y) for ˜y˜Y and define ˜E(˜y)=E(ϕ(˜y)), ˜S(˜y)=S(ϕ(˜y)), ˜L(˜y)=Dϕ(˜y)1L(ϕ(˜y))Dϕ(˜y), and ˜K(˜y)=Dϕ(˜y)1K(ϕ(˜y))Dϕ(˜y), where Dϕ(˜y):˜YY is the adjoint of the inverse of Dϕ(˜y):˜YY. Then, the quintuple (˜Y,˜E,˜S,˜L,˜K) satisfies the above structural assumptions and the GENERIC structure (2.1) can be rewritten as ˜y=˜L(˜y)D˜E(˜y)+˜K(˜y)˜S(˜y).

    We now reconsider the discussion leading to (1.2) and specify it further by remarking that relation (2.1) is actually equivalent to the inequality

    S(y(t))+t0Ψ(y,yL(y)DE(y))dr+t0Ψ(y,S(y))dr+S(y(0))0 (2.6)

    for all t[0,T]. The equivalence between (2.1) and (2.6) follows from Fenchel's relations. Applied to the entropy-production potential Ψ, these relations read

    Ψ(y,η)+Ψ(y,ξ)ξ,ηy,ηY, ξY, (2.7)
    Ψ(y,η)+Ψ(y,ξ)=ξ,ηξΨ(y,η), (2.8)

    where the subdifferential is taken in the second variable only. By noting that Ψ(y,S(y))=K(y)S(y) and using (2.7) and (2.8), one can prove the equivalences

    (2.1)  yL(y)DE(y)=Ψ(y,S(y))  a.e.(2.8)Ψ(y,yL(y)DE(y))+Ψ(y,S(y))yL(y)DE(y),S(y)0  a.e.(2.4)Ψ(y,yL(y)DE(y))+Ψ(y,S(y))ddtS(y)0  a.e.(2.7)(2.6).

    In particular, the last left-to-right implication follows by integration while the right-to-left counterpart from the nonnegativity of the integrand, given (2.4) and (2.7).

    The minimizing-movements scheme corresponds to a discretization of inequality (2.6). To each time partition 0=t0<t1<<tN=T, we associate the time steps τi=titi1 and the diameter τ=maxτi. Given the vector {yi}Ni=0YN+1, we introduce the backward piecewise constant and piecewise linear interpolations ¯y:[0,T]Y and ˆy:[0,T]Y,

    ¯y(0)=ˆy(0), ¯y(t)=yi, ˆy(t)=ttitτiyi1+titτiyi,t(ti1,ti], i=1,,N.

    We define the incremental functional by G:(0,)×Y×Y(,] as

    G(τ,η;y)=S(y)+τΨ(yϕ,yητL(y)DE(yϕ))+τΨ(yϕ,S(y))+S(η)

    where yϕ=ϕy+(1ϕ)η and ϕ[0,1] is fixed throughout. By letting y0=y0 we find the discrete solution {yi}Ni=0 by subsequently solving the minimization problem

    minyG(τi,yi1;y)for i=1,,N. (2.9)

    The state dependence in Ψ, Ψ, and DE can be made implicit or explicit suitably choosing the parameter ϕ[0,1], and our analysis is independent from such a specific choice.

    For all τ>0 and yi1Y with S(yi1)>, the map yG(τi,yi1;y) is strongly lower semicontinuous because of the lower semicontinuity of S, the lower semicontinuity (2.3) of Ψ+Ψ, the weak-strong continuity of DE and S, and the continuity (2.2) of L. In order to solve problem (2.9) one has hence to check that yG(τi,yi1;y) is strongly coercive.

    The main result of this section is the following conditional convergence result.

    Theorem 2.1 (Conditional convergence). Under the above assumptions, let a sequence of partitions 0=tn0<<tnNn=T be given with τn0 as n, and let {yni}Nni=1 be such that yn0=y0, not necessarily being minimizers from (2.9). Assume that

    Nni=1(G(τni,yni1;yni))+0  as n, (2.10)
    {ˆyn} is bounded in H1(0,T;Y) and takes values in K⊂⊂Y, (2.11)
    {L(¯yn)DE(¯yϕn)} is bounded in L2(0,T;Y), (2.12)
    {S(¯yn)} is bounded in L2(0,T;Y). (2.13)

    Then, up to a subsequence, ˆyny in H1(0,T;H), where y is a solution of the GENERIC system (2.1) in the sense of inequality (2.6).

    Proof of Theorem 2.1. We infer from the Aubin-Lions lemma [49,Cor. 7], upon extracting not relabeled subsequences, that

    ˆyny  in  H1(0,T;Y),¯yn,ˆyn,¯yϕny  in  C([0,T];Y),L(¯yn)DE(¯yϕn)  in  L2(0,T;Y), (2.14)
    S(¯yn)s  in  L2(0,T;Y). (2.15)

    As S is strongly-weakly closed, we have that s=S(y) almost everywhere. On the other hand, the strong-weak continuity of DE and the continuity (2.2) of L imply that L(¯yn)DE(¯yϕn)L(y)DE(y) pointwise in time, so that =L(y)DE(y) almost everywhere.

    Fix any t(0,T] and let tnm be such that t(tnm1,tnm]. The discrete sequence of solutions {yni}Nni=0 fulfills

    S(¯yn(t))+tnm0Ψ(¯yϕn,(ˆyn)L(¯yn)DE(¯yϕn))dr+tnm0Ψ(¯yϕn,S(¯yn))dr+S(y0)=mi=1G(τni,yni1;yni). (2.16)

    As Ψ(˜y,)Ψ(˜y,0)=0 for all ˜y, we conclude that Ψ0 as well. By restricting integrals to the interval [0,t][0,tnm] equation (2.16) implies that

    S(¯yn(t))+t0Ψ(¯yϕn,(ˆyn)L(¯yn)DE(¯yϕn))dr+t0Ψ(¯yϕn,S(¯yn))dr+S(y0)Nni=1(G(τni,yni1;yni))+. (2.17)

    We now pass to the limit inferior as n in relation (2.17). By using convergence (2.10), the lower semicontinuity of S, convergences (2.14) and (2.15), and the lower semicontinuity (2.3), we readily obtain that the limit y fulfills inequality (2.6).

    The conditional convergence result of Theorem 2.1 relies on the possibility of solving the inequality G(τni,yni1;yni)0 up to a small, controllable error, and establishing some a priori bounds on the discrete solution. The validity of these conditions has to be checked on the specific problem at hand. In the upcoming Section 3 we give an example of a situation where (2.10)–(2.13) actually hold.

    In case S is convex, an example of {yni}Nni=0 fulfilling (2.10) are the solutions of the Euler scheme

    yniyniτni=L(yni)DE(yϕni)+K(yϕni)S(yni) (2.18)

    with yϕni=ϕyni+(1ϕ)yni1. Indeed, such {yni}Nni=0 fulfills

    G(τni,yni1;yni)τniΨ(yϕni,yniyniτniL(yni)DE(yϕni))+τniΨ(yϕni,S(yni))S(yni),yniyni1=0, (2.19)

    where the last equality follows from (2.8) and (2.18).

    In the purely dissipative case L=0, the existence of a solution to the minimization problem (2.9) is ensured as soon as τ>0 exists, such that, for all τ(0,τ) and yi1Y with S(yi1)>, the function

    yS(y)+τΨ(ϕy+(1ϕ)yi1,yyi1τ)

    is strongly coercive. The latter follows whenever S has strongly compact sublevels. Note however that the specific example of the damped harmonic oscillator from Section 3 does not fall into this class, as it is not purely dissipative and S does not have strongly compact sublevels. Still, existence for the minimization problem (2.9) can be directly checked.

    Let us mention that, in specific applications, the bounds (2.11)–(2.13) may follow from (2.10). For instance, this would be the case if the coercivity

    Ψ(˜y,η)+Ψ(˜y,ξ)cη2Y+cξ2Y1c (2.20)

    holds for some c>0 and all ˜y,ηY and ξY, namely in case K(˜y) is positive definite and bounded, uniformly with respect to ˜y. This however does not apply to the example of the damped harmonic oscillator from Section 3, for K is singular there.

    The quadratic nature of the entropy-production potential could be generalized to the case of p-growth with p>1 without any specific intricacy. In particular, one could consider the polynomial case Ψ(˜y,ξ)=K(˜y)ξp2Yξ and coercivity (2.20) would then read

    Ψ(˜y,η)+Ψ(˜y,ξ)cηpY+cξpY1c

    for 1/p+1/p=1. This setting would correspond to the case of doubly-nonlinear GENERIC continuous dynamics, namely

    y=L(y)DE(y)+K(y)S(y)p2YS(y).

    Again, under the noninteraction condition (2.4), suitably regular solutions conserve energy, since DE(y),Ψ(y,S(y))=S(y)p2YS(y),K(y)DE(y)=0, and have entropy rate (d/dt)(S(y))=S(y)p2YS(y),K(y)DS(y)0.

    Let us further remark that the conditional convergence result of Theorem 2.1 can serve as an a-posteriori tool to check the convergence of time-discrete approximations {yni}, regardless of the method used to generate them. Indeed, as already mentioned in the statement of Theorem 2.1, {yni} need not be a minimizer. In particular, the case of suitable approximate minimizers can be treated as well. In essence, relation (2.10) can be seen as a sort of a-posteriori error estimator.

    In this section, we analyze the simplest GENERIC system fulfilling conditions (2.10)–(2.13). We consider the case of the damped harmonic oscillator, namely

    mq"+νq+κq+λθ=0, (3.1)
    cθ=ν(q)2+λqθ. (3.2)

    Here, qR represents the position of the harmonic oscillator and θ>0 is its absolute temperature. Note that the case of θ being the given, constant temperature of a surrounding heat bath is considered in [38] instead. The positive constants m, ν, κ, λ and c are the mass of the oscillator, the viscosity of the medium, the elastic modulus, the thermal-exchange coefficient, and the heat capacity, respectively.

    By letting pR be the momentum of the harmonic oscillator, we rewrite (3.1)–(3.2) as the first-order system

    q=pm, (3.3)
    p=νpmκqλθ, (3.4)
    θ=νp2m2c+λpθmc. (3.5)

    System (3.3)–(3.5) can be written in the GENERIC form (2.1) by letting y=(q,p,θ)Y=R2×(0,) represent the state of the harmonic oscillator and by defining the total energy and the total entropy as

    E(q,p,θ)=p22m+cθ+κq22  and  S(q,p,θ)=λq+c+clnθ.

    These definitions comply with the classical Helmholtz relations under the choice for the free energy

    Φ(q,p,θ)=κq22+λqθcθlnθ.

    In particular, S=θΦ and E=p2/(2m)+Φ+θS. Note that both E and S are smooth, S is convex, and S(y)> iff θ>0.

    The operators L,K:R3R3×3 are given by

    L(x)=(01010λθ/c0λθ/c0)  and   K(x)=νθ(00001p/(mc)0p/(mc)p2/(mc)2).

    One readily checks that L is antisymmetric and a tedious computation shows that the Jacobi identity holds. On the other hand, K is symmetric and positive semidefinite, while not being invertible. Note that K(y) has rank two for all p0, so that the construction in [38] does not apply here.

    By computing the gradients DE(y)=(κq,p/m,c) and DS(y)=(λ,0,c/θ), one can easily check that the noninteraction condition (2.4) holds.

    Letting ξ=(ξq,ξp,ξθ)R3, η=(ηq,ηp,ηθ)R3, and ˜y=(˜q,˜p,˜θ)R2×(0,), the entropy-production potential and its dual read

    Ψ(˜y,ξ)=12ξK(˜y)ξ=ν˜θ2(ξp˜pmcξθ)2,Ψ(˜y,η)={12η2pν˜θif ηq=0 and ˜pηp+mcηθ=0,otherwise.

    We now choose ϕ=1/2, so that the state dependences in Ψ, Ψ, and DE are evaluated at the midpoint. With the definitions above, the incremental functional G:(0,)×R3×R3(,] for y=(q,p,θ) and ¯y=(¯q,¯p,¯θ) takes the form

    G(τ,¯y;y)=λqclnθ+τ2ν2(θ+¯θ)(p¯pτ+κq+¯q2+λθ)2+τν2θ+¯θ2(p+¯p2)2(1mθ)2λ¯q+cln¯θif  θ>0,  q¯qτ=p+¯p2m,  and  p22m+cθ+κq22=¯p22m+c¯θ+κ¯q22,G(τ,¯y;y)=otherwise.

    Assume to be given the time partition 0=t0<<tN=T with τi=titi1 for i=1,,N. By letting ¯y=(qi1,pi1,θi1) and defining (q0,p0,θ0)=(q0,p0,θ0) for some given initial state (q0,p0,θ0)R2×(0,), the incremental minimization problem (2.9) becomes

    min(qi,pi,θi){λqiclnθi+τiν(θi+θi1)(pipi1τi+κqi+qi12+λθi)2+τiν2θi+θi12(pi+pi12)21m2θ2iλqi1+clnθi1} (3.6)
    under the constraintsθi>0, (3.7)
    qiqi1τi=pi+pi12m, (3.8)
    p2i2m+cθi+κq2i2=p2i12m+cθi1+κq2i12,for i=1,,N. (3.9)

    Note in particular that the constraint (3.9) is nothing but the conservation of energy, namely

    E(qi,pi,θi)=E(qi1,pi1,θi1).

    This property is of course of great physical relevance and plays a crucial role in the a priori estimates. In particular, one has that

    E(qi,pi,θi)=E(q0,p0,θ0).

    This entails the bound

    |pi|+|θi|+|qi|Ci=1,,N, (3.10)

    where, here and in the following, the symbol C denotes a generic positive constant depending on the initial data y0 and on material parameters, but not on the time partition and, in particular, not on i. In addition, bound (3.10) entails that there exists θmin>0 independent of the time partition such that θiθmin, see (3.19) below. Note that energy conservation is specific of the current choice ϕ=1/2 and hinges on the quadratic nature of E.

    We devote the remainder of this section to prove that the convergence result of Theorem 2.1 applies to the scheme (3.6)–(3.9), namely that conditions (2.10)–(2.13) are fulfilled. In particular, we check that the minimization problem admits a solution {yi}Ni=0={(qi,pi,θi)}Ni=0 (Subsection 3.1) and that each such solution fulfills condition (2.10) in the stronger sense G(τi,yi1;yi)0 (Subsection 3.2), and that the bounds (2.11)–(2.13) can be established, independently of the partition (Subsection 3.3). Eventually, numerical simulations are presented in Subsection 3.4.

    Assume to be given ¯y=(qi1,pi1,θi1)R2×(0,). As yG(τi,¯y;y) is smooth on its domain, in order to prove that the scheme (3.6)–(3.9) admits a solution, we just need to check coercivity, namely that sublevels of yG(τi,¯y;y) are bounded.

    By using the constraints (3.8)–(3.9) we can express qi and θi in terms of pi and the values (qi1,pi1,θi1) in the form

    qi=τipi+pi12m+qi1  and  θi=f(pi), (3.11)

    where the second-order polynomial f in pi is defined by

    f(pi)=(12mcκτ2i8m2c)p2i+(κτ2ipi14m2cκτiqi12mc)pi+(p2i12mc+θi1κτ2ip2i18m2cκτipi1qi12mc).

    By substituting the two expressions in the minimum problem (3.6)–(3.9), we can reduce it to a minimization in the variable pi only. Indeed, problem (3.6)–(3.9) is equivalent to

    minpiF(pi)  under the constraint  f(pi)>0, (3.12)

    where we have defined

    F(pi):=τiλ(pi+pi1)2mclnf(pi):xx+τiν(f(pi)+θi1)(pipi1τi+τiκ2m(pi+pi12)+κqi1+λf(pi))2:xx+τiν2(f(pi)+θi12)(pi+pi12)21m2(f(pi))2.

    As F is smooth on its domain, in order to solve the minimization problem (3.12) we just need to prove that F has bounded sublevels. This follows from the fact that F(pi) if f(pi)0+. Hence, it remains to prove that f(pi)>0 holds on a bounded interval only, depending on (qi1,pi1,θi1), for τi chosen to be small enough. Since f has the form f(pi)=αp2i+βpi+γ, it suffices to check that α<0 and γ>0, for τi small enough. In fact, α<0 as soon as τ2<4m/κ. On the other hand, under the same upper bound on τi we have

    γ=(12mcκτ2i8m2c)p2i1+θi1κτipi1qi12mcθi1κτipi1qi12mc(3.10)θminκτiC22mc,

    where C is the constant in (3.10). In particular, γ>0 if τi is chosen such that τi<2mcθmin/(κC2).

    We have hence proved that, for all given (qi1,pi1,θi1)R2×(0,), we can find a solution pi of (3.12). We conclude from (3.11) that for τi small enough there exists (qi,pi,θi) solving (3.6)–(3.9). In particular, we have that θi>0.

    Moving from relation (2.19), the convexity of S ensures that condition (2.10) holds, for minimizers can be compared with solutions of the Euler scheme (2.18)

    qiqi1τi=pi+pi12m, (3.13)
    pipi1τi=ν(pi+pi1)(θi+θi1)4mκqi+qi12λθi, (3.14)
    θiθi1τi=νm2cθi+θi12θi(pi+pi12)2+λ(pi+pi1)θi2mc (3.15)

    for i=1,,N.

    By adapting the argument from Subsection 3.1, an elementary but tedious computation ensures that the Euler scheme (3.13)–(3.15) admits indeed a solution yei=(qi,pi,θi).

    Let now yi be a solution to the incremental minimization problem (3.6). Owing to (2.19), we conclude that

    G(τi,yi1;yi)=minyG(τi,yi1;y)G(τi,yi1;yei)(2.19)=0,

    where we have also used the fact that yei fulfills the constraints (3.7)–(3.9). In particular, condition (2.10) holds in the even stronger form

    G(τi,yi1;yi)0for  i=1,,N. (3.16)

    We now prove that condition (2.11)–(2.13) of Theorem 2.1 hold. This will follow by checking a priori bounds on minimizers {yi}Ni=0 of (3.6), independently from the time partition.

    Bound (3.10) and constraint (3.8) imply that

    |qiqi1τi|Ci=1,,N. (3.17)

    Moreover, we readily check that

    |L(yi)DE(yi+yi12)|=|(pi+pi12m,κqi+qi12λθi,λ(pi+pi1)θi2mc)|(3.10)Ci=1,,N. (3.18)

    Let us take the sum for i=1,,j for jN in (3.16), getting

    S(yj)+ψj+ψjS(y0),

    where

    ψj=ji=1τiν(θi+θi1)(pipi1τi+κqi+qi12+λθi)2,ψj=ji=1τiν2θi+θi12(pi+pi12)21m2θ2i.

    As ψj and ψj are nonnegative, we infer that S(yi) is nondecreasing. In particular,

    clnθjS(y0)λqj+c(3.10)Cj=1,,N.

    and consequently,

    θiθmin>0i=1,,N, (3.19)

    for some given θmin depending just on y0 and the material parameters. This ensures that

    |DS(yi)|=|(λ,0,c/θi)|Ci=1,,N. (3.20)

    It follows from the bound (3.10) on qi and θi that S(yi)=λqi+c+clnθiC for all i=1,,N. We conclude that ψNS(y0)+S(yN)C and

    Ni=1τiν(θi+θi1)|pipi1τi|2=ψNNi=1τiν(θi+θi1)(κqi+qi12+λθi)22Ni=1τiν(θi+θi1)(pipi1τi)(κqi+qi12+λθi)(3.10)C+Ni=1τi2ν(θi+θi1)|pipi1τi|2. (3.21)

    Using the fact that θi is uniformly bounded by (3.10), we can hence bound

    Ni=1τi|pipi1τi|24ν(maxiθi)Ni=1τi2ν(θi+θi1)|pipi1τi|2(3.10)+(3.21)C. (3.22)

    Eventually, we infer from constraint (3.9) that

    cNi=1τi|θiθi1τi|2Ni=1τi|12m(pi+pi1)pipi1τiκ2(qi+qi1)qiqi1τi|2C, (3.23)

    where the last inequality follows from bounds (3.10), (3.17), and (3.22).

    Bounds (3.10)–(3.17) and (3.22)–(3.23) imply condition (2.11) from Theorem 2.1. On the other hand, bounds (3.18) and (3.20) imply conditions (2.12) and (2.13), respectively. Hence, the convergence statement of Theorem 2.1 holds. In particular, given initial values (q0,p0,θ0)R3 with θ0>0, a sequence of partitions 0=tn0<<tnNn=T with τn=maxi(tnitni1)0, and corresponding minimizers {(qni,pni,θni)}Ni=0 of (3.6), it follows that

    (ˆqn,ˆpn,ˆθn)(q,p,θ)  in  H1(0,T;R3), (3.24)

    where (q,p,θ) solves the damped harmonic oscillator system (3.3)–(3.5) with initial value (q(0),p(0),θ(0))=(q0,p0,θ0). Note that solutions of (3.3)–(3.5) are unique. Hence, convergence (3.24) holds for the whole sequence of discrete solutions, not just for a subsequence.

    We record here some illustration of the performance of the minimizing-movements scheme (3.6)–(3.9). All computations are made in Matlab. In the following, we choose

    m=ν=κ=λ=c=1,y0=(q0,p0,θ0)=(1,1,1),T=15. (3.25)

    Given a uniform time-partition with time step τ=T/N, we find a solution {yi}Ni=0 of the minimizing-movements scheme (3.6)–(3.9) via Newton's method. The numerical reference solution ty(t) is calculated by means of the Matlab solver ode45 choosing 104 for the maximal time step and 108 for the absolute tolerance.

    Figures 13 illustrate the numerical reference solution and the time-discrete solution for τ=1/4. As expected, the minimizing-movement scheme conserves energy, see Figure 3 left, while entropy is nondecreasing, see Figure 3 right. A second set of experiments is illustrated in Figure 4. For the same choices in (3.25) and different time steps

    τn=2nfor  n=1,0,,11, (3.26)
    Figure 1.  Position q (left) and momentum p (right) with respect to time for the numerical reference solution (red, dotted) and for the minimizing-movements scheme (solid). The curves differ only slightly.
    Figure 2.  Temperature with respect to time for the numerical reference solution (red, dotted) and for the minimizing-movements scheme (solid).
    Figure 3.  Energy (left) and entropy (right) as function of time for the numerical reference solution (red dotted) and for the minimizing-movements scheme (solid).
    Figure 4.  Error maxt[0,T]|p(t)ˆp(t)| with respect to 1/τn. The (green) dotted lines indicate the order of convergence 1.

    we compute the uniform errors of the temperature component of the solution with respect to the numerical reference solution.

    As τ converges to 0, our computations confirm that the minimizing-movements is of order τ. Let us mention that a proof of first-order convergence for the minimizing-movements scheme in the nondissipative regime (L=0, K independent of the state) is given in [25,Prop. 4.3].

    The Authors gratefully acknowledge a remark by Alexander Mielke on an earlier version of this work, which eventually inspired the design of the numerical scheme for the damped harmonic oscillator. This research is supported by Austrian Science Fund (FWF) project F 65 and W 1245. AJ is supported by the FWF projects P 30000 and P 33010. US is supported by the FWF projects I 4354 and P 32788 and by the Vienna Science and Technology Fund (WWTF) through Project MA14-009.

    The authors declare no conflict of interest.



    [1] L. Ambrosio, Minimizing movements, Rend. Accad. Naz. Sci. XL Mem. Mat. Appl., 19 (1995), 191–246.
    [2] L. Ambrosio, N. Gigli, G. Savaré, Gradient flows in metric spaces and in the space of probability measures, 2 Eds., Basel: Birkhäuser Verlag, 2008.
    [3] F. Auricchio, E. Boatti, A. Reali, U. Stefanelli, Gradient structures for the thermomechanics of shape-memory materials, Comput. Methods Appl. Mech. Engrg., 299 (2016), 440–469. doi: 10.1016/j.cma.2015.11.005
    [4] A. Bacho, E. Emmrich, A. Mielke, An existence result and evolutionary Γ-convergence for perturbed gradient systems, J. Evol. Equ., 19 (2019), 479–522. doi: 10.1007/s00028-019-00484-x
    [5] P. Betsch, M. Schiebl, GENERIC-based formulation and discretization of initial boundary value problems for finite strain thermoelasticity, Comput. Mech., 65 (2020), 503–531. doi: 10.1007/s00466-019-01781-5
    [6] P. Betsch, M. Schiebl, Energy–momentum–entropy consistent numerical methods for large strain thermoelasticity relying on the GENERIC formalism, Internat. J. Numer. Methods Engrg., 119 (2019), 1216–1244.
    [7] H. Brézis, Opérateurs maximaux monotones et semi-groupes de contractions dans les espaces de Hilbert, Amsterdam: North-Holland, 1973.
    [8] E. Celledoni, V. Grimm, R. McLachlan, D. McLaren, D. O'Neale, B. Owren, et al., Preserving energy resp. dissipation in numerical PDEs using the "Average Vector Field" method, J. Comput. Phys., 231 (2012), 6770–6789. doi: 10.1016/j.jcp.2012.06.022
    [9] S. Conde Martín, P. Betsch, J. C. García Orden, A temperature based thermodynamically consistent integration scheme for discrete thermo-elastodynamics, Commun. Nonlinear Sci. Numer. Simul., 32 (2016), 63–80. doi: 10.1016/j.cnsns.2015.08.006
    [10] S. Conde Martín, J. C. García Orden, On energy–entropy– momentum integration methods for discrete thermo-viscoelastodynamics, Comput. Struct., 181 (2017), 3–20. doi: 10.1016/j.compstruc.2016.05.010
    [11] E. De Giorgi, A. Marino, M. Tosques, Problems of evolution in metric spaces and maximal decreasing curve, Atti Accad. Naz. Lincei Rend. Cl. Sci. Fis. Mat. Natur. (8), 68 (1980), 180–187.
    [12] P. Dondl, T. Frenzel, A. Mielke, A gradient system with a wiggly energy and relaxed EDP-convergence, ESAIM Control Optim. Calc. Var., 25 (2019), 68. doi: 10.1051/cocv/2018058
    [13] M. H. Duong, Formulation of the relativistic heat equation and the relativistic kinetic Fokker-Planck equations using GENERIC, Phys. A, 439 (2015), 34–43. doi: 10.1016/j.physa.2015.07.019
    [14] M. H. Duong, M. A. Peletier, J. Zimmer, GENERIC formalism of a Vlasov–Fokker–Planck equation and connection to large-deviation principles, Nonlinearity, 26 (2013), 2951–2971. doi: 10.1088/0951-7715/26/11/2951
    [15] J. C. García Orden, I. Romero, Energy-Entropy-Momentum integration of discrete thermo-visco-elastic dynamics, Eur. J. Mech. Solids, 32 (2012), 76–87. doi: 10.1016/j.euromechsol.2011.09.007
    [16] O. Gonzalez, Time integration and discrete Hamiltonian systems, J. Nonlin. Sci., 6 (1996), 449–467. doi: 10.1007/BF02440162
    [17] O. Gonzalez, Exact energy-momentum conserving algorithms for general models in nonlinear elasticity, Comput. Methods Appl. Mech. Engrg., 190 (2000), 1763–1783. doi: 10.1016/S0045-7825(00)00189-4
    [18] M. Grmela, H. C. Öttinger, Dynamics and thermodynamics of complex fluids. I. Development of a general formalism, Phys. Rev. E, 56 (1997), 6620–6632.
    [19] E. Hairer, C. Lubich, G. Wanner, Geometric numerical integration, 2 Eds., Berlin: Springer, 2006.
    [20] M. Hoyuelos, GENERIC framework for the Fokker-Planck equation, Phys. A, 442 (2016), 350–358. doi: 10.1016/j.physa.2015.09.035
    [21] M. Hütter, T. A. Tervoort, Finite anisotropic elasticity and material frame indifference from a nonequilibrium thermodynamics perspective, J. Non-Newton. Fluid., 152 (2008), 45–52. doi: 10.1016/j.jnnfm.2007.10.009
    [22] M. Hütter, B. Svendsen, On the formulation of continuum thermodynamic models for solids as general equations for non-equilibrium reversible–irreversible coupling, J. Elast., 104 (2011), 357–368. doi: 10.1007/s10659-011-9327-4
    [23] M. Hütter, B. Svendsen, Thermodynamic model formulation for viscoplastic solids as general equations for nonequilibrium reversible–irreversible coupling, Contin. Mech. Thermodyn., 24 (2012), 211–227. doi: 10.1007/s00161-011-0232-7
    [24] A. Jelić, M. Hütter, H. C. Öttinger, Dissipative electromagnetism from a nonequilibrium thermodynamics perspective, Phys. Rev. E, 74 (2006), 041126. doi: 10.1103/PhysRevE.74.041126
    [25] A. Jüngel, U. Stefanelli, L. Trussardi, Two structure-preserving time discretizations for gradient flows, Appl. Math. Optim., 80 (2020), 733–764.
    [26] R. C. Kraaij, A. Lazarescu, C. Maes, M. Peletier, Fluctuation symmetry leads to GENERIC equations with non-quadratic dissipation, Stochastic Process. Appl., 130 (2020), 139–170. doi: 10.1016/j.spa.2019.02.001
    [27] R. Kraaij, A. Lazarescu, C. Maes, M. Peletier, Deriving GENERIC from a generalized fluctuation symmetry, J. Stat. Phys., 170 (2018), 492–508. doi: 10.1007/s10955-017-1941-5
    [28] M. Krüger, M. Groß, P. Betsch, An energy–entropy-consistent time stepping scheme for nonlinear thermo-viscoelastic continua, ZAMM Z. Angew. Math. Mech., 96 (2016), 141–178. doi: 10.1002/zamm.201300268
    [29] M. Liero, A. Mielke, M. A. Peletier, D. R. M. Renger, On microscopic origins of generalized gradient structures, Discrete Contin. Dyn. Syst. Ser. S, 10 (2017), 1–35.
    [30] R. McLachlan, G. Quispel, N. Robidoux, Geometric integration using discrete gradients, Phil. Trans. R. Soc. Lond. A, 357 (1999), 1021–1045. doi: 10.1098/rsta.1999.0363
    [31] A. Mielke, Formulation of thermoelastic dissipative material behavior using GENERIC, Contin. Mech. Thermodyn., 23 (2011), 233–256. doi: 10.1007/s00161-010-0179-0
    [32] A. Mielke, Dissipative quantum mechanics using GENERIC, In: Proc. of the conference on recent trends in dynamical systems, Springer, 2013,555–585.
    [33] A. Mielke, A. Montefusco, M. Peletier, Exploring families of energy-dissipation landscapes via tilting–three types of EDP convergence, Contin. Mech. Thermodyn., 33 (2021), 611–637. doi: 10.1007/s00161-020-00932-x
    [34] A. Mielke, R. Rossi, G. Savaré, BV solutions and viscosity approximations of rate-independent systems, ESAIM Control Optim. Calc. Var., 18 (2012), 36–80. doi: 10.1051/cocv/2010054
    [35] A. Montefusco, F. Consonni, G. Beretta, Essential equivalence of the general equation for the nonequilibrium reversible-irreversible coupling (GENERIC) and steepest-entropy-ascent models of dissipation for nonequilibrium thermodynamics, Phys. Rev. E, 91 (2015), 042138. doi: 10.1103/PhysRevE.91.042138
    [36] B. S. Mordukhovich, Variational analysis and generalized differentiation. I. Basic theory, Berlin: Springer-Verlag, 2006.
    [37] H. C. Öttinger, Beyond equilibrium thermodynamics, New Jersey: John Wiley, 2005.
    [38] H. C. Öttinger, Generic integrators: structure preserving time integration for thermodynamic systems, J. Non-Equil. Thermody., 43 (2018), 89–100. doi: 10.1515/jnet-2017-0034
    [39] D. Portillo, J. C. García Orden, I. Romero, Energy–entropy– momentum integration schemes for general discrete non-smooth dissipative problems in thermomechanics, Internat. J. Numer. Methods Engrg., 112 (2017), 776–802. doi: 10.1002/nme.5532
    [40] L. Portinale, U. Stefanelli, Penalization via global functionals of optimal-control problems for dissipative evolution, Adv. Math. Sci. Appl., 28 (2019), 425–447.
    [41] G. Quispel, D. McLaren, A new class of energy-preserving numerical integration methods, J. Phys. A-Math. Theor., 41 (2008), 045206. doi: 10.1088/1751-8113/41/4/045206
    [42] T. Roche, R. Rossi, U. Stefanelli, Stability results for doubly nonlinear differential inclusions by variational convergence, SIAM J. Control Optim., 52 (2014), 1071–1107. doi: 10.1137/130909391
    [43] I. Romero, Thermodynamically consistent time-stepping algorithms for non-linear thermomechanical systems, Internat. J. Numer. Methods Engrg., 79 (2009), 706–732. doi: 10.1002/nme.2588
    [44] I. Romero, Algorithms for coupled problems that preserve symmetries and the laws of thermodynamics. Part I: monolithic integrators and their application to finite strain thermoelasticity, Comput. Methods Appl. Mech. Engrg., 199 (2010), 1841–1858. doi: 10.1016/j.cma.2010.02.014
    [45] I. Romero, Algorithms for coupled problems that preserve symmetries and the laws of thermodynamics. Part II: fractional step methods, Comput. Methods Appl. Mech. Engrg., 199 (2010), 2235–2248. doi: 10.1016/j.cma.2010.03.016
    [46] R. Rossi, G. Savaré, {Gradient flows of non convex functionals in Hilbert spaces and applications}, ESAIM Control Optim. Calc. Var., 12 (2006), 564–614. doi: 10.1051/cocv:2006013
    [47] R. Rossi, A, Mielke, G. Savaré, A metric approach to a class of doubly nonlinear evolution equations and applications, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 7 (2008), 97–169.
    [48] S. Sato, T. Matsuo, H. Suzuki, D. Furihata, A Lyapunov-type theorem for dissipative numerical integrators with adaptive time-stepping, SIAM J. Numer. Anal., 53 (2015), 2505–2518. doi: 10.1137/140996719
    [49] J. Simon, Compact sets in the space Lp(0,T;B), Ann. Mat. Pura Appl. (4), 146 (1987), 65–96.
    [50] Y. Suzuki, M. Ohnawa, GENERIC formalism and discrete variational derivative method for the two-dimensional vorticity equation, J. Comput. Appl. Math., 296 (2016), 690–708. doi: 10.1016/j.cam.2015.10.018
  • This article has been cited by:

    1. Serena Dipierro, Luca Lombardini, Partial differential equations from theory to applications: Dedicated to Alberto Farina, on the occasion of his 50th birthday, 2023, 5, 2640-3501, 1, 10.3934/mine.2023050
    2. Matthias Erbar, Zihui He, A variational approach to a fuzzy Boltzmann equation, 2025, 38, 0951-7715, 055019, 10.1088/1361-6544/adcb7e
  • Reader Comments
  • © 2022 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(2717) PDF downloads(182) Cited by(2)

Figures and Tables

Figures(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog