Loading [MathJax]/extensions/TeX/boldsymbol.js

Stochastic two-scale convergence and Young measures

  • Received: 01 June 2021 Revised: 01 January 2022 Published: 22 February 2022
  • Primary: 74Q05, 47J30; Secondary: 49J45

  • In this paper we compare the notion of stochastic two-scale convergence in the mean (by Bourgeat, Mikelić and Wright), the notion of stochastic unfolding (recently introduced by the authors), and the quenched notion of stochastic two-scale convergence (by Zhikov and Pyatnitskii). In particular, we introduce stochastic two-scale Young measures as a tool to compare mean and quenched limits. Moreover, we discuss two examples, which can be naturally analyzed via stochastic unfolding, but which cannot be treated via quenched stochastic two-scale convergence.

    Citation: Martin Heida, Stefan Neukamm, Mario Varga. Stochastic two-scale convergence and Young measures[J]. Networks and Heterogeneous Media, 2022, 17(2): 227-254. doi: 10.3934/nhm.2022004

    Related Papers:

    [1] Mogtaba Mohammed, Mamadou Sango . Homogenization of nonlinear hyperbolic stochastic partial differential equations with nonlinear damping and forcing. Networks and Heterogeneous Media, 2019, 14(2): 341-369. doi: 10.3934/nhm.2019014
    [2] Patrick Henning . Convergence of MsFEM approximations for elliptic, non-periodic homogenization problems. Networks and Heterogeneous Media, 2012, 7(3): 503-524. doi: 10.3934/nhm.2012.7.503
    [3] Erik Grandelius, Kenneth H. Karlsen . The cardiac bidomain model and homogenization. Networks and Heterogeneous Media, 2019, 14(1): 173-204. doi: 10.3934/nhm.2019009
    [4] Hakima Bessaih, Yalchin Efendiev, Florin Maris . Homogenization of the evolution Stokes equation in a perforated domain with a stochastic Fourier boundary condition. Networks and Heterogeneous Media, 2015, 10(2): 343-367. doi: 10.3934/nhm.2015.10.343
    [5] Alexander Mielke, Sina Reichelt, Marita Thomas . Two-scale homogenization of nonlinear reaction-diffusion systems with slow diffusion. Networks and Heterogeneous Media, 2014, 9(2): 353-382. doi: 10.3934/nhm.2014.9.353
    [6] Catherine Choquet, Ali Sili . Homogenization of a model of displacement with unbounded viscosity. Networks and Heterogeneous Media, 2009, 4(4): 649-666. doi: 10.3934/nhm.2009.4.649
    [7] Liselott Flodén, Jens Persson . Homogenization of nonlinear dissipative hyperbolic problems exhibiting arbitrarily many spatial and temporal scales. Networks and Heterogeneous Media, 2016, 11(4): 627-653. doi: 10.3934/nhm.2016012
    [8] Grigor Nika, Adrian Muntean . Hypertemperature effects in heterogeneous media and thermal flux at small-length scales. Networks and Heterogeneous Media, 2023, 18(3): 1207-1225. doi: 10.3934/nhm.2023052
    [9] Jean-Yves Le Boudec . The stationary behaviour of fluid limits of reversible processes is concentrated on stationary points. Networks and Heterogeneous Media, 2013, 8(2): 529-540. doi: 10.3934/nhm.2013.8.529
    [10] Markus Gahn, Maria Neuss-Radu, Peter Knabner . Effective interface conditions for processes through thin heterogeneous layers with nonlinear transmission at the microscopic bulk-layer interface. Networks and Heterogeneous Media, 2018, 13(4): 609-640. doi: 10.3934/nhm.2018028
  • In this paper we compare the notion of stochastic two-scale convergence in the mean (by Bourgeat, Mikelić and Wright), the notion of stochastic unfolding (recently introduced by the authors), and the quenched notion of stochastic two-scale convergence (by Zhikov and Pyatnitskii). In particular, we introduce stochastic two-scale Young measures as a tool to compare mean and quenched limits. Moreover, we discuss two examples, which can be naturally analyzed via stochastic unfolding, but which cannot be treated via quenched stochastic two-scale convergence.



    In this paper we compare quenched stochastic two-scale convergence [38] with the notion of stochastic unfolding [30,19], which is equivalent to stochastic two-scale convergence in the mean [6]. In particular, we introduce the concept of stochastic two-scale Young measures to relate quenched stochastic two-scale limits with the mean limit and discuss examples of convex homogenization problems that can be treated with two-scale convergence in the mean, but not conveniently in the quenched setting of two-scale convergence.

    Two-scale convergence has been introduced in [32,1,25] for homogenization problems (partial differential equations or variational problems) with periodic coefficients. The essence of two-scale convergence is that the two-scale limit of an oscillatory sequence captures oscillations that emerge along the sequence and that are to leading order periodic on a definite microscale, typically denoted by ε>0. It is especially well-suited for problems where oscillations of solutions solely stem from prescribed oscillations of the coefficients or the data. For instance, this is the case for equations with a monotone structure or convex variational problems. In contrast, problems that feature pattern formation to leading order (e.g., nonconvex variational problems or singular partial differential equations with non-convex domain) typically cannot be conveniently treated with two-scale convergence. Another well established method for periodic homogenization is periodic unfolding, see [9,35,27,10] as well as [36,3] for the periodic modulation method, which is related. These methods build on an isometric operator---the periodic unfolding (or dilation) operator. It allows us to embed oscillatory sequences into a larger two-scale space and to transform an oscillatory problem into an "unfolded" problem on the two-scale space. The latter often features a better separation of macro- and microscopic properties, which often is convenient for the analysis. We refer to [14,7,28,8,15,24,26] for various interesting applications of this method. Both notions are closely linked, since weak convergence of "unfolded" sequence in the two-scale space is equivalent to weak two-scale convergence, see [5].

    In this paper we are interested in stochastic homogenization, i.e. problems with random coefficients with a stationary distribution. The first stochastic homogenization result has been obtained by Papanicolaou and Varadhan in [33] (and independently by Kozlov [23]) for linear, elliptic equations with stationary and ergodic random coefficients on Rd. In their seminal paper, Papanicolaou and Varadhan introduce a functional analytic framework, which, by now, is the standard way to model random coefficients. We briefly recall it in the special case of convex integral functionals with quadratic growth: Let (Ω,F,P) denote a probability space of parameter fields ωΩ and let τx:ΩΩ, xRd, denote a measure preserving and ergodic group action, see Assumption 2.1 for details. A standard model for a convex, integral functional with a stationary, ergodic, random microstructure on scale ε>0 is then given by the functional Eωε:H1(Q)R{},

    Eωε(u)=QV(τxεω,u(x))f(x)u(x)dx

    where QRd denotes an open and bounded domain, fL2(Q), and V(ω,F) is an integrand that is measurable in ωΩ, convex in FRd, and satisfies a quadratic growth condition. A classical result [11] shows that in the homogenization limit ε0, the functionals Γ-converge to the homogenized functional Ehom:H1(Q)R{}, given by

    Ehom(u)=QVhom(u(x))f(x)u(x)dx,

    where Vhom is a deterministic, convex integrand and characterized by a homogenization formula, see (31) below. There are different natural choices for the topology when passing to this limit:

    ● In the mean setting, minimizers uωε of Eωε, ωΩ, are viewed as random fields (ω,x)uωε(x) in L2(Ω;H1(Q)) and one considers Γ-convergence of the averaged functional L2(Ω;H1(Q))uΩEε(u)dP w.r.t. strong convergence in L2(Ω×Q). In fact, the first result in stochastic homogenization [33] establishes convergence of solutions in this mean sense.

    ● In the quenched setting, one studies the limiting behavior of a minimizer uεH1(Q) of Eωε for fixed ωΩ. One then considers Γ-convergence of Eωε w.r.t. strong convergence in L2(Q) for P-a.a. ωΩ.

    Similarly, two variants of stochastic two-scale convergence have been introduced as generalizations of periodic two-scale convergence (for the sake of brevity, we restrict the following review to the Hilbert-space case p=2, and note that the following extends to Lp(Ω×Q) with p(1,)):

    ● In [6,2] the mean variant has been introduced as follows: We say that a sequence of random fields (uε)L2(Ω×Q) stochastically two-scale converges in the mean to uL2(Ω×Q), if

    limε0Ω×Quε(ω,x)φ(τxεω,x)dP(ω)dx=Ω×Qu(ω,x)φ(ω,x)dP(ω)dx, (1)

    for all admissible test functions φL2(Ω×Q), see Remark 1 for details.

    ● More recently, Zhikov and Pyatnitskii introduced in [38] a quenched variant: We say that a sequence (uε)L2(Q) quenched stochastically two-scale converges to uL2(Ω×Q) w.r.t. to a fixed parameter field ω0Ω, if

    limε0Quε(x)φ(τxεω0,x)dx=Ω×Qu(ω,x)φ(ω,x)dP(ω)dx,

    for all admissible test functions φL2(Ω×Q). Note that the two-scale limit u a priori depends on ω0. In fact, in [37] (see also [16]) quenched two-scale convergence has been introduced in a very general setting that includes the case of integration against random, rapidly oscillating measures, which naturally emerge when describing coefficients defined relative to random geometries. In this work, we restrict our considerations to the simplest case where the random measure is the Lebesgue measure.

    Similarly to the periodic case, stochastic two-scale convergence in the mean can be rephrased with help of a transformation operator, see [34,19,30], where the stochastic unfolding operator Tε:L2(Ω×Q)L2(Ω×Q),

    Tεu(ω,x)=u(τxεω,x), (2)

    has been introduced. As in the periodic case, it is a linear isometry and it turns out that for a bounded sequence (uε)L2(Ω×Q), stochastic two-scale convergence in the mean is equivalent to weak convergence of the unfolded sequence Tεuε. As we demonstrate below in Section 4.1, the stochastic unfolding method leads to a very economic and streamlined analysis of convex homogenization problems. Moreover, it allows us to derive two-scale functionals of the form E(u,χ)=ΩQV(ω,u(x)+χ(ω,x))dxdP as a Γ-limit of Eε, see Theorem 4.1 for details. In contrast to the periodic case, where the unfolding operator is an isometry from L2(Rd) to L2(Y×Rd) (with Y denoting the unit torus), in the random case it is not possible to interpret (2) as a continuous operator from L2(Q) to L2(Ω×Q). Therefore, quenched two-scale convergence cannot be characterized via stochastic unfolding directly.

    In the present paper we compare the different notions of stochastic two-scale convergence. Although the mean and quenched notion of two-scale convergence look quite similar, it is non-trivial to relate both. As a main result, we introduce stochastic two-scale Young measures as a tool to compare quenched and mean limits, see Theorem 3.12. The construction invokes a metric characterization of quenched stochastic two-scale convergence, which is a tool of independent interest, see Lemma 3.6. As an application we demonstrate how to lift a mean two-scale homogenization result to a quenched statement, see Section 4.3. Moreover, we present two examples that can only be conveniently treated with the mean notion of two-scale convergence. In the first example, see Section 4.1, the assumption of ergodicity is dropped (as it is natural in the context of periodic representative volume approximation schemes). In the second example we consider a model that invokes a mean field interaction in form of a variance-type regularization of a convex integral functional with degenerate growth, see Section 4.2.

    Structure of the paper. In the following section we present the standard setting for stochastic homogenization. In Section 3 we provide the main properties of the stochastic unfolding method, present the most important facts about quenched two-scale convergence and present our results about Young measures. In Section 4 we present examples of stochastic homogenization and applications of the methods developed in this paper.

    In the following we briefly recall the standard setting for stochastic homogenization. Throughout the entire paper we assume the following:

    Assumption 2.1. Let (Ω,F,P) be a complete and separable probability space. Let τ={τx}xRd denote a group of invertible measurable mappings τx:ΩΩ such that:

    (i)(Group property). τ0=Id and τx+y=τxτy for all x,yRd.

    (ii)(Measure preservation). P(τxE)=P(E) for all EF and xRd.

    (iii)(Measurability). (ω,x)τxω is (FL(Rd),F)-measurable, where L(Rd) denotes the Lebesgue σ-algebra.

    We write to denote the expectation ΩdP. By the separability assumption on the measure space it follows that Lp(Ω) is separable for p1. The proof of the following lemma is a direct consequence of Assumption 2.1, thus we omit it.

    Lemma 2.2 (Stationary extension). Let φ:ΩR be F-measurable. Let QRd be open and denote by L(Q) the corresponding Lebesgue σ-algebra. Then Sφ:Ω×QR, Sφ(ω,x):=φ(τxω) defines an FL(Q)-measurable function – called the stationary extension of φ. Moreover, if Q is bounded, for all 1p< the map S:Lp(Ω)Lp(Ω×Q) is a linear injection satisfying

    SφLp(Ω×Q)=|Q|1pφLp(Ω).

    We say (Ω,F,P,τ) is ergodic ( is ergodic), if

     every shift invariant AF (i.e. τxA=A for all xRd) satisfies P(A){0,1}.

    In this case the celebrated Birkhoff's ergodic theorem applies, which we recall in the following form:

    Theorem 2.3(Birkhoff's ergodic Theorem [12,Theorem 10.2.II]). Let be ergodic and φ:ΩR be integrable. Then for P-a.a. ωΩ it holds: Sφ(ω,) is locally integrable and for all open, bounded sets QRd we have

    limε0QSφ(ω,xε)dx=|Q|φ. (3)

    Furthermore, if φLp(Ω) with 1p, then for P-a.a. ωΩ it holds: Sφ(ω,)Lploc(Rd), and provided p< it holds Sφ(ω,ε)φ weakly in Lploc(Rd) as ε0.

    Stochastic gradient. For p(1,) consider the group of isometric operators {Ux:xRd} on Lp(Ω) defined by Uxφ(ω)=φ(τxω). This group is strongly continuous (see [22,Section 7.1]). For i=1,...,d, we consider the 1-parameter group of operators {Uhei:hR} and its infinitesimal generator Di:DiLp(Ω)Lp(Ω)

    Diφ=limh0Uheiφφh,

    which we refer to as stochastic derivative. Di is a linear and closed operator and its domain Di is dense in Lp(Ω). We set W1,p(Ω)=di=1Di and define for φW1,p(Ω) the stochastic gradient as Dφ=(D1φ,...,Ddφ). In this way, we obtain a linear, closed and densely defined operator D:W1,p(Ω)Lp(Ω)d, and we denote by

    Lppot(Ω):=¯R(D)Lp(Ω)d (4)

    the closure of the range of D in Lp(Ω)d. We denote the adjoint of D by D:DLq(Ω)dLq(Ω) where here and below q:=pp1 denotes the dual exponent. It is a linear, closed and densely defined operator (D is the domain of D). We define the subspace of shift invariant functions in Lp(Ω) by

    Lpinv(Ω)={φLp(Ω):Uxφ=φfor all xRd},

    and denote by Pinv:Lp(Ω)Lpinv(Ω) the conditional expectation with respect to the σ-algebra of shift invariant sets {AF:τxA=A for all xRd}. Pinv a contractive projection and for p=2 it coincides with the orthogonal projection onto L2inv(Ω). The following well-known equivalence holds:

     is ergodic  Lpinv(Ω)R  Pinvf=f.

    Random fields. We introduce function spaces for functions defined on Ω×Q as follows: For closed subspaces XLp(Ω) and YLp(Q), we denote by XY the closure of

    XaY:={ni=1φiηi:φiX,ηiY,nN}

    in Lp(Ω×Q). Note that in the case X=Lp(Ω) and Y=Lp(Q), we have XY=Lp(Ω×Q). Up to isometric isomorphisms, we may identify Lp(Ω×Q) with the Bochner spaces Lp(Ω;Lp(Q)) and Lp(Q;Lp(Ω)). Slightly abusing the notation, for closed subspaces XLp(Ω) and YW1,p(Q), we denote by XY the closure of

    XaY:={ni=1φiηi:φiX,ηiY,nN}

    in Lp(Ω;W1,p(Q)). In this regard, we may identify uLp(Ω)W1,p(Q) with the pair (u,u)Lp(Ω×Q)1+d. We mostly focus on the space Lp(Ω×Q) and the above notation is convenient for keeping track of its various subspaces.

    In the following we first discuss two notions of stochastic two-scale convergence and their connection through Young measures. In particular, Section 3.1 is devoted to the introduction of the stochastic unfolding operator and its most important properties. In Section 3.2 we discuss quenched two-scale convergence and its properties. Section 3.3 presents the results about Young measures.

    In the following we briefly introduce the stochastic unfolding operator and provide its main properties, for the proofs and detailed studies we refer to [19,30,31,34].

    Lemma 3.1([19,Lemma 3.1]). Let ε>0, 1<p<, q=pp1, and QRd be open. There exists a unique linear isometric isomorphism

    Tε:Lp(Ω×Q)Lp(Ω×Q)

    such that

    uLp(Ω)aLp(Q):(Tεu)(ω,x)=u(τxεω,x)a.e.inΩ×Q.

    Moreover, its adjoint is the unique linear isometric isomorphism Tε:Lq(Ω×Q)Lq(Ω×Q) that satisfies (Tεu)(ω,x)=u(τxεω,x) a.e. in Ω×Q for all uLq(Ω)aLq(Q), q:=pp1.

    Definition 3.2 (Unfolding and two-scale convergence in the mean). The operator Tε:Lp(Ω×Q)Lp(Ω×Q) in Lemma 3.1 is called the stochastic unfolding operator. We say that a sequence (uε)Lp(Ω×Q) weakly (strongly) two-scale converges in the mean in Lp(Ω×Q) to uLp(Ω×Q) if (as ε0)

    Tεuεu weakly (strongly) in Lp(Ω×Q).

    In this case we write uε2u (uε2u) in Lp(Ω×Q).

    Remark 1 (Equivalence to stochastic two-scale convergence in the mean). Stochastic two-scale convergence in the mean was introduced in [6]. In particular, it is said that a sequence of random fields uεLp(Ω×Q) stochastically two-scale converges in the mean if

    limε0Quε(ω,x)φ(τxεω,x)dx=Qu(ω,x)φ(ω,x)dx, (5)

    for any φLq(Ω×Q), q=pp1, that is admissible, i.e., in the sense that the transformation (ω,x)φ(τxεω,x) is well-defined. For a bounded sequence uεLp(Ω×Q), (5) is equivalent to Tεuεu weakly in Lp(Ω×Q), i.e., to weak stochastic two-scale convergence in the mean. Indeed, with help of Tε (and its adjoint) we might rephrase the integral on the left-hand side in (5) as

    Quε(Tεφ)dx=Q(Tεuε)φdx, (6)

    which proves the equivalence.

    We summarize some of the main properties:

    Proposition 1 (Main properties). Let p(1,), q=pp1 and QRd be open.

    (i)(Compactness, [19,Lemma 3.4].) If lim supε0uεLp(Ω×Q)<, then there exists a subsequence ε and uLp(Ω×Q) such that uε2u in Lp(Ω×Q).

    (ii)(Limits of gradients, [19,Proposition 3.7]) Let (uε) be a bounded sequence in Lp(Ω)W1,p(Q). Then, there exist uLpinv(Ω)W1,p(Q) and χLppot(Ω)Lp(Q) such that (up to a subsequence)

    uε2uinLp(Ω×Q),uε2u+χinLp(Ω×Q)d. (7)

    If, additionally, is ergodic, then u=Pinvu=uW1,p(Q) and uεu weakly in W1,p(Q).

    (iii)(Recovery sequences, [19,Lemma 4.3]) Let uLpinv(Ω)W1,p(Q) and χLppot(Ω)Lp(Q). There exists uεLp(Ω)W1,p(Q) such that

    uε2u,uε2u+χinLp(Ω×Q).

    If additionally uLpinv(Ω)W1,p0(Q), we have uεLp(Ω)W1,p0(Q).

    In this section, we recall the concept of quenched stochastic two-scale convergence (see [38,16]). The notion of quenched stochastic two-scale convergence is based on the individual ergodic theorem, see Theorem 2.3. We thus assume throughout this section that

    isergodic.

    Moreover, throughout this section we fix exponents p(1,), q:=pp1, and an open and bounded domain QRd. We denote by (Bp,Bp) the Banach space Lp(Ω×Q) and the associated norm, and we write (Bp) for the dual space. For the definition of quenched two-scale convergence we need to specify a suitable space of test-functions in Bq that is countably generated. To that end we fix sets DΩ and DQ such that

    DΩ is a countable set of bounded, measurable functions on (Ω,F) that contains the identity 1Ω1 and is dense in L1(Ω) (and thus in Lr(Ω) for any 1r<).

    DQC(¯Q) is a countable set that contains the identity 1Q1 and is dense in L1(Q) (and thus in Lr(Q) for any 1r<).

    We denote by

    A:={φ(ω,x)=φΩ(ω)φQ(x):φΩDΩ,φQDQ}

    the set of simple tensor products (a countable set), and by D0 the Q-linear span of A, i.e.

    D0:={mj=1λjφj:mN,λ1,,λmQ,φ1,,φmA}.

    We finally set

    D:=spanA=spanD0and¯D:=span(DQ)

    (the span of DQ seen as a subspace of D), and note that D and D0 are dense subsets of Bq, while the closure of ¯D in Bq is isometrically isomorphic to Lq(Q). Let us anticipate that D serves as our space of test-functions for stochastic two-scale convergence. As opposed to two-scale convergence in the mean, "quenched" stochastic two-scale convergence is defined relative to a fixed "admissible" realization ω0Ω. Throughout this section we denote by

    Ω0the set of admissible realizations;

    it is a set of full measure determined by the following lemma:

    Lemma 3.3. There exists a measurable set Ω0Ω with P(Ω0)=1 s.t. for all φ,φA, all ω0Ω0, and r{p,q} we have with (Tεφ)(ω,x):=φ(τxεω,x),

    lim supε0(Tεφ)(ω0,)Lr(Q)φBrandlimε0QTε(φφ)(ω0,x)dx=Q(φφ)(ω0,x)dx.

    Proof. This is a simple consequence of Theorem 2.3 and the fact that A is countable.

    For the rest of the section Ω0 is fixed according to Lemma 3.3.

    The idea of quenched stochastic two-scale convergence is similar to periodic two-scale convergence: We associate with a bounded sequence (uε)Lp(Q) and ω0Ω0, a sequence of linear functionals (uε) defined on D. We can pass (up to a subsequence) to a pointwise limit U, which is again a linear functional on D and which (thanks to Lemma 3.3) can be uniquely extended to a bounded linear functional on Bq. We then define the weak quenched ω0-two-scale limit of (uε) as the Riesz-representation uBp of U(Bq).

    Definition 3.4 (quenched two-scale limit, cf. [38,17]). Let (uε) be a sequence in Lp(Q), and let ω0Ω0 be fixed. We say that uε converges (weakly, quenched) ω0-two-scale to uBp, and write uε2ω0u, if the sequence uε is bounded in Lp(Q), and for all φD we have

    limε0Quε(x)(Tεφ)(ω0,x)dx=ΩQu(x,ω)φ(ω,x)dxdP(ω). (8)

    Lemma 3.5 (Compactness). Let (uε) be a bounded sequence in Lp(Q) and ω0Ω0. Then there exists a subsequence (still denoted by ε) and uBp such that uε2ω0u and

    uBplim infε0uεLp(Q), (9)

    and uεu weakly in Lp(Q).

    (For the proof see Section 3.2.1).

    For our purpose it is convenient to have a metric characterization of two-scale convergence.

    Lemma 3.6 (Metric characterization). (i)Let {φj}jN denote an enumeration of the countable set {φφBq:φD0}. The vector space Lin(D):={U:DRlinear} endowed with the metric

    d(U,V;Lin(D)):=jN2j|U(φj)V(φj)||U(φj)V(φj)|+1

    is complete and separable.

    (ii)Let ω0Ω0. Consider the maps

    Jω0ε:Lp(Q)Lin(D),(Jω0εu)(φ):=Qu(x)(Tεφ)(ω0,x)dx,J0:BpLin(D),(J0u)(φ):=Quφ.

    Then for any bounded sequence uε in Lp(Q) and any uBp we have uε2ω0u if and only if Jω0εuεJ0u in Lin(D).

    (For the proof see Section 3.2.1).

    Remark 2. Convergence in the metric space (Lin(D),d(,,Lin(D)) is equivalent to pointwise convergence. (Bq) is naturally embedded into the metric space by means of the restriction J:(Bq)Lin(D), JU=U|D. In particular, we deduce that for a bounded sequences (Uk) in (Bq) we have UkU if and only if JUkJU in the metric space. Likewise, Bp (resp. Lp(Q)) can be embedded into the metric space Lin(D) via J0 (resp. Jω0ε with ε>0 and ω0Ω0 arbitrary but fixed), and for a bounded sequence (uk) in Bp (resp. Lp(Q)) weak convergence in Bp (resp. Lp(Q)) is equivalent to convergence of (J0uk) (resp. (Jω0εuk)) in the metric space.

    Lemma 3.7 (Strong convergence implies quenched two-scale convergence). Let (uε) be a strongly convergent sequence in Lp(Q) with limit uLp(Q). Then for all ω0Ω0 we have uε2ω0u.

    (For the proof see Section 3.2.1).

    Definition 3.8 (set of quenched two-scale cluster points). For a bounded sequence (uε) in Lp(Q) and ω0Ω0 we denote by CP(ω0,(uε)) the set of all ω0-two-scale cluster points, i.e. the set of uBp with J0uk=1¯{Jω0εuε:ε<1k} where the closure is taken in the metric space (Lin(D),d(,;Lin(D))).

    We conclude this section with two elementary results on quenched stochastic two-scale convergence:

    Lemma 3.9 (Approximation of two-scale limits). Let uBp.Then for all ω0Ω0, there exists a sequenceuεLp(Q) such that uε2ω0u as ε0.

    (For the proof see Section 3.2.1).

    Similar to the slightly different setting in [17] one can prove the following result:

    Lemma 3.10 (Two-scale limits of gradients). Let (uε) be a sequence in W1,p(Q) and ω0Ω0. Then there exist a subsequence (not relabeled) and functions uW1,p(Q) and χLppot(Ω)Lp(Q) such that uεu weakly in W1,p(Q) and

    uε2ω0uanduε2ω0u+χasε0.

    Proof of Lemma 3.5. Set C0:=lim supε0uεLp(Q) and note that C0<. By passing to a subsequence (not relabeled) we may assume that C0=lim infε0uεLp(Q). Fix ω0Ω0. Define linear functionals UεLin(D) via

    Uε(φ):=Quε(x)(Tεφ)(ω0,x)dx.

    Note that for all φA, (uε(φ)) is a bounded sequence in R. Indeed, by Hölder's inequality and Lemma 3.3,

    lim supε0|uε(φ)|lim supε0uεLp(Q)Tεφ(ω0,)Lq(Q)C0φBq. (10)

    Since A is countable we can pass to a subsequence (not relabeled) such that uε(φ) converges for all φA. By linearity and since D=span(A), we conclude that uε(φ) converges for all φD, and U(φ):=limε0uε(φ) defines a linear functional on D. In view of (10) we have |U(φ)|C0φBq, and thus U admits a unique extension to a linear functional in (Bq). Let uBp denote its Riesz-representation. Then uε2ω0u, and

    uBp=U(Bq)C0=lim infε0uεLp(Q).

    Since 1ΩDΩ we conclude that for all φQDQ we have

    Quε(x)φQ(x)dx=uε(1ΩφQ)U(1ΩφQ)=Qu(ω,x)φQ(x)dx=Qu(x)φQ(x)dx.

    Since (uε) is bounded in Lp(Q), and DQLp(Q) is dense, we conclude that uεu weakly in Lp(Q).

    Proof of Lemma 3.6. We use the following notation in this proof A1:={φφBq:φD0}.

    (i) Argument for completeness: If (Uj) is a Cauchy sequence in Lin(D), then for all φA1, (Uj(φ)) is a Cauchy sequence in R. By linearity of the Uj's this implies that (Uj(φ)) is Cauchy in R for all φD. Hence, UjU pointwise in D and it is easy to check that U is linear. Furthermore, UjU pointwise in A1 implies UjU in the metric space.

    Argument for separability: Consider the (injective) map J:(Bq)Lin(D) where J(U) denotes the restriction of U to D. The map J is continuous, since for all U,V(Bq) and φA1 we have |(JU)(φ)(JV)(φ)|UV(Bq)φBq=UV(Bq) (recall that the test functions in A1 are normalized). Since (Bq) is separable (as a consequence of the assumption that F is countably generated), it suffices to show that the range R(J) of J is dense in Lin(D). To that end let ULin(D). For kN we denote by Uk(Bq) the unique linear functional that is equal to U on the the finite dimensional (and thus closed) subspace span{φ1,,φk}Bq (where {φj} denotes the enumeration of A1), and zero on the orthogonal complement in Bq. Then a direct calculation shows that d(U,J(Uk);Lin(D))j>k2j=2k. Since kN is arbitrary, we conclude that R(J)Lin(D) is dense.

    (ii) Let uε denote a bounded sequence in Lp(Q) and uBp. Then by definition, uε2ω0u is equivalent to Jω0εuεJ0u pointwise in D, and the latter is equivalent to convergence in the metric space Lin(D).

    Proof of Lemma 3.7. This follows from Hölder's inequality and Lemma 3.3, which imply for all φA the estimate

    limsupε0Q|(uε(x)u(x))Tεφ(ω0,x)|dxlimsupε0(uεuLp(Q)(Q|Tεφ(ω0,x)|qdx)1q)=0.

    Proof of Lemma 3.9. Since D (defined as in Lemma 3.6) is dense in Bp, for any δ>0 there exists vδD0 with uvδBpδ. Define vδ,ε(x):=Tεvδ(ω0,x). Let φD. Since vδ and φ (resp. vδφ) are by definition linear combinations of functions (resp. products of functions) in A, we deduce from Lemma 3.3 that (vδ,ε)ε is bounded in Lp(Q), and that

    Qvδ,εTεφ(ω0,x)=QTε(vδφ)(ω0,x)Qvδφ.

    By appealing to the metric characterization, we can rephrase the last convergence statement as d(Jω0εvδ,ε,J0vδ;Lin(D))0. By the triangle inequality we have

    d(Jω0εvδ,ε,J0u;Lin(D))d(Jω0εvδ,ε,J0vδ;Lin(D))+d(J0vδ,J0u;Lin(D)).

    The second term is bounded by vδuBpδ, while the first term vanishes for ε0. Hence, there exists a diagonal sequence uε:=vδ(ε),ε (bounded in Lp(Q)) such that there holds d(Jω0εuε,J0u;Lin(D))0. The latter implies uε2ω0u by Lemma 3.6.

    In this section we establish a relation between quenched two-scale convergence and two-scale convergence in the mean (in the sense of Definition 3.2). The relation is established by Young measures: We show that any bounded sequence uε in Bp – viewed as a functional acting on test-functions of the form Tεφ – generates (up to a subsequence) a Young measure on Bp that (a) concentrates on the quenched two-scale cluster points of uε, and (b) allows to represent the two-scale limit (in the mean) of uε. In entire Section 3.3 we assume that

     is ergodic.

    Also, throughout this section we fix exponents p(1,), q:=pp1, and an open and bounded domain QRd. Furthermore, we frequently use the objects and notations introduced in Section 3.2.

    Definition 3.11. We say ν:={νω}ωΩ is a Young measure on Bp, if for all ωΩ, νω is a Borel probability measure on Bp (equipped with the weak topology) and

    ωνω(B)is measurable for all BB(Bp),

    where B(Bp) denotes the Borel-σ-algebra on Bp (equipped with the weak topology).

    Theorem 3.12. Let uε denote a bounded sequence in Bp.Then there exists a subsequence (still denoted by ε) and a Young measure ν on Bpsuch that for all ω0Ω0,

    νω0is concentrated onCP(ω0,(uε(ω0,))),

    and

    lim infε0uεpBpΩ(BpvpBpdνω(v))dP(ω).

    Moreover, we have

    uε2uwhereu:=ΩBpvdνω(v)dP(ω).

    Finally, if there exists ˆu:ΩBp measurable and νω=δˆu(ω) for P-a.a. ωΩ, then up to extraction of a further subsequence (still denoted by ε) we have

    uε(ω)2ωˆu(ω)forPa.a.ωΩ.

    (For the proof see Section 3.3.1).

    In the opposite direction we observe that quenched two-scale convergence implies two-scale convergence in the mean in the following sense:

    Lemma 3.13. Consider a family {(uωε)}ωΩ of sequences (uωε) in Lp(Q) and suppose that:

    (i)There exists uBp s.t. for P-a.a. ωΩ we have uωε2ωu.

    (ii)There exists a sequence (˜uε) in Bp s.t. uωε(x)=˜uε(ω,x) for a.a. (ω,x)Ω×Q.

    (iii)There exists a bounded sequence (χε) in Lp(Ω) such that uωεLp(Q)χε(ω) for a.a. ωΩ.

    Then ˜uε2u weakly two-scale (in the mean).

    (For the proof see Section 3.3.1).

    To compare homogenization of convex integral functionals w.r.t. stochastic two-scale convergence in the mean and in the quenched sense, we appeal to the following result:

    Lemma 3.14. Let h:Ω×Q×RdR be such that for all ξRd, h(,,ξ) is FB(Rd)-measurable and for a.a. (ω,x)Ω×Q, h(ω,x,) is convex. Let (uε) denote a bounded sequence in Bp that generates a Young measure ν on Bp in the sense of Theorem 3.12.Suppose that hε:ΩR, hε(ω):=Qmin{0,h(τxεω,x,uε(ω,x))}dx is uniformly integrable. Then

    liminfε0ΩQh(τxεω,x,uε(ω,x))dxdP(ω)ΩBp(ΩQh(˜ω,x,v(˜ω,x))dxdP(˜ω))dνω(v)dP(ω). (11)

    (For the proof see Section 3.3.1).

    Remark 3. In [18,Lemma 5.1] it is shown that h satisfying the assumptions of Lemma 3.14 satisfies the following property: For P-a.a. ω0Ω0 we have: For any sequence (uε) in Lp(Q) it holds

    uε2ω0ulim infε0Qh(τxεω0,x,uε(x))dxΩQh(ω,x,u(ω,x))dxdP(ω). (12)

    We first recall some notions and results of Balder's theory for Young measures [4]. Throughout this section M is assumed to be a separable, complete metric space with metric d(,;M).

    Definition 3.15. ● We say a function s:ΩM is measurable, if it is FB(M)-measurable where B(M) denotes the Borel-σ-algebra in M.

    ● A function h:Ω×M(,+] is called a normal integrand, if h is FB(M)-measurable, and for all ωΩ the function h(ω,):M(,+] is lower semicontinuous.

    ● A sequence sε of measurable functions sε:ΩM is called tight, if there exists a normal integrand h such that for every ωΩ the function h(ω,) has compact sublevels in M and lim supε0Ωh(ω,sε(ω))dP(ω)<.

    ● A Young measure in M is a family μ:={μω}ωΩ of Borel probability measures on M such that for all BB(M) the map Ωωμω(B)R is F-measurable.

    Theorem 3.16.([4,Theorem I]). Let sε:ΩM denote a tight sequence of measurable functions. Then there exists a subsequence, still indexed by ε, and a Young measure μ:ΩM such that for every normal integrand h:Ω×M(,+] we have

    lim infε0Ωh(ω,sε(ω))dP(ω)ΩMh(ω,ξ)dμω(ξ)dP(ω), (13)

    provided that the negative part hε()=|min{0,h(,sε())}| is uniformly integrable.Moreover, for P-a.a. ωΩ0 the measure μω is supported in the set of all cluster points of sε(ω), i.e. in k=1¯{sε(ω):ε<1k} (where the closure is taken in M).

    In order to apply the above theorem we require an appropriate metric space in which two-scale convergent sequences and their limits embed:

    Lemma 3.17. (i)>We denote by M the set of all triples (U,ε,r) with ULin(D), ε0, r0. M endowed with the metric

    d((U1,ε1,r1),(U2,ε2,r2);M):=d(U1,U2;Lin(D))+|ε1ε2|+|r1r2|

    is a complete, separable metric space.

    (ii)For ω0Ω0 we denote by Mω0 the set of all triples (U,ε,r)M such that

    U={Jω0εufor someuLp(Q)withuLp(Q)rin the caseε>0,J0ufor someuBpwithuBprin the caseε=0. (14)

    Then Mω0 is a closed subspace of M.

    (iii)Let ω0Ω0, and (U,ε,r)Mω0. Then the function u in the representation (14) of U is unique, and

    {uLp(Q)=supφ¯D, φBq1|U(φ)|ifε>0,uBp=supφD, φBq1|U(φ)|ifε=0. (15)

    (iv)For ω0Ω0 the function ω0:Mω0[0,),

    (U,ε,r)ω0:={(supφ¯D, φBq1|U(φ)|p+ε+rp)1pif(U,ε,r)Mω0,ε>0,(supφD, φBq1|U(φ)|p+rp)1pif(U,ε,r)Mω0,ε=0,

    is lower semicontinuous on Mω0.

    (v)For all (u,ε) with uLp(Q) and ε>0 we have s:=(Jω0εu,ε,uLp(Q))Mω0 and sω0=(2upLp(Q)+ε)1p. Likewise, for all (u,ε) with uBp and ε=0 we have s=(J0u,ε,uBp) and sω0=21puBp.

    (vi)For all R< the set {(U,ε,r)Mω0:(U,ε,r)ω0R} is compact in M.

    (vii)Let ω0Ω0 and let uε denote a bounded sequence in Lp(Q). Then the triple sε:=(Jω0εuε,ε,uεLp(Q)) defines a sequence in Mω0. Moreover, we have sεs0 in M as ε0 if and only if s0=(J0u,0,r) for some uBp, ruBp, and uε2ω0u.

    Proof.(i)This is a direct consequence of Lemma 3.6 (i) and the fact that the product of complete, separable metric spaces remains complete and separable.

    (ii)Let sk:=(Uk,εk,rk) denote a sequence in Mω0 that converges in M to some s0=(U0,ε0,r0). We need to show that s0Mω0. By passing to a subsequence, it suffices to study the following three cases: εk>0 for all kN0, εk=0 for all kN0, and ε0=0 while εk>0 for all kN.

    Case 1: εk>0 for all kN0.

    W.l.o.g. we may assume that infkεk>0. Hence, there exist ukLp(Q) with Uk=Jω0εkuk. Since rkr, we conclude that (uk) is bounded in Lp(Ω). We thus may pass to a subsequence (not relabeled) such that uku0 weakly in Lp(Q), and

    u0Lp(Q)lim infkukLp(Q)limkrk=r0. (16)

    Moreover, UkU in the metric space Lin(D) implies pointwise convergence on D, and thus for all φQDQ we have Uk(1ΩφQ)=QukφQQu0φQ. We thus conclude that U0(1ΩφQ)=Qu0φQ. Since DQLq(Q) dense, we deduce that uku0 weakly in Lp(Q) for the entire sequence. On the other hand the properties of the shift τ imply that for any φΩDΩ we have φΩ(τεkω0)φΩ(τε0ω0) in Lq(Q). Hence, for any φΩDΩ and φQDQ we have

    Uk(φΩφQ)=Quk(x)φQ(x)φΩ(τxεkω0)dxQu0(x)φQ(x)φΩ(τxε0ω0)dx=Jω0ε0(φΩφQ)

    and thus (by linearity) U0=Jω0ε0u0.

    Case 2: εk=0 for all kN0.

    In this case there exist a bounded sequence uk in Bp with Uk=J0uk for kN. By passing to a subsequence we may assume that uku0 weakly in Bp for some u0Bp with

    u0Bplim infkuεkBpr0. (17)

    This implies that Uk=J0ukJ0u0 in Lin(D). Hence, U0=J0u0 and we conclude that s0Mω0.

    Case 3: εk>0 for all kN and ε0=0.

    There exists a bounded sequence uk in Lp(Q). Thanks to Lemma 3.5, by passing to a subsequence we may assume that uk2ω0u0 for some uBp with

    u0Bplim infkukLp(Q)r0. (18)

    Furthermore, Lemma 3.6 implies that Jω0εkukJ0u0 in Lin(D), and thus U0=J0u0. We conclude that s0Mω0.

    (iii)We first argue that the representation (14) of U by u is unique. In the case ε>0 suppose that u,vLp(Q) satisfy U=Jω0εu=Jω0εv. Then for all φQDQ we have Q(uv)φQ=Jω0εu(1ΩφQ)Jω0εv(1ΩφQ)=U(1ΩφQ)U(1ΩφQ)=0, and since DQLq(Q) dense, we conclude that u=v. In the case ε=0 the statement follows by a similar argument from the fact that D is dense Bq.

    To see (15) let u and U be related by (14). Since ¯D (resp. D) is dense in Lq(Q) (resp. Bq), we have

    {uLp(Q)=supφ¯D, φBq1|QuφdxdP|=supφ¯D, φBq1|U(φ)|if ε>0,uBp=supφD, φBq1|ΩQuφdxdP|=supφD, φBq1|U(φ)|if ε=0.

    (iv)Let sk=(Uk,εk,rk) denote a sequence in Mω0 that converges in M to a limit s0=(U0,ε0,r0). By (ii) we have s0Mω0. For kN0 let uk in Lp(Q) or Bp denote the representation of Uk in the sense of (14). We may pass to a subsequence such that one of the three cases in (ii) applies and (as in (ii)) either uk weakly converges to u0 (in Lp(Q) or Bp), or uk2ω0u0. In any of these cases the claimed lower semicontinuity of ω0 follows from εkε0, rkr0, and (15) in connection with one of the lower semicontinuity estimates (16) – (18).

    (v)This follows from the definition and duality argument (15).

    (vi)Let sk denote a sequence in Mω0. Let uk in Lp(Q) or Bp denote the (unique) representative of Uk in the sense of (14). Suppose that skω0R. Then (rk) and (εk) are bounded sequences in R0, and supkuksupkrk< (where stands short for either Lp(Q) or Bp). Thus we may pass to a subsequence such that rkr0, εkε0, and one of the following three cases applies:

    ● Case 1: infkN0εk>0. In that case we conclude (after passing to a further subsequence) that uku0 weakly in Lp(Q), and thus UkU0=Jω0ε0u0 in Lin(D).

    ● Case 2: εk=0 for all kN0. In that case we conclude (after passing to a further subsequence) that uku0 weakly in Bp(Q), and thus UkU0=J0u0 in Lin(D).

    ● Case 3: εk>0 for all kN and ε0=0. In that case we conclude (after passing to a further subsequence) that uk2ω0u0, and thus UkU0=J0u0 in Lin(D).

    In all of these cases we deduce that s0=(U0,ε0,r0)Mω0, and sks0 in M.

    (vii)This is a direct consequence of (ii) – (vi), and Lemma 3.6.

    Now we are in position to prove Theorem 3.12

    Proof of Theorem 3.12. Let M, Mω0, Jω0ε etc. be defined as in Lemma 3.17.

    Step 1. (Identification of (uε) with a tight M-valued sequence). Since uεBp, by Fubini's theorem, we have uε(ω,)Lp(Q) for P-a.a. ωΩ. By modifying uε on a null-set in Ω×Q (which does not alter two-scale limits in the mean), we may assume w.l.o.g. that uε(ω,) Lp(Q) for all ωΩ. Consider the measurable function sε:ΩM defined as

    sε(ω):={(Jωεuε(ω,),ε,uε(ω,)Lp(Q))if ωΩ0(0,0,0)else.

    We claim that (sε) is tight. To that end consider the integrand h:Ω×M(,+] defined by

    h(ω,(U,ε,r)):={(U,ε,r)pωif ωΩ0 and (U,ε,r)Mω,+else.

    From Lemma 3.17 (iv) and (vi) we deduce that h is a normal integrand and h(ω,) has compact sublevels for all ωΩ. Moreover, for all ω0Ω0 we have sε(ω0)Mω0 and thus h(ω0,sε(ω0))=2uε(ω0,)pLp(Q)+ε. Hence,

    Ωh(ω,sε(ω))dP(ω)=2uεpBp+ε.

    We conclude that (sε) is tight.

    Step 2. (Compactness and definition of ν). By appealing to Theorem 3.16 there exists a subsequence (still denoted by ε) and a Young measure μ that is generated by (sε). Let μ1 denote the first component of μ, i.e. the Young measure on Lin(D) characterized for ωΩ by

    Lin(D)f(ξ)dμ1,ω(ξ)=Mf(ξ1)dμω(ξ),

    for all f:Lin(D)R continuous and bounded, where Mξ=(ξ1,ξ2,ξ3)ξ1Lin(D) denotes the projection to the first component. By Balder's theorem, μω is concentrated on the limit points of (sε(ω)). By Lemma 3.17 we deduce that for all ωΩ0 any limit point s0(ω) of sε(ω) has the form s0(ω)=(J0u,0,r) where 0r< and uBp is a ω-two-scale limit of a subsequence of uε(ω,). Thus, μ1,ω is supported on {J0u:uCP(ω,(uε(ω,))} which in particular is a subset of (Bq). Since J0:Bp(Bq) is an isometric isomorphism (by the Riesz-Frechét theorem), we conclude that ν={νω}ωΩ, νω(B):=μ1,ω(J0B) (for all Borel sets BBp where Bp is equipped with the weak topology) defines a Young measure on Bp, and for all ωΩ0, νω is supported on CP(ω,(uε(ω,)).

    Step 3. (Lower semicontinuity estimate). Note that h:Ω×M[0,+],

    h(ω,(U,ε,r)):={supφ¯D,φBq1|U(φ)|pif ωΩ0 and (U,ε,r)Mω,ε>0,supφD,φBq1|U(φ)|pif ωΩ0 and (U,ε,r)Mω,ε=0,+else.

    defines a normal integrand, as can be seen as in the proof of Lemma 3.17. Thus Theorem 3.16 implies that

    lim infε0Ωh(ω,sε(ω))dP(ω)ΩMh(ω,ξ)dμω(ξ)dP(ω).

    In view of Lemma 3.17 we have supφ¯D,φBq1|(Jωεuε)(ω,))(φ)|=uε(ω,)Lp(Q) for ωΩ0, and thus the left-hand side turns into lim infε0uεpBp. Thanks to the definition of ν the right-hand side turns into ΩBpvpBpdνω(v)dP(ω).

    Step 4. (Identification of the two-scale limit in the mean). Let φD0. Then h:Ω×M[0,+],

    h(ω,(U,ε,r)):={U(φ)if ωΩ0,(U,ε,r)Mω,+else.

    defines a normal integrand. Since h(ω,sε(ω))=Quε(ω,x)Tεφ(ω,x)dx for P-a.a. ωΩ, we deduce that |h(,sε())| is uniformly integrable. Thus, (13) applied to ±h and the definition of ν imply that

    limε0ΩQuε(ω,x)(Tεφ)(ω,x)dxdP(ω)=limε0Ωh(ω,sε(ω))dP(ω)=ΩBph(ω,v)dνω(v)dP(ω)=ΩBpQvφdνω(v)dP(ω).

    Set u:=ΩBpvdνω(v)dP(ω)Bp. Then Fubini's theorem yields

    limε0ΩQuε(ω,x)(Tεφ)(ω,x)dxdP(ω)=Quφ.

    Since span(D0)Bq dense, we conclude that uε2u.

    Step 5. Recovery of quenched two-scale convergence. Suppose that νω is a delta distribution on Bp, say νω=δ˜u(ω) for some measurable ˜u:ΩBp. Note that h:Ω×M[0,+],

    h(ω,(U,ε,r)):=d(U,J0˜u(ω);Lin(D))

    is a normal integrand and |h(,sε())| is uniformly integrable. Thus, (13) yields

    lim supε0Ωd(Jωεuε(ω,),J0˜u(ω);Lin(D))dP(ω)=lim infε0Ωh(ω,sε(ω))dP(ω)ΩBph(ω,J0v)dνω(v)dP(ω)=Ωh(ω,J0˜u(ω))dP(ω)=0.

    Thus, there exists a subsequence (not relabeled) such that d(Jωεuε(ω,),J0˜u(ω);Lin(D))0 for a.a. ωΩ0. In view of Lemma 3.6 this implies that uε2ω˜u(ω) for a.a. ωΩ0.

    Proof of Lemma 3.14. Step 1. Representation of the functional by a lower semicontinuous integrand on M.

    For all ω0Ω0 and s=(U,ε,r)Mω0 we write πω0(s) for the unique representation u in Bp (resp. Lp(Q)) of U in the sense of (14). We thus may define for ωΩ0 and sMω0 the integrand

    ¯h(ω0,s):={Qh(τxεω,x,(πω0s)(x))dxif s=(U,ε,s) with ε>0,ΩQh(ω,x,(πω0s)(x))dxdP(ω)if s=(U,ε,s) with ε=0.

    We extend ¯h(ω0,) to M by +, and define ¯h(ω,)0 for ωΩΩ0. We claim that ¯h(ω,):M(,+] is lower semicontinuous for all ωΩ. It suffices to consider ω0Ω0 and a convergent sequence sk=(Uk,εk,rk) in Mω0. For brevity we only consider the (interesting) case when εkε0=0. Set uk:=πω0(sk). By construction we have

    ¯h(ω0,sk)=Qh(τxεkω0,uk(ω0,x))dx,

    and

    ¯h(ω0,s0)=ΩQh(ω,x,u0(ω,x))dxdP(ω).

    Since sks0 and εk0, Lemma 3.17 (vi) implies that uk2ω0u0, and since h satisfies 12 from Remark 3, we conclude that lim infk¯h(ω0,sk)¯h(ω0,s0), and thus ¯h is a normal integrand.

    Step 2. Conclusion. As in Step 1 of the proof of Theorem 3.12 we may associate with the sequence (uε) a sequence of measurable functions sε:ΩM that (after passing to a subsequence that we do not relabel) generates a Young measure μ on M. Since by assumption uε generates the Young measure ν on Bp, we deduce that the first component μ1 satisfies νω(B)=μω(J0B) for any Borel set B. Applying (13) to the integrand ¯h of Step 1, yields

    lim infε0ΩQh(τxεω0,uε(ω0,x))dxdP(ω)=lim infε0Ω¯h(ω,sε(ω))dP(ω)ΩM¯h(ω,ξ)dμω(ξ)dP(ω)=ΩBp(ΩQh(˜ω,x,v(˜ω,x))dxdP(˜ω))dνω(v)dP(ω).

    Proof of Lemma 3.13. By (ii) and (iii) the sequence (˜uε) is bounded in Bp and thus we can pass to a subsequence such that (˜uε) generates a Young measure ν. Set ˜u:=ΩBpvdνω(v)dP(ω) and note that Theorem 3.12 implies that ˜uε2˜u weakly two-scale in the mean. On the other hand the theorem implies that νω concentrates on the quenched two-scale cluster points of (uωε) (for a.a. ωΩ). Hence, in view of (i) we conclude that for a.a. ωΩ the measure νω is a Dirac measure concentrated on u, and thus ˜u=u a.e. in Ω×Q.

    In this section we revisit a standard model example of stochastic homogenization of integral functionals from the viewpoint of stochastic two-scale convergence and unfolding. In particular, we discuss two examples of convex homogenization problems that can be treated with stochastic two-scale convergence in the mean, but not with the quenched variant. In the first example in Section 4.1 the randomness is nonergodic and thus quenched two-scale convergence does not apply. In the second example, in Section 4.2, we consider a variance-regularization to treat a convex minimization problem with degenerate growth conditions. In these two examples we also demonstrate the simplicity of using the stochastic unfolding operator. Furthermore, in Section 4.3 we use the results of Section 3.3 to further reveal the structure of the previously obtained limits in the classical ergodic case with non-degenerate growth with help of Young measures. In particular, we show how to lift mean homogenization results to quenched statements.

    In this section we consider a nonergodic stationary medium. Such random ensembles arise naturally, e.g., in the context of periodic representative volume element (RVE) approximations, see [13]. For example, we may consider a family of i.i.d. random variables {¯ω(z)}zZd. A realization of a stationary and ergodic random checkerboard is given by

    ω:RdR,ω(x)=iZd1i+y+(x)¯ω(x),

    where xZd is the integer part of x and y is the center of the checkerboard chosen uniformly from =[0,1)d. For LN, we may consider the map πL:ωωL given by πLω(x)=ω(x) for x[0,L)d and πLω is L-periodically extended. The push forward of the map πL defines a stationary and nonergodic probability measure, that is a starting point in the periodic RVE method. Another standard example of a nonergodic structure may be obtained by considering a medium with a noncompatible quasiperiodic microstructure, see [38,Example 1.2].

    In this section we consider the following situation. Let p(1,) and QRd be open and bounded. We consider V:Ω×Q×RdR and assume:

    (A1)V(,,F) is FL(Q)-measurable for all FRd.

    (A2)V(ω,x,) is convex for a.a. (ω,x)Ω×Q.

    (A3)There exists C>0 such that

    1C|F|pCV(ω,x,F)C(|F|p+1)

    for a.a. (ω,x)Ω×Q and all FRd.

    We consider the functional

    Eε:Lp(Ω)W1,p0(Q)R,Eε(u)=QV(τxεω,x,u(ω,x))dx. (19)

    Under assumptions (A1)-(A3), in the limit ε0 we obtain the two-scale functional

    E0:(Lpinv(Ω)W1,p0(Q))×(Lppot(Ω)Lp(Q))R,E0(u,χ)=QV(ω,x,u(ω,x)+χ(ω,x))dx. (20)

    Theorem 4.1 (Two-scale homogenization). Let p(1,) and QRd be open and bounded. Assume (A1)-(A3).

    (i)(Compactness and liminf inequality.) Let uεLp(Ω)W1,p0(Q) be such thatlim supε0Eε(uε)<.There exist (u,χ)(Lpinv(Ω)W1,p0(Q))×(Lppot(Ω)Lp(Q)) and a subsequence (not relabeled) such that

    uε2uinLp(Ω×Q),uε2u+χinLp(Ω×Q), (21)
    lim infε0Eε(uε)E0(u,χ). (22)

    (ii)(Limsup inequality.) Let (u,χ)(Lpinv(Ω)W1,p0(Q))×(Lppot(Ω)Lp(Q)). There exists a sequence uεLp(Ω)W1,p0(Q) such that

    uε2uinLp(Ω×Q),uε2u+χinLp(Ω×Q), (23)
    lim supε0Eε(uε)E0(u,χ). (24)

    Proof of Theorem 4.1. (i) The Poincaré inequality and (A3) imply that uε is bounded in Lp(Ω)W1,p(Q). By Proposition 1 (ii) there exist uLpinv(Ω)W1,p(Q) and χLppot(Ω)Lp(Q) such that (21) holds. Also, note that Pinvuεu weakly in Lp(Ω)W1,p(Q) and PinvuεLpinv(Ω)W1,p0(Q), which implies that u also has 0 boundary values, i.e., uLpinv(Ω)W1,p0(Q). Finally, we note that, see [19,Proposition 3.5 (i)],

    QV(τxεω,x,v(ω,x))=QV(ω,x,Tεv(ω,x))for any vLp(Ω×Q), (25)

    and thus using the convexity of V we conclude

    lim infε0Eε(uε)=lim infε0QV(ω,x,Tεuε)E0(u,χ).

    (ii) The existence of a sequence uε with (23) follows from Proposition 1 (iii). Furthermore, (25) and the growth assumption (A3) yield

    limε0Eε(uε)=limε0QV(ω,x,Tεuε)=E0(u,χ).

    This concludes the claim, in particular, we even show a stronger result stating convergence of the energy.

    Remark 4 (Convergence of minimizers). We consider the setting of Theorem 4.1. Let uεLp(Ω)W1,p0(Q) be a minimizer of the functional

    Iε:Lp(Ω)W1,p0(Q)R,Iε(u)=Eε(u)Quεfεdx,

    where fεLq(Ω×Q) and fε2f with fLq(Q) and 1p+1q=1. By a standard argument from the theory of Γ-convergence Theorem 4.1 (cf. [34,Corollary 7.2]) implies that there exist a subsequence (not relabeled), uLpinv(Ω)×W1,p0(Q), and χLppot(Ω)Lp(Q) such that uε2u in Lp(Ω×Q), uε2u+χ in Lp(Ω×Q), and

    limε0minIε=limε0Iε(uε)=I0(u,χ)=minI0,

    where I0:Lpinv(Ω)W1,p0(Q)R is given by I0(u)=E0(u)Qfudx. This, in particular, rigorously justifies the formal two-scale expansion uε(x)u0(ω,x)+χ(τxεω,x).

    Remark 5 (Uniqueness). If V(ω,x,) is strictly convex the minimizers are unique and the convergence in the above remark holds for the entire sequence.

    In this section we consider homogenization of convex functionals with degenerate growth. More precisely, we consider an integrand that satisfies (A1), (A2) and the following assumption (as a replacement of (A3)):

    (A3')There exists and a random variable such that

    (26)

    and

    for a.a. and all .

    Moreover, we assume that is ergodic. For we consider the following functional

    for and otherwise. Here denotes the closure of w.r.t. the weighted norm

    Recently, in [29,20,21] it shown that Mosco-converges to the functional

    for and otherwise, where is given by the homogenization formula,

    (27)

    for and . Moreover, it is shown that is a normal convex integrand that satisfies a standard -growth condition. Note that the assumption (A3') in comparison to (A3) makes a genuine difference in regard to the homogenization formula (27). In particular, in the setting of assumption (A3) minimizers are attained due to the coercivity of the underlying functional in . It is thus easy to see that the homogenized integrand satisfies -growth condition as well, see Section 4.3 below. On the other hand, in the setting of this section assuming (A3'), (27) is a degenerate minimization problem and a priori minimizers will only have finite first moments. An additional argument is required to infer that in (27) is non-degenerate, in particular, in [29,Theorem 3.1] it is shown that there exists a constant such that for all and we have

    (28)

    One of the difficulties in the proof of the homogenization result for is due to the fact that the domain of the functionals are -dependent. Moreover, assumption (A3') only yields equicoercivity in , while the limit is properly defined on . Therefore, in practice it is convenient to regularize the problem: For we consider the regularized homogenization formula

    It is simple to show that the infimum on the right-hand side is attained by a unique minimizer. We also consider the corresponding regularized homogenized integral functional

    for and otherwise. Furthermore, thanks to (A3'), it is relatively easy to see that this regularization is consistent:

    Lemma 4.2. Let and be open and bounded. Assume (A1), (A2) and (A3'). Then, for all and , we have

    (29)

    Moreover, Mosco converges to as , i.e., the following statements hold:

    (i)If weakly in , then

    (ii)For any there exists a sequence such that

    Proof. Let and . Since , we have . On the other hand, we consider a minimizing sequence in (27), e.g.,

    We have

    Letting first and then , we conclude (29).

    We further consider a sequence such that weakly in as . We assume without loss of generality that. This, in particular, with the help of (28) and the Poincaré inequality implies that . Thus, up to a subsequence, we have weakly in . Using this, we obtain

    The first inequality follows by (29) and the second is a consequence of the fact that is convex and of Fatou's Lemma. We conclude that (i) holds.

    If , we simply choose . On the other hand, for , (29) and the dominated convergence theorem yield

    This means that (ii) holds.

    In the following we introduce a variance regularization of the original functional that removes the degeneracy of the problem and thus can be analyzed by the standard strategy of Section 4.1. For , we consider ,

    (30)

    for and otherwise. Due to the structure of the additional term, we call it a variance-regularization and we note that it only becomes active for non-deterministic functions. For fixed , the functional is equicoercive on :

    Lemma 4.3. Let and be open and bounded. Assume (A1) and (A3'). Then there exists such that, for all , it holds

    Proof. By Jensen's and Hölder's inequalities we have

    where we use the notation . Furthermore, using (A3'), we conclude that

    In the end, using the variance-regularization we obtain

    This concludes the proof.

    The regularization on the -level is also consistent. In particular, we show that in the limit , we recover . We discuss the mean functionals and , since the former does not admit a well-defined pointwise evaluation in for the reason of the nonlocal variance term. Also, for the same reason the quenched version of stochastic two-scale convergence is not suitable for this setting and we apply the unfolding procedure. On the other hand, the homogenization of can be conducted on the level of typical realizations, that was in fact studied in [29,20,21].

    Lemma 4.4. Let and be open and bounded. Assume (A1), (A2) and (A3'). Then, Mosco converges to as i.e., the following statements hold:

    (i)If weakly in , then

    (ii)For any there exists a sequence such that

    Proof. (i) Let be a sequence such that weakly in . Without loss of generality we assume that . This and the proof of Lemma 4.3 imply that the sequence is bounded in with the notation . This means that, up to a subsequence, we have weakly in for some . Thus, for an arbitrary , we have

    This means that converges weakly in and since weakly in we may conclude that weakly in . This yields

    (ii) For an arbitrary , we find a sequence such that, for ,

    Using this and the dominated convergence theorem, we conclude that

    This in turn yields

    We extract a diagonal sequence as such that satisfies strongly in and . This concludes the proof.

    The homogenization of the regularized functional boils down to a very similar simple argumentation as in Section 4.1.

    Theorem 4.5. Let and be open and bounded. Assume (A1), (A2) and (A3'). For all , as , Mosco converges to in the following sense:

    (i)Let be such that. Then there exist and a subsequence (not relabeled) such that

    (ii)If , and weakly in , then

    (iii)For any , there exists a sequence such that

    Proof. (i) The statement follows analogously to the proof of Theorem 4.1 (i).

    (ii) Let weakly in . We may assume without loss of generality that . In this case, Lemma 4.3 implies that is bounded in . We may proceed analogously to Theorem 4.1 and Remark 5 to obtain

    (ii) This part is analogous to Theorem 4.1 and Remark 5.

    The results of Lemmas (4.2) and (4.4), Theorem (4.5) and [29, 20,21] can be summarized in the following commutative diagram:

    The arrows denote Mosco convergence in the corresponding convergence regimes.

    In this section we demonstrate how to lift homogenization results w.r.t. two-scale convergence in the mean to quenched statements at the example of Section 4.1. Throughout this section we assume that is ergodic. For we define ,

    with satisfying (A1)-(A3). The goal of this section is to relate two-scale limits of "mean"-minimizers, i.e. functions that minimize , with limits of "quenched"-minimizers, i.e. families of minimizers to in . We also remark that if is strictly convex and may be identified since minimizers of both functionals and are unique.

    Before presenting the main result of this section, we remark that in the ergodic case, the limit functional (20) reduces to a single-scale energy

    where the homogenized integrand is given for and by

    (31)

    In particular, we may obtain an analogous statement to Theorem 4.1 where we replace with . The proof of this follows analogously with the only difference that in the construction of the recovery sequence we first need to find such that . This is done by a usual measurable selection argument, cf. [34,Theorem 7.6].

    Theorem 4.6. Let , be open and bounded, and be ergodic. Assume (A1)-(A3). Let be a minimizer of . Then there exists a subsequence such that generates a Young measure in in the sense of Theorem 3.12, and for -a.a. , concentrates on the set of minimizers of the limit functional. Moreover, if is strictly convex for all and -a.a. , then the minimizer of and the minimizer of are unique, and for -a.a. we have (for a not relabeled subsequence)

    Remark 6 (Identification of quenched two-scale cluster points). If we combine Theorem 4.6 with the identification of the support of the Young measure in Theorem 3.12 we conclude the following: There exists a subsequence such that two-scale converges in the mean to a limit of the form with , and for a.a. the set of quenched -two-scale cluster points is contained in . In the strictly convex case we further obtain that where is the unique minimizer to . Note, however, that our argument (that extracts quenched two-scale limits from the sequence of "mean" minimizers) involves an exceptional -null-set that a priori depends on the selected subsequence. This is in contrast to the classical result in [11] which is based on a subadditive ergodic theorem and states that there exists a set of full measure such that for all the minimizer to weakly converges in to the deterministic minimizer of the reduced functional for any sequence .

    In the proof of Theorem 4.6 we combine homogenization in the mean in form of Theorem 4.1, the connection to quenched two-scale limits via Young measures in form of Theorem 3.12, and a recent result described in Remark 3 by Nesenenko and the first author.

    Proof of Theorem 4.6. Step 1. (Identification of the support of ).

    Since is a sequence of minimizers, by Corollary 4 there exists a subsequence (not relabeled) and minimizers of such that that , , and

    (32)

    In particular, the sequence is bounded in . By Theorem 3.12 we may pass to a further subsequence (not relabeled) such that generates a Young measure on . Since is supported on the set of quenched -two-scale cluster points of , we deduce from Lemma 3.10 that the support of is contained in which is a closed subspace of . Moreover, thanks to the relation of the generated Young measure and stochastic two-scale convergence in the mean, we have . Furthermore, Lemma 3.14 implies that

    In view of (32) and the fact that is supported in , we conclude that

    Since , we have , and thus we conclude that for -a.a. , concentrates on .

    Step 2. (The strictly convex case).

    The uniqueness of and is clear. From Step 1 we thus conclude that where . Theorem 3.12 implies that (for -a.a. ). By Lemma 3.14 we have for -a.a. ,

    On the other hand, since minimizes , we deduce by a standard argument that for -a.a. ,

    The authors thank Alexander Mielke for fruitful discussions and valuable comments. MH has been funded by Deutsche Forschungsgemeinschaft (DFG) through grant CRC 1114 "Scaling Cascades in Complex Systems", Project C05 "Effective models for materials and interfaces with multiple scales". SN and MV acknowledge funding by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – project number 405009441.



    [1] Homogenization and two-scale convergence. SIAM J. Math. Anal. (1992) 23: 1482-1518.
    [2] Stochastic homogenization of elliptic boundary-value problems with -data. Asymptot. Anal. (1998) 17: 165-184.
    [3] Derivation of the double porosity model of single phase flow via homogenization theory. SIAM J. Math. Anal. (1990) 21: 823-836.
    [4] A general approach to lower semicontinuity and lower closure in optimal control theory. SIAM J. Control Optim. (1984) 22: 570-598.
    [5]

    A. Bourgeat, S. Luckhaus and A. Mikelić, A rigorous result for a double porosity model of immiscible two-phase flow, Comptes Rendusa l'Académie des Sciences, 320 (1994), 1289–1294.

    [6] Stochastic two-scale convergence in the mean and applications. J. Reine Angew. Math. (1994) 456: 19-51.
    [7] Homogenization of nonlinear integrals via the periodic unfolding method. C. R. Math. (2004) 339: 77-82.
    [8] The periodic unfolding method in domains with holes. SIAM J. Math. Anal. (2012) 44: 718-760.
    [9] Periodic unfolding and homogenization. C. R. Math. (2002) 335: 99-104.
    [10] The periodic unfolding method in homogenization. SIAM J. Math. Anal. (2008) 40: 1585-1620.
    [11] Nonlinear stochastic homogenization.. Ann. Mat. Pura Appl. (1986) 144: 347-389.
    [12]

    D. Daley and D. Vere-Jones, An Introduction to the Theory of Point Processes, Springer Series in Statistics. Springer-Verlag, New York, 1988.

    [13] Optimal homogenization rates in stochastic homogenization of nonlinear uniformly elliptic equations and systems. Arch. Ration. Mech. Anal. (2021) 242: 343-452.
    [14] Error estimate and unfolding for periodic homogenization. Asymptot. Anal. (2004) 40: 269-286.
    [15] Homogenization in gradient plasticity. Math. Models Methods Appl. Sci. (2011) 21: 1651-1684.
    [16] An extension of the stochastic two-scale convergence method and application. Asymptot. Anal. (2011) 72: 1-30.
    [17] Stochastic homogenization of rate-independent systems and applications. Contin. Mech. Thermodyn. (2017) 29: 853-894.
    [18] Stochastic homogenization of rate-dependent models of monotone type in plasticity. Asymptot. Anal. (2019) 112: 185-212.
    [19] Stochastic homogenization of -convex gradient flows. Discrete Contin. Dyn. Syst. Ser. S (2021) 14: 427-453.
    [20]

    H. Hoppe, Homogenization of Rapidly Oscillating Riemannian Manifolds, Dissertation, TU Dresden, 2020, https://nbn-resolving.org/urn:nbn:de:bsz:14-qucosa2-743766.

    [21]

    H. Hoppe, S. Neukamm and M. Schäffner, Stochastic homogenization of non-convex integral functionals with degenerate growth, (in preparation), 2021.

    [22]

    V. V. Jikov, S. M. Kozlov and O. A. Oleinik, Homogenization of Differential Operators and Integral Functionals, Springer-Verlag, Berlin, 1994.

    [23] Averaging of random operators. Mat. Sb. (1979) 109: 188-202.
    [24]

    M. Liero and S. Reichelt, Homogenization of Cahn–Hilliard-type equations via evolutionary -convergence, NoDEA Nonlinear Differential Equations Appl., 25 (2018), Paper No. 6, 31 pp.

    [25] Two-scale convergence. Int. J. Pure Appl. Math. (2002) 2: 35-86.
    [26] Two-scale homogenization of nonlinear reaction-diffusion systems with slow diffusion. Netw. Heterog. Media (2014) 9: 353-382.
    [27] Two-scale homogenization for evolutionary variational inequalities via the energetic formulation. SIAM J. Math. Anal. (2007) 39: 642-668.
    [28]

    S. Neukamm, Homogenization, linearization and dimension reduction in elasticity with variational methods, Technische Universität München, 2010.

    [29] Stochastic homogenization of nonconvex discrete energies with degenerate growth. SIAM J. Math. Anal. (2017) 49: 1761-1809.
    [30] Stochastic unfolding and homogenization of spring network models. Multiscale Model. Simul. (2018) 16: 857-899.
    [31] Two-scale homogenization of abstract linear time-dependent PDEs. Asymptot. Anal. (2021) 125: 247-287.
    [32] A general convergence result for a functional related to the theory of homogenization. SIAM J. Math. Anal. (1989) 20: 608-623.
    [33] Boundary value problems with rapidly oscillating random coefficients. Random Fields (1979) 1: 835-873.
    [34]

    M. Varga, Stochastic Unfolding and Homogenization of Evolutionary Gradient Systems, Dissertation, TU Dresden, 2019, https://nbn-resolving.org/urn:nbn:de:bsz:14-qucosa2-349342.

    [35] Towards a two-scale calculus. ESAIM Control Optim. Calc. Var. (2006) 12: 371-397.
    [36]

    C. Vogt, A homogenization theorem leading to a Volterra-integrodifferential equation for permeation chromotography, Preprint No 155, SFB 123, Heidelberg, 1982.

    [37] On an extension of the method of two-scale convergence and its applications. Sb. Math. (2000) 191: 973-1014.
    [38] Homogenization of random singular structures and random measures. Izv. Math. (2006) 70: 19-67.
  • This article has been cited by:

    1. Thuyen Dang, Yuliya Gorb, Silvia Jiménez Bolaños, Homogenization of a NonLinear Strongly Coupled Model of Magnetorheological Fluids, 2023, 55, 0036-1410, 102, 10.1137/22M1476939
    2. S. Yu. Dobrokhotov, V. E. Nazaikinskii, Averaging method for problems on quasiclassical asymptotics, 2024, 70, 2949-0618, 53, 10.22363/2413-3639-2024-70-1-53-76
    3. Patrick Dondl, Yongming Luo, Stefan Neukamm, Steve Wolff-Vorbeck, Efficient Uncertainty Quantification for Mechanical Properties of Randomly Perturbed Elastic Rods, 2024, 22, 1540-3459, 1267, 10.1137/23M1574233
    4. Sabine Haberland, Patrick Jaap, Stefan Neukamm, Oliver Sander, Mario Varga, Representative Volume Element Approximations in Elastoplastic Spring Networks, 2024, 22, 1540-3459, 588, 10.1137/23M156656X
    5. S. Yu. Dobrokhotov, V. E. Nazaikinskii, Homogenization Method for Problems on Quasiclassical Asymptotics, 2024, 1072-3374, 10.1007/s10958-024-07490-6
  • 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(1355) PDF downloads(157) Cited by(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog