Loading [MathJax]/jax/element/mml/optable/MiscTechnical.js
Research article

Automatic segmentation of breast cancer histological images based on dual-path feature extraction network


  • The traditional manual breast cancer diagnosis method of pathological images is time-consuming and labor-intensive, and it is easy to be misdiagnosed. Computer-aided diagnosis of WSIs gradually comes into people*s sight. However, the complexity of high-resolution breast cancer pathological images poses a great challenge to automatic diagnosis, and the existing algorithms are often difficult to balance the accuracy and efficiency. In order to solve these problems, this paper proposes an automatic image segmentation method based on dual-path feature extraction network for breast pathological WSIs, which has a good segmentation accuracy. Specifically, inspired by the concept of receptive fields in the human visual system, dilated convolutional networks are introduced to encode rich contextual information. Based on the channel attention mechanism, a feature attention module and a feature fusion module are proposed to effectively filter and combine the features. In addition, this method uses a light-weight backbone network and performs pre-processing on the data, which greatly reduces the computational complexity of the algorithm. Compared with the classic models, it has improved accuracy and efficiency and is highly competitive.

    Citation: Xi Lu, Xuedong Zhu. Automatic segmentation of breast cancer histological images based on dual-path feature extraction network[J]. Mathematical Biosciences and Engineering, 2022, 19(11): 11137-11153. doi: 10.3934/mbe.2022519

    Related Papers:

    [1] Martin Heida . Stochastic homogenization on perforated domains Ⅰ – Extension Operators. Networks and Heterogeneous Media, 2023, 18(4): 1820-1897. doi: 10.3934/nhm.2023079
    [2] Arianna Giunti . Convergence rates for the homogenization of the Poisson problem in randomly perforated domains. Networks and Heterogeneous Media, 2021, 16(3): 341-375. doi: 10.3934/nhm.2021009
    [3] Daniel Peterseim . Robustness of finite element simulations in densely packed random particle composites. Networks and Heterogeneous Media, 2012, 7(1): 113-126. doi: 10.3934/nhm.2012.7.113
    [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] Patrizia Donato, Olivier Guibé, Alip Oropeza . Corrector results for a class of elliptic problems with nonlinear Robin conditions and L1 data. Networks and Heterogeneous Media, 2023, 18(3): 1236-1259. doi: 10.3934/nhm.2023054
    [6] Shin-Ichiro Ei, Toshio Ishimoto . Effect of boundary conditions on the dynamics of a pulse solution for reaction-diffusion systems. Networks and Heterogeneous Media, 2013, 8(1): 191-209. doi: 10.3934/nhm.2013.8.191
    [7] Brahim Amaziane, Leonid Pankratov, Andrey Piatnitski . Homogenization of variational functionals with nonstandard growth in perforated domains. Networks and Heterogeneous Media, 2010, 5(2): 189-215. doi: 10.3934/nhm.2010.5.189
    [8] 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
    [9] Manuel Friedrich, Bernd Schmidt . On a discrete-to-continuum convergence result for a two dimensional brittle material in the small displacement regime. Networks and Heterogeneous Media, 2015, 10(2): 321-342. doi: 10.3934/nhm.2015.10.321
    [10] Iryna Pankratova, Andrey Piatnitski . Homogenization of convection-diffusion equation in infinite cylinder. Networks and Heterogeneous Media, 2011, 6(1): 111-126. doi: 10.3934/nhm.2011.6.111
  • The traditional manual breast cancer diagnosis method of pathological images is time-consuming and labor-intensive, and it is easy to be misdiagnosed. Computer-aided diagnosis of WSIs gradually comes into people*s sight. However, the complexity of high-resolution breast cancer pathological images poses a great challenge to automatic diagnosis, and the existing algorithms are often difficult to balance the accuracy and efficiency. In order to solve these problems, this paper proposes an automatic image segmentation method based on dual-path feature extraction network for breast pathological WSIs, which has a good segmentation accuracy. Specifically, inspired by the concept of receptive fields in the human visual system, dilated convolutional networks are introduced to encode rich contextual information. Based on the channel attention mechanism, a feature attention module and a feature fusion module are proposed to effectively filter and combine the features. In addition, this method uses a light-weight backbone network and performs pre-processing on the data, which greatly reduces the computational complexity of the algorithm. Compared with the classic models, it has improved accuracy and efficiency and is highly competitive.



    Soon after the groundbreaking introduction of stochastic homogenization by Papanicolaou and Varadhan [22] and Kozlov [12], research developed a natural interest in the homogenization of randomly perforated domains. A good summary of the existing methods up to 1994 can be found in [13]. By the same time, Zhikov [28] provided a homogenization result for linear parabolic equations on stationary randomly perforated domains. Bourgeat et al. [27] introduced a concept of two-scale convergence in the mean for stochastic problems, where the two-scale limit is performed simultaneously over all realizations. This method was the first to apply the two-scale convergence idea to a stochastic setting, but it was not capable to deal with nonlinearities that require strong convergence of solutions (see also [10]). Hence, we cannot benefit of this approach in our work. It then became silent for a decade. In [30], Zhikov and Piatnitsky reopened the case by introducing stochastic two-scale convergence as a generalization of [20,1,29] to the stochastic setting, particularly to random measures that comprise random perforations and random lower-dimensional structures in a natural way. The method was generalized to various applications in discrete and continuous homogenization [16,3,4] and recently also to an unfolding method [19,10].

    Concerning the homogenization on randomly perforated domains, there seems to be few results in the literature, with [7,5,23] being the closest related work from the PDE point of view. We emphasize that there is a further discipline in stochastic homogenization, studying critical regimes of scaling for holes in a perforated domain of the Stokes equation; see [6] and references therein.

    In this work, we focus on the geometric aspects in the homogenization of quasi-linear parabolic equations and go beyond any recent assumptions on the random geometry. Given ε>0, we consider a bounded domain QRd perforated by a random set Gε and write Qε:=QGε. Typically, GεεG where G is a stationary random set and Gε is additionally regularized close to Q [7,5,23]. Also, we assume that Qε is connected, for simplicity of calculations and presentation, and our geometric model will be regularized in such a way; see also Remark 5.

    We then study the following PDE on Qε for the time interval I=[0,T]:

    tuε(A(uε)uε)=fin I×QεA(uε)uεν=0on I×QA(uε)uεν=εh(uε)on I×QεQuε(0,x)=u0(x)in Qε (1.1)

    with ν being the outer normal vector.

    In case of a fully linear PDE, i.e., h()=const and A()=const, this problem was homogenized already in the aforementioned [28] and later reconsidered in [30]. In this linear case, one benefits from the regularity of the limit solution and the weak convergence of the ε-solutions that is given a priori.

    However, the nonlinear case is more difficult. Weak convergence of solutions is no longer sufficient. Thus, one needs to establish strong convergence of the uε. As we will discuss below, a lack of a uniformly continuous family of extension operators or, more generally, a degeneration of the homogenized matrix will cause the whole argumentation to break down. Hence, typical assumptions in the literature, such as minimal smoothness (see Definition 23) of G and uniform boundedness of the hole sizes, ensure the existence of uniformly bounded extension operators Uε,:W1,2(Qε)W1,2(Q) [7]. This in turn implies weak compactness of Uε,uε in W1,2(Q), a property of uttermost importance to pass to the homogenization limit in the nonlinear terms. Other approaches are conceivable, e.g., exploiting the Frechet–Riesz–Kolmogorov compactness theorem, but in application the prerequisites are hard to prove.

    If all limit passages go through, the homogenized limit as ε0 reads for some positive definite matrix A(G) as

    C1,(G)tudiv(A(u)A(G)u)C2,(G)h(u)=C1,(G)fin I×QA(u)A(G)uν=0on I×Qu(0,x)=C1,(G)u0(x)in Q, (1.2)

    which represents the macroscopic behavior of our object. We note at this point that positivity of A(G) is, in general, nontrivial but can be shown for minimally smooth domain examples (see Sections 8 and 9).

    Unfortunately, canonical perforation models are neither minimally smooth nor is the size of their holes uniformly bounded. Our toy model of choice will be the Boolean model ΞXpoi:=¯xXpoiBr(x) (see Definition 1) driven by a Poisson point process Xpoi. It clearly reveals the following general issues for the homogenization analysis:

    i) ΞXpoi=RdΞXpoi is not connected due to areas that are encircled.

    ii) Two distinct balls can lie arbitrarily close to each other or – in case they intersect – have arbitrary small overlap. This implies that

    ● the connected components in ΞXpoi develop arbitrarily large local Lipschitz constants: Two balls of equal radius intersecting at an angle α have the Lipschitz constant tan((πα)/2) at the points of intersection, and

    ● there is no δ>0 such that for every pΞXpoi the surface Bδ(p)ΞXpoi is a graph of a function: If x,yXpoi with |xy|=2r+η and |px|=r, |py|=r+η, Bδ(p)ΞXpoi can be a graph only if δ<η.

    The first issue can be fixed by considering a "filled-up model" Xpoi in Definition 1. Furthermore, we will see that – under mild regularity conditions – the procedure of filling up a perforation does not change the effective conductivity A(G) at all (Theorem 58). Unfortunately, the second issue poses an actual problem. In a recent work [9], one of the authors has shown that in some cases an extension operator Uε,:W1,p(Qε)W1,q(Q), 1q<p, can be constructed for some geometries including the Boolean model (strictly speaking, this was shown for an extension from the balls to the complement in the percolation case). However, [9] also suggests that the Boolean model for the Poisson point process requires p>2 in order for Uε, to be properly defined for some q1.

    Due to these severe analytical difficulties, we need other approaches to the problem. We call our approach 'regularized homogenization' and it consists of an approximation of the random geometry G by Gn, performing homogenization, and afterward letting n. In our particular example, it consists of the following steps:

    i) Given a general stationary ergodic (admissible) random point process X, we construct a regularization X(n):=FnX (see Definition 3) such that the set X(n) is uniformly minimally smooth for given nN.

    ii) Given nN, we perform homogenization for the smoothed geometry X(n) instead of X (see Lemma 7).

    iii) We pass to the limit n along a subsequence to obtain the anticipated homogenized limit problem (see Theorem 56), where the coefficients are independent from the intermediate regularizations (Corollary 39). This happens under the assumption that X is statistically connected (see Definition 15).

    iv) We show that the Poisson point process in the subcritical regime is a valid example for our general homogenization result (see Section 9).

    We are thus in a position to prove an indirect homogenization result. This seems to us an appropriate intermediate step on the way to a full homogenization result, which may be achieved in the future using further developed hom ogenization techniques based on a better understanding of the interaction of geometry and homogenization. Let us note that we focus on fixed radii r>0 for the sake of presentation. In fact, random radii pose no issue for the procedure as briefly mentioned in Remark 18 as long as the remaining conditions are satisfied.

    This paper is structured as follows:

    ● In Section 3, we introduce the core objects and state the main result. This includes the thinned point processes X(n) and its filled-up Boolean model X(n).

    ● In Section 4, we prove relevant properties of the thinning map and the thinned point processes, most importantly, minimal smoothness of X(n) (Theorem 25) and X(n)X in a certain sense (Lemma 29).

    ● Section 5 deals with the cell solutions and the definition of the effective conductivity A.

    ● The homogenization theory for minimally smooth holes is sketched in Section 6 on the basis of stochastic two-scale convergence. Due to the considerations in Section 5, the underlying probability space is a compact separable metric space.

    ● In Section 7, we show that the homogenized solutions to Eq (1.2) for G=X(n) converge and that their limit is a solution to the anticipated limit problem for G=X.

    ● Section 8 establishes a criterion for statistical connectedness (nondegeneracy of the effective conductivity A) using percolation channels. We follow the ideas in [13, Chapter 9] where a discrete model was considered.

    ● In Section 9, we show that the Poisson point process Xpoi is indeed admissible, which follows from readily available percolation results. Showing statistical connectedness of Xpoi is much harder. We do so using the criterion established in Section 8 and a version of [11, Theorem 11.1]. As the original [11, Theorem 11.1] is a statement about percolation channels on the Z2-lattice, we need to adjust both the statement and the proof to our setting.

    General notation

    M(Rd): Space of Radon measures on Rd equipped with the vague topology

    S(Rd)M(Rd): Space of boundedly finite point clouds/point measures in Rd

    A: Complement of a set A

    B(X): Borel-σ-algebra of the topological space X

    Ld: d-dimensional Lebesgue-measure

    Hd: d-dimensional Hausdorff-measure

    H: Restriction of {\mathcal{H}}^{d} to A , i.e. {\mathcal{H}}_{\llcorner A}^{d}(B): = {\mathcal{H}}^{d}(B\cap A)

    o: = 0_{{\mathbb{R}}^{d}}\in{\mathbb{R}}^{d} : Origin in {\mathbb{R}}^{d}

    {\mathbb{1}}_{A} : Indicator/characteristic function of a set A

    Specific notation introduced later

    \mathbb{B}_{r}{\left(A\right)} : Open r -neighborhood around A . (Definition 1)

    \Xi\mathbb{x} and {\boldsymbol{\boxminus}}\mathbb{x} : Boolean model of \mathbb{x} and its filled version (Definition 1)

    {\mathcal{C}}_{\mathbb{x}}(x) : Cluster of x in \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) (Definition 3)

    \mathbb{x}^{(n)} for \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) : \mathbb{x}^{(n)} = {\mathtt{F}_{n}}\mathbb{x} with thinning map {\mathtt{F}_{n}} (Definition 3)

    Q_{\mathbb{x}}^{{\varepsilon}} and J^{{\varepsilon}}(Q, \mathbb{x}) : Perforated domain and index set generating perforations (Definition 4)

    \tau_{x}\colon {\mathcal{M}}({\mathbb{R}}^{d})\to{\mathcal{M}}({\mathbb{R}}^{d}) : Shift-operator in {\mathcal{M}}({\mathbb{R}}^{d}) (Definition 8)

    \lambda(\mu) : Intensity of random measure \mu (Definition 8)

    \mu_{\mathbb{x}} : {\mathcal{H}}^{d-1} restricted to {\partial}{\boldsymbol{\boxminus}}\mathbb{x} (Definition 30)

    {\mathcal{A}} and \alpha_{{\mathcal{A}}} : Effective conductivity and smallest eigenvalue of {\mathcal{A}} (Definition 37)

    {\mathcal{U}} and {\mathcal{T}} : Extension and trace operators (Theorem 43 and Theorem 51)

    \mu^{{\varepsilon}} : Scaled measure (Assumption 46)

    We start by introducing some concepts from the theory of point processes. We will not formulate the concepts in full generality but only as general as needed for our purpose. Let d\geq2 and let {\mathcal{S}}({\mathbb{R}}^{d}) be the set of boundedly finite point clouds in {\mathbb{R}}^{d} (i.e., point clouds without accumulation points) and {\mathcal{M}}({{{\mathbb{R}}}^{d}}) the space of Radon measures with the vague topology, that is, the smallest topology on {\mathcal{M}}({\mathbb{R}}^{d}) such that

    \mu\mapsto\int_{{\mathbb{R}}^{d}}f\text{d}\mu

    is continuous for every f\in C_{c}^{\infty}({\mathbb{R}}^{d}) . Every \mathbb{x}\in{\mathcal{S}}({{{\mathbb{R}}}^{d}}) can be identified with a Borel measure through the correspondence

    \mathbb{x}(A) = \sum\limits_{x\in\mathbb{x}}\delta_{x}(A).

    Hence, we identify {\mathcal{S}}({\mathbb{R}}^{d})\subset{\mathcal{M}}({\mathbb{R}}^{d}) .

    Our perforation model of interest is the Boolean model driven by a point cloud \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) . While it is a natural way to generate perforations, we require its complement to be connected for suitable \mathbb{x} . Hence, the perforation \Xi\mathbb{x} needs to be filled up in order to remove all finite-sized connected components from its complement. These can be easily identified as they do not admit a path of infinite diameter.

    Definition 1 (Boolean model \Xi of a point cloud and filled-up model {\boldsymbol{\boxminus}} (see Figure 1)). Let \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) . The Boolean model of \mathbb{x} for a radius r > 0 is

    Figure 1.  Initial Boolean model \Xi\mathbb{x} vs filled-up Boolean model {\boldsymbol{\boxminus}}\mathbb{x} .
    \Xi\mathbb{x}: = \overline{\bigcup\limits_{x\in\mathbb{x}}\mathbb{B}_{r}{\left(x\right)}} = \overline{\mathbb{B}_{r}{\left({\mathbb{x}}\right)}},

    where \mathbb{B}_{r}{\left(x\right)} is the open ball of radius r around x and \mathbb{B}_{r}{\left(A\right)}: = \bigcup_{x\in A}\mathbb{B}_{r}{\left(x\right)} . We define the filled-up Boolean model {\boldsymbol{\boxminus}}\mathbb{x} of \mathbb{x} for radius r through its complement, i.e.,

    \begin{align*} {\boldsymbol{\boxminus}}\mathbb{x}{}^{\complement}: = \big\{ x\in{\mathbb{R}}^{d}\, \vert\, & \exists\gamma\colon[0, \infty)\to\Xi\mathbb{x}{}^{\complement}{\text{continuous and}}\gamma(0) = x, \, \limsup\limits_{t\to\infty}{\big|}\gamma(t){\big|} = \infty\big\}. \end{align*}

    Remark 2. We observe that

    \Xi(\mathbb{x}+x) = \Xi(\mathbb{x})+x\qquad{\text{and}}\qquad{\boldsymbol{\boxminus}}(\mathbb{x}+x) = {\boldsymbol{\boxminus}}\mathbb{x}+x.

    As discussed in the introduction, we need to "smoothen" the geometry in order to be able to apply standard homogenization methods. Given a Lipschitz domain P\subset{{{\mathbb{R}}}^{d}} , we define for p\in\partial P

    \delta{\left(p\right)}: = \frac{1}{2}\sup\limits_{\delta' > 0}\left\{ \, \partial P{\text{ is }}{\text{ Lipschitz-graph in }}\mathbb{B}_{{\delta'}}{\left(p\right)}\right\},

    and because \delta\colon \partial P\to{\mathbb{R}}_{\geq0} is continuous [9], we can define for bounded P

    \delta(P): = \min\limits_{p\in\partial P}\delta(p).

    Definition 3 (Thinning maps {\mathtt{F}_{n}} (see Figure 2)). Let \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) be a point cloud. We denote the cluster of x in \mathbb{x} by

    Figure 2.  Thinning of point clouds under {\mathtt{F}_{n}} pictured via the Boolean model \Xi . From left to right are \mathbb{x} , \mathbb{x}^{(2)} , and \mathbb{x}^{(5)} .
    {\mathcal{C}}_{\mathbb{x}}(x): = \{y\in\mathbb{x}\, \vert\, \exists{\text{path from }}x{\text{ to }}y{\text{ inside }}\Xi\mathbb{x}\}.

    We set

    \begin{align} {\mathtt{F}}_{1, n}\mathbb{x} & : = \left\{ x\in\mathbb{x}\, \vert\, \forall y\in\mathbb{x}\colon d(x, y)\notin\big(0, \tfrac{1}{n}\big)\cup\big(2r-\tfrac{1}{n}, \, 2r+\tfrac{1}{n}\big)\right\}, \end{align} (3.1)
    \begin{align} {\mathtt{F}}_{2, n}\mathbb{x} & : = \left\{ x\in\mathbb{x}\, \vert\, \#\mathcal{C}_{\mathbb{x}}(x)\leq n, \, \delta{\left({\mathbb{B}_{r}{\left({{\mathcal{C}}_{\mathbb{x}}(x)}\right)}}\right)}\geq\tfrac{1}{n}\right\}, \end{align} (3.2)

    and define the thinning map {\mathtt{F}_{n}}

    {\mathtt{F}_{n}}\colon {\mathcal{S}}({\mathbb{R}}^{d})\to{\mathcal{S}}({\mathbb{R}}^{d}), \qquad{\mathbb{x}^{(n)}}: = {\mathtt{F}_{n}}\mathbb{x}: = \big({\mathtt{F}}_{2, n}\circ{\mathtt{F}}_{1, n}\big)\mathbb{x}.

    {\mathtt{F}_{n}} can be understood as a generalization of the classical Matérn construction [15,26]. For an arbitrary \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) , we see that ({\boldsymbol{\boxminus}}{\mathbb{x}^{(n)}})^{\complement} is always minimally smooth (see Definition 23). Furthermore, if {\mathbb{X}} is a stationary point process (as defined later), then the same holds for {\mathbb{X}}^{(n)} = {\mathtt{F}_{n}}{\mathbb{X}} . We note that {\mathtt{F}_{n}} is in general not monotone in n , i.e., {\mathtt{F}}_{m}\mathbb{x}\not\subset{\mathtt{F}_{n}}\mathbb{x} for m\leq n .

    Given a scale {\varepsilon} > 0 , we define the perforation domain Q^{{\varepsilon}} such that the perforations have some minimal distance from the boundary {\partial} Q :

    Definition 4 (Perforation of domain Q^{{\varepsilon}} ). Let \mathbb{x}\in{\mathcal{S}}({{{\mathbb{R}}}^{d}}) . We set

    \begin{align*} J^{{\varepsilon}}(\mathbb{x}, Q) & : = \left\{ x\in\mathbb{x}\, \vert\, {\mathrm{dist}}{\left({{\varepsilon}\, {\mathcal{C}}_{\mathbb{x}}(x), \, Q^{\complement}}\right)} > 2{\varepsilon} r\right\} , \qquad G_{\mathbb{x}}^{{\varepsilon}}: = {\varepsilon}{\boldsymbol{\boxminus}}(J^{{\varepsilon}}(\mathbb{x}, Q)) \end{align*}

    as well as the perforated domain

    Q_{\mathbb{x}}^{{\varepsilon}}: = Q\, \backslash\, G_{\mathbb{x}}^{{\varepsilon}}.

    One quickly verifies that Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}} is minimally smooth (Definition 23); see Theorem 25.

    Remark 5. By construction, our perforation model G_{\mathbb{x}}^{{\varepsilon}}: = {\varepsilon}{\boldsymbol{\boxminus}}(J^{{\varepsilon}}(\mathbb{x}, Q)) ensures connectedness of its complement {{{\mathbb{R}}}^{d}}\setminus G_{\mathbb{x}}^{{\varepsilon}} , resp., of Q_{\mathbb{x}}^{{\varepsilon}} . We prefer this approach of "filling the holes of the holes" because it allows for an easy use of extension operators in a way that we can safely apply compact embeddings. Otherwise, once {{{\mathbb{R}}}^{d}}\setminus G_{\mathbb{x}}^{{\varepsilon}} becomes disconnected, we would have to rely on additional regularity assumptions on the initial data in those holes, which is then carried on by the evolution equation and would lead to additional tedious calculations. Otherwise, we would have to introduce a second PDE on the holes with suitable coupling boundary conditions. While we do not claim that a treatment of such a situation is impossible, we claim that the additional effort would not be justified by the additional expected insight.

    We make the following parameter assumptions on our partial differential Eq (1.1).

    Assumption 6 (Parameters of PDE). Let I = [0, T]\subset{\mathbb{R}} and Q\subset{\mathbb{R}}^{d} be a bounded, connected open domain. We assume that

    u_{0}\in W^{1, 2}(Q)

    f\in L^{2}(I; L^{2}(Q))

    h\colon {\mathbb{R}}\to{\mathbb{R}} is Lipschitz continuous with Lipschitz constant L_{h}

    A\colon {\mathbb{R}}\to{\mathbb{R}} is continuous with 0 < \inf(A) and \sup(A) < \infty .

    Generalized time derivatives will always be considered under the evolution triple W^{1, 2}(Q)\hookrightarrow L^{2}(Q)\hookrightarrow W^{1, 2}(Q)^{*} or W^{1, 2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})\hookrightarrow L^{2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})\hookrightarrow W^{1, 2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})^{*} in the case of a perforated domain {Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}} .

    Lemma 7 (Solution to PDE for minimally smooth holes). Let \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) and n\in{\mathbb{N}} . Under Assumption 6, we have on {Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}} : There exists a weak solution {u^{{\varepsilon}}}\in L^{2}(I; W^{1, 2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})) with generalized time derivative {\partial}_{t}{u^{{\varepsilon}}}\in L^{2}(I; W^{1, 2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})^{*}) to Eq (1.1) on {Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}} instead of Q^{\varepsilon} .

    This u^{{\varepsilon}} satisfies for some C > 0 independent of {\varepsilon} depending only on Q , n , f and u_{0} but not on {\varepsilon}

    \begin{align*} \underset{t \in I}{\operatorname{ess} \sup }{\|}{u^{{\varepsilon}}}(t){\|}_{L^{2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})}+{\|}{u^{{\varepsilon}}}{\|}_{L^{2}(I;W^{1, 2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}}))} & \leq C\\ {\|}{\partial}_{t}{u^{{\varepsilon}}}{\|}_{L^{2}(I;W^{1, 2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})^{*})} & \leq C. \end{align*}

    The proof is given in Section 6 (Theorem 53).

    The next step is passing to the limit {\varepsilon}\to0 . We do so in the case of \mathbb{x} being a realization of a stationary ergodic point process {\mathbb{X}} as defined below:

    Definition 8 (Random measure and shift-operator \tau_{x} ). A random measure \mu_{\bullet} is a random variable with values in {\mathcal{M}}({\mathbb{R}}^{d}) . It induces a probability distribution {\mathbb{P}} on {\mathcal{M}}({{{\mathbb{R}}}^{d}}) . Given the continuous map

    \begin{equation} \tau_{x}\colon{\mathcal{M}}({{{\mathbb{R}}}^{d}})\to{\mathcal{M}}({{{\mathbb{R}}}^{d}}), \qquad\tau_{x}\xi(A): = \xi(A+x), \end{equation} (3.3)

    a random measure is stationary if {\mathbb{P}}(F) = {\mathbb{P}}(\tau_{x}F) for every F\in{\mathcal{B}}({\mathcal{M}}({{{\mathbb{R}}}^{d}})) and every x\in{{{\mathbb{R}}}^{d}} . In line with the above setting, a random point process {\mathbb{X}} is a random measure with {\mathbb{P}}({\mathcal{S}}({\mathbb{R}}^{d})) = 1 , and one quickly verifies that {\mathbb{X}} is stationary if for every N\in{\mathbb{N}} , x\in{{{\mathbb{R}}}^{d}} , and bounded open A\subset{{{\mathbb{R}}}^{d}} , it holds

    {\mathbb{P}}\left(\mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d})\colon\mathbb{x}{\left(A\right)} = N\right) = {\mathbb{P}}\left(\mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d})\colon\mathbb{x}{\left({A+x}\right)} = N\right).

    We call a stationary random measure \mu_{\bullet} ergodic if the \sigma -algebra of \tau -invariant sets is trivial under its distribution {\mathbb{P}} .

    Remark 9 (Compatibility of thinning with shifts). The thinning map {\mathtt{F}_{n}} is compatible with the shift \tau_{x} , i.e., on {\mathcal{S}}({\mathbb{R}}^{d})

    {\mathtt{F}_{n}}\circ\tau_{x} = \tau_{x}\circ{\mathtt{F}_{n}}.

    Lemma 10 (Homogenized PDE for minimally smooth domains). Let {\mathbb{X}} be a stationary ergodic point process and n\in{\mathbb{N}} be fixed. For almost every realization \mathbb{x} of {\mathbb{X}} , we have under Assumption 6:

    For {\varepsilon} > 0 , let {u^{{\varepsilon}}}\in L^{2}(I; W^{1, 2}({Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}})) be a solution to Eq (1.1) on {Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}} instead of Q^{\varepsilon} . Given any sequence {\varepsilon}\to0 , we find a subsequence (still denoted as {\varepsilon}\to0 ) and corresponding \tilde{u}^{{\varepsilon}}\in L^{2}(I; W^{1, 2}(Q)) with \tilde{u}^{{\varepsilon}}|_{{Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}}}} = u^{{\varepsilon}} , such that \tilde{u}^{{\varepsilon}}\to u_{n} strongly in L^{2}(I; L^{2}(Q)) for some u_{n}\in L^{2}(I; W^{1, 2}(Q)) with generalized time derivative {\partial}_{t}u_{n}\in L^{2}(I; W^{1, 2}(Q)^{*}) . This u_{n} is a weak solution to

    \begin{align} C_{1, {\mathbb{P}}^{(n)}}{\partial}_{t}u_{n}-\nabla\cdot\big(A(u_{n}){\mathcal{A}}^{(n)}\nabla u_{n}\big)-C_{2, {\mathbb{P}}^{(n)}}h(u_{n}) & = C_{1, {\mathbb{P}}^{(n)}}f & & \mathit{\text{ in }}I\times Q \\ A(u_{n}){\mathcal{A}}^{(n)}\nabla u_{n}\cdot\nu & = 0 & & \mathit{\text{ on }}I\times{\partial} Q \\ u_{n}(0, x) & = C_{1, {\mathbb{P}}^{(n)}}u_{0}(x) & & \mathit{\text{ in }}Q, \end{align} (3.4)

    with constants C_{i, {\mathbb{P}}^{(n)}} > 0 depending on the distribution {\mathbb{P}}^{(n)} of {\mathbb{X}}^{(n)} and {\mathcal{A}}^{(n)} being a symmetric positive semi-definite matrix – the so-called effective conductivity based on the event that the origin is not covered by {\boldsymbol{\boxminus}}{\mathbb{X}}^{(n)} (see Definition 37).

    Proof. This is shown in Section 6 (Theorem 55) using two-scale convergence.

    A more general result, which immediately is implied by the proof of the previous lemma is the following Lemma 11. There are many random geometries that do not have an extension operator W^{1, 2}(Q^{\varepsilon})\to W^{1, 2}(Q) but still an extension operator W^{1, 2}(Q^{\varepsilon})\to W^{1, q}(Q) for some 1\leq q < 2 , see [9]. For such geometries, the desired homogenization result could be established without the bypass that we will take below. However, results in [9] will provide sufficient conditions for homogenization, but they do not provide necessary conditions.

    Lemma 11. Assume that Q^{\varepsilon} is such that there exists q\in(1, 2] such that for every family v^{\varepsilon}\in W^{1, 2}(Q^{\varepsilon}) with \sup_{\varepsilon}(\left\Vert {\nabla v^{\varepsilon}}\right\Vert_{L^2(Q^{\varepsilon})}+\left\Vert {v^{\varepsilon}}\right\Vert_{L^2(Q^{\varepsilon})}) < +\infty , there exists a V^{\varepsilon}\in L^q(Q) , such that v^{\varepsilon} = {\mathbb{1}}_{Q^{\varepsilon}}V^{\varepsilon} and such that V^{\varepsilon} is pre-compact in L^q(Q) . Then, every convergent subsequence of the solutions u^{\varepsilon} of Eq (1.1) converges to a weak solution of Eq (3.5) given below.

    Remark 12 (Estimate). Observe that the situation in Lemma 7 provides us with an estimate \sup_{\varepsilon}(\left\Vert {\nabla u^{\varepsilon}}\right\Vert_{L^2((0, T]\times Q^{\varepsilon})}+\left\Vert {u^{\varepsilon}}\right\Vert_{L^2((0, T]\times Q^{\varepsilon})}) < +\infty on the solutions u^{\varepsilon} , and the compact operator is a combination of the extension operator for minimally smooth domains and the standard Sobolev embedding.

    Remark 13 (Compactness). It has to be noted that the claim of compactness is somewhat natural and for most nonlinearities, even necessary. Indeed, unless the nonlinearities are monotone operators, we are not aware of an existence theory for equations of Type (1.1), which would not rely on compactness of a sequence of approximate solutions in some L^p -space. However, if existence theory is already heavily relying on this compactness assumption, we have no hope that homogenization could cope with less.

    When it comes to the final homogenization result, we will need the following assumptions on the point process {\mathbb{X}} .

    Definition 14 (Admissible point process). We call a point cloud \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) admissible the following holds (with r > 0 from Definition 1):

    i) Equidistance Property: \forall x, y\in\mathbb{x}\colon \left|x-y\right|\neq2r .

    ii) Finite Clusters: For every x\in\mathbb{x} , we have \#{\mathcal{C}}_{\mathbb{x}}(x) < \infty .

    A stationary ergodic boundedly finite point process {\mathbb{X}} is called admissible if its realizations are almost surely admissible.

    Definition 15 (Statistical connectedness). The random set {\boldsymbol{\boxminus}}{\mathbb{X}}^{\complement} is statistically connected if the effective conductivity {\mathcal{A}} (Definition 37) based on the event that the origin is covered by {\boldsymbol{\boxminus}}{\mathbb{X}}^{\complement} is strictly positive definite.

    In our setting of Boolean models of admissible point processes, the procedure of filling up a perforation does not change its effective conductivity {\mathcal{A}} (Theorem 58). In particular, {\boldsymbol{\boxminus}}{\mathbb{X}}^{\complement} is statistically connected if, and only if, the same holds for \Xi{\mathbb{X}}^{\complement} .

    Remark 16 (On statistical connectedness). Let us note that the existence of an infinite connected component in {\boldsymbol{\boxminus}}{\mathbb{X}}^\complement is a necessary requirement for {\mathcal{A}} > 0 , as can be seen later in Section 8. There, we also give a criterion for statistical connectedness: the existence of sufficiently many so-called percolation channels. The procedure is based on [13, Lemma 9.7], which we adjust to the continuum setting.

    We may now state the main theorem of this work.

    Theorem 17 (Homogenized limit for admissible point processes). Let {\mathbb{X}} be an admissible point process and {\boldsymbol{\boxminus}}{\mathbb{X}}{}^{\complement} statistically connected. Under Assumption 6, we have for almost every realization \mathbb{x} of {\mathbb{X}} : For every n\in{\mathbb{N}} , let u_{n} be a homogenized limit in Lemma 10. For every subsequence of (u_n)_{n\in{\mathbb{N}}} , we can extract a subsequence (u_{n_k})_{k\in{\mathbb{N}}} such that there exists a u\in L^{2}(I; W^{1, 2}(Q)) with generalized time-derivative {\partial}_{t}u\in L^{2}(I; W^{1, 2}(Q)^{*})

    \begin{align*} u_{n} & \mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(I;W^{1, 2}(Q))}u\qquad\mathit{\text{and}}\qquad{\partial}_{t}u_{n}\mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(I;W^{1, 2}(Q)^{*})}{\partial}_{t}u, \end{align*}

    and u is a weak solution to

    \begin{equation} \begin{aligned} C_{1, {\mathbb{P}}}{\partial}_{t}u-\nabla\cdot\big(A(u){\mathcal{A}}\nabla u\big)-C_{2, {\mathbb{P}}}h(u) & = C_{1, {\mathbb{P}}}f & & \mathit{\text{ in }}I\times Q\\ A(u){\mathcal{A}}\nabla u\cdot\nu & = 0 & & \mathit{\text{ on }}I\times{\partial} Q\\ u(0, x) & = C_{1, {\mathbb{P}}}u_{0}(x) & & \mathit{\text{ in }}Q, \end{aligned} \end{equation} (3.5)

    with constants C_{i, {\mathbb{P}}} > 0 only depending on the distribution {\mathbb{P}} of {\mathbb{X}} and {\mathcal{A}} being a symmetric positive definite matrix – the so-called effective conductivity {\mathcal{A}} based on the event that the origin is not covered by {\boldsymbol{\boxminus}}{\mathbb{X}} (Definition 37). The limit u may depend on the chosen subsequence if the solution to Eq (3.5) is not unique. Otherwise, the whole sequence (u_n)_{n\in{\mathbb{N}}} converges to u .

    Proof. This is proved in Theorem 56.

    Regarding the derivation of Eq (3.5), we only need local convergence of {\boldsymbol{\boxminus}}\mathbb{x}^{(n)}\to{\boldsymbol{\boxminus}}\mathbb{x} (Lemma 29). Hence, Eq (3.5) is independent of the thinning procedure as long as local convergence is satisfied.

    Remark 18 (Random radii). For simplicity, we have chosen the Boolean model with fixed radius r as our underlying model. One can easily generalize the procedure to random independent radii. Given a marked point process \mathcal{X} = \bigcup_i (x_i, r_i) whose marks represent the radius r_i of the ball around x_i , we need to adjust Point 1 in Definition 14 accordingly, i.e.,

    \forall (x, r_x), (y, r_y)\in\mathcal{X}\colon |x-y|\neq r_x+r_y.

    The thinning maps {\mathtt{F}_{n}} (Definition 3) also need to be modified to ensure minimal smoothness (Definition 23), e.g., balls with especially large/small radii need to be removed. Eq (3.1) has the purpose to ensure that two balls in the regularized geometry either have a minimal distance 1/n or, if they intersect, the intersection angle is bounded away from 0 . This also has to be modified depending on radii r_1 and r_2 of two intersecting balls. Local convergence is preserved if all clusters remain finite. Hence, the rest of our procedure essentially follows as is under the assumption of {\mathcal{A}} > 0 .

    Remark 19 (Homogenization procedure). For fixed {\varepsilon} > 0 , solutions u^{{\varepsilon}} = u_{\mathbb{x}}^{{\varepsilon}} to Eq (1.1) exist for admissible \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) as Q_{{\mathbb{x}^{(n)}}}^{{\varepsilon}} = Q_{\mathbb{x}}^{{\varepsilon}} for n large enough (Lemma 29). If \mathbb{x} is a realization of some admissible point process {\mathbb{X}} , then this is still not sufficient to pass to the limit {\varepsilon}\to0 . The missing regularity of {\boldsymbol{\boxminus}}{\mathbb{X}} still prevents us from establishing a priori estimates. All in all, our procedure yields the following diagram:

    (3.6)

    Statistical connectedness of {\boldsymbol{\boxminus}}{\mathbb{X}}{}^{\complement} is crucial to establish W^{1, 2}(Q) -estimates for u_{n} . This indicates that the direct limit passing u^{{\varepsilon}}\to u might only rely on the statistical connectedness property, but we cannot answer that as of now. On the other hand, if the assumptions of Lemma 11 hold, then the diagram commutes.

    In order to demonstrate that the class of point process satisfying our assumptions is not empty, we show in Section 9 that the Poisson point process {\mathbb{X}}_{{\mathrm{poi}}} is indeed suitable for our framework. We obtain the following.

    Theorem 20 (Admissibility and statistical connectedness for {\mathbb{X}}_{{\mathrm{poi}}} ). In the subcritical regime (see Assumption 62), we have for the Poisson point process {\mathbb{X}}_{{\mathrm{poi}}} that

    {\mathbb{X}}_{{\mathrm{poi}}} is an admissible point process.

    {\boldsymbol{\boxminus}}{\mathbb{X}}_{{\mathrm{poi}}}^{\complement} is statistically connected.

    While admissibility is easily proven, statistical connectedness is much harder to deal with. Most of Section 9 is dedicated to this proof. It also builds up on Section 8 in which we show that so-called percolation channels yield statistical connectedness.

    In what follows, we collect examples where stochastic homogenization is well understood and where our new ansatz of regularized stochastic homogenization is consistent in the sense that the Diagram (3.6) commutes for these examples.

    Let G be a stationary ergodic random set and let G_n be jointly stationary sets with G where G_n\subset G = :G_\infty and where G_n\to G pointwise as n\to\infty . We then define Q^{\varepsilon}: = Q\setminus {\varepsilon} G as before and additionally Q^{\varepsilon}_n: = Q\setminus {\varepsilon} G_n . Then, Q^{\varepsilon}_n\supset Q^{\varepsilon} by definition, and we have the following result.

    Theorem 21. Under the above assumptions, let f_n^{\varepsilon}\in L^2(Q_n^{\varepsilon}) be extended by 0 to Q and \lambda > 0 .

    i) For each {\varepsilon} > 0 , assume that f_n^{\varepsilon}{\rightharpoonup} f_\infty^{\varepsilon} = :f^{\varepsilon} as n\to\infty . Then, the solutions u^{\varepsilon}_n of the problem

    \begin{equation} \begin{aligned}-\nabla\cdot\nabla u^{{\varepsilon}}_n + \lambda u^{\varepsilon}_n & = f_n^{\varepsilon} & & \mathit{\text{ in }} Q^{{\varepsilon}}_n\\ u^{{\varepsilon}}_n & = 0 & & \mathit{\text{ on }}{\partial} Q\\ \nabla u^{{\varepsilon}}_n\cdot\nu & = 0 & & \mathit{\text{ on }}{\partial} Q^{{\varepsilon}}_n\backslash{\partial} Q \end{aligned} \end{equation} (3.7)

    satisfy u^{\varepsilon}_n{\rightharpoonup} u^{\varepsilon}_\infty = :u^{\varepsilon} as n\to\infty , where u^{\varepsilon} is the solution to Eq (3.7) on Q_\infty^{\varepsilon}: = Q^{\varepsilon} .

    ii) For each n\in{\mathbb{N}}\cup\{+\infty\} , there exists a positive semi-definite {\mathcal{A}}^{(n)} such that f^{\varepsilon}_n{\rightharpoonup} f_n as {\varepsilon}\to0 implies u_n^{\varepsilon}{\rightharpoonup} u_n in L^2(Q) , where u_n solves

    -\nabla\cdot {\mathcal{A}}^{(n)}\nabla u_n +\lambda u_n = f_n \mathit{\text{ in }} Q.

    iii) If f_n{\rightharpoonup} f_\infty = : f as n\to\infty , then {\mathcal{A}}^{(n)}\to {\mathcal{A}}^{(\infty)} = \colon A and u_n{\rightharpoonup} u_\infty = : u weakly in L^2(Q) . Note from Statement 2 that u solves

    -\nabla\cdot {\mathcal{A}}\nabla u +\lambda u = f \mathit{\text{ in }} Q.

    In other words, Diagram 3.6 commutes for linear equations.

    We now sketch the homogenization of Eq (1.1) on a geometry that is not minimally smooth but allows for homogenization. This sample geometry was introduced in [9]: We start from a Poisson point process and erase all points that are closer to each other than a given distance threshold s > 0 . We then construct the Delaunay triangulation and assign a pipe of random diameter 0 < D < s/2 with a distribution {\mathbb{P}}(D < x) < \exp(-1/x) to every edge, for all x > 0 . Furthermore, we enrich the geometry by the balls of radius s/4 around the remaining points. This union of pipes and balls is stationary and ergodic and we then define Q^{\varepsilon} as the intersection of Q with the system of balls and pipes scaled by a factor {\varepsilon} . In three dimensions, the complement of Q^{\varepsilon} is pathwise connected and unbounded, but in two dimensions, the complement consists of bounded sets, where there is no upper bound on the diameter of these sets. Also, there is no upper bound on the local Lipschitz constant.

    Writing W^{1, 2}_{(0), \partial Q}(Q^{\varepsilon}) for the functions in W^{1, 2} with value zero on \partial Q then, in [9, Theorem 1.15], it is shown that, for every 1 < q < 2 , there almost-surely exists an extension operator {\mathcal{U}}^{\varepsilon}\colon W^{1, 2}_{(0), \partial Q}(Q^{\varepsilon})\to W^{1, q}_0(\mathbb{B}_{{{\varepsilon}^\beta}}{\left(Q\right)}) that is continuous with \beta\in(0, 1) depending only on the random geometry, such that, for every {\varepsilon} > 0 and u^{\varepsilon}\in W^{1, 2}_{(0), \partial Q}(Q^{\varepsilon}) , it holds that

    \begin{equation} \left\Vert {\nabla {\mathcal{U}}^{\varepsilon} u^{\varepsilon}}\right\Vert_{L^q({\mathbb{R}}^d)}\leq C\left\Vert {\nabla u^{\varepsilon}}\right\Vert_{L^2(Q^{\varepsilon})}\qquad\text{ and }\qquad \left\Vert { {\mathcal{U}}^{\varepsilon} u^{\varepsilon}}\right\Vert_{L^q({\mathbb{R}}^d)}\leq C\left\Vert {u^{\varepsilon}}\right\Vert_{L^2(Q^{\varepsilon})}. \end{equation} (3.8)

    An inspection of the proof of [9, Theorem 1.7] reveals that it also holds for q = p = d = 2 therein, that is, there exists a constant C > 0 , independent from {\varepsilon} , such that

    \begin{equation} \left\Vert {{\mathcal{T}}^{\varepsilon} u^{\varepsilon}}\right\Vert_{L^2(\partial Q^{\varepsilon})}\leq C\left\Vert {\nabla u^{\varepsilon}}\right\Vert_{L^2(Q^{\varepsilon})}. \end{equation} (3.9)

    The important insight is that {\mathcal{U}}^{\varepsilon} u^{\varepsilon} is bounded in W^{1, q}_0(\mathbb{B}_{{1}}{\left(Q\right)}) and hence, pre-compact in L^2({\mathbb{R}}^2) , which in turn yields two-scale pre-compactness of the traces {\mathcal{T}}^{\varepsilon} u^{\varepsilon} , in a similar way as below. Thus, we can argue as in our proof of Lemma 10 to obtain the respective homogenization result. The needed two-scale convergence methods have been introduced in [8].

    On the other hand, in {\mathbb{R}}^2 , we can regularize the geometry by filling up holes with a diameter larger than a prescribed threshold and we can additionally prescribe a minimal thickness for the pipes. Since the generating points of the Delaunay triangulation have a minimal mutual distance, the filling of large holes provides minimal smoothness. However, Lemma 10 guarantees homogenization. Together with the proof of Theorem 17, we again find that Diagram 3.6 commutes.

    In the previous example, we explicitly constructed the approximation in a way that Q_n^{\varepsilon}\supset Q^{\varepsilon} . However, this is not strictly necessary. The following result shows that the homogenization limits would also exist and look the same as above, as long as the distributions of several geometric quantities do not worsen (significantly), with the respective conditions to be taken from [9].

    Theorem 22. Let G_n be jointly stationary sets with G where G_n\to G pointwise as n\to\infty . Assume further that there exists C > 0 , independent of {\varepsilon} and n , such that Eq (3.8) holds for every {\varepsilon} and n . Then, the assertion in Theorem 21 still holds.

    Without going further into detail, we mention that our main results in form of Lemma 10 and Theorem 56 can be reproduced for any family of geometries satisfying Eqs (3.8) and (3.9) using Theorem 22.

    It is currently not clear whether the Boolean model introduced above has a family of extension operators such as in the pipe example. Furthermore, if such a family does exist, then the diagram commutes. Hence, we propose the approximation method to derive an educated guess for a homogenized model.

    We first establish some properties of {\mathtt{F}_{n}}\colon{\mathcal{S}}({\mathbb{R}}^{d})\to{\mathcal{S}}({\mathbb{R}}^{d}) , most importantly the minimal smoothness of {\boldsymbol{\boxminus}}\mathbb{x}^{(n)} .

    Definition 23 (Minimal smoothness [25]). An open set P\subset{\mathbb{R}}^{d} is called minimally smooth with constants (\delta, N, M) if we may cover {\partial} P by a countable sequence of open sets (U_{i})_{i} such that

    i) \forall x\in{\mathbb{R}}^{d}\colon \#\{U_{i}\, \vert\, x\in U_{i}\}\leq N .

    ii) \forall x\in{\partial} P\, \exists U_{i}\colon \mathbb{B}_{{\delta}}{\left(x\right)}\subset U_{i} .

    iii) For every i , {\partial} P\cap U_{i} agrees (in some Cartesian system of coordinates) with the graph of a Lipschitz function whose Lipschitz semi-norm is at most M .

    Lemma 24 (Uniform \delta on individual clusters). Let \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) be an admissible point cloud. Then, for every x\in\mathbb{x}

    \delta{\left({\Xi{\left({{\mathcal{C}}_{\mathbb{x}}(x)}\right)}}\right)} > 0.

    Proof. Let \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) and assume \delta{\left({\Xi{\left({{\mathcal{C}}_{\mathbb{x}}(x)}\right)}}\right)} = 0 for some x\in\mathbb{x} . Then, there must be some p\in{\partial}\Xi_{\mathbb{x}} with \delta(p) = 0 . This together with bounded finiteness gives x_{p}, y_{p}\in\mathbb{x} such that p\in B_{r}(x_{p})\cap B_{r}(y_{p}) , in particular |x_{p}-y_{p}| = 2r . This contradicts the equidistance property of \mathbb{x} .

    The thinning maps {\mathtt{F}_{n}} have been constructed just to yield the following theorem.

    Theorem 25 (Minimal smoothness of thinned point clouds). For every \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) , both (\Xi\mathbb{x}^{(n)})^{\complement} and ({\boldsymbol{\boxminus}}\mathbb{x}^{(n)})^{\complement} are minimally smooth with \delta = 1/n , M = \sqrt{2nr} . Furthermore, every connected component of \Xi\mathbb{x}^{(n)} or {\boldsymbol{\boxminus}}\mathbb{x}^{(n)} has diameter less than 2nr .

    Proof. It remains to verify the estimate on M . Let x = o = 0_{{\mathbb{R}}^{d}} and y = \left(2r-n^{-1}, 0, \dots, 0\right) . Then the Lipschitz constant at the intersection of the two balls \mathbb{B}_{r}{\left(x\right)} and \mathbb{B}_{r}{\left(y\right)} is less than \sqrt{2nr} .

    Theorem 26 (Further properties of {\mathtt{F}_{n}} ). The set {\mathcal{S}}_{{\mathcal{A}}}({{{\mathbb{R}}}^{d}}) of admissible point clouds is measurable in the vague \sigma -algebra. Given n\in{\mathbb{N}} , it holds that {\mathtt{F}_{n}}\colon{\mathcal{S}}({{{\mathbb{R}}}^{d}})\to{\mathcal{S}}({{{\mathbb{R}}}^{d}}) is measurable, {\mathcal{S}}^{(n)}: = {\mathtt{F}_{n}}{\mathcal{S}}({{{\mathbb{R}}}^{d}}) is compact in the vague topology and the following three properties of \mathbb{x}\in{\mathcal{S}}({{{\mathbb{R}}}^{d}}) are equivalent:

    i) {\mathtt{F}_{n}}\mathbb{x} = \mathbb{x} ,

    ii) \mathbb{x}\in{\mathcal{S}}^{(n)} ,

    iii) Eqs (4.1) and (4.2) hold:

    \begin{align} \forall x, y\in\mathbb{x}, x\neq y\colon \quad & d(x, y)\notin(0, 1/n)\cup(2r-1/n, \, 2r+1/n), \end{align} (4.1)
    \begin{align} \forall x\in\mathbb{x}\colon \quad & \#\mathcal{C}_{\mathbb{x}}(x)\leq n, \, \delta{\left({\mathbb{B}_{r}{\left({{\mathcal{C}}_{\mathbb{x}}(x)}\right)}}\right)}\geq 1/n. \end{align} (4.2)

    Proof. {\mathtt{F}_{n}}\mathbb{x} = \mathbb{x} implies \mathbb{x}\in{\mathcal{S}}^{(n)} since {\mathtt{F}_{n}}\mathbb{x}\in{\mathcal{S}}^{(n)} , and vice versa, \mathbb{x}\in{\mathcal{S}}^{(n)} implies {\mathtt{F}_{n}}\mathbb{x} = \mathbb{x} by definition of {\mathtt{F}_{n}} . By construction of {\mathtt{F}_{n}} it follows that Eqs (4.1) and (4.2) hold if, and only if, \mathbb{x}\in{\mathcal{S}}^{(n)} .

    Consider the space of (non-simple) counting measures {\mathcal{N}}({\mathbb{R}}^{d})\subset{\mathcal{M}}({\mathbb{R}}^{d}) , i.e.,

    {\mathcal{N}}({\mathbb{R}}^{d}): = \Big\{\mu\in{\mathcal{M}}({\mathbb{R}}^{d})\colon \mu = \sum\limits_{k\in\mathcal{I}\subset{\mathbb{N}}}a_{k}\delta_{x_{k}}\text{ such that }a_{k}\in{\mathbb{N}}\text{ and }x_{k}\in{\mathbb{R}}^{d}\Big\}.

    We see, e.g., in [2], that

    {\mathcal{S}}({\mathbb{R}}^{d}) and {\mathcal{N}}({\mathbb{R}}^{d}) are both measurable w.r.t. the Borel- \sigma -algebra of {\mathcal{M}}({\mathbb{R}}^{d}) .

    {\mathcal{S}}({\mathbb{R}}^{d})\subset{\mathcal{N}}({\mathbb{R}}^{d}) and {\mathcal{N}}({\mathbb{R}}^{d}) is closed in {\mathcal{M}}({\mathbb{R}}^{d}) . In particular, {\mathcal{N}}({\mathbb{R}}^{d}) is also complete under the Prokhorov metric.

    Now {\mathcal{S}}^{(n)} is pre-compact because of the characterization of pre-compact sets in the vague topology: For every bounded open A\subset{{{\mathbb{R}}}^{d}} , it holds that \sup_{\mathbb{x}\in{\mathcal{S}}^{(n)}}\mathbb{x}(A)\leq C\left(\mathrm{diam}\, A\right)^{d} with C depending only on n . It remains to show that {\mathcal{S}}^{(n)} is closed as a subset of {\mathcal{N}}({\mathbb{R}}^{d}) . Let \left(\mathbb{x}_{j}\right)_{j\in{\mathbb{N}}}\subset{\mathcal{S}}^{(n)} be a converging sequence with limit \mathbb{x}\in{\mathcal{N}}({{{\mathbb{R}}}^{d}}) . One checks that (4.1) (namely d(x, y)\notin\big(0, 1/n\big) ) ensures \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) , e.g., in a procedure similar to the proof of [2, Lemma 9.1.V]. We observe that for every x, y\in\mathbb{x} , there exist x_{j}, y_{j}\in\mathbb{x}_{j} such that x_{j}\to x , y_{j}\to y as j\to\infty . This implies by a limit in Eq (4.1) that \mathbb{x} still satisfies Eq (4.1).

    For x\in\mathbb{x} , one checks that Eq (4.1) (namely d(x, y)\notin\big(2r-1/n, 2r+1/n\big) ) implies \#{\mathcal{C}}_{\mathbb{x}}(x)\leq n .

    Let p\in\partial\Xi(\mathbb{x}) and let \{ x^{(1)}, \dots, x^{(K)}\} = \mathbb{B}_{{10r}}{\left(p\right)}\cap\mathbb{x} with sequences x_{j}^{(k)}\to x^{(k)} , x_{j}^{(k)}\in\mathbb{x}_{j} . Given \eta > 0 , let J\in{\mathbb{N}} such that for all j > J and k = 1, \dots, K it holds |x^{(k)}-x_{j}^{(k)}| < \eta . Then there exists p_{j}\in\partial\Xi(\mathbb{x}_{j}) such that |p_{j}-p| < \eta and \partial\Xi(\mathbb{x}_{j}) is a Lipschitz graph in the ball \mathbb{B}_{{2\delta}}{\left({p_{j}}\right)} for every \delta < 1/n . Hence \partial\Xi(\mathbb{x}_{j}) is a Lipschitz graph in the ball \mathbb{B}_{{2\delta-\eta}}{\left(p\right)} . Because the Lipschitz regularity of \partial\Xi(\mathbb{x}_{j}) changes continuously under slight shifts of the balls, there exists \eta_{0} such that for \eta < \eta_{0} and \partial\Xi(\mathbb{x}) is a Lipschitz graph in \mathbb{B}_{{2\delta-2\eta}}{\left(p\right)} . Since \eta is arbitrary, we find \partial\Xi(\mathbb{x}) is Lipschitz graph in \mathbb{B}_{{2\delta}}{\left(p\right)} for every \delta < 1/n , implying \delta(p)\geq1/n . Since this holds for every p , we conclude Eq (4.2) and {\mathcal{S}}^{(n)} is compact.

    To see that {\mathcal{S}}_{{\mathcal{A}}}({{{\mathbb{R}}}^{d}}) is measurable, consider for \mathbb{x}\in{\mathcal{S}}({{{\mathbb{R}}}^{d}})

    \begin{align*} \mathbb{x}_{1, m, R} & : = \{ x\in\mathbb{x}\colon x\notin\mathbb{B}_{R}{\left(o\right)}\text{ or }\, d(x, y)\notin(0, 1/m)\cup(2r-1/m, \, 2r+1/m)\ \forall y\in\mathbb{x}\} , \\ \mathbb{x}_{2, l, R} & : = \{ x\in\mathbb{x}\, \vert\, x\notin\overline{\mathbb{B}_{R}{\left(o\right)}}\text{ or }\#\mathcal{C}_{\mathbb{x}}(x)\leq l, \, \delta{\left({\mathbb{B}_{r}{\left({{\mathcal{C}}_{\mathbb{x}}(x)}\right)}}\right)} > 1/l\} , \end{align*}

    and define

    \begin{align*} {\mathtt{F}}_{1, m, R}\mathbb{x} & : = \mathbb{x}_{1, m, R} & {\mathcal{S}}^{(1, m, R)} & : = {\mathtt{F}}_{1, m, R}{\mathcal{S}}({\mathbb{R}}^{d})\\ {\mathtt{F}}_{2, l, R}\mathbb{x} & : = \mathbb{x}_{2, l, R}. & {\mathcal{S}}^{(2, m, R)} & : = {\mathtt{F}}_{2, m, R}{\mathcal{S}}({\mathbb{R}}^{d}). \end{align*}

    We check that {\mathcal{S}}^{(1, m, R)} is a closed subset inside {\mathcal{S}}({\mathbb{R}}^{d}) (repeat the arguments above), i.e., \overline{{\mathcal{S}}^{(1, m, R)}}\cap{\mathcal{S}}({\mathbb{R}}^{d}) = {\mathcal{S}}^{(1, m, R)} . In particular, {\mathcal{S}}^{(1, m, R)} is measurable w.r.t. the vague topology of {\mathcal{M}}({\mathbb{R}}^{d}) . Similarly, one shows that {\mathcal{S}}({\mathbb{R}}^{d})\backslash{\mathcal{S}}^{(2, m, R)} is closed as a subset inside {\mathcal{S}}({\mathbb{R}}^{d}) . Again, this shows that {\mathcal{S}}^{(2, m, R)} is measurable. Consider now the measurable sets

    {\mathcal{S}}^{(1, \infty, \infty)}: = \bigcap\limits_{R\in{\mathbb{N}}}\bigcup\limits_{m\in{\mathbb{N}}}{\mathcal{S}}^{(1, m, R)}\qquad\text{and}\qquad{\mathcal{S}}^{(2, \infty, \infty)}: = \bigcap\limits_{R\in{\mathbb{N}}}\bigcup\limits_{m\in{\mathbb{N}}}{\mathcal{S}}^{(2, m, R)}.

    We see that

    i) \mathbb{x}\in{\mathcal{S}}^{(1, \infty, \infty)} if, and only if, for all x, y\in\mathbb{x} , it holds d(x, y)\neq 2r .

    ii) \mathbb{x}\in{\mathcal{S}}^{(2, \infty, \infty)} if, and only if, for every x\in\mathbb{x} , it holds that \#\mathcal{C}_{\mathbb{x}}(x) < \infty and \delta{\left({\mathbb{B}_{r}{\left({{\mathcal{C}}_{\mathbb{x}}(x)}\right)}}\right)} > 0 .

    Therefore,

    {\mathcal{S}}_{{\mathcal{A}}}({{{\mathbb{R}}}^{d}}) = {\mathcal{S}}^{(1, \infty, \infty)}\cap{\mathcal{S}}^{(2, \infty, \infty)},

    is measurable.

    To see that {\mathtt{F}_{n}}\colon {\mathcal{S}}({{{\mathbb{R}}}^{d}})\to{\mathcal{S}}({{{\mathbb{R}}}^{d}}) is measurable, recall {\mathtt{F}_{n}} = {\mathtt{F}}_{2, n}\circ{\mathtt{F}}_{1, n} from Definition 3. It, therefore, suffices to show that the following maps are measurable:

    {\mathtt{F}}_{1, n}\colon {\mathcal{S}}({\mathbb{R}}^{d})\to{\mathtt{F}}_{1, n}{\mathcal{S}}({\mathbb{R}}^{d})\subset{\mathcal{S}}({\mathbb{R}}^{d})\qquad\text{and}\qquad{\mathtt{F}}_{2, n}\colon {\mathtt{F}}_{1, n}{\mathcal{S}}({\mathbb{R}}^{d})\to{\mathcal{S}}({\mathbb{R}}^{d}).

    For f\in C_{c}({{{\mathbb{R}}}^{d}}) , consider the evaluation by f , i.e.,

    M_{f}\colon{\mathcal{S}}({{{\mathbb{R}}}^{d}})\to{\mathbb{R}}, \qquad\mathbb{x}\mapsto\int_{{{{\mathbb{R}}}^{d}}}f\, {\mathrm{d}}\mathbb{x}.

    If f\geq0 , we observe the upper semi-continuity of

    M_{f}\circ{\mathtt{F}}_{1, n}\colon {\mathcal{S}}({\mathbb{R}}^{d})\to{\mathbb{R}}\qquad\text{and}\qquad M_{f}\circ{\mathtt{F}}_{2, n}\colon {\mathtt{F}}_{1, n}{\mathcal{S}}({\mathbb{R}}^{d})\to{\mathbb{R}}.

    We have lower semi-continuity for f\leq0 since M_{-f} = -M_{f} . Therefore M_{f}\circ{\mathtt{F}}_{i, n} with i\in\{1, 2\} is measurable in the cases f\geq0 and f\leq0 and hence, in general. Since the vague topology is generated by \big(M_{f}\big)_{f\in C_{c}({\mathbb{R}}^{d})} , we conclude that {\mathtt{F}}_{1, n} and {\mathtt{F}}_{2, n} are measurable.

    Remark 27 (Fine details of Theorem 26).

    ● For {\mathcal{S}}^{(n)}: = {\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) , we have that

    \bigcup\limits_{n\in{\mathbb{N}}}{\mathcal{S}}^{(n)}\subsetneq{\mathcal{S}}_{{\mathcal{A}}}({\mathbb{R}}^{d})\subsetneq\big\{\mathbb{x}\colon\lim\limits_{n\to\infty}{\mathtt{F}_{n}}\mathbb{x} = \mathbb{x}\big\}\subsetneq{\mathcal{S}}({\mathbb{R}}^{d})\subsetneq\overline{\bigcup\limits_{n\in{\mathbb{N}}}{\mathcal{S}}^{(n)}} = {\mathcal{N}}({\mathbb{R}}^{d}).

    M_{f}\circ{\mathtt{F}}_{2, n} is not upper semi-continuous on {\mathcal{S}}({\mathbb{R}}^{d}) (in contrast to {\mathtt{F}}_{1, n}{\mathcal{S}}({\mathbb{R}}^{d}) ): The condition that d(x, y)\notin(2r-1/n, \, 2r+1/n)\ \forall x, y\in{\mathtt{F}}_{1, n}\mathbb{x} is crucial to ensure that clusters do not change sizes.

    Definition 28. We define the events that the origin is not covered by the filled-up Boolean model, i.e.,

    {\mathbf{G}}: = \{ \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d})\, \vert\, o\notin{\boldsymbol{\boxminus}}\mathbb{x}\} \qquad{\text{and}}\qquad{\mathbf{G}}_{n}: = {\mathtt{F}_{n}}{}^{-1}({\mathbf{G}}) = \{ \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d})\, \vert\, o\notin{\boldsymbol{\boxminus}}\mathbb{x}^{(n)}\} .

    This gives us, for x\in{\mathbb{R}}^{d} , that

    {\mathbb{1}}_{{\boldsymbol{\boxminus}}\mathbb{x}^{\complement}}(x) = {\mathbb{1}}_{{\mathbf{G}}}(\tau_{x}\mathbb{x}).

    We will later consider the effective conductivities based on these events.

    Lemma 29 (Approximation properties). Let \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d}) be an admissible point cloud.

    i) For every bounded domain \Lambda , there exists an N(\mathbb{x}, \Lambda)\in{\mathbb{N}} such that for every n\geq N(\mathbb{x}, \Lambda)

    \mathbb{x}^{(n)}\cap\Lambda = \mathbb{x}\cap\Lambda, \qquad\mathit{\text{in particular}}\qquad\mathbb{x} = \bigcup\limits_{n\in{\mathbb{N}}}\mathbb{x}^{(n)}.

    ii) For every bounded domain \Lambda , there exists an \tilde{N}(\mathbb{x}, \Lambda)\in{\mathbb{N}} such that for every n\geq\tilde{N}(\mathbb{x}, \Lambda)

    {\boldsymbol{\boxminus}}\mathbb{x}{}^{(n)}\cap\Lambda = {\boldsymbol{\boxminus}}\mathbb{x}\cap\Lambda, \qquad\mathit{\text{in particular}}\qquad{\boldsymbol{\boxminus}}\mathbb{x} = \bigcup\limits_{n\in{\mathbb{N}}}{\boldsymbol{\boxminus}}\mathbb{x}{}^{(n)}.

    iii) There exists an N = N(\mathbb{x})\in{\mathbb{N}} such that for every n\geq N :

    o\notin{\boldsymbol{\boxminus}}\mathbb{x}^{(n)}\iff o\notin{\boldsymbol{\boxminus}}\mathbb{x}.

    In particular, \bigcap_{n\in{\mathbb{N}}}{\mathbf{G}}_{n}\backslash{\mathbf{G}} only consists of non-admissible point clouds.

    Proof. i) Boundedness of \Lambda implies that there are only finitely many mutually disjoint clusters {\mathcal{C}}_{\mathbb{x}}(x_{i}) , i = 1, \dots, N_{{\mathcal{C}}} that intersect with \Lambda . Furthermore, because \#\left(\mathbb{x}\cap\mathbb{B}_{r}{\left({\Lambda}\right)}\right) < \infty and because of Property 1 of admissible point clouds, we know

    \min\left\{ \left|\left|x-y\right|-2r\right|\colon x, y\in\mathbb{x}\cap\mathbb{B}_{r}{\left({\Lambda}\right)}, \, x\neq y\right\} > 0

    and Lemma 24 yields

    \min\left\{ \delta{\left({\Xi{\mathcal{C}}_{\mathbb{x}}(x_{i})}\right)}\colon{\mathcal{C}}_{\mathbb{x}}(x_{i})\cap\Lambda\neq\emptyset\right\} > 0.

    This implies the first statement.

    ii) By making \Lambda larger, we may assume \Lambda = [-k, k]^{d} for some k\in{\mathbb{N}} . For n\geq N(\mathbb{B}_{r}{\left({[-k, k]^{d}}\right)}) ,

    [-k, k]^{d}\backslash\Xi\mathbb{x}^{(n)} = [-k, k]^{d}\backslash\Xi\mathbb{x}.

    [-k, k]^{d}\backslash\Xi\mathbb{x} only has finitely many connected components \mathcal{C}_{i} . Take one of these connected components \mathcal{C}_{i} and suppose it lies in {\boldsymbol{\boxminus}}\mathbb{x} . Then, it has to be encircled by finitely many balls \mathbb{B}_{r}{\left(x\right)} in \Xi\mathbb{x} . Let n_{i} be large enough such that all these x lie in \mathbb{x}^{(n_{i})} . Then, \mathcal{C}_{i}\subset{\boldsymbol{\boxminus}}\mathbb{x}^{(n_{i})} . We may do so for every \mathcal{C}_{i} . Take

    \tilde{N}(\mathbb{x}, \Lambda): = \max\big\{ n_{i}, \, N(\mathbb{B}_{r}{\left({[-k, k]^{d}}\right)})\big\}.

    For every n\geq\tilde{N}(\mathbb{x}, \Lambda) , the connected components \mathcal{C}_{i} of [-k, k]^{d}\backslash\Xi\mathbb{x}^{(n)} and [-k, k]^{d}\backslash\Xi\mathbb{x} are identical since [-k, k]^{d}\backslash\Xi\mathbb{x}^{(n)} = [-k, k]^{d}\backslash\Xi\mathbb{x} . Therefore, we get the claim

    \begin{align*} [-k, k]^{d}\backslash{\boldsymbol{\boxminus}}\mathbb{x}^{(n)} & = \Big([-k, k]^{d}\backslash\Xi\mathbb{x}^{(n)}\Big)\backslash\bigcup\limits_{C_{i}\subset{\boldsymbol{\boxminus}}\mathbb{x}^{(n)}}\mathcal{C}_{i}\\ & = \Big([-k, k]^{d}\backslash\Xi\mathbb{x}\Big)\backslash\bigcup\limits_{C_{i}\subset{\boldsymbol{\boxminus}}\mathbb{x}}\mathcal{C}_{i}\quad = [-k, k]^{d}\backslash{\boldsymbol{\boxminus}}\mathbb{x}. \end{align*}

    iii) This is a direct consequence of Point 2. If \mathbb{x}\in\bigcap_{n\in{\mathbb{N}}}{\mathbf{G}}_{n}\backslash{\mathbf{G}} , then o\notin{\boldsymbol{\boxminus}}\mathbb{x}^{(n)} for every n but o\in{\boldsymbol{\boxminus}}\mathbb{x} . Therefore, \mathbb{x} cannot be admissible by Point 2.

    Definition 30 (Surface measure of {\boldsymbol{\boxminus}}\mathbb{x} ). We define the surface measure for \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d})

    \mu_{\mathbb{x}}(A): = {\mathcal{H}}_{\llcorner{\partial}{\boldsymbol{\boxminus}}\mathbb{x}}^{d-1}(A) = {\mathcal{H}}^{d-1}(A\cap{\partial}{\boldsymbol{\boxminus}}\mathbb{x}).

    Note that \mu_{\mathbb{x}}{\left({[0, 1]^{d}}\right)}\leq{\mathcal{H}}^{d-1}{\left({\mathbb{B}_{r}{\left(o\right)}}\right)}\cdot\mathbb{x}{\left({\mathbb{B}_{r}{\left({[0, 1]^{d}}\right)}}\right)} .

    Definition 31 (Intensity of random measure). Given a stationary random measure \tilde{\mu} , we define its intensity

    \lambda(\tilde{\mu}): = {\mathbb{E}}\big[\tilde{\mu}\big([0, 1)^{d}\big)\big].

    We define the intensity of point processes by identifying them as random measures.

    Lemma 32 (Convergence of intensities). Let {\mathbb{X}} be an admissible stationary point process with finite intensity \lambda({\mathbb{X}}) . Then,

    \lim\limits_{n\to\infty}\lambda({\mathbb{X}}^{(n)}) = \lambda({\mathbb{X}})\qquad\mathit{\text{and}}\qquad\lim\limits_{n\to\infty}\lambda(\mu_{{\mathbb{X}}^{(n)}}) = \lambda(\mu_{{\mathbb{X}}}).

    Proof. "Almost surely" is to be understood w.r.t. the distribution {\mathbb{P}} of {\mathbb{X}} .

    i) By Lemma 29, we have almost surely {\mathbb{X}}^{(n)}{\left({[0, 1]^{d}}\right)}\to{\mathbb{X}}{\left({[0, 1]^{d}}\right)} as n\to\infty . Dominated convergence with majorant {\mathbb{X}}{\left({[0, 1]^{d}}\right)} yields

    \lambda({\mathbb{X}}^{(n)}) = {\mathbb{E}}[\mathbb{x}^{(n)}{\left({[0, 1]^{d}}\right)}]\to{\mathbb{E}}[\mathbb{x}{\left({[0, 1]^{d}}\right)}] = \lambda({\mathbb{X}}).

    ii) Again, by Lemma 29, we have almost surely {\boldsymbol{\boxminus}}{\mathbb{X}}^{(n)}\cap[0, 1]^{d}\to{\boldsymbol{\boxminus}}{\mathbb{X}}\cap[0, 1]^{d} , in particular

    \mu_{{\mathbb{X}}^{(n)}}{\left({[0, 1]^{d}}\right)} = {\mathcal{H}}_{\llcorner{\partial}{\boldsymbol{\boxminus}}{\mathbb{X}}^{(n)}}^{d-1}{\left({[0, 1]^{d}}\right)}\to{\mathcal{H}}_{\llcorner{\partial}{\boldsymbol{\boxminus}}{\mathbb{X}}}^{d-1}{\left({[0, 1]^{d}}\right)} = \mu_{{\mathbb{X}}}{\left({[0, 1]^{d}}\right)}.

    Dominated convergence yields, again, convergence of intensities.

    Remark 33 (Local convergence). The convergence in Lemma 29 is much stronger than what is actually needed to prove the convergence of intensities. Indeed, we could prove convergence even for so-called tame and local functions f\colon {\mathcal{M}}({\mathbb{R}}^{d})\to{\mathbb{R}} , among which the intensity \lambda is just one special case f(\mathbb{x}): = \mathbb{x}{\left({[0, 1]^{d}}\right)} .

    The structure ({\mathcal{M}}({\mathbb{R}}^{d}), {\mathcal{B}}({\mathcal{M}}({\mathbb{R}}^{d})), {\mathbb{P}}, \tau) as in Definition 8 is a dynamical system:

    Definition 34 (Dynamical system, stationarity, ergodicity). Let (\Omega, {\mathcal{F}}, {\mathcal{P}}) be a separable metric probability space. A dynamical system \tau = (\tau_{x})_{x\in{\mathbb{R}}^{d}} is a family of measurable mappings \tau_{x}\colon \Omega\to\Omega satisfying

    ● Group property: \tau_{0} = \mathit{\text{id}}_{\Omega} and \tau_{x+y} = \tau_{x}\circ\tau_{y} for any x, \, y\in\mathbb{R}^{d} .

    ● Measure preserving: For any x\in\mathbb{R}^{d} and any F\in\mathcal{F} , we have {\mathcal{P}}\left(\tau_{x}(F)\right) = {\mathcal{P}}\left(F\right) .

    ● Continuity: The map \mathcal{T}\colon \Omega\times\mathbb{R}^{d}\to\Omega , \left(\omega, \, x\right)\mapsto\tau_{x}(\omega) is continuous w.r.t. the product topology on \Omega\times{\mathbb{R}}^{d} .

    \tau is called ergodic if the \sigma -algebra of \tau -invariant sets is trivial under {\mathcal{P}} .

    Our practical setting will always be some \Omega\subset{\mathcal{M}}({\mathbb{R}}^{d}) , but we will still work with abstract dynamical systems in Sections 5 and 6. In this context, let us mention that continuity of \tau was not needed in the first place [22], but turned out to be very useful in the proofs concerning two-scale convergence [30]. Since we frequently use results from [30] and since we get continuity of \tau for free in our applied setting, we will simply rely on this property.

    Let (\Omega, {\mathcal{B}}(\Omega), {\mathcal{P}}, \tau) be a dynamical system. We write L^{2}(\Omega): = L^{2}(\Omega, \, {\mathcal{P}}) . The dynamical system \tau introduces a strongly continuous group action on L^{2}(\Omega)\to L^{2}(\Omega) through T_{x}f(\omega): = f(\tau_{x}\omega) with the d independent generators

    {\mathrm{D}}_{i}f: = \lim\limits_{t\to0}\frac{1}{t}\left(f-f(\tau_{te_{i}}\bullet)\right),

    with domain \mathcal{D}_{i} where \left(e_{i}\right)_{i = 1, \dots, d}\subset{\mathbb{R}}^{d} is the canonical Euclidean basis. Introducing

    H^{1}(\Omega): = \bigcap\limits_{i = 1}^{d}\mathcal{D}_{i}\subset L^{2}(\Omega),

    and the gradient \nabla_{\omega}f: = \left({\mathrm{D}}_{1}f, \dots, {\mathrm{D}}_{d}f\right)^{\top} , we can define the space of potential vector fields

    \mathcal{V}_{{\mathrm{pot}}}^{2}(\Omega): = \big\{\nabla_{\tilde{\omega}}f\, \vert\, f\in H^{1}(\Omega)\text{ and }\int_{\Omega}\nabla_{\tilde{\omega}}f\text{d}{\mathcal{P}}(\omega) = 0_{{\mathbb{R}}^{d}}\big\}.

    Defining L_{{\mathrm{sol}}}^{2}(\Omega): = {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega)^{\bot} , we find with u_{\omega}(x): = u(\tau_{x}\omega) that

    \begin{align} L_{{\mathrm{pot}}}^{2}(\Omega) & : = \left\{ u\in L^{2}(\Omega;{{{\mathbb{R}}}^{d}})\colon u_{\omega}\in L_{{\mathrm{pot}}, {\mathrm{loc}}}^{2}({{{\mathbb{R}}}^{d}})\, \, \text{for }{\mathcal{P}}-\text{a.e. }\omega\in\Omega\right\} , \\ L_{{\mathrm{sol}}}^{2}(\Omega) & = \left\{ u\in L^{2}(\Omega;{{{\mathbb{R}}}^{d}})\colon u_{\omega}\in L_{{\mathrm{sol}}, {\mathrm{loc}}}^{2}({{{\mathbb{R}}}^{d}})\, \, \text{for }{\mathcal{P}}-\text{a.e. }\omega\in\Omega\right\} , \\ {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega) & = \Big\{ u\in L_{{\mathrm{pot}}}^{2}(\Omega)\colon\int_{\Omega}u\, {\mathrm{d}}{\mathcal{P}} = 0\Big\}, \end{align} (5.1)

    because \Omega is a separable metric [8] where

    \begin{align*} L_{{\mathrm{pot}}, {\mathrm{loc}}}^{2}({{{\mathbb{R}}}^{d}}) & : = \left\{ u\in L_{{\mathrm{loc}}}^{2}({{{\mathbb{R}}}^{d}};{{{\mathbb{R}}}^{d}})\colon\forall U\, \, \text{bounded domain }\exists\varphi\in W^{1, 2}(U)\colon u = \nabla\varphi\right\} , \\ L_{{\mathrm{sol}}, {\mathrm{loc}}}^{2}({{{\mathbb{R}}}^{d}}) & : = \Big\{ u\in L_{{\mathrm{loc}}}^{2}({{{\mathbb{R}}}^{d}};{{{\mathbb{R}}}^{d}})\colon\int_{{{{\mathbb{R}}}^{d}}}u\cdot\nabla\varphi\text{d} x = 0\, \, \forall\varphi\in C_{c}^{1}({{{\mathbb{R}}}^{d}})\Big\}. \end{align*}

    For A\subset\Omega measurable, we define

    {\mathcal{V}}_{{\mathrm{pot}}}^{2}(A\vert\Omega): = cl_{L^{2}(A)^{d}}\left\{ v_{\vert_{A}}\colon v\in\mathcal{V}_{{\mathrm{pot}}}^{2}(\Omega)\right\} .

    Definition 35 (Cell solutions). Let (\Omega, {\mathcal{B}}(\Omega), {\mathcal{P}}) be a separable metric probability space with dynamical system \tau and let {\mathcal{Q}}\in{\mathcal{B}}(\Omega) . We notice that for every unit vector e_{i} , the map

    v\mapsto-\int_{{\mathcal{Q}}}e_{i}\cdot v\text{ d}{\mathcal{P}}(\omega) = -{\big\langle} e_{i}, v{\big\rangle}_{L^{2}({\mathcal{Q}})^{d}},

    is a bounded linear functional on the Hilbert space {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) . Using the Riesz representation theorem, we obtain a unique w_{i}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) such that {\big\langle} w_{i}, v{\big\rangle}_{{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega)} = -{\big\langle} e_{i}, v{\big\rangle}_{L^{2}({\mathcal{Q}})^{d}} or equivalently

    \forall v\in\mathcal{V}_{{\mathrm{pot}}}^{2}(\Omega)\colon \qquad\int_{{\mathcal{Q}}}[w_{i}+e_{i}]\cdot v\text{ d}{\mathcal{P}}(\omega) = 0.

    w_{i} is called the i -th cell solution. The cell solutions satisfy

    {\|} w_{i}{\|}_{L^{2}({\mathcal{Q}})^{d}}\leq\sqrt{{\mathcal{P}}({\mathcal{Q}})}\leq1,

    and can be grouped in the matrix

    W_{{\mathcal{Q}}}: = (w_{1}, \dots, w_{d}).

    Remark 36. We observe that by definition, w_i is the minimizer in {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) of the functional

    \begin{equation} \mathcal{E}_i(w) = \int_{\mathcal{Q}}\frac12(w_i+e_i)^2\text{ d}{\mathcal{P}}. \end{equation} (5.2)

    Definition 37 (Effective conductivity {\mathcal{A}} ). Let w_{i} be the cell solution on {\mathcal{Q}}\in{\mathcal{B}}(\Omega) . The effective conductivity \mathcal{A} based on the event {\mathcal{Q}} is defined as

    \begin{align} {\mathcal{A}} & : = \int_{{\mathcal{Q}}}(I_{d}+W_{{\mathcal{Q}}})^{t}(I_{d}+W_{{\mathcal{Q}}})\text{ d}{\mathcal{P}}(\omega), \end{align} (5.3)

    with I_{d} being the identity matrix. We observe for the entries \left(\mathcal{A}_{i, j}\right)_{i, j = 1, \dots, d} of {\mathcal{A}} that

    \begin{align} \mathcal{A}_{i, j} & = \int_{{\mathcal{Q}}}\left[e_{i}+w_{i}(\omega)\right]\cdot\left[e_{j}+w_{j}(\omega)\right]\text{ d}{\mathcal{P}}(\omega) = \int_{{\mathcal{Q}}}\left[e_{i}+w_{i}(\omega)\right]\cdot e_{j}\text{ d}{\mathcal{P}}(\omega). \end{align} (5.4)

    We write \alpha_{{\mathcal{A}}}\geq0 for its smallest eigenvalue.

    Lemma 38 (Convergence of cell solutions). Let \left({\mathcal{Q}}_{n}\right)_{n\in{\mathbb{N}}}\subset{\mathcal{B}}(\Omega) and {\mathcal{Q}}\in{\mathcal{B}}(\Omega) such that {\mathbb{1}}_{{\mathcal{Q}}_{n}}\to{\mathbb{1}}_{{\mathcal{Q}}} \; {\mathcal{P}} -almost everywhere as n\to\infty and

    either {\mathcal{Q}}_n\supset{\mathcal{Q}} for every n , or

    there exist C > 0 , r\in(1, 2] independent from n and continuous operators {\mathcal{U}}_n\colon {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega)\to{\mathcal{V}}_{{\mathrm{pot}}}^{r}(\Omega) such that, for every n and every w\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}_n\vert\Omega) , it holds that \left\Vert {{\mathcal{U}}_n w}\right\Vert_{L^r(\Omega)}\leq C\left\Vert {w}\right\Vert_{{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}_n\vert\Omega)} .

    Then, the sequence of cell solutions w_{i}^{(n)} to the cell problem on {\mathcal{Q}}_{n} satisfies

    w_{i}^{(n)}{\rightharpoonup} w_{i}\qquad\mathit{\text{ in }}L^2(\Omega)^d\mathit{\text{ as }}n\to\infty,

    where w_{i}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) is the i -th cell solution on {\mathcal{Q}}

    \forall v\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega)\colon \qquad\int_{{\mathcal{Q}}}\left[w_{i}+e_{i}\right]\cdot v\text{ d}{\mathcal{P}}(\omega) = 0.

    Proof. We first check that the limit satisfies \int_{{\mathcal{Q}}}\left[w_{i}+e_{i}\right]\cdot v\text{d}{\mathcal{P}}(\omega) = 0 and then w_{i}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) .

    i) In both cases, the a priori estimate yields an L^{2} -weakly convergent subsequence of w_{i}^{(n)}\rightharpoonup w_{i}\in L^{2}(\Omega)^{d} after extending w_{i}^{(n)} to the whole of \Omega via 0. Let v\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega) . We have {\mathbb{1}}_{{\mathcal{Q}}_{n}}\to{\mathbb{1}}_{{\mathcal{Q}}} \; {\mathcal{P}} -almost everywhere, so dominated convergence yields

    \begin{align*} \lim\limits_{n\to\infty}\int_{{\mathcal{Q}}_{n}}e_{i}\cdot v\text{d}{\mathcal{P}}(\omega) & = \int_{{\mathcal{Q}}}e_{i}\cdot v\text{d}{\mathcal{P}}(\omega), \end{align*}

    while weak convergence yields

    \begin{align*} \lim\limits_{n\to\infty}\int_{{\mathcal{Q}}_{n}}w_{i}^{(n)}\cdot v\text{d}{\mathcal{P}}(\omega) & = \int_{\Omega}w_{i}\cdot v\text{d}{\mathcal{P}}(\omega). \end{align*}

    We also have

    {\mathbb{1}}_{{\mathcal{Q}}_{n}}w_{i}^{(n)} = w_{i}^{(n)}\mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(\Omega)^d}w_{i},

    which implies

    {\mathbb{1}}_{{\mathcal{Q}}}w_{i} = w_{i}.

    Therefore, with w_{i}\in L^{2}({\mathcal{Q}})^{d} :

    \begin{align*} 0 & = \lim\limits_{n\to\infty}\int_{{\mathcal{Q}}_{n}}[w_{i}^{(n)}+e_{i}]\cdot v\text{d}{\mathcal{P}}(\omega) = \int_{{\mathcal{Q}}}\left[w_{i}+e_{i}\right]\cdot v\text{d}{\mathcal{P}}(\omega). \end{align*}

    ii) In both cases of the lemma, the space {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega)\subset L^{2}({\mathcal{Q}})^{d} is closed and convex, so it is also weakly closed. We construct a weakly converging sequence in {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) that converges to w_{i} . Since w_{i}^{(n)}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}_{n}\vert\Omega) , we find v^{(n)}\in\mathcal{V}_{{\mathrm{pot}}}^{2}(\Omega) such that

    \|w_{i}^{(n)}-{\mathbb{1}}_{{\mathcal{Q}}_{n}}v^{(n)}\|_{L^{2}({\mathcal{Q}})^{d}}\leq\frac{1}{n}.

    Since w_{i}^{(n)}\rightharpoonup w_{i} , we get

    {\mathbb{1}}_{{\mathcal{Q}}_{n}}v^{(n)}\mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(\Omega)^{d}}w_{i}.

    Note that \left({\mathbb{1}}_{{\mathcal{Q}}_{n}}-{\mathbb{1}}_{{\mathcal{Q}}}\right)v^{(n)} is a bounded sequence that is weakly converging to 0 because for every \phi\in L^2(\Omega) we find by dominated convergence that \left({\mathbb{1}}_{{\mathcal{Q}}_{n}}-{\mathbb{1}}_{{\mathcal{Q}}}\right)\phi\to 0 strongly in L^2(\Omega) . Therefore, we also obtain

    {\mathbb{1}}_{{\mathcal{Q}}}v^{(n)}\mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(\Omega)^{d}}w_{i}.

    In the first case of the lemma, {\mathbb{1}}_{{\mathcal{Q}}}v^{(n)}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) , so we get that w_{i}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) . In the second case, we obtain {\mathbb{1}}_{{\mathcal{Q}}}v^{(n)}\in{\mathcal{V}}_{{\mathrm{pot}}}^{r}({\mathcal{Q}}\vert\Omega) and hence w_{i}\in{\mathcal{V}}_{{\mathrm{pot}}}^{r}({\mathcal{Q}}\vert\Omega) . Since being a L_{{\mathrm{pot}}}^{r} -function is distinguished from being a pure L^r -function only by the Condition (5.1), the integrability of w_i then yields w_{i}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega) .

    Corollary 39 (Convergence of effective conductivities). Let \left({\mathcal{Q}}_{n}\right)_{n\in{\mathbb{N}}}\subset{\mathcal{B}}(\Omega) and {\mathcal{Q}}\in{\mathcal{B}}(\Omega) such that {\mathbb{1}}_{{\mathcal{Q}}_{n}}\to{\mathbb{1}}_{{\mathcal{Q}}} \; {\mathcal{P}} -almost surely and

    either {\mathcal{Q}}_n\supset{\mathcal{Q}} for every n , or

    there exist C > 0 , r\in(1, 2] independent from n and continuous operators {\mathcal{U}}_n\colon {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}\vert\Omega)\to{\mathcal{V}}_{{\mathrm{pot}}}^{r}(\Omega) such that, for every n and every w\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}_n\vert\Omega) , it holds that \left\Vert {{\mathcal{U}}_n w}\right\Vert_{L^r(\Omega)}\leq C\left\Vert {w}\right\Vert_{{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}}_n\vert\Omega)} .

    Let {\mathcal{A}}^{(n)} be the effective conductivity of {\mathcal{Q}}_{n} and {\mathcal{A}} be the effective conductivity of {\mathcal{Q}} . Then,

    \mathcal{A}^{(n)}\xrightarrow{n\to\infty}\mathcal{A}.

    Proof. This follows from Eq (5.4) and weak convergence w_{i}^{(n)}\rightharpoonup w_{i} .

    Remark 40 (Variational formulation). There is another way to define {\mathcal{A}} : For \eta\in{\mathbb{R}}^{d} , W_{{\mathcal{Q}}}\eta (see Definition 35) is the unique minimizer to

    \min\limits_{\varphi\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}})}\int_{{\mathcal{Q}}}{\big|}\eta+\varphi{\big|}^{2}\text{ d}{\mathcal{P}}(\omega),

    and therefore

    \eta^{t}{\mathcal{A}}\eta = \int_{{\mathcal{Q}}}{\big|}(I_{d}+W_{{\mathcal{Q}}})\eta{\big|}^{2}\text{ d}{\mathcal{P}}(\omega) = \min\limits_{\varphi\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{Q}})}\int_{{\mathcal{Q}}}{\big|}\eta+\varphi{\big|}^{2}\text{ d}{\mathcal{P}}(\omega).

    This equality is related to Theorem 57.

    We will prove both theorems at once. The first part follows using the weak formulation, weak convergence of both u_n^{\varepsilon} and \nabla u_n^{\varepsilon} as n\to\infty as well as standard PDE arguments. The second part is proved in [30] using our notation and also before in [28]. The convergence of {\mathcal{A}}^{(n)} in Part three is given by Corollary 39. Hence, it only remains to prove the convergence of u_n .

    When it comes to the latter, we have to distinguish between three cases:

    i) There exists C_0 > 0 such that, for every \xi\in{\mathbb{R}}^d\setminus\{0\} and every n\in{\mathbb{N}}\cup\{\infty\} , it holds

    |\xi|^2\leq C_0 \xi\cdot {\mathcal{A}}^{(n)}\xi.

    Then, the proof is straightforward using the resulting uniform estimates on \left\Vert {\nabla u}\right\Vert_{L^2(Q)}+\left\Vert {u}\right\Vert_{L^2(Q)} .

    ii) {\mathcal{A}} = 0 : Then, convergence follows after assuming that f_n\in H^1(Q) is uniformly bounded as well as testing with \lambda u_n-f_n . This will eventually lead to an estimate

    \left\Vert {\lambda u_n-f_n}\right\Vert_{L^2(Q)}^2\leq \frac{\tilde C}\lambda \left(\int_Q\nabla f_n\cdot {\mathcal{A}}^{(n)}\nabla f_n\right)\to 0.

    Afterward, we may use an approximation argument for f_n bounded in L^2(Q) using a standard mollifier: f_n^\delta = f_n\ast \eta_\delta with \|a-a\ast\eta_\delta\|_{L^2(Q)}\leq \delta \|a\|_{L^2(Q)} . Then,

    \int_Q \nabla(u_n^\delta-u_n){\mathcal{A}}^{(n)} \nabla(u_n^\delta-u_n) + \frac\lambda2\|u_n^\delta-u_n\|_{L^2(Q)}^2\leq \frac1{2\lambda}\|f_n^\delta-f_n\|_{L^2(Q)}^2.

    iii) \xi\cdot {\mathcal{A}}\xi = 0 along some \xi\in{{{\mathbb{R}}}^{d}}\setminus\{0\} . Then, these \xi form a linear subspace and we may restrict our testing functions to a dependence in the orthogonal direction, only. More precisely, we can consider testing functions with support concentrated on the orthogonal plane to \xi as n\to\infty .

    In Section 6, we will use two-scale convergence to homogenize Eq (1.1) for fixed n . This process is more convenient to handle if the underlying probability space is compact. Here we show that we may take {\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) as the underlying probability space instead of {\mathcal{S}}({\mathbb{R}}^{d}) .

    Lemma 41. Let {\mathcal{P}} be a distribution on {\mathcal{S}}({{{\mathbb{R}}}^{d}}) and let {\mathcal{S}}^{(n)}: = {\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) with the push-forward measure \tilde{{\mathcal{P}}}_{n}: = {\mathcal{P}}\circ{\mathtt{F}_{n}}^{-1} . Recall {\mathbf{G}}_{n}: = \left\{ \mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d})\colon o\notin{\boldsymbol{\boxminus}}\mathbb{x}^{(n)}\right\} and let \tilde{{\mathbf{G}}}_{n}: = \left\{ \mathbb{x}\in{\mathcal{S}}^{(n)}\colon o\notin{\boldsymbol{\boxminus}}\mathbb{x}\right\} = {\mathtt{F}_{n}}{\mathbf{G}}_{n} . Let w_{i}^{(n)} be the cell solutions on {\mathbf{G}}_{n} and \tilde{w}_{i}^{(n)} the cell solutions on \tilde{{\mathbf{G}}}_{n} for their respective dynamical systems. Then, for every i, j\in\left\{ 1, \dots, d\right\} , it holds that

    \begin{equation} \int_{{\mathbf{G}}_{n}}\big[w_{i}^{(n)}+e_{i}\big]\cdot e_{j}\text{ d}{\mathcal{P}} = \int_{\tilde{{\mathbf{G}}}_{n}}\big[\tilde{w}_{i}^{(n)}+e_{i}\big]\cdot e_{j}\text{ d}\tilde{{\mathcal{P}}}_{n}. \end{equation} (5.5)

    Lemma 42 (Properties of pull-back functions). Let (\Omega, {\mathcal{F}}, {\mathcal{P}}, \tau) , (\tilde{\Omega}, \tilde{{\mathcal{F}}}, \tilde{{\mathcal{P}}}, \tilde{\tau}) be dynamical systems, \phi\colon \Omega\to\tilde{\Omega} measurable such that \tilde{{\mathcal{P}}} = {\mathcal{P}}\circ\phi^{-1} and such that for every x\in{\mathbb{R}}^{d}

    \begin{equation} \phi\circ\tau_{x} = \tilde{\tau}_{x}\circ\phi. \end{equation} (5.6)

    Then, the following holds: For every \tilde{f}\in L^{2}(\tilde{\Omega})^{d} , we have f: = \tilde{f}\circ\phi\in L^{2}(\Omega)^{d} with {\|} f{\|}_{L^{2}(\Omega)^{d}} = {\|}\tilde{f}{\|}_{L^{2}(\tilde{\Omega})^{d}} . If \tilde{f}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}(\tilde{\Omega}) , then f\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega) . If \tilde{f}\in L_{{\mathrm{sol}}}^{2}(\tilde{\Omega}) , then f\in L_{{\mathrm{sol}}}^{2}(\Omega) . f is called the pull-back of \tilde{f} .

    Proof. Due to \tilde{{\mathcal{P}}} = {\mathcal{P}}\circ\phi^{-1} , we immediately obtain for arbitrary measurable \tilde{g}\in L^{1}(\tilde{\Omega})^{d} and its pull-back g

    \begin{align} \int_{\tilde{\Omega}}\tilde{g}\text{d}\tilde{{\mathcal{P}}} & = \int_{\Omega}\tilde{g}\circ\phi\text{d}{\mathcal{P}} = \int_{\Omega}g\text{d}{\mathcal{P}}. \end{align} (5.7)

    Therefore {\|} f{\|}_{L^{2}(\Omega)^{d}} = {\|}\tilde{f}{\|}_{L^{2}(\tilde{\Omega})^{d}} and (5.6) yields \tilde{f}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}(\tilde{\Omega})\implies f\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega) . For \tilde{f}\in L_{{\mathrm{sol}}}^{2}(\tilde{\Omega}) , f\in L_{{\mathrm{sol}}}^{2}(\Omega), follows from \phi\circ\tau_{x} = \tilde{\tau}_{x}\circ\phi and checking

    \int_{\Lambda}f(\tau_{x}\omega)\cdot\nabla\varphi(x)\text{d} x = \int_{\Lambda}\tilde{f}(\tilde{\tau}_{x}\omega)\cdot\nabla\varphi(x)\text{d} x = 0,

    for {\mathcal{P}} -almost every \omega and every \varphi\in C_{c}^{1}(\Lambda) on a bounded domain \Lambda\subset{\mathbb{R}}^{d} .

    Proof of Lemma 41. We use Lemma 42 for \phi = {\mathtt{F}_{n}} with (\Omega, {\mathcal{F}}, {\mathcal{P}}, \tau) = ({\mathcal{S}}({\mathbb{R}}^{d}), {\mathcal{B}}({\mathcal{S}}({\mathbb{R}}^{d})), {\mathbb{P}}, \tau) , where \tau_{x} is the shift-operator on {\mathcal{S}}({\mathbb{R}}^{d}) and {\mathcal{B}}({\mathcal{S}}({\mathbb{R}}^{d})) is the Borel- \sigma -algebra generated by the vague topology, and (\tilde{\Omega}, \tilde{{\mathcal{F}}}, \tilde{{\mathcal{P}}}, \tilde{\tau}) = ({\mathcal{S}}^{(n)}), {\mathcal{B}}({\mathcal{S}}^{(n)}), {\mathbb{P}}\circ{\mathtt{F}_{n}}^{-1}, \tau) again with shift-operator and Borel- \sigma -algebra. Let \overline{w}_{i} be the pull-back of \tilde{w}_{i}^{(n)} according to Lemma 42. We see {\mathtt{F}_{n}}^{-1}\tilde{{\mathbf{G}}}_{n} = {\mathbf{G}}_{n} , so \overline{w}_{i} has support in {\mathbf{G}}_{n} . Let \tilde{v}_{k}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathcal{S}}^{(n)}) with {\|}\tilde{w}_{i}^{(n)}-\tilde{v}_{k}{\|}_{L^{2}(\tilde{{\mathbf{G}}}_{n})^{d}}\leq\frac{1}{k} . The pull-back v_{k}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega) of \tilde{v}_{k} satisfies {\|}\overline{w}_{i}-v_{k}{\|}_{L^{2}({\mathbf{G}}_{n})^{d}}\leq\frac{1}{k} and hence, \overline{w}_{i}\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathbf{G}}_{n}\vert{\mathcal{S}}({{{\mathbb{R}}}^{d}})) . We observe \big(\tilde{w}_{i}^{(n)}+e_{i}\big){\mathbb{1}}_{\tilde{{\mathbf{G}}}_{n}}\in L_{{\mathrm{sol}}}^{2}({\mathcal{S}}^{(n)}) with the pull-back \big(\overline{w}_{i}+e_{i}\big){\mathbb{1}}_{{\mathbf{G}}_{n}}\in L_{{\mathrm{sol}}}^{2}({\mathcal{S}}({{{\mathbb{R}}}^{d}})) . This implies \overline{w}_{i} = w_{i}^{(n)} . Eq (5.7) yields Eq (5.5).

    We first collect all the tools needed to prove the homogenization result for minimally smooth domains (Lemma 10).

    Theorem 43 (Extending beyond holes and trace operator). There exists a constant C > 0 depending only on n\in{\mathbb{N}} and M_{0} > 1 such that the following holds: Assume that Q\subset{{{\mathbb{R}}}^{d}} is a bounded Lipschitz domain with Lipschitz constant M_{0} , \mathbb{x}\in{\mathtt{F}_{n}}{\mathcal{S}}({{{\mathbb{R}}}^{d}}) and \mathbb{x}_{Q}\subset\mathbb{x} such that for every x\in\mathbb{x}_{Q} it holds \mathbb{B}_{{2r}}{\left(x\right)}\subset Q . Then, there exists an extension operator

    {\mathcal{U}}_{\mathbb{x}_{Q}}\colon W^{1, 2}(Q\setminus{\boldsymbol{\boxminus}}\mathbb{x}_{Q})\to W^{1, 2}(Q),

    such that ({\mathcal{U}}_{\mathbb{x}_{Q}}u)_{\vert Q\setminus{\boldsymbol{\boxminus}}\mathbb{x}_{Q}} = u and

    \begin{equation} \Vert {{\mathcal{U}}_{\mathbb{x}_{Q}}u}\Vert_{L^{2}(Q)}\leq C\Vert u_\Vert{L^{2}(Q\setminus{\boldsymbol{\boxminus}}\mathbb{x}_{Q})}, \qquad\Vert {\nabla{\mathcal{U}}_{\mathbb{x}_{Q}}u}\Vert_{L^{2}(Q)}\leq C\Vert {\nabla u}\Vert_{L^{2}(Q\setminus{\boldsymbol{\boxminus}}\mathbb{x}_{Q})}. \end{equation} (6.1)

    Furthermore, there exists a trace operator

    {\mathcal{T}}_{\mathbb{x}_{Q}}\colon W^{1, 2}(Q\setminus{\boldsymbol{\boxminus}}\mathbb{x}_{Q})\to L^{2}(\partial{\boldsymbol{\boxminus}}\mathbb{x}_{Q}): = L^{2}(\partial{\boldsymbol{\boxminus}}\mathbb{x}_{Q}, \, {\mathcal{H}}^{d-1}),

    such that {\mathcal{T}}_{\mathbb{x}_{Q}}u = u|_{\partial{\boldsymbol{\boxminus}}\mathbb{x}_{Q}} for every u\in C_{c}^{1}(Q) and

    \begin{equation} {\|}{\mathcal{T}}_{\mathbb{x}_{Q}}u{\|}_{L^{2}(\partial{\boldsymbol{\boxminus}}\mathbb{x}_{Q})}\leq C(\Vert u_\Vert{L^{2}(Q\setminus{\boldsymbol{\boxminus}}(\mathbb{x}_{Q}))}+\Vert {\nabla u}\Vert_{L^{2}(Q\setminus{\boldsymbol{\boxminus}}\mathbb{x}_{Q})}). \end{equation} (6.2)

    Proof. For every \mathbb{x}_{Q}\subset\mathbb{x} with \mathbb{x}\in{\mathtt{F}_{n}}{\mathcal{S}}({{{\mathbb{R}}}^{d}}) , the set Q\setminus{\boldsymbol{\boxminus}}\mathbb{x}_{Q} is minimally smooth with \delta = \min\left\{ 1/n, r\right\} and, M = \max\left\{ \sqrt{2nr}, M_{0}\right\} . Furthermore, the connected components of {\boldsymbol{\boxminus}}\mathbb{x}_{Q} have a diameter less than 2nr . The existence of {\mathcal{U}}_{\mathbb{x}_{Q}} satisfying Eq (6.1) follows from [5, Lemma 2.4] (actually, this was pointed out before by [7, Section 3], but the implications there are not obvious). The existence of {\mathcal{T}}_{\mathbb{x}_{Q}} satisfying Eq (6.2) is provided in [9].

    Definition 44 (Stationary and ergodic random measures). A random measure \mu_{\bullet}\colon \Omega\to{\mathcal{M}}({\mathbb{R}}^{d}) with underlying dynamical system (\Omega, {\mathcal{F}}, {\mathcal{P}}, \tau) is called stationary if

    \mu_{\tau_{x}\omega}(A) = \mu_{\omega}(A+x),

    for every measurable A\subset{\mathbb{R}}^{d} , x\in{\mathbb{R}}^{d} , and {\mathcal{P}} -almost every \omega\in\Omega . \mu_{\bullet} is called ergodic if it is stationary and \tau is ergodic.

    Definition 44 is compatible with Definition 8 given in Section 3 by considering the canonical underlying probability space ({\mathcal{M}}({\mathbb{R}}^{d}), {\mathcal{B}}({\mathcal{M}}({\mathbb{R}}^{d})), {\mathbb{P}}_{\mu}, \tau) with {\mathbb{P}}_{\mu} being the distribution of \mu .

    Theorem 45 (Palm theorem (for finite intensity) [17]). Let \mu_{\bullet} be a stationary random measure with underlying dynamical system (\Omega, {\mathcal{F}}, {\mathcal{P}}, \tau) of finite intensity \lambda(\mu_{\bullet}) .

    Then, there exists a unique finite measure \mu_{{\mathcal{P}}} on (\Omega, {\mathcal{F}}) such that for every g\colon {\mathbb{R}}^{d}\times\Omega\to{\mathbb{R}} measurable and either g\geq0 or g\in L^{1}\left({\mathbb{R}}^{d}\times\Omega, \, {\mathcal{L}}^{d}\otimes\mu_{{\mathcal{P}}}\right) :

    \int_{\Omega}\int_{{\mathbb{R}}^{d}}g(x, \tau_{x}\omega)\text{ d}\mu_{\omega}(x)\text{ d}{\mathcal{P}}(\omega) = \int_{{\mathbb{R}}^{d}}\int_{\Omega}g(x, \omega)\text{ d}\mu_{{\mathcal{P}}}(\omega)\text{ d} x.

    For arbitrary f\in L^{1}({\mathbb{R}}^{d}) with \int_{{\mathbb{R}}^{d}}f\text{ d} x = 1 , we have that

    \mu_{{\mathcal{P}}}(A) = \int_{\Omega}\int_{{\mathbb{R}}^{d}}f(x){\mathbb{1}}_{A}(\tau_{x}\omega)\text{ d}\mu_{\omega}(x)\text{ d}{\mathcal{P}}(\omega),

    in particular, \mu_{{\mathcal{P}}}(\Omega) = \lambda(\mu) . Furthermore, for every \phi\in C_{c}({{{\mathbb{R}}}^{d}}) and g\in L^{1}(\Omega; {\mu_{\mathcal{P}}}) , the ergodic limit

    \begin{equation} \lim\limits_{{\varepsilon}\to0}\int_{{{{\mathbb{R}}}^{d}}}\phi(x)\, g{\left({\tau_{\frac{x}{{\varepsilon}}}\omega}\right)}\, \text{ d} \mu_{\omega}(x) = \int_{{{{\mathbb{R}}}^{d}}}\int_{\Omega}\phi(x)\, g(\tilde{\omega})\, {\mathrm{d}}{\mu_{\mathcal{P}}}(\tilde{\omega})\, {\mathrm{d}} x, \end{equation} (6.3)

    holds for {\mathcal{P}} -almost every \omega . We call \mu_{{\mathcal{P}}} the Palm measure of \mu_{\bullet} .

    For the rest of this subsection, we use the following assumptions.

    Assumption 46. \Omega is a compact metric space with a probability measure {\mathcal{P}} and continuous dynamical system \left(\tau_{x}\right)_{x\in{{{\mathbb{R}}}^{d}}} . Furthermore, \mu_{\bullet}\colon \Omega\to{\mathcal{M}}({\mathbb{R}}^{d}) is a stationary ergodic random measure with Palm measure {\mu_{\mathcal{P}}} . We define {\mu_{\omega}^{{\varepsilon}}}(A): = {\varepsilon}^{d}{\mu_{\omega}}({\varepsilon}^{-1}A) .

    According to [30] (by an application of Eq (6.3)), almost every \omega\in\Omega is typical, i.e., for such an \omega , it holds for every \phi\in C(\Omega) that

    \lim\limits_{{\varepsilon}\to0}|Q|^{-1}\int_{Q}\phi{\left({\tau_{\frac{x}{{\varepsilon}}}\omega}\right)}\, {\mathrm{d}} x = \int_{\Omega}\phi\, {\mathrm{d}}{\mathcal{P}}.

    Definition 47 (Two-scale convergence). Let Assumption 46 hold and let \omega\in\Omega be typical. Let \left({u^{{\varepsilon}}}\right)_{{\varepsilon} > 0} be a sequence {u^{{\varepsilon}}}\in L^{2}(Q, {\mu_{\omega}^{{\varepsilon}}}) and let u\in L^{2}(Q; L^{2}(\Omega, {\mu_{\mathcal{P}}})) such that

    \sup\limits_{{\varepsilon} > 0}\left\Vert {{u^{{\varepsilon}}}}\right\Vert_{L^{2}(Q, {\mu_{\omega}^{{\varepsilon}}})} < \infty,

    and such that for every \varphi\in C_{c}^{\infty}(Q) , \psi\in C(\Omega)

    \begin{equation} \lim\limits_{{\varepsilon}\to0}\int_{Q}{u^{{\varepsilon}}}(x)\varphi(x)\psi{\left({\tau_{\frac{x}{{\varepsilon}}}\omega}\right)}\, {\mathrm{d}}{\mu_{\omega}^{{\varepsilon}}}(x) = \int_{Q}\int_{\Omega}u(x, \tilde{\omega})\varphi(x)\psi{\left({\tilde{\omega}}\right)}\, {\mathrm{d}}{\mu_{\mathcal{P}}}(\tilde{\omega})\, {\mathrm{d}} x. \end{equation} (6.4)

    Then, {u^{{\varepsilon}}} is said to be (weakly) two-scale convergent to u , written {u^{{\varepsilon}}}{\stackrel{2s}{\rightharpoonup}} u .

    Remark 48 (Extending the space of test functions).

    ● For K\in{\mathbb{N}} , let \chi_{1}, \dots, \chi_{K}\in L^{\infty}(\Omega, {\mu_{\mathcal{P}}}) . The original proof of the following Lemma 49 in [30] shows that we can equally define two-scale convergence additionally claiming Eq (6.4) has to hold for every \psi\in C(\Omega) and \psi = \chi_{k}\tilde{\psi} , where \tilde{\psi}\in C(\Omega) . In particular, given a fixed \chi\in L^{\infty}(\Omega, {\mu_{\mathcal{P}}}) , we can w.l.o.g. say that two-scale convergence of {u^{{\varepsilon}}}{\stackrel{2s}{\rightharpoonup}} u implies {u^{{\varepsilon}}}\chi(\tau_{\frac{\cdot}{{\varepsilon}}}\omega){\stackrel{2s}{\rightharpoonup}} u\chi .

    ● Using a standard approximation argument, we can extend the class of test functions from \varphi\in C_{c}^{\infty}(Q) to \varphi\in L^{2}(Q) , provided \mu_{\omega} is uniformly continuous w.r.t. the Lebesgue measure. Then, strong L^{2}(Q) -convergence implies two-scale convergence for \mu_{\omega}\equiv{\mathcal{L}}^{d} .

    Lemma 49 ([30, Lemma 5.1]). Let Assumption 46 hold. Let \omega\in\Omega be typical and {u^{{\varepsilon}}}\in L^{2}(Q, {\mu_{\omega}^{{\varepsilon}}}) be a sequence of functions such that \left\Vert {{u^{{\varepsilon}}}}\right\Vert_{L^{2}(Q, {\mu_{\omega}^{{\varepsilon}}})}\leq C for some C > 0 independent of {\varepsilon} . Then, there exists a subsequence of (u^{{\varepsilon}'})_{{\varepsilon}'\to0} and u\in L^{2}(Q; L^{2}(\Omega, {\mu_{\mathcal{P}}})) such that u^{{\varepsilon}'}\stackrel{2s}{{\rightharpoonup}}u and

    \begin{equation} \Vert u_\Vert{L^{2}(Q;L^{2}(\Omega, {\mu_{\mathcal{P}}}))}\leq\liminf\limits_{{\varepsilon}'\to0}\|u^{{\varepsilon}'}\|_{L^{2}(Q, {\mu_{\omega}^{{\varepsilon}}})}. \end{equation} (6.5)

    Theorem 50 (Two-scale convergence in W^{1, 2}(Q) [30]). Under Assumption 46, for every typical \omega\in\Omega the following holds: If u^{{\varepsilon}}\in W^{1, 2}(Q; {{{\mathbb{R}}}^{d}}) for all {\varepsilon} and if

    \sup\limits_{{\varepsilon} > 0}\big(\left\Vert {u^{{\varepsilon}}}\right\Vert_{L^{2}(Q)}+\left\Vert {\nabla u^{{\varepsilon}}}\right\Vert_{L^{2}(Q)}\big) < \infty,

    then, there exists a u\in W^{1, 2}(Q) with u^{{\varepsilon}}{\rightharpoonup} u weakly in W^{1, 2}(Q) and there exists v\in L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega)) such that \nabla{u^{{\varepsilon}}}{\stackrel{2s}{\rightharpoonup}}\nabla u+v weakly in two scales.

    Due to Theorem 26, the set {\mathtt{F}_{n}}{\mathcal{S}}({{{\mathbb{R}}}^{d}})\subset{\mathcal{M}}({{{\mathbb{R}}}^{d}}) is compact, hence the above two-scale convergence method can be applied for the stationary ergodic point process {\mathbb{X}}^{(n)} taking values in {\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) only. To be more precise, we consider the compact metric probability space \Omega = {\mathtt{F}_{n}}{\mathcal{S}}({{{\mathbb{R}}}^{d}}) and a random variable {\mathbb{X}}^{n}\colon \Omega\to{\mathcal{S}}({\mathbb{R}}^{d}) such that {\mathbb{X}}^{n} and {\mathbb{X}}^{(n)} have the same distribution. By the considerations made in Subsection 5.4, they will both result in the same partial differential equation.

    Theorem 51 (Extension and trace estimates on Q_{\mathbb{x}}^{{\varepsilon}} for \mathbb{x}\in{\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) ). Let Q\subset{{{\mathbb{R}}}^{d}} be a bounded domain, n\in{\mathbb{N}} be fixed. Let {\mathbb{X}} be an admissible point process with values in {\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) . For almost every realization \mathbb{x} of {\mathbb{X}} , we have: Let Q_{\mathbb{x}}^{{\varepsilon}} and G_{\mathbb{x}}^{{\varepsilon}} be defined according to Definition 4.

    i) There exists a C > 0 depending only on Q and n and a family of extension and trace operators

    {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}\colon W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})\to W^{1, 2}(Q), \qquad{\mathcal{T}}_{{\varepsilon}, \mathbb{x}}\colon W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})\to L^{2}(\partial G_{\mathbb{x}}^{{\varepsilon}}),

    such that for every u\in W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}}) it holds

    \begin{align*} \Vert {{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}u}\Vert_{W^{1, 2}(Q)} & \leq C\Vert u_\Vert{W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})}, \\ {\varepsilon}\Vert {{\mathcal{T}}_{{\varepsilon}, \mathbb{x}}u}\Vert_{L^{2}(\partial G_{\mathbb{x}}^{{\varepsilon}})}^{2} & \leq C(\Vert u_\Vert{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}^{2}+{\varepsilon}^{2}\Vert {\nabla u}\Vert_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}^{2}). \end{align*}

    ii) If u^{{\varepsilon}}\in W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}}) is a sequence satisfying \sup_{{\varepsilon}}\left\Vert {{u^{{\varepsilon}}}}\right\Vert_{W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})} < \infty , then, there exists a u\in W^{1, 2}(Q) and a subsequence still indexed by {\varepsilon} such that {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}{\rightharpoonup} u weakly in W^{1, 2}(Q) and there exists v\in L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega)) such that

    \nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}{\stackrel{2s}{\rightharpoonup}}\nabla u+v, \qquad\nabla{u^{{\varepsilon}}}{\stackrel{2s}{\rightharpoonup}}{\mathbb{1}}_{{\mathbf{G}}_{n}}\left(\nabla u+v\right),

    where \mathbf{G}_{n}: = \left\{ \mathbb{x}\in{\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d})\, \vert\, o\notin{\boldsymbol{\boxminus}}\mathbb{x}\right\} . Furthermore, for some C > 0 depending only on Q and n

    \begin{equation} {\varepsilon}\left\Vert {{\mathcal{T}}_{{\varepsilon}, \mathbb{x}}({u^{{\varepsilon}}}-u)}\right\Vert_{L^{2}(\partial G_{\mathbb{x}}^{{\varepsilon}})}^{2}\leq C\left(\left\Vert {{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}-u}\right\Vert_{L^{2}(Q)}^{2}+{\varepsilon}^{2}\left\Vert {\nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}-\nabla u}\right\Vert_{L^{2}(Q)}^{2}\right). \end{equation} (6.6)

    Proof. i) follows from using Theorem 43 on {\varepsilon}^{-1}G_{\mathbb{x}}^{{\varepsilon}} and rescaling the inequalities (6.1) and (6.2).

    ii) is a bit more lengthy. The existence of a subsequence and u\in W^{1, 2}(Q) and v\in L^{2}(Q; L_{{\mathrm{pot}}}^{2}(\Omega)) such that {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}{\rightharpoonup} u and \nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}{\stackrel{2s}{\rightharpoonup}}\nabla u+v follows from Theorem 50. We observe that {\mathbb{1}}_{\mathbf{G}_{n}}(\tau_{x}\mathbb{x}) = {\mathbb{1}}_{{\boldsymbol{\boxminus}}\mathbb{x}^{\complement}}(x) and {\mathbb{1}}_{{\boldsymbol{\boxminus}}\mathbb{x}^{\complement}}(\frac{x}{{\varepsilon}}) = {\mathbb{1}}_{{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}^{\complement}}(x) . Therefore, {\mathbb{1}}_{{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}^{\complement}}\nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}{\stackrel{2s}{\rightharpoonup}}{\left({\nabla u+v}\right)}{\mathbb{1}}_{\mathbf{G}_{n}} (Remark 48(i)). Furthermore, we observe with Q_{n, r}^{{\varepsilon}}: = \left\{ x\in Q\colon {\mathrm{dist}}(x, \partial Q)\leq{\varepsilon} nr\right\} such that we have \left({\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}\cap Q\right)\backslash G_{\mathbb{x}}^{{\varepsilon}}\subset Q_{n, r}^{{\varepsilon}} and

    {\big|}{\mathbb{1}}_{{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}}-{\mathbb{1}}_{G_{\mathbb{x}}^{{\varepsilon}}}{\big|}\leq{\mathbb{1}}_{Q_{n, r}^{{\varepsilon}}}\xrightarrow{{\varepsilon}\to0}0\text{ pointwise a.e..}

    Therefore, {\mathbb{1}}_{{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}}-{\mathbb{1}}_{G_{\mathbb{x}}^{{\varepsilon}}}\to0 strongly in L^{p}(Q) , p\in[1, \infty) , and hence, taking any arbitrary \phi\in C(\Omega) , \psi\in C(\overline{Q}) , we find

    \int_{Q}{\left({{\mathbb{1}}_{{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}}-{\mathbb{1}}_{G_{\mathbb{x}}^{{\varepsilon}}}}\right)}\nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}\phi(\tau_{\frac{\bullet}{{\varepsilon}}}\mathbb{x})\psi\text{d} x \leq\left\Vert {{\mathbb{1}}_{{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}}-{\mathbb{1}}_{G_{\mathbb{x}}^{{\varepsilon}}}}\right\Vert_{L^{2}(Q)}\left\Vert {\nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}}\right\Vert_{L^{2}(Q)}\left\Vert {\phi}\right\Vert_{\infty}\left\Vert {\psi}\right\Vert_{\infty}\to0,

    which means that {\mathbb{1}}_{{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}}\nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}} and \nabla u^{{\varepsilon}} = {\mathbb{1}}_{G_{\mathbb{x}}^{{\varepsilon}}}\nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}} have the same two-scale limit

    \nabla u^{{\varepsilon}}\mathop \rightharpoonup \limits_{{\varepsilon}\to0}^{2s}{\mathbb{1}}_{{\mathbf{G}}_{n}}{\left({\nabla u+v}\right)}.

    Due to the absolutely bounded diameter of the connected components of {\boldsymbol{\boxminus}}\mathbb{x} , there exists a domain B\supset Q big enough such that, with the notation of Definition 4,

    Q\cap{\varepsilon}{\boldsymbol{\boxminus}}(J_{{\varepsilon}}(\mathbb{x}, B)) = Q\cap{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x}\quad\forall{\varepsilon}\in(0, 1).

    Now let {\mathcal{U}}_{Q}\colon W^{1, 2}(Q)\to W^{1, 2}(B) be the canonical extension operator satisfying

    \left\Vert {{\mathcal{U}}_{Q}u}\right\Vert_{L^{2}(B)}\leq C\left\Vert {u}\right\Vert_{L^{2}(Q)}\qquad\text{and}\qquad\left\Vert {\nabla{\mathcal{U}}_{Q}u}\right\Vert_{L^{2}(B)}\leq C\left\Vert {\nabla u}\right\Vert_{L^{2}(Q)}.

    Reapplying Theorem 43 to the trace on {\varepsilon}^{-1}\left(B\backslash{\varepsilon}{\boldsymbol{\boxminus}}(J_{{\varepsilon}}(\mathbb{x}, B))\right) , we find for some constant C independent from {\varepsilon} and {\mathbb{X}} but depending on Q , B , and n and varying from line to line:

    \begin{align*} {\varepsilon}\left\Vert {{\mathcal{T}}_{{\varepsilon}, \mathbb{x}}({u^{{\varepsilon}}}-u)}\right\Vert_{L^{2}(\partial G_{\mathbb{x}}^{{\varepsilon}})}^{2} & \leq{\varepsilon}\left\Vert {{\mathcal{T}}_{{\varepsilon}, \mathbb{x}}({\mathcal{U}}_{Q}{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}-{\mathcal{U}}_{Q}u)}\right\Vert_{L^{2}({\varepsilon}\partial{\boldsymbol{\boxminus}}(J_{{\varepsilon}}(\mathbb{x}, B)))}^{2}\\ & \leq C\left(\left\Vert {{\mathcal{U}}_{Q}{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}-{\mathcal{U}}_{Q}u}\right\Vert_{L^{2}(B)}^{2}+{\varepsilon}^{2}\left\Vert {\nabla{\mathcal{U}}_{Q}{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}-\nabla{\mathcal{U}}_{Q}u}\right\Vert_{L^{2}(B)}^{2}\right)\\ & \leq C\left(\left\Vert {{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}-u}\right\Vert_{L^{2}(Q)}^{2}+{\varepsilon}^{2}\left\Vert {\nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}-\nabla u}\right\Vert_{L^{2}(Q)}^{2}\right), \end{align*}

    as desired.

    Due to the perforations, \partial_{t}{u^{{\varepsilon}}} cannot be embedded in a common space in a convenient way for the application of the Aubin–Lions theorem. Hence, we use the following general characterization of compact sets.

    Theorem 52 (Characterization of compact sets in L^{p}(I; V) [24, Theorem 1]). Let V be a Banach space, p\in[1, \infty) and \Lambda\subset L^{p}(I; V) . \Lambda is relatively compact in L^{p}(I; V) if, and only if,

    \begin{align} & \left\{ \int_{t_{1}}^{t_{2}}v(t)\text{ d} t\, \vert\, v\in\Lambda\right\} \mathit{\text{ is relatively compact in }}V\ \forall\, 0 < t_{1} < t_{2} < T, \end{align} (6.7)
    \begin{align} & \sup\limits_{v\in\Phi}{\|}\mathfrak{s}_{h}[v]-v{\|}_{L^{p}(0, T-h;V)}\to0\mathit{\text{ as }}h\to0, \end{align} (6.8)

    where \mathfrak{s}_{h}[v(\, \cdot\, )]: = v(\, \cdot\, +h) is the shift by h\in\left(0, \, T\right) .

    We can now establish the existence of a solution for fixed {\varepsilon} > 0 to our partial differential equation.

    Theorem 53 (Existence of solution on perforated domains and a priori estimate). Let \mathbb{x}\in{\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) . Under Assumption 6 and with Q_{\mathbb{x}}^{{\varepsilon}} as defined in Definition 4, we have that there exists a solution {u^{{\varepsilon}}}\in L^{2}(I; W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})) , with generalized time derivative {\partial}_{t}{u^{{\varepsilon}}}\in L^{2}(I; W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})^{*}) , to Eq (1.1), i.e.,

    \begin{equation} \begin{aligned}{\partial}_{t}{u^{{\varepsilon}}}-\nabla\cdot\left(A({u^{{\varepsilon}}})\, \nabla{u^{{\varepsilon}}}\right) & = f & & \mathit{\text{ in }}I\times Q_{\mathbb{x}}^{{\varepsilon}}\\ A({u^{{\varepsilon}}})\, \nabla{u^{{\varepsilon}}}\cdot\nu & = 0 & & \mathit{\text{ on }}I\times{\partial} Q\\ A({u^{{\varepsilon}}})\, \nabla{u^{{\varepsilon}}}\cdot\nu & = {\varepsilon} h({u^{{\varepsilon}}}) & & \mathit{\text{ on }}I\times{\partial} Q_{\mathbb{x}}^{{\varepsilon}}\backslash{\partial} Q\\ {u^{{\varepsilon}}}(0, x) & = u_{0}(x) & & \mathit{\text{ in }}Q_{\mathbb{x}}^{{\varepsilon}}, \end{aligned} \end{equation} (6.9)

    which satisfies the following a priori estimates for {\varepsilon} small enough

    \begin{align} \underset{t \in I}{\operatorname{ess} \sup }{\|}{u^{{\varepsilon}}}(t){\|}_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}^{2} & \leq\exp(C_{1})\big[{\|} u_{0}{\|}_{L^{2}(Q)}^{2}+C_{2}\big] \\ {\|}\nabla{u^{{\varepsilon}}}{\|}_{L^{2}(I;L^{2}(Q_{\mathbb{x}}^{{\varepsilon}}))}^{2} & \leq\frac{1}{\inf(A)}\big(1+C_{1}\exp(C_{1})\big)\big[{\|} u_{0}{\|}_{L^{2}(Q)}^{2}+C_{2}\big] \\ {\|}{\partial}_{t}{u^{{\varepsilon}}}{\|}_{L^{2}(I;W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})^{*})}^{2} & \leq\tilde{C}, \end{align} (6.10)

    where

    C_{1}: = T(1+3CL_{h})\qquad\mathit{\text{and}}\qquad C_{2}: = TL_{h}h(0)^{2}{\mathcal{L}}^{d}(Q)+{\|} f{\|}_{L^{1}(I;L^{2}(Q))}^{2},

    C is from Theorem 51 depending only on Q and n and where \tilde{C} > 0 is independent of {\varepsilon} .

    Proof. We will only sketch the proof. There are 3 main steps: Deriving a priori estimates, existence of Galerkin solutions and the limit passing.

    i) Testing Eq (6.9) with {u^{{\varepsilon}}} and using

    \langle{\partial}_{t}{u^{{\varepsilon}}}, \, {u^{{\varepsilon}}}\rangle_{W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})^{*}, W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})} = \frac{1}{2}\frac{\text{d}}{\text{d} t}{\|}{u^{{\varepsilon}}}{\|}_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}^{2},

    yields

    \begin{align*} \frac{1}{2}\frac{\text{d}}{\text{d} t}{\|}{u^{{\varepsilon}}}{\|}_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}^{2}+A({u^{{\varepsilon}}})\, {\|}\nabla{u^{{\varepsilon}}}{\|}_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}^{2}-{\varepsilon}\big(h({u^{{\varepsilon}}}), \, {u^{{\varepsilon}}}\big)_{L^{2}({\partial} G_{\mathbb{x}}^{{\varepsilon}})} & = \big(f, \, {u^{{\varepsilon}}}\big){}_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}. \end{align*}

    The a priori estimate then follows from the Gronwall inequality and the trace estimate in Theorem 51. For the a priori estimate in {\partial}_{t}{u^{{\varepsilon}}} , one simply uses

    \langle{\partial}_{t}{u^{{\varepsilon}}}, \, \varphi\rangle = \big(A({u^{{\varepsilon}}})\, \nabla{u^{{\varepsilon}}}, \, \nabla\varphi\big){}_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}+{\varepsilon}\big(h({u^{{\varepsilon}}}), \, \varphi\big){}_{L^{2}({\partial} G_{\mathbb{x}}^{{\varepsilon}})}+\big(f, \, \varphi\big){}_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}.

    ii) Let (V_{m})_{m\in{\mathbb{N}}} be a family of finite-dimensional vector spaces, V_{m}\nearrow W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}}) . Solutions to Eq (1.1) exist in V_{m} , that is,

    \begin{align*} {\partial}_{t}u_{(m)}^{{\varepsilon}}-\nabla\cdot\left(A(u_{(m)}^{{\varepsilon}})\, \nabla u_{(m)}^{{\varepsilon}}\right) & = \mathbf{P}_{m}f & & \text{in }I\times Q^{{\varepsilon}}\\ A(u_{(m)}^{{\varepsilon}})\, \nabla u_{(m)}^{{\varepsilon}}\cdot\nu & = 0 & & \text{on }I\times{\partial} Q\\ A(u_{(m)}^{{\varepsilon}})\, \nabla u_{(m)}^{{\varepsilon}}\cdot\nu & = {\varepsilon} h(u_{(m)}^{{\varepsilon}}) & & \text{on }I\times{\partial} Q^{{\varepsilon}}\backslash{\partial} Q\\ u_{(m)}^{{\varepsilon}}(0, x) & = \mathbf{P}_{m}u_{0}(x) & & \text{in }Q^{{\varepsilon}}, \end{align*}

    where \mathbf{P}_{m}\colon W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})\to V_{m} is the orthogonal projection. This can be shown via, e.g., [21, Theorem 3.7] which yields solutions for some I': = [0, T']\subset[0, T] . In this reference, their y\in W^{1, p}(I') corresponds to the parameters of the Galerkin approximation, i.e., u_{(m)}^{{\varepsilon}} = \sum_{i = 1}^{m}y_{i}v_{i} with v_{i}\in V_{m} being fixed orthonormal base vectors. These solutions u_{(m)}^{{\varepsilon}} satisfy the a priori estimate in Eq (6.10) as well as

    \begin{equation} \sup\limits_{m\in{\mathbb{N}}}{\|}{\partial}_{t}u_{(m)}^{{\varepsilon}}{\|}_{L^{2}(I';V_{m}^{*})} < \infty, \end{equation} (6.11)

    in particular, we may extend the solution u_{(m)}^{{\varepsilon}} to the whole time interval I (with still uniform bounds in m ).

    iii) The a priori estimates yield a L^{2}(I; L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})) -weakly convergent subsequence to some {u^{{\varepsilon}}}\in L^{2}(I; L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})) . Theorem 52 and Eq (6.11) imply pre-compactness of (u_{(m)}^{{\varepsilon}})_{m\in{\mathbb{N}}}\subset L^{2}(I; L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})) as well as pre-compactness of ({\mathcal{T}}_{{\varepsilon}, \mathbb{x}}u_{(m)}^{{\varepsilon}})_{m\in{\mathbb{N}}}\subset L^{2}(I; L^{2}({\partial} G_{\mathbb{x}}^{{\varepsilon}})) , see Remark 54. Testing with functions in L^{2}(I; V_{m}) and passing to the limit m\to\infty finishes the proof since \bigcup_{m\in{\mathbb{N}}}V_{m} is dense in W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}}) .

    Remark 54 (Procedure of Simon's theorem). We will use Simon's theorem (Theorem 52) on multiple occasions. The general procedure will always be the same. We will exemplaily prove the following result: Let I = [0, T] , U\subset{\mathbb{R}}^{d} be some bounded Lipschitz-domain and {\mathcal{T}}\colon W^{1, 2}(U)\to L^{2}({\partial} U) the trace operator. For each k\in{\mathbb{N}} , let u_{k}\in L^{2}(I; W^{1, 2}(U)) with generalized time-derivative {\partial}_{t}u_{k}\in L^{2}(I; W^{1, 2}(U)^{*}) via W^{1, 2}(U)\hookrightarrow L^{2}(U)\hookrightarrow W^{1, 2}(U)^{*} . Assume that

    C: = \sup\limits_{k\in{\mathbb{N}}}{\|} u_{k}{\|}_{L^{2}(I;W^{1, 2}(U))} < \infty\qquad{\text{and}}\qquad\tilde{C}: = \sup\limits_{k\in{\mathbb{N}}}{\|}{\partial}_{t}u_{k}{\|}_{L^{2}(I;V_{k}^{*})} < \infty,

    for either the situation that W^{1, 2}(U)\subset V_{k}\subset L^{2}(U) with \left\Vert {\cdot}\right\Vert_{V_{k}}\leq\left\Vert {\cdot}\right\Vert_{W^{1, 2}(U)} and uniformly continuous injective maps {\mathcal{U}}_{k}\colon V_{k}\to W^{1, 2}(U) or for the situation that V_{k}\subset W^{1, 2}(U) . We further claim u_{k}(t)\in V_{k} for almost every t\in I . Then,

    ({\mathcal{U}}_{k}u_{k})_{k\in{\mathbb{N}}}\subset L^{2}(U)\mathit{\text{resp.}}(u_{k})_{k\in{\mathbb{N}}}\subset L^{2}(U)\quad{\text{and}}\quad({\mathcal{T}} u_{k})_{k\in{\mathbb{N}}}\subset L^{2}({\partial} U),

    are relatively compact.

    Exemplarily proof for the procedure of Simon's theorem. We need to show Conditions (6.7) and (6.8) from Theorem 52.

    i) Condition (6.7) usually relies on compactness results for the stationary setting. Since

    \sup\limits_{k\in{\mathbb{N}}}\left\Vert {\int_{t_{1}}^{t_{2}}u_{k}\text{d} t}\right\Vert_{W^{1, 2}(Q)}\leq\sup\limits_{k\in{\mathbb{N}}}\sqrt{T}{\|} u_{k}{\|}_{L^{2}(I;W^{1, 2}(U))} < \infty,

    compactness of {\mathcal{T}} yields pre-compactness of \big(\int_{t_{1}}^{t_{2}}{\mathcal{T}} u_{k}\text{d} t\big)_{k\in{\mathbb{N}}} = \big({\mathcal{T}}\int_{t_{1}}^{t_{2}}u_{k}\text{d} t\big)_{k\in{\mathbb{N}}}\subset L^{2}({\partial} U) , so we have shown Condition (6.7).

    ii) Condition (6.8) will additionally require some a priori estimate on {\partial}_{t}u_{k} . We have

    u_{k}(t_{2}) = u_{k}(t_{1})+\int_{t_{1}}^{t_{2}}{\partial}_{t}u_{k}\text{d} s,

    as elements of W^{1, 2}(U)^{*} . Using the Cauchy–Schwarz inequality twice, we get for h\in(0, T) :

    \begin{align*} {\|}\mathfrak{s}_{h}[u_{k}]-u_{k} & {\|}_{L^{2}((0, T-h);L^{2}(U))}^{2} = \int_{0}^{T-h}\big(u_{k}(t+h)-u_{k}(t), \, u_{k}(t+h)-u_{k}(t)\big)_{L^{2}(U)}\text{d} t\\ & = \int_{0}^{T-h}{\big\langle}\int_{t}^{t+h}{\partial}_{t}u_{k}(s)\text{d} s, \, u_{k}(t+h)-u_{k}(t){\big\rangle}_{W^{1, 2}(U)^{*}, W^{1, 2}(U)}\text{d} t\\ & \leq\int_{0}^{T-h}{\|}\int_{t}^{t+h}{\partial}_{t}u_{k}(s)\text{d} s{\|}_{L^{2}(V_{k}^{*})}\, {\|}{\mathcal{U}}_{k}u_{k}(t+h)-{\mathcal{U}}_{k}u_{k}(t){\|}_{W^{1, 2}(U)}\text{d} t\\ & \leq h{\|}{\partial}_{t}u_{k}{\|}_{L^{2}(I;V_{k}^{*})}\, 2{\|} u_{k}{\|}_{L^{2}(I;W^{1, 2}(U))}\leq2hC\tilde{C}. \end{align*}

    Compactness of {\mathcal{T}} implies that for every \delta > 0 , there exists a C_{\delta} > 0 such that

    {\|}{\mathcal{T}} v{\|}_{L^{2}({\partial} U)}^{2}\leq C_{\delta}{\|} v{\|}_{L^{2}(U)}^{2}+\delta{\|}\nabla v{\|}_{L^{2}(U)}^{2}\ \forall v\in W^{1, 2}(U).

    Therefore,

    \begin{align*} {\|}\mathfrak{s}_{h}[{\mathcal{T}} u_{k}] & -{\mathcal{T}} u_{k}{\|}_{L^{2}((0, T-h);L^{2}({\partial} U))}^{2} = {\|}{\mathcal{T}}[\mathfrak{s}_{h}u_{k}-u_{k}]{\|}_{L^{2}((0, T-h);L^{2}({\partial} U))}^{2}\\ \leq & C_{\delta}{\|}\mathfrak{s}_{h}u_{k}-u_{k}{\|}_{L^{2}({\partial} U)}^{2}+\delta{\|}\nabla\mathfrak{s}_{h}u_{k}-\nabla u_{k}{\|}_{L^{2}(U)}^{2}\\ \leq & 2hC_{\delta}C\tilde{C}+2\delta\tilde{C}. \end{align*}

    The estimate is independent of the chosen u_{k} , and Condition (6.8) holds.

    We have shown both conditions and concluded.

    We can now pass to the limit {\varepsilon}\to0 for the homogenized system. Some extra care has to be taken since Q_{\mathbb{x}}^{{\varepsilon}}\neq Q\backslash{\varepsilon}{\boldsymbol{\boxminus}}\mathbb{x} , especially in the boundary term. However, we show that the difference becomes negligible for the two-scale convergence as {\varepsilon}\to0 .

    Theorem 55 (Homogenized system for {\boldsymbol{\boxminus}}{\mathbb{X}}^{(n)} ). Let {\mathbb{X}} be a stationary ergodic point process with values in {\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d}) . Recall the surface measure \mu_{\mathbb{x}} from Definition 30

    \mu_{\mathbb{x}}: = {\mathcal{H}}_{\llcorner{\partial}{\boldsymbol{\boxminus}}\mathbb{x}}^{d-1}.

    Under Assumption 6, we have for almost every realization \mathbb{x} of {\mathbb{X}} and with Q_{\mathbb{x}}^{{\varepsilon}} as defined in Definition 4: Let u^{{\varepsilon}} be a solution to Eq (6.9) and let {\mathcal{U}}_{{\varepsilon}, \mathbb{x}} be given as in Theorem 51. There exists a u_{n}\in L^{2}(I; W^{1, 2}(Q)) with generalized time derivative {\partial}_{t}u_{n}\in L^{2}(I; W^{1, 2}(Q)^{*}) such that for a subsequence

    \begin{align*} {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}u^{{\varepsilon}} \xrightarrow[{\varepsilon}\to0]{L^{2}(I;L^{2}(Q))}u_{n}\qquad \mathit{\text{and}}\qquad {\partial}_{t}u^{{\varepsilon}} \mathop \rightharpoonup \limits_{{\varepsilon}\to0}^{L^{2}(I;W^{1, 2}(Q)^{*})}{\mathbb{P}}({\mathbf{G}}_{n}){\partial}_{t}u_{n}, \end{align*}

    and u_{n} is a (not necessarily unique) solution to

    \begin{align} {\mathbb{P}}({\mathbf{G}}_{n}){\partial}_{t}u_{n}-\nabla\cdot\left(A(u_{n}){\mathcal{A}}\, \nabla u_{n}\right)-\lambda(\mu_{{\mathbb{X}}})h(u_{n}) & = {\mathbb{P}}({\mathbf{G}}_{n})f & & \mathit{\text{ in }}I\times Q \\ A(u_{n}){\mathcal{A}}^{(n)}\, \nabla u_{n}\cdot\nu & = 0 & & \mathit{\text{ on }}I\times{\partial} Q \\ u_{n}(0, x) & = {\mathbb{P}}({\mathbf{G}}_{n})u_{0}(x) & & \mathit{\text{ in }}Q, \end{align} (6.12)

    with {\mathcal{A}}^{(n)} being the effective conductivity based on the event {\mathbf{G}}_{n} = \big\{\mathbb{x}\in{\mathtt{F}_{n}}{\mathcal{S}}({\mathbb{R}}^{d})\, \vert\, o\notin{\boldsymbol{\boxminus}}\mathbb{x}\big\} defined in Definition 37. Furthermore, u_{n} satisfies the following a priori estimates

    \begin{align*} \underset{t \in I}{\operatorname{ess} \sup }{\|} u_{n}(t){\|}_{L^{2}(Q)}^{2} & \leq\exp(C_{1}^{(n)})\big[{\|} u_{0}{\|}_{L^{2}(Q)}^{2}+C_{2}^{(n)}\big]\\ {\|}\nabla u_{n}{\|}_{L^{2}(I;L^{2}(Q))}^{2} & \leq\frac{{\mathbb{P}}({\mathbf{G}}_{n})}{2\alpha_{{\mathcal{A}}^{(n)}}\inf(A)}\big(1+C_{1}^{(n)}\exp(C_{1}^{(n)})\big)\big[{\|} u_{0}{\|}_{L^{2}(Q)}^{2}+C_{2}^{(n)}\big], \end{align*}

    for

    \begin{align*} C_{1}^{(n)} : = T\big(1+\frac{\lambda(\mu_{{\mathbb{X}}})}{{\mathbb{P}}({\mathbf{G}}_{n})}(1+2L_{h})\big)\qquad\mathit{\text{and}}\qquad C_{2}^{(n)} : = {\|} f{\|}_{L^{2}(I;L^{2}(Q))}^{2}+2T\frac{\lambda(\mu_{{\mathbb{X}}})}{{\mathbb{P}}({\mathbf{G}}_{n})}{\big|} h(0){\big|}^{2}. \end{align*}

    Proof. The a priori estimates in Eq (6.10) and Theorem 51 tell us that

    \begin{align*} {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}u^{{\varepsilon}}{\stackrel{2s}{\rightharpoonup}} u_{n}, \qquad \nabla{\mathcal{U}}_{{\varepsilon}, \mathbb{x}}u^{{\varepsilon}} {\stackrel{2s}{\rightharpoonup}}\nabla u_{n}+v, \qquad u^{{\varepsilon}} {\stackrel{2s}{\rightharpoonup}}{\mathbb{1}}_{{\mathbf{G}}_{n}}u_{n}, \qquad \nabla u^{{\varepsilon}} {\stackrel{2s}{\rightharpoonup}}{\mathbb{1}}_{{\mathbf{G}}_{n}}(\nabla u_{n}+v), \end{align*}

    for some u_{n}\in L^{2}(I; W^{1, 2}(Q)) and v\in L^{2}(I; L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega))) where the two-scale convergence is with respect to the Lebesgue measure {\mathcal{L}}^{d} . The uniform bound for {\partial}_{t}u^{{\varepsilon}} in Eq (6.10) together with Theorem 52 yields (for yet another subsequence)

    \begin{equation} {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}\xrightarrow[{\varepsilon}\to0]{L^{2}(I;L^{2}(Q))}u_{n}, \end{equation} (6.13)

    compared to, e.g., Remark 54.

    For \varphi_{1}, \varphi_{2}\in C^{1}([0, T]\times\overline{Q}) with \varphi_{1}(T, \cdot) = 0 and \psi\in H^{1}(\Omega) with \int_{\Omega}\nabla_{\tilde{\omega}}\psi\text{d} {\mathbb{P}} = 0 , we use \varphi^{{\varepsilon}}(t, x): = \varphi_{1}(t, x)+{\varepsilon}\varphi_{2}(t, x)\psi(\tau_{\frac{x}{{\varepsilon}}}\mathbb{x}) as a test function and pass to the limit using two-scale convergence. Furthermore, we use

    \begin{equation} A(u^{{\varepsilon}})\nabla u^{{\varepsilon}}{\stackrel{2s}{\rightharpoonup}} A(u_{n}){\mathbb{1}}_{{\mathbf{G}}_{n}}\left(\nabla u_{n}+v\right)\qquad\text{and}\qquad h(u^{{\varepsilon}})\mathop \rightharpoonup \limits^{2s, \, \mu_{{\mathbb{X}}}} h(u_{n}), \end{equation} (6.14)

    which we will prove below. We then obtain the two equations

    \begin{align} {-}\int_{0}^{T}\int_{Q}u_{n}\partial_{t}\varphi_{1}\text{d} x\text{d} t+\int_{Q}u_{0}\varphi_{1}\text{d} x +\int_{0}^{T}\int_{Q}\int_{{\mathbf{G}}_{n}}\nabla\varphi_{1}\cdot& A(u_{n})\left(\nabla u_{n} +v\right)\text{d} {\mathbb{P}}(\mathbb{x}) \text{d} x\text{d} t \\ +\int_{0}^{T}\int_{Q}h(u_{n})\varphi_{1}\text{d} x\text{d} t\int_{\Omega}{\mathrm{d}}{\mu_{\mathcal{P}}}(x) & = \int_{0}^{T}\int_{Q}f\varphi_{1}\text{d} x\text{d} t, \end{align} (6.15)
    \begin{align} \int_{0}^{T}\int_{Q}\int_{{\mathbf{G}}_{n}}\varphi_{2}\nabla_{\tilde{\omega}}\psi\cdot A(u_{n})\left(\nabla u_{n}+v\right)\text{d}{\mathbb{P}}(\mathbb{x})\text{d} x \text{d} t & = 0. \end{align} (6.16)

    The second equation holds true for every choice of \varphi_{2} and \psi as above if we make the standard ansatz v = \sum_{i = 1}^{d}\partial_{i}u_{n}w_{i}^{(n)} (see the remark following this proof), where w_{i}^{(n)} are the cell solutions from Definition 35 for \Omega = {\mathtt{F}_{n}}{\mathcal{S}}({{{\mathbb{R}}}^{d}}) and {\mathcal{P}} = {\mathbb{P}} being the distribution of {\mathbb{X}} . Plugging this information into the first equation yields (6.12). The a priori estimate follows from testing Eq (6.12) with u_{n} and the Gronwall inequality (see, e.g., the proof of Theorem 53). It only remains to prove Eq (6.14).

    Now, we show the first part of Eq (6.14). By Remark 48, we know that

    A(u_{n})\nabla u^{{\varepsilon}}\mathop \rightharpoonup \limits_{{\varepsilon}\to0}^{2s}{\mathbb{1}}_{{\mathbf{G}}_{n}}A(u_{n}){\left({\nabla u_{n}+v}\right)}.

    Using dominated convergence and Eq (6.13) yields a subsequence such that A({\mathcal{U}}_{{\varepsilon}, \mathbb{x}}u^{{\varepsilon}})\to A(u_{n}) in L^{p}(0, T; L^{p}(Q)) for every 1\leq p < \infty . Using test functions \phi\in C(\overline{Q}) and \psi\in C(\Omega) , we observe that \left(A({\mathcal{U}}_{{\varepsilon}, \mathbb{x}}u^{{\varepsilon}})-A(u_{n})\right)\nabla u^{{\varepsilon}}{\stackrel{2s}{\rightharpoonup}}0 , so A(u^{{\varepsilon}})\nabla u^{{\varepsilon}}{\stackrel{2s}{\rightharpoonup}}{\mathbb{1}}_{{\mathbf{G}}_{n}}A(u_{n}){\left({\nabla u_{n}+v}\right)} .

    The second part of Eq (6.14) is more difficult. Given \varphi\in C^{1}(\overline{Q}) and \psi\in C(\Omega) , we set \psi^{{\varepsilon}, \mathbb{x}}(x): = \psi(\tau_{\frac{x}{{\varepsilon}}}\mathbb{x}) and find

    \begin{align*} & \left|{\varepsilon}\int_{{\partial} G_{\mathbb{x}}^{{\varepsilon}}}h\big({u^{{\varepsilon}}}(x)\big)\varphi(x)\psi^{{\varepsilon}, \mathbb{x}}(x){\mathrm{d}}{\mathcal{H}}^{d-1}(x)-\int_{Q}\int_{\Omega}h\big(u_{n}(x)\big)\, \varphi(x)\, \psi(\mathbb{x})\, {\mathrm{d}}{\mu_{\mathcal{P}}}(\mathbb{x})\, {\mathrm{d}} x\right|\\ & \leq\quad\left|{\varepsilon}\int_{{\partial} G_{\mathbb{x}}^{{\varepsilon}}}h\big({u^{{\varepsilon}}}(x)\big)\varphi(x)\psi^{{\varepsilon}, \mathbb{x}}(x){\mathrm{d}}{\mathcal{H}}^{d-1}(x)-{\varepsilon}\int_{{\partial} G_{\mathbb{x}}^{{\varepsilon}}}h\big(u_{n}(x)\big)\varphi(x)\psi^{{\varepsilon}, \mathbb{x}}(x){\mathrm{d}}{\mathcal{H}}^{d-1}(x)\right|\\ & \quad+\left|{\varepsilon}\int_{{\partial} G_{\mathbb{x}}^{{\varepsilon}}}h\big(u_{n}(x)\big)\varphi(x)\psi^{{\varepsilon}, \mathbb{x}}(x){\mathrm{d}}{\mathcal{H}}^{d-1}(x)-{\varepsilon}\int_{Q\cap{\varepsilon}\partial{\boldsymbol{\boxminus}}\mathbb{x}}h\big(u_{n}(x)\big)\varphi(x)\psi^{{\varepsilon}, \mathbb{x}}(x){\mathrm{d}}{\mathcal{H}}^{d-1}(x)\right|\\ & \quad+\left|{\varepsilon}\int_{Q\cap{\varepsilon}\partial{\boldsymbol{\boxminus}}\mathbb{x}}h\big(u_{n}(x)\big)\varphi(x)\psi^{{\varepsilon}, \mathbb{x}}(x){\mathrm{d}}{\mathcal{H}}^{d-1}(x)-\int_{Q}\int_{\Omega}h\big(u_{n}(x)\big)\, \varphi(x)\, \psi(\mathbb{x})\, {\mathrm{d}}{\mu_{\mathcal{P}}}(\mathbb{x})\, {\mathrm{d}} x\right|. \end{align*}

    We will show that all these terms go to 0 as {\varepsilon}\to0 . Due to the Lipschitz continuity of h and Stampacchia's lemma, we find

    \begin{align*} \Vert {\nabla h(u^{{\varepsilon}})}\Vert_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})} & \leq\Vert h_\Vert{C^{0, 1}}\Vert {\nabla u^{{\varepsilon}}}\Vert_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}, & \Vert {h(u^{{\varepsilon}})}\Vert_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})} & \leq\Vert h_\Vert{C^{0, 1}}(\Vert {u^{{\varepsilon}}}\Vert_{L^{2}(Q_{\mathbb{x}}^{{\varepsilon}})}+1), \\ \Vert {\nabla h(u_{n})}\Vert_{L^{2}(Q)} & \leq\Vert h_\Vert{C^{0, 1}}\Vert {\nabla u_{n}}\Vert_{L^{2}(Q)}, & \Vert {h(u_{n})}\Vert_{L^{2}(Q)} & \leq\Vert h_\Vert{C^{0, 1}}(\Vert {u_{n}}\Vert_{L^{2}(Q)}+1). \end{align*}

    Furthermore, {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}\to u_{n} strongly in L^{2}(I; L^{2}(Q)) and weakly in L^{2}(I; W^{1, 2}(Q)) implies that h\left({\mathcal{U}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}\right)\to h(u_{n}) in the same topologies. G_{\mathbb{x}}^{{\varepsilon}} as in Definition 4 fulfills {\partial} G_{\mathbb{x}}^{{\varepsilon}} = {\partial} Q_{\mathbb{x}}^{{\varepsilon}}\backslash{\partial} Q . Eq (6.6) together with the strong convergence of {\mathcal{U}}_{{\varepsilon}, \mathbb{x}}u^{{\varepsilon}} tells us

    {\varepsilon}\left\Vert {h\big({\mathcal{T}}_{{\varepsilon}, \mathbb{x}}{u^{{\varepsilon}}}\big)-h\big({\mathcal{T}}_{{\varepsilon}, \mathbb{x}}u_{n}\big)}\right\Vert_{L^{2}(I;L^{2}(\partial G_{\mathbb{x}}^{{\varepsilon}}))}^{2}\leq L_{h}{\varepsilon}\left\Vert {{\mathcal{T}}_{{\varepsilon}, \mathbb{x}}({u^{{\varepsilon}}}-u_{n})}\right\Vert_{L^{2}(I;L^{2}(\partial G_{\mathbb{x}}^{{\varepsilon}}))}^{2}\to0,

    which already shows convergence in the first summand. Similar considerations to the proof of Eq (6.6) tell us that {\mathcal{T}}_{{\varepsilon}, \mathbb{x}}\colon W^{1, 2}(Q_{\mathbb{x}}^{{\varepsilon}})\to L^{2}({\varepsilon}{\partial}{\boldsymbol{\boxminus}}\mathbb{x}\cap Q) is a bounded linear operator, so we can consider the trace not only on \partial G_{\mathbb{x}}^{{\varepsilon}} but even for clusters close to the boundary. We have, with C > 0 changing from line to line but independent of {\varepsilon} ,

    \begin{align*} {\big|}{\varepsilon}\int_{I}\int_{Q} & \big({\mathbb{1}}_{{\varepsilon}{\partial}{\boldsymbol{\boxminus}}\mathbb{x}}-{\mathbb{1}}_{{\partial} G_{\mathbb{x}}^{{\varepsilon}}}\big)h(u^{{\varepsilon}})\varphi\psi^{{\varepsilon}, \mathbb{x}}\text{d}{\mathcal{H}}^{d-1}\text{d} t{\big|}^{2}\\ & \leq C{\varepsilon}{\|} h(u^{{\varepsilon}}){\|}_{L^{2}(I;L^{2}(({\varepsilon}{\partial}{\boldsymbol{\boxminus}}\mathbb{x})\backslash{\partial} G_{\mathbb{x}}^{{\varepsilon}}))}^{2}\cdot{\varepsilon}{\|}1{\|}_{L^{2}(I;L^{2}(({\varepsilon}{\partial}{\boldsymbol{\boxminus}}\mathbb{x})\backslash{\partial} G_{\mathbb{x}}^{{\varepsilon}}))}^{2}\\ & \leq C\big\{{\|}{\mathbb{1}}_{Q_{n, r}^{{\varepsilon}}}h(u^{{\varepsilon}}){\|}_{L^{2}(I;L^{2}(Q))}^{2}+{\varepsilon}{\|}{\mathbb{1}}_{Q_{n, r}^{{\varepsilon}}}\nabla h(u^{{\varepsilon}}){\|}_{L^{2}(I;L^{2}(Q))}^{2}\big\}\cdot{\mathcal{L}}^{d}(Q_{n, r}^{{\varepsilon}}), \end{align*}

    where

    Q_{n, r}^{{\varepsilon}}: = \big\{ x\in Q\, \vert\, {\mathrm{dist}}(x, {\partial} Q)\leq{\varepsilon} nr\big\}.

    We observe that {\mathcal{L}}^{d}(Q_{n, r}^{{\varepsilon}})\to0 as {\varepsilon}\to0 , i.e., {\mathbb{1}}_{Q_{n, r}^{{\varepsilon}}}\to0 point-wise {\mathcal{L}}^{d} -almost everywhere. We know that h(u^{{\varepsilon}})\to h(u) strongly in L^{2}(I; L^{2}(Q)) , so dominated convergence yields that the second summand also converges to 0 .

    The third summand follows from two-scale convergence, i.e., h(u_{n})\mathop \rightharpoonup \limits^{2s, \, \mu_{\mathbb{x}}}h(u_{n}) for almost every \mathbb{x} .

    Remark (On the special solution of Eq (6.16)). Our special ansatz v = \sum_{i = 1}^{d}\partial_{i}u_{n}w_{i}^{(n)} as a solution to Eq (6.16) for a fixed, given u has a long tradition, where we mention [5] for perforated domains or the more "historical" papers [22,30]. For readers familiar with periodic homogenization, let us mention that due to v being {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega) problem (6.16) is the probabilistic equivalent of

    \forall\varphi\in C_{c}(Q), \psi\in H_{per}^{1}([0, 1)^{d})\colon \qquad\int_{[0, 1)^d}\int_{Q}\varphi(x)\nabla_{y}\psi(y)A(\nabla u+\nabla_{y}u_{1})\text{d} x \text{d} y = 0 .

    Now given a u_{n}\in L^{2}(I; W^{1, 2}(Q)) , we are looking for a solution v\in L^{2}(I; L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega))) to

    \int_{0}^{T}\int_{Q}\int_{{\mathbf{G}}_{n}}\varphi_{2}\nabla_{\tilde{\omega}}\psi\cdot A(u_{n})\left(\nabla u_{n}+v\right)\text{d}{\mathbb{P}}(\mathbb{x})\text{d} x\text{d} t = 0, \,

    for arbitrary \varphi_{2}\in C^{1}([0, T]\times\overline{Q}) and \psi\in H^{1}(\Omega) with \int_{\Omega}\nabla_{\tilde{\omega}}\psi\text{d}{\mathbb{P}} = 0 . Since functions of the form \nabla_{\tilde{\omega}}\psi\vert_{{\mathbf{G}}_{n}} are dense in {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathbf{G}}_{n}\vert\Omega) for \psi\in H^{1}(\Omega) with \int_{\Omega}\nabla_{\tilde{\omega}}\psi = 0 (by definition of {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega) ), this equation actually has to hold for arbitrary \phi\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathbf{G}}_{n}\vert\Omega) . By the fundamental lemma of calculus of variations, this solution must satisfy for almost every t\in I and almost every x\in Q

    \phi\in{\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathbf{G}}_{n}\vert\Omega)\colon \quad\int_{{\mathbf{G}}_{n}}\phi\cdot A(u_{n})\left(\nabla u_{n}+v\right)\text{d}{\mathbb{P}} = 0.

    Next, we see that v: = \sum_{i = 1}^{d}\partial_{i}u_{n}w_{i}^{(n)}\in L^{2}(I; L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathbf{G}}_{n}\vert\Omega))) is indeed a solution since the w_{i}^{(n)} are cell solutions

    \begin{align*} \int_{{\mathbf{G}}_{n}}\phi\cdot A(u_{n}) & \left(\nabla u_{n}+v\right)\text{d}{\mathbb{P}} = \sum\limits_{i = 1}^{d}{\partial}_{i}u_{n}\, A(u_{n})\int_{{\mathbf{G}}_{n}}\big[e_{i}+w_{i}\big]\cdot\phi\text{d}{\mathbb{P}} = 0. \end{align*}

    On the other hand, such a solution is unique: First, observe that functions of the form \varphi_{2}\cdot\nabla_{\tilde{\omega}}\psi are dense in L^{2}(I; L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega))) . Then, we consider the coercive bilinear form on L^{2}(I; L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}(\Omega))) :

    B(v_{1}, v_{2}): = \int_{0}^{T}\int_{Q}\int_{{\mathbf{G}}_{n}}v_{1}\cdot A(u_{n})v_{2} \text{d}{\mathbb{P}}(\mathbb{x})\text{d} x\text{d} t.

    Lax–Milgram tells us that the solution v\in L^{2}(I; L^{2}(Q; {\mathcal{V}}_{{\mathrm{pot}}}^{2}({\mathbf{G}}_{n}\vert\Omega))) to

    B(\, \cdot, v) = \Big[v_{1}\mapsto-\int_{0}^{T}\int_{Q}\int_{{\mathbf{G}}_{n}}v_{1}\cdot A(u_{n})\nabla u_{n}\text{d}{\mathbb{P}}(\mathbb{x})\text{d} x\text{d} t\Big],

    is unique, which is equivalent to

    0 = B(\phi, v)+\int_{0}^{T}\int_{Q}\int_{{\mathbf{G}}_{n}}\phi\cdot A(u_{n})\nabla u_{n} = \int_{0}^{T}\int_{Q}\int_{{\mathbf{G}}_{n}}\phi\cdot A(u_{n})\big[\nabla u_{n}+v\big],

    which justifies the standard ansatz.

    Theorem 17 is a consequence of the following.

    Theorem 56 (Main theorem: homogenized limit of admissible point processes). Let {\mathbb{X}} be a stationary ergodic admissible point process with distribution {\mathbb{P}} such that {\boldsymbol{\boxminus}}{\mathbb{X}}^{\complement} is statistically connected. Under Assumption 6, let u_{n}\in L^{2}(I; W^{1, 2}(Q)) be a homogenized solution from Theorem 55 for the thinned point process {\mathbb{X}}^{(n)} .

    For any subsequence of (u_{n})_{n\in{\mathbb{N}}} , we are able to extract yet another subsequence that converges to a u\in L^{2}(I; W^{1, 2}(Q)) . This u is a (not necessarily unique) weak solution to the initial value problem

    \begin{align*} {\mathbb{P}}({\mathbf{G}}){\partial}_{t}u-\nabla\cdot\left(A(u){\mathcal{A}}\, \nabla u\right)-\lambda(\mu_{{\mathbb{X}}})h(u) & = {\mathbb{P}}({\mathbf{G}})f & & \mathit{\text{ in }}I\times Q, \\ A(u){\mathcal{A}}\nabla u\cdot\nu & = 0 & & \mathit{\text{ on }}I\times{\partial} Q, \\ u(0, x) & = {\mathbb{P}}({\mathbf{G}})u_{0}(x) & & \mathit{\text{ in }}Q. \end{align*}

    Here {\mathcal{A}} is the effective conductivity defined in Definition 37 based on the event {\mathcal{Q}} = {\mathbf{G}} = \big\{\mathbb{x}\in{\mathcal{S}}({\mathbb{R}}^{d})\, \vert\, o\notin{\boldsymbol{\boxminus}}\mathbb{x}\big\} , \Omega = {\mathcal{S}}({{{\mathbb{R}}}^{d}}) , {\mathcal{P}} = {\mathbb{P}} and \lambda(\mu_{{\mathbb{X}}}) is the intensity of \mu_{{\mathbb{X}}}: = {\mathcal{H}}_{\llcorner{\partial}{\boldsymbol{\boxminus}}{\mathbb{X}}}^{d-1} . Furthermore, with \alpha_{{\mathcal{A}}} > 0 being the smallest eigenvalue of {\mathcal{A}} and L_{h} being the Lipschitz constant of h

    \begin{align*} \underset{t \in I}{\operatorname{ess} \sup }{\|} u(t){\|}_{L^{2}(Q)}^{2} & \leq\exp(C_{1})\big[{\|} u_{0}{\|}_{L^{2}(Q)}^{2}+C_{2}\big]\\ {\|}\nabla u{\|}_{L^{2}(I;L^{2}(Q))}^{2} & \leq\frac{{\mathbb{P}}({\mathbf{G}})}{2\alpha_{{\mathcal{A}}}\inf(A)}\big(1+\exp(C_{1})\big)\big[{\|} u_{0}{\|}_{L^{2}(Q)}^{2}+C_{2}\big], \end{align*}

    where

    C_{1}: = T\big(1+\frac{\lambda(\mu_{{\mathbb{X}}})}{{\mathbb{P}}({\mathbf{G}})}(1+2L_{h})\big)\qquad\mathit{\text{and}}\qquad C_{2}: = {\|} f{\|}_{L^{2}(I;L^{2}(Q))}^{2}+2T\frac{\lambda(\mu_{{\mathbb{X}}})}{{\mathbb{P}}({\mathbf{G}})}{\big|} h(0){\big|}^{2}.

    Proof. We note that {\mathcal{A}}^{(n)} from Theorem 55 is defined with cell solutions on \Omega = {\mathtt{F}_{n}}{\mathcal{S}}({{{\mathbb{R}}}^{d}}) and the push-forward measure {\mathbb{P}}\circ{\mathtt{F}_{n}}^{-1} . We use the pull-back result from Lemma 41 to obtain a representation of {\mathcal{A}}^{(n)} in terms of \Omega = {\mathcal{S}}({{{\mathbb{R}}}^{d}}) and the original probability distribution.

    Lemmas 32 and 29 and Corollary 39 yield respectively

    \lambda(\mu_{{\mathbb{X}}^{(n)}})\to\lambda(\mu_{{\mathbb{X}}}), \qquad{\mathbb{P}}({\mathbf{G}}_{n})\to{\mathbb{P}}({\mathbf{G}}) > 0, \qquad{\mathcal{A}}^{(n)}\to{\mathcal{A}}, \qquad\alpha_{{\mathcal{A}}^{(n)}}\to\alpha_{{\mathcal{A}}} > 0,

    for {\mathbf{G}}_{n}: = \big\{\mathbb{x}\, \vert\, o\notin{\boldsymbol{\boxminus}}\mathbb{x}^{(n)}\big\} and {\mathbf{G}}: = \big\{\mathbb{x}\, \vert\, o\notin{\boldsymbol{\boxminus}}\mathbb{x}\big\} . From the a priori estimates in Theorem 55, we furthermore find

    \begin{align*} \limsup\limits_{n\to\infty}\left(\|u_{n}(t)\|_{L^{\infty}(0, T;L^{2}(Q))}^{2}+\|\nabla u_{n}\|_{L^{2}(I;L^{2}(Q))}^{2}\right) & < \infty, \\ \limsup\limits_{n\to\infty}{\|}{\partial}_{t}u_{n}{\|}_{L^{2}(I;W^{1, 2}(Q)^{*})} & < \infty, \end{align*}

    and Aubin–Lions (or more general, Theorem 52) yields pre-compactness. These uniform bounds together with compactness arguments yield the existence of u\in L^{2}(I; W^{1, 2}(Q)) with generalized time derivative {\partial}_{t}u\in L^{2}(I; W^{1, 2}(Q))^{*} , such that for a subsequence

    \begin{align*} u_{n} \mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(I;W^{1, 2}(Q))}u, \quad {\partial}_{t}u_{n} & \mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(I;W^{1, 2}(Q))^{*}}{\partial}_{t}u, \quad u_{n} \xrightarrow[n\to\infty]{L^{2}(I;L^{2}(Q))}u, \quad h(u_{n}) \xrightarrow[n\to\infty]{L^{2}(I;L^{2}(Q))}h(u), \end{align*}

    as well as

    A(u_{n}){\mathcal{A}}^{(n)}\nabla u_{n}\mathop \rightharpoonup \limits_{n\to\infty}^{L^{2}(I;L^{2}(Q))}A(u){\mathcal{A}}\nabla u.

    From here we conclude.

    In this chapter, we will establish a criterion for {\boldsymbol{\boxminus}}{\mathbb{X}}{}^{\complement} to be statistically connected (Definition 15), that is Theorem 60. To be precise, we will show that

    e_{1}^{t}{\mathcal{A}} e_{1} > 0,

    as all other directions \eta\in{\mathbb{R}}^{d} can be shown analogously via rotation. The procedure will be based on [13, Chapter 9]. The matrix {\mathcal{A}} corresponds to the matrix {\mathcal{A}}^{0} there. We will also see that {\boldsymbol{\boxminus}}{\mathbb{X}}^{\complement} is statistically connected if \Xi{\mathbb{X}}^{\complement} is statistically connected.

    Notation. Given a fixed admissible point process {\mathbb{X}} , we write in this section

    {\boldsymbol{\boxminus}}: = {\boldsymbol{\boxminus}}{\mathbb{X}}\qquad\Xi: = \Xi{\mathbb{X}}.

    Most arguments work for more general random perforations \Xi and their filled-up versions as long as \Xi has no infinite connected component (Theorem 58 needs additionally that almost surely, the bounded connected components of {\mathbb{R}}^{d}\backslash\Xi have non-zero distance to the infinite connected components). We refrain from doing so since we would need to introduce the notion of stationary random sets and the main focus here lies on point processes.

    The following theorem gives us a different point of view on the effective conductivity {\mathcal{A}} :

    Theorem 57 (Variational formulation [13, Theorem 9.1]). For every ergodic admissible point process, we have almost surely and for every \eta\in{\mathbb{R}}^{d} :

    \begin{align*} \eta^{t}{\mathcal{A}}\eta = \lim\limits_{n\to\infty}n^{-d}\inf\limits_{v\in C_{0}^{\infty}([0, n]^{d})}\int_{[0, n]^{d}\backslash\Xi}{\big|}\eta-\nabla v{\big|}^{2}\text{ d} x, \, \end{align*}

    where {\mathcal{A}} is the effective conductivity based on the event \{o\notin\Xi\} .

    The first observation we can make is that the effective conductivity depends monotonously on the domain: The larger the set of holes, the lower the effective conductivity. The question arises in which cases this term becomes 0. This should only happen if {\mathbb{R}}^{d}\backslash\Xi is "insufficiently connected". Intuitively, we want v\approx-\eta\cdot x+\text{const} , but at the same time, v needs to be 0 at the boundaries. If our region is badly connected, we can hide large gradients inside the holes, see, e.g., Figure 3. As in [13], we will see that the existence of sufficiently many "channels" connecting the left to the right side of a box [0, n]^{d} will ensure e_{1}^{t}{\mathcal{A}} e_{1} > 0 . Before we do that, we establish an important fact:

    Figure 3.  High versus low conductivity. The balls represent \Xi . The white area corresponds to v\approx-\eta\cdot x+const . Black lines indicate large contributions to \int_{[0, n]^{d}\backslash\Xi}{\big|}\eta+\nabla v{\big|}^{2}\text{d} x .

    We have defined statistical connectedness (Definition 15) via the filled-up Boolean model {\boldsymbol{\boxminus}} . Unfortunately, filling up holes is non-local (depending on the size of holes), which is troublesome on the stochastic side. However, an analogue of [13, Lemma 9.7] tells us that the effective conductivity of both the Boolean model \Xi and its filled-up version {\boldsymbol{\boxminus}} are the same.

    Theorem 58 (Filling up holes preserves the effective conductivity). For every ergodic admissible point process, we have almost surely

    \begin{align*} \eta^{t}{\mathcal{A}}\eta & = \lim\limits_{n\to\infty}n^{-d}\inf\limits_{v}\int_{[0, n]^{d}\backslash{\boldsymbol{\boxminus}}}{\big|}\eta-\nabla v{\big|}^{2}\text{ d} x = \lim\limits_{n\to\infty}n^{-d}\inf\limits_{v}\int_{[0, n]^{d}\backslash\Xi}{\big|}\eta-\nabla v{\big|}^{2}\text{ d} x, \end{align*}

    where the infimum is over v\in C_{0}^{\infty}([0, n]^{d}) .

    Proof. As mentioned before, this is a variation of [13, Lemma 9.7] fitted to our purpose. Let

    K_{n}^{s} be the set of islands (i.e., connected components in {\mathbb{R}}^{d}\backslash\Xi of finite diameter) of diameter \leq s that intersect but do not lie inside [0, n]^{d} .

    L_{n}^{s} be the set of islands of diameter > s that do not completely lie inside [0, n]^{d} and that are encircled by a \Xi -cluster of size larger than s .

    All the islands in K_{n}^{s} and L_{n}^{s} belong to connected components of {\mathbb{R}}^{d}\backslash\Xi different from {\boldsymbol{\boxminus}}^{\complement} (the unique unbounded connected component). Since {\mathbb{X}} is admissible, almost surely they all have non-zero distance to {\boldsymbol{\boxminus}}^{\complement} . Therefore, the following infimum decomposes, with all the infima being over v\in C_{0}^{\infty}([0, n]^{d})

    \begin{align*} \inf\limits_{v}\int_{[0, n]^{d}\backslash\Xi}{\big|}\eta-\nabla v{\big|}^{2}\text{d} x & = \inf\limits_{v}\int_{([0, n]^{d}\backslash\Xi)\backslash(K_{n}^{s}\cup L_{n}^{s})}{\big|}\eta-\nabla v{\big|}^{2}\text{d} x+\inf\limits_{v}\int_{K_{n}^{s}\cup L_{n}^{s}}{\big|}\eta-\nabla v{\big|}^{2}\text{d} x\\ & = \inf\limits_{v}\int_{([0, n]^{d}\backslash{\boldsymbol{\boxminus}})\backslash(K_{n}^{s}\cup L_{n}^{s})}{\big|}\eta-\nabla v{\big|}^{2}\text{d} x+\inf\limits_{v}\int_{K_{n}^{s}\cup L_{n}^{s}}{\big|}\eta-\nabla v{\big|}^{2}\text{d} x\\ & = \inf\limits_{v}\int_{[0, n]^{d}\backslash{\boldsymbol{\boxminus}}}{\big|}\eta-\nabla v{\big|}^{2}\text{d} x+C, \end{align*}

    with

    {\big|} C{\big|}\leq{\big|}\eta{\big|}^2\left[{\mathcal{L}}^{d}(K_{n}^{s})+{\mathcal{L}}^{d}(L_{n}^{s})\right],

    and where the second equality comes from the fact that filling up islands that lie completely inside [0, n]^{d} does not change the value of the infimum. Now, we observe

    {\mathcal{L}}^{d}(K_{n}^{s})\sim O(n^{d-1}) for fixed s , so \lim_{n\to\infty}n^{-d}{\mathcal{L}}^{d}(K_{n}^{s}) = 0 and

    ● denoting by L^{s} all islands of diameter greater than s , we have almost surely

    \lim\limits_{n\to\infty}n^{-d}{\mathcal{L}}^{d}(L_{n}^{s})\leq\lim\limits_{n\to\infty} n^{-d}{\mathcal{L}}^d(L^s\cap [0, n]^d) = \colon \text{density}(L^{s}) = {\mathbb{P}}(o\in L^s).

    All islands are of finite size, yielding \bigcap_{s\in{\mathbb{N}}}L^{s} = \emptyset . Hence, \lim_{s\to\infty}\text{density}(L^{s}) = 0 .

    Choosing s sufficiently large finishes the proof.

    Definition 59 (Percolation channels (see Figure 4)). Fix a k_{\mathrm{scale}}\in{\mathbb{N}} . We consider the lattice {\mathbb{Z}}_{n}^{d}\subset{\mathbb{Z}}^{d} and the cube with corner z = (z_{1}, \dots, z_{d})\in{\mathbb{Z}}^{d}

    Figure 4.  Percolation channels for different k_{\mathrm{scale}} .
    \begin{align*} {\mathbb{Z}}_{n}^{d}: & = {\mathbb{Z}}^{d}\cap[0, n)^{d}\qquad{\text{and}}\qquad{\mathcal{K}}_{z}: = \mathop \times \limits_{i = 1}^{d} [z_{i}, z_{i}+1] \end{align*}

    and call two vertices z, z' neighbors if their l^{1} -distance is equal to 1 .

    We call z open if

    \Xi\cap k_{\mathrm{scale}}^{-1}{\mathcal{K}}_{z} = \emptyset.

    An open left-right crossing \gamma = (z^{(1)}, \dots, z^{(l)}) of {\mathbb{Z}}_{n}^{d} is called a percolation channel in {\mathbb{Z}}_{n}^{d} , i.e.,

    i) z^{(i)}, z^{(i+1)} are neighbors for each i < l ,

    ii) all the z^{(i)} are open, and

    iii) z_{1}^{(1)} = 0 and z_{1}^{(l)} = n-1 .

    We define the quantity (depending on the random \Xi and on k_{{{{{\mathrm{scale}}}}}} )

    \begin{align*} \mathbf{N}(n): & = \max\big\{ j\, \vert\, \gamma_{1}, \dots, \gamma_{j}{\text{ are disjoint percolation channels in }}{\mathbb{Z}}_{n}^{d}\big\}\\ & = {\text{"maximal number of disjoint percolation channels in }} {{\mathbb{Z}}_{n}^{d}}" \end{align*}

    and the tube L(\gamma) corresponding to the path \gamma = (z^{(1)}, \dots, z^{(l)}) as

    L(\gamma): = \bigcup\limits_{i\leq l}k_{{{{{\mathrm{scale}}}}}}^{-1}{\mathcal{K}}_{z^{(i)}}.

    Statistical connectedness of {\mathbb{R}}^{d}\backslash\Xi then reads as follows:

    Theorem 60 (Percolation channels imply conductivity). For almost every realization \mathbb{x} of an ergodic admissible point process, we have for \Xi = \Xi\mathbb{x}

    \lim\limits_{n\to\infty}n^{-d}\inf\limits_{v\in C_{0}^{\infty}([0, n]^{d})}\int_{[0, n]^{d}\backslash\Xi}{\big|} e_{1}-\nabla v{\big|}^{2}\text{ d} x\geq\limsup\limits_{n\to\infty}\Big(\frac{\mathbf{N}(n)}{n^{d-1}}\Big)^{2}.

    In particular, the effective conductivity is strictly positive if almost surely

    \begin{equation} \limsup\limits_{n\to\infty}\frac{\mathbf{N}(n)}{n^{d-1}} > 0. \end{equation} (8.1)

    Proof. This is an analogue of [13, Theorem 9.11] and relies on defining a suitable vector field \overrightarrow{F}_{\gamma}\colon [0, k_{{{{\mathrm{scale}}}}}^{-1}n]^{d}\to{\mathbb{R}}^{d} inside channels \gamma = (z^{(1)}, \dots, z^{(l)}) on {\mathbb{Z}}_{n}^{d} . We want \overrightarrow{F}_{\gamma} to satisfy the following

    {\big|}\overrightarrow{F}_{\gamma}(x){\big|} = 1 for every x inside the tube L(\gamma) and \overrightarrow{F}_{\gamma}(x) = 0 outside.

    \overrightarrow{F}_{\gamma} is parallel to {\partial} L(\gamma) except on corners as well as {\partial} L(\gamma)_{-}: = {\mathcal{K}}_{z^{(1)}}\cap\{x_{1} = 0\} and {\partial} L(\gamma)_{+}: = {\mathcal{K}}_{z^{(l)}}\cap\{x_{1} = k_{{{{\mathrm{scale}}}}}^{-1}n\} .

    \overrightarrow{F}_{\gamma}(x) = e_{1} for x\in{\partial} L(\gamma)_{-}\cup{\partial} L(\gamma)_{+} .

    ● For the standard normal vector \nu to {\partial} L(\gamma) :

    \begin{equation} \int_{L(\gamma)}(e_{1}-\nabla v)\cdot\overrightarrow{F}_{\gamma}\text{d} x = \int_{{\partial} L(\gamma)}(x_{1}-v)\overrightarrow{F}_{\gamma}\cdot\nu\text{d} \mathcal{H}^{d-1}(x). \end{equation} (8.2)

    Thus, the vector field was deliberately chosen such that the Gauß divergence theorem only yields contributions from the "starting" and "ending" surfaces. Figure 5 illustrates how \overrightarrow{F}_{\gamma} can be chosen to satisfy these properties.

    Figure 5.  The vector field \overrightarrow{F} simply follows along the direction of the path \gamma . Whenever the tube makes a turn, we divide the corresponding cube diagonally as depicted with the vector field remaining constant in each respective half. The resulting \overrightarrow{F} is piecewise constant and Eq (8.2) holds via the Gauß divergence theorem as contributions on the diagonal surfaces cancel out: We have a positive contribution from the incoming arrows and a negative contribution from the outgoing one.

    The rest is simple. Take \gamma_{1}, \dots, \gamma_{\mathbf{N}(n)} disjoint nonself-intersecting channels in {\mathbb{Z}}_{n}^{d} . Set

    T: = \bigcup\limits_{i\leq\mathbf{N}(n)}L(\gamma_{i})\subset[0, k_{{{{\mathrm{scale}}}}}^{-1}n]^{d}, \qquad\overrightarrow{F}: = \sum\limits_{i\leq\mathbf{N}(n)}\overrightarrow{F}_{\gamma_{i}}.

    Then,

    \begin{align*} \int_{[0, k_{{{{\mathrm{scale}}}}}^{-1}n]^{d}\backslash\Xi}{\big|} e_{1}-\nabla v{\big|}^{2}\text{d} x & \geq\int_{T}{\big|} e_{1}-\nabla v{\big|}^{2}\text{d} x\geq\int_{T}{\big|}(e_{1}-\nabla v)\cdot\overrightarrow{F}{\big|}^{2}\text{d} x\\ & \geq\frac{1}{{\mathcal{L}}^{d}(T)}\left(\int_{T}(e_{1}-\nabla v)\cdot\overrightarrow{F}\text{d} x\right)^{2}\geq\frac{k_{{{{\mathrm{scale}}}}}^{d}}{n^{d}}\left(\int_{T}(e_{1}-\nabla v)\cdot\overrightarrow{F}\text{d} x\right)^{2}. \end{align*}

    For a fixed tube L = L(\gamma_{i}) , we have

    \begin{align*} \int_{L}(e_{1}-\nabla v)\cdot\overrightarrow{F}\text{d} x & = \int_{{\partial} L}(x_{1}-v)\overrightarrow{F}\cdot\nu\text{d}{\mathcal{H}}^{d-1}(x) = \int_{{\partial} L_{-}\cup{\partial} L_{+}}(x_{1}-v)\overrightarrow{F}\cdot\nu\text{d}{\mathcal{H}}^{d-1}(x)\\ & = \int_{{\partial} L_{+}}k_{{{{\mathrm{scale}}}}}^{-1}ne_{1}\cdot e_{1}\text{d}{\mathcal{H}}^{d-1}(x) = k_{{{{\mathrm{scale}}}}}^{-1}n{\mathcal{H}}^{d-1}({\partial} L_{+}) = k_{{{{\mathrm{scale}}}}}^{-d}n. \end{align*}

    Therefore,

    \begin{align*} \int_{[0, k_{{{{\mathrm{scale}}}}}^{-1}n]^{d}\backslash\Xi)}{\big|} e_{1}-\nabla v{\big|}^{2}\text{d} x & \geq\frac{k_{{{{\mathrm{scale}}}}}^{d}}{n^{d}}\big(k_{{{{\mathrm{scale}}}}}^{-d}n\mathbf{N}(n)\big)^{2}, \end{align*}

    and so

    \big(k_{{{{\mathrm{scale}}}}}^{-1}n\big)^{-d}\int_{[0, k_{{{{\mathrm{scale}}}}}^{-1}n]^{d}\backslash\Xi}{\big|} e_{1}-\nabla v{\big|}^{2}\text{d} x\geq\left(\frac{\mathbf{N}(n)}{n^{d-1}}\right)^{2}.

    Passing to the \limsup finishes the proof.

    Remark 61 ( d = 2 and bottom-top crossings). Let \mathbf{L}(n) be the minimal number of open vertices that a l^{\infty} -bottom-top crossing of {\mathbb{Z}}_{n}^{2} must have. It turns out that in d = 2

    \mathbf{L}(n) = \mathbf{N}(n)\,

    (see Lemma 73). We will use this to show Eq (8.1) for the Poisson point process {\mathbb{X}}_{{\mathrm{poi}}} .

    The driving force behind this work has been a stationary Poisson point process {\mathbb{X}}_{{\mathrm{poi}}} . It is known that the Poisson point process is ergodic (even mixing) and its high spatial independence makes it the canonical random point process. As pointed out before though, \Xi{\mathbb{X}}_{{\mathrm{poi}}} gives rise to numerous analytical issues which prevent the usage of the usual homogenization tools.

    The main theorem (Theorem 56) tells us that homogenization is still reasonable for highly irregular filled-up Boolean models {\boldsymbol{\boxminus}}{\mathbb{X}} driven by admissible point processes {\mathbb{X}} .

    It is known for {\mathbb{X}}_{{\mathrm{poi}}} that there exists some critical radius r_{c}: = r_{c}[\lambda({\mathbb{X}}_{{\mathrm{poi}}})]\in(0, \, \infty) such that

    {\boldsymbol{\boxminus}}{\mathbb{X}}_{{\mathrm{poi}}} only consists of finite clusters for r < r_{c} (subcritical regime) and

    {\boldsymbol{\boxminus}}{\mathbb{X}}_{{\mathrm{poi}}} has a unique infinite cluster for r > r_{c} (supercritical regime).

    The behavior at criticality r = r_{c} is still a point of research. For details, we refer to [14] for the Poisson point process {\mathbb{X}}_{{\mathrm{poi}}} and [18, Chapter 3] for the Boolean model \Xi{\mathbb{X}}_{{\mathrm{poi}}} .

    We will see in the subcritical regime that

    i) {\mathbb{X}}_{{\mathrm{poi}}} is an ergodic admissible point process and

    ii) \Xi{\mathbb{X}}_{{\mathrm{poi}}}^{\complement} is statistically connected, which is equivalent to {\boldsymbol{\boxminus}}{\mathbb{X}}_{{\mathrm{poi}}}^{\complement} being statistically connected (see Theorem 58).

    We therefore make the following assumption for the rest of this section:

    Assumption 62 (subcritical regime). We assume that

    r < r_{c}.

    Remark 63 (Scaling relation). r_{c} has the following scaling relation

    r_{c}[k^{d}\cdot\lambda({\mathbb{X}}_{{\mathrm{poi}}})] = r_{c}[\lambda(k^{-1}{\mathbb{X}}_{{\mathrm{poi}}})] = k^{-1}r_{c}[\lambda({\mathbb{X}}_{{\mathrm{poi}}})].

    The Mecke–Slivnyak theorem tells us that the Palm probability measure (Theorem 45) of a stationary Poisson point process is just a Poisson point process with a point added in the origin. This gives us the following lemma:

    Lemma 64 (Equidistance property). The stationary Poisson point process {\mathbb{X}}_{{\mathrm{poi}}} satisfies the equidistance property for arbitrary r > 0 , i.e.,

    {\mathbb{P}}{\left({\exists x, y\in{\mathbb{X}}_{{\mathrm{poi}}}\, \vert\, d(x, y) = 2r}\right)} = 0.

    Proof. This follows from using the Palm theorem Theorem 45 on

    f(x, \mathbb{x}): = \sum\limits_{x_{i}\in\mathbb{x}}{\mathbb{1}}\left\{ d(x, x_{i}) = 2r\right\},

    and the Mecke–Slivnyak theorem [14, Theorem 9.4].

    Corollary 65 ( {\mathbb{X}}_{{\mathrm{poi}}} is admissible). Under Assumption 62, {\mathbb{X}}_{{\mathrm{poi}}} is an admissible point process.

    Proof. {\mathbb{X}}_{{\mathrm{poi}}} is not just ergodic, but even mixing (see [14, Theorem 8.13]). The equidistance property has been proven in Corollary 64. Finiteness of clusters follows from the subcritical regime (Assumption 62).

    Proving the statistical connectedness of \Xi{\mathbb{X}}_{{\mathrm{poi}}}{}^{\complement} (Definition 15) is much harder and does not immediately follow from readily available results. Our procedure is as follows:

    i) We employ the criterion from Section 8. Therefore, we will check that there are sufficiently many percolation channels for \Xi{\mathbb{X}}_{{\mathrm{poi}}} .

    ii) Using the spatial independence of the Poisson point process {\mathbb{X}}_{{\mathrm{poi}}} , we show that it is sufficient to only consider 2 -dimensional slices.

    iii) We show the statement in d = 2 using ideas in [11, Chapter 11]. There, the result has been proven for certain iid fields on planar graphs, including {\mathbb{Z}}^{2} .

    Additionally to Assumption 62, we need sufficient discretization for the percolation channels:

    Assumption 66 (Sufficient scaling). Let k_{\mathrm{scale}}\in{\mathbb{N}} be large enough such that, for the critical radius r_{c} ,

    (r_{c}-r)/2 > \sqrt{d}k_{\mathrm{scale}}^{-1},

    e.g., k_{\mathrm{scale}}: = \lceil2\frac{\sqrt{d}}{r_{c}-r}\rceil+1 .

    Definition 67 (Recap and random field (X_{z})_{z\in{\mathbb{Z}}^{d}} ). Recall Definition 59, most importantly

    \begin{align*} {\mathbb{Z}}_{n}^{i}: & = {\mathbb{Z}}^{i}\cap[0, n)^{i}\qquad{\text{and}}\qquad{\mathcal{K}}_{z}: = \mathop \times \limits_{i = 1}^{d} [z_{i}, z_{i}+1], \end{align*}

    as well as the notion of percolation channels for k_{\mathrm{scale}} and

    \begin{align*} \mathbf{N}(n): & = {\text{"maximal number of disjoint percolation channels in }} {{\mathbb{Z}}_{n}^{d}}". \end{align*}

    We define the random field

    (X_{z})_{z\in{\mathbb{Z}}^{d}}: = \Big({\mathbb{1}}\{\Xi{\mathbb{X}}_{{\mathrm{poi}}}\cap k_{\mathrm{scale}}^{-1}{\mathcal{K}}_{z} = \emptyset\}\Big){}_{z\in{\mathbb{Z}}^{d}}.

    We say that z\in{\mathbb{Z}}^{d} is blocked if X_{z} = 0 and open if X_{z} = 1 (this is consistent with Definition 59).

    Theorem 68 (Percolation channels of the Poisson point process). Under Assumptions 62 and 66, there is a C > 0 such that Eq (8.1) holds, i.e.,

    \begin{align*} {\mathbb{P}}\big(\limsup\limits_{n\to\infty}n^{1-d}\mathbf{N}(n)\geq C\big) = 1. \end{align*}

    In particular, {\boldsymbol{\boxminus}}{\mathbb{X}}_{{\mathrm{poi}}}{}^{\complement} is statistically connected (see Theorem 60).

    The rest of the section deals with the proof of Theorem 68. It will follow as a direct consequence of Lemma 70 (reduction to d = 2 ) and Lemma 74 (main result for d = 2 ) which are given later.

    For disjoint U_{1}, \, U_{2}, \dots\, \subset{\mathbb{R}}^{d} and events A_{i} only depending on {\mathbb{X}}_{{\mathrm{poi}}} inside U_{i} , we know that (A_{i})_{i} is an independent family. This is one of the striking properties of a Poisson point process and we will heavily make use of it. The Boolean model \Xi{\mathbb{X}}_{{\mathrm{poi}}} for radius r still retains this property in a slightly weaker form and correspondingly the random field (X_{z})_{z\in{\mathbb{Z}}^{d}} :

    Lemma 69 (Independence in large distances). Let A, B\subset{\mathbb{Z}}^{d} such that

    \begin{equation} d^{\infty}(A, B): = \min\limits_{z_{a}\in A, z_{b}\in B}{\|} z_{b}-z_{a}{\|}_{\infty}\geq2rk_{\mathrm{scale}}+1. \end{equation} (9.1)

    Then, (X_{z})_{z\in A} and (X_{z})_{z\in B} are independent.

    Proof. (X_{z})_{z\in A} is only affected by points of {\mathbb{X}}_{{\mathrm{poi}}} inside

    U_{A}: = \bigcup\limits_{z\in A}\mathbb{B}_{r}{\left({k_{{{{\mathrm{scale}}}}}^{-1}{\mathcal{K}}_{z}}\right)}.

    The same holds for (X_{z})_{z\in B} and we check that Eq (9.1) implies U_{A}\cap U_{B} = \emptyset .

    Lemma 70 ( 2 -dimensional percolation channels imply channel property for d > 2 ). For \tilde{z}\in{\mathbb{Z}}^{d-2} , we define (compare to Definition 59)

    \begin{align*} \mathbf{N}_{\tilde{z}}^{(2)}(n): & = \mathit{\text{"maximal number of disjoint percolation channels in }} {{\mathbb{Z}}_{n}^{2}\times\tilde{z}}". \end{align*}

    If there are \tilde{C}, \, p_{0} > 0 such that for some \tilde{z}\in{\mathbb{Z}}^{d-2} ,

    \begin{equation} \limsup\limits_{n\to\infty}{\mathbb{P}}{\left({\mathbf{N}_{\tilde{z}}^{(2)}(n)\geq\tilde{C}n}\right)} > p_{0} > 0, \end{equation} (9.2)

    then there exists a C > 0 such that

    \limsup\limits_{n\to\infty}{\mathbb{P}}{\left({\mathbf{N}(n)\geq Cn^{d-1}}\right)} = {\mathbb{P}}\big(\limsup\limits_{n\to\infty}n^{1-d}\mathbf{N}(n)\geq C > 0\big) = 1.

    (This proof heavily relies on the independence structure of the Poisson point process, i.e., Lemma 69.)

    Proof. {\mathbb{X}}_{{\mathrm{poi}}} is stationary, so for distinct \tilde{z}_{1}, \, \tilde{z}_{2}\in{\mathbb{Z}}^{d-2} ,

    p(n): = {\mathbb{P}}{\left({\mathbf{N}_{\tilde{z}_{1}}^{(2)}(n)\geq\tilde{C}n}\right)} = {\mathbb{P}}{\left({\mathbf{N}_{\tilde{z}_{2}}^{(2)}(n)\geq\tilde{C}n}\right)}.

    Let k: = \lceil2rk_{{{{\mathrm{scale}}}}}\rceil+1 . By Lemma 69, the events on {\mathbb{Z}}^{2}\times(k\tilde{z}_{1}) are independent from the events on {\mathbb{Z}}^{2}\times(k\tilde{z}_{2}) . Therefore, \big({\mathbb{1}}\{\mathbf{N}_{k\tilde{z}}^{(2)}(kn)\geq\tilde{C}n\}\big)_{\tilde{z}\in{\mathbb{Z}}^{d-2}} is an iid family of Bernoulli random variables with parameter p(n) . Then,

    \begin{align*} {\mathbb{P}}\Big(\mathbf{N}(kn) \geq\frac{\tilde{C}p_{0}}{2k^{d-2}}(kn)^{d-1}\Big) &\geq {\mathbb{P}}\big(\text{For at least } {p_{0}/2} \text{ of the }\tilde{z}\in{\mathbb{Z}}_{n}^{(d-2)}\colon \mathbf{N}_{\tilde{z}}^{(2)}(kn)\geq\tilde{C}kn\big)\\ & = {\mathbb{P}}\Big(\frac{1}{\#{\mathbb{Z}}_{n}^{(d-2)}}\sum\limits_{\tilde{z}\in{\mathbb{Z}}_{n}^{(d-2)}}{\mathbb{1}}\{\mathbf{N}_{k\tilde{z}}^{(2)}(kn)\geq\tilde{C}n\}\geq\frac{1}{2}p_{0}\Big). \end{align*}

    By Eq (9.2) and the law of large numbers, we get

    \limsup\limits_{n\to\infty}{\mathbb{P}}\Big(\mathbf{N}(n)\geq\frac{\tilde{C}p_{0}}{2k^{d-2}}n{}^{d-1}\Big)\geq\limsup\limits_{n\to\infty}{\mathbb{P}}\Big(\mathbf{N}(kn)\geq\frac{\tilde{C}p_{0}}{2k^{d-2}}(kn)^{d-1}\Big) = 1.

    Setting C = \frac{\tilde{C}p_{0}}{2k^{d-2}} , we obtain Eq (8.1) after checking

    {\mathbb{P}}\big(\limsup\limits_{n\to\infty}n^{1-d}\mathbf{N}(n)\geq C\big) = \limsup\limits_{n\to\infty}{\mathbb{P}}\big(n^{1-d}\mathbf{N}(n)\geq C\big) = 1,

    which finishes the proof.

    Remark 71. Spatial independence is needed to move from d = 2 to d\geq3 . The strong independence properties of {\mathbb{X}}_{{\mathrm{poi}}} allow far weaker conditions on \mathbf{N}^{(2)}(n) (positive probability) than on \mathbf{N}(n) (probability 1). Either way, Lemma 74 shows that {\mathbb{P}}(\mathbf{N}^{(2)}(n) < Cn) drops exponentially in n .

    As shown before, we may limit ourselves to a fixed lattice {\mathbb{Z}}^{2}\times0_{{\mathbb{Z}}^{d-2}}\simeq{\mathbb{Z}}^{2} . Therefore, we will often suppress the "anchor point" 0_{{\mathbb{Z}}^{d-2}} and just act like we are in {\mathbb{Z}}^{2} . Our random field from Definition 67 is then by abuse of notation

    (X_{z})_{z\in{\mathbb{Z}}^{2}}\simeq(X_{z})_{z\in{\mathbb{Z}}^{2}\times0_{{\mathbb{Z}}^{d-2}}}.

    Definition 72 (Vertical crossings). Consider the ({\mathbb{Z}}^{2}, l^{\infty}) -lattice, that is z, z' are neighbors if {\|} z-z'{\|}_{\infty} = \; 1 .

    An l^{\infty} -bottom-top crossing in {\mathbb{Z}}_{n}^{2} is called a vertical crossing. We call a path blocked if all its vertices are blocked. We define the quantity

    \mathbf{L}(n): = {\text{"minimal number of open vertices in a vertical crossing in }} {{\mathbb{Z}}_{n}^{2}}".

    (The percolation channels lie on the l^{1} -graph, while the vertical crossings lie on the l^{\infty} -graph.)

    We may work with single vertical crossings instead of collections of percolation channels:

    Lemma 73 (Percolation channels vs vertical crossings (see Figure 6)). It holds that

    \mathbf{N}(n) = \mathbf{L}(n).
    Figure 6.  Disjoint percolation channels vs. vertical crossings. On the left side, we see that we can only have at most two l^{1} -channels. The right figure shows that any l^{\infty} -vertical crossing must contain at least two open vertices.

    Proof. See the proof of [11, Theorem 11.1] based on Menger's theorem and [11, Proposition 2.2].

    The main work is proving the following equivalent of [11, Proposition 11.1]:

    Lemma 74 (Open vertices in vertical crossings). Under Assumptions 62 and 66, there are C_{i} > 0 such that

    \begin{align*} & {\mathbb{P}}\big(\exists o\leadsto{\mathbb{Z}}\times\{n\}\mathit{\text{ with at most }}C_{1}n\mathit{\text{ open vertices }}\big)\leq C_{2}\exp\big(-C_{3}n\big), \end{align*}

    in particular,

    {\mathbb{P}}\big(\mathbf{N}(n)\geq C_{1}n\big)\geq1-C_{2}n\exp(-C_{3}n).

    The proof relies on a reduction scheme of the path \gamma\colon o\leadsto{\mathbb{Z}}\times\{n\} . We divide \gamma into several segments which must either contain an open vertex or contain a blocked path of large diameter. Since we are in the subcritical regime, the probability of such paths decreases exponentially in their diameter:

    Lemma 75 (Diameter of blocked paths). Let z\in{\mathbb{Z}}^{2} . Under Assumption 62 and k_{\mathrm{scale}}\in{\mathbb{N}} as in Assumption 66, there are C_{i} > 0 such that

    {\mathbb{P}}\big(\exists\mathit{\text{blocked path}}\gamma, \, z\in\gamma, \, \mathrm{diam}(\gamma)\geq n\big)\leq C_{1}\exp\big(-C_{2}n\big),

    where

    \mathrm{diam}(\gamma): = \max\limits_{z_{1}, z_{2}\in\gamma}{\|} z_{1}-z_{2}{\|}_{2}.

    Proof. Consider the Boolean model for radius R: = \frac{1}{2}(r+r_{c}) < r_{c} , i.e.,

    \Xi^{(R)}{\mathbb{X}}_{{\mathrm{poi}}}: = \mathbb{B}_{R}{\left({{\mathbb{X}}_{{\mathrm{poi}}}}\right)}.

    Let \gamma = (z^{(1)}, \dots, z^{(l)}) be a blocked path in {\mathbb{Z}}^{2} containing z with diameter \geq n . Since \gamma is blocked and R-r > \sqrt{d}k_{{{{\mathrm{scale}}}}}^{-1} (Assumption 66), we find for every 1\leq i\leq l some x_{i}\in{\mathbb{X}}_{{\mathrm{poi}}} such that

    \begin{align*} {\mathcal{K}}_{z^{(i)}}\cap & k_{{{{\mathrm{scale}}}}}\mathbb{B}_{r}{\left({x_{i}}\right)}\neq\emptyset, \end{align*}

    and therefore

    {\mathcal{K}}_{z^{(i)}}\subset k_{{{{\mathrm{scale}}}}}\mathbb{B}_{R}{\left({x_{i}}\right)}.

    Connecting all the z^{(i)} by a straight line, we obtain a continuous path inside k_{{{{\mathrm{scale}}}}}\Xi^{(R)}{\mathbb{X}}_{{\mathrm{poi}}} . In particular, they all belong to the same k_{{{{\mathrm{scale}}}}}\Xi^{(R)}{\mathbb{X}}_{{\mathrm{poi}}} -cluster. Then,

    \begin{align*} {\mathbb{P}}\big(\exists & \text{closed path }\gamma, \, z\in\gamma\text{ and }\mathrm{diam}(\gamma)\geq n\big)\\ & \leq{\mathbb{P}}\big(z\text{ lies in a cluster in }k_{{{{\mathrm{scale}}}}}\Xi^{(R)}{\mathbb{X}}_{{\mathrm{poi}}}\text{ of diameter }\geq n\big)\\ & \leq C_{1}\exp(-C_{2}n), \end{align*}

    since the occurrence of large clusters drops exponentially in the diameter ([18, Lemma 2.4]).

    Let n\in{\mathbb{N}} . As pointed out before, we follow the procedure in [11, Proposition 11.1] adjusted to the continuum setting. We define A(z, k) for z\in{\mathbb{Z}}^{2} and k\in{\mathbb{N}} as

    \begin{align*} A(z, k): = \big\{ & \exists l^{\infty}\text{-path }z\leadsto{\mathbb{Z}}\times\{n\}\text{ with at most }k\text{ open vertices}\big\}.\text{ } \end{align*}

    The idea is to break up the path o\leadsto{\mathbb{Z}}\times\{n\} into multiple segments (see Figure 7). In each segment, we can either reduce k by 1 or employ Lemma 75. We set

    \begin{align*} \tilde{s}: & = \lceil2rk_{{{{\mathrm{scale}}}}}\rceil+1\\ B_{1}^{\infty}(z, s): & = \big\{ v\in{\mathbb{Z}}^{2}\, \vert\, {\|} z-v{\|}_{\infty}\leq s\big\}\\ B_{2}^{\infty}(z, s): & = \big\{ v\in{\mathbb{Z}}^{2}\, \vert\, {\|} z-v{\|}_{\infty}\leq s+\tilde{s}\big\}\\ D^{\infty}(z, s): & = \big\{ v\in{\mathbb{Z}}^{2}\, \vert\, {\|} z-v{\|}_{\infty} = s+\tilde{s}+1\big\} = \text{"boundary of }B_{2}^{\infty}(z, s)". \end{align*}
    Figure 7.  B_{1}^{\infty}, \, B_{2}^{\infty}, \, D^{\infty} and decomposition of paths (left-right crossing instead of top-bottom).

    These boxes are defined so that the following holds: For fixed z\in{\mathbb{Z}}^{2} , we have by Lemma 69 that the random variables (X_{v})_{v\in B_{1}^{\infty}(z, s)} and (X_{v})_{v\in{\mathbb{Z}}^{2}\backslash B_{2}^{\infty}(z, s)} are independent. That means the state of the vertices in B_{1}^{\infty}(z, s) is independent from the state of the vertices in {\mathbb{Z}}^{2}\backslash B_{2}^{\infty}(z, s) = B_{2}^{\infty}(z, s)^{\complement} . Additionally, we define the probability

    \begin{align*} g(z, s): = & {\mathbb{P}}\big(\exists z\leadsto B_{1}^{\infty}(z, s)^{\complement}\text{ blocked inside }B_{1}^{\infty}(z, s)\big). \end{align*}

    The key inequality for the iteration in k is the following

    \begin{equation} {\mathbb{P}}\big(A(z, k)\big)\leq\sum\limits_{v\in D^{\infty}(z, s)}\Big[g(z, s){\mathbb{P}}\big(A(v, k)\big)+{\mathbb{P}}\big(A(v, k-1)\big)\Big], \end{equation} (9.3)

    for z = (z_{1}, z_{2})\in{\mathbb{Z}}^{2} whenever z_{2} < n-(s+\tilde{s}) .

    Proof of Eq (9.3). Consider the event that for some v\in D^{\infty}(z, s) , we find a path v\leadsto{\mathbb{Z}}\times\{n\} that has at most k-1 open vertices, i.e.,

    E: = \bigcup\limits_{v\in D^{\infty}(z, s)}A(v, k-1).

    Now assume that the event A(z, k)\backslash E happens. Take a path \gamma = (z, v^{(1)}, \dots, v^{(j)}) with v_{2}^{(j)} = n and at most k of the v^{(i)} being open. Let i_{1} be the last index with v^{(i_{1})}\in D^{\infty}(z, s) . This i_{1} exists since z_{2} < n-(s+\tilde{s}) , so \gamma has to pass by D^{\infty}(z, s) to reach {\mathbb{Z}}\times\{n\} . For this i_{1} , we know that (v^{(i_{1})}, \dots, v^{(j)}) completely lies in B_{2}^{\infty}(z, s)^{\complement} . Since E does not happen, it must have k open vertices. (z, v^{(1)}, \dots, v^{(i_{1})}) is a path from z to B_{1}^{\infty}(z, s)^{\complement} that is blocked everywhere except its end. Therefore,

    \begin{align*} A(z, k)\backslash E\subset\bigcup\limits_{v\in D^{\infty}(z, s)}\big\{ & \exists z\leadsto B_{1}^{\infty}(z, s)^{\complement}\text{ blocked in }B_{1}^{\infty}(z, s)\text{ and}\\ & \exists v\leadsto{\mathbb{Z}}\times\{n\}\text{ in }B_{2}^{\infty}(z, s)^{\complement}\text{ with }\text{at most }k\text{ open vertices}\big\}. \end{align*}

    As previously mentioned, the events in B_{1}^{\infty}(z, s) and B_{2}^{\infty}(z, s)^{\complement} are independent from each other. This gives us

    \begin{align*} {\mathbb{P}}\big(A(z, k)\backslash E\big)\leq\sum\limits_{v\in D^{\infty}(z, s)}{\mathbb{P}}\big( & \exists z\leadsto B_{1}^{\infty}(z, s)^{\complement}\text{ blocked in }B_{1}^{\infty}(z, s)\text{ and}\\ & \exists v\leadsto{\mathbb{Z}}\times\{n\}\text{ in }B_{2}^{\infty}(z, s)^{\complement}\text{ with }\text{at most }k\text{ open vertices}\big)\\ = \sum\limits_{v\in D^{\infty}(z, s)}{\mathbb{P}}\big( & \exists z\leadsto B_{1}^{\infty}(z, s)^{\complement}\text{ blocked in }B_{1}^{\infty}(z, s)\big)\\ \times{\mathbb{P}}\big( & \exists v\leadsto{\mathbb{Z}}\times\{n\}\text{ in }B_{2}^{\infty}(z, s)^{\complement}\text{ with }\text{at most }k\text{ open vertices}\big)\\ \leq\sum\limits_{v\in D^{\infty}(z, s)}g( & z, s){\mathbb{P}}\big(A(v, k)\big). \end{align*}

    Now,

    \begin{align*} {\mathbb{P}}\big(A(z, k)\big) & \leq{\mathbb{P}}\big(A(z, k)\backslash E\big)+{\mathbb{P}}(E)\leq\sum\limits_{v\in D^{\infty}(z, s)}\Big[g(z, s){\mathbb{P}}\big(A(v, k)\big)+{\mathbb{P}}\big(A(v, k-1)\big)\Big], \end{align*}

    concludes the proof of Eq (9.3).

    Observe that the reduction in k can only happen until k = 0 , so more g(z, s) -terms have to show up at some point. Since any path z\leadsto B_{1}^{\infty}(z, s)^{\complement} has diameter of at least s , Lemma 75 tells us that g(z, s)\leq C_{1}\exp\big(-C_{2}s\big) for C_{1}, \, C_{2} > 0 independent of z and s . Choose s large such that

    \begin{align*} g(z, s)\cdot\#D^{\infty}(z, s) & \leq C_{1}\exp\big(-C_{2}s\big)\, \cdot8(s+\tilde{s}+1)\leq 1/4. \end{align*}

    For simplicity, we introduce

    \begin{align*} D^{\infty} & : = D^{\infty}(o, s)\qquad\text{and}\qquad h(z, y): = \begin{cases} g(z, s) & \text{if }y = 0\\ 1 & \text{if }y = 1, \end{cases} \end{align*}

    and rewrite Eq (9.3) into

    \begin{align} {\mathbb{P}}\big(A(z, k)\big) & \leq\sum\limits_{v^{(1)}\in D^{\infty}, \, y_{1}\in\{0, 1\}}h(z, y_{1}){\mathbb{P}}\big(A(z+v^{(1)}, k-y_{1})\big). \end{align} (9.4)

    We now iteratively use Eq (9.4) up to l times as it is only applicable when z_{2} < n-(s+\tilde{s}) . All the v^{(i)} are summed over D^{\infty} and all the y_{i} over \{0, 1\} .

    \begin{align*} {\mathbb{P}}\big(A & (o, k)\big)\\ \leq\ \ & \sum\limits_{\substack{i\leq l, v^{(i)}, y_{i}\\ v_{2}^{(1)}+\dots+v_{2}^{(l)} < n-(s+\tilde{s})\\ y_{1}+\dots+y_{l}\leq k} }{\mathbb{P}}(A(v^{(1)}+\dots+v^{(l)}, k-y_{1}-\dots-y_{l}))\prod\limits_{m\leq l}h(v^{(1)}+\dots+v^{(m-1)}, y_{m})\\ + & \sum\limits_{\frac{n-(s+\tilde{s})}{s+\tilde{s}+1}\leq j\leq l}\sum\limits_{\substack{i\leq j, v^{(i)}, y_{i}\\ v_{2}^{(1)}+\dots+v_{2}^{(j-1)} < n-(s+\tilde{s})\\ v_{2}^{(1)}+\dots+v_{2}^{(j)}\geq n-(s+\tilde{s})\\ y_{1}+\dots+y_{j}\leq k} }{\mathbb{P}}(A(v^{(1)}+\dots+v^{(j)}, k-y_{1}-\dots-y_{j}))\prod\limits_{m\leq j}h(v^{(1)}+\dots+v^{(m-1)}, y_{m})\\ \leq\ \ & \sum\limits_{\substack{y_{1}, \dots, y_{l}\\ y_{1}+\dots+y_{l}\leq k} }\Big(\#D^{\infty}\sup\limits_{\substack{v\in{\mathbb{Z}}^{2}\\ v_{2} < n-(s+\tilde{s})} }h(v, y_{m})\Big)^{l}+\sum\limits_{\frac{n-(s+\tilde{s})}{s+\tilde{s}+1}\leq j\leq l}\sum\limits_{\substack{y_{1}, \dots, y_{j}\\ y_{1}+\dots+y_{j}\leq k} }\Big(\#D^{\infty}\sup\limits_{\substack{v\in{\mathbb{Z}}^{2}\\ v_{2} < n-(s+\tilde{s})} }h(v, y_{m})\Big)^{j}. \end{align*}

    We iterate as long as 0+v_{2}^{(1)}+\dots+ v_{2}^{(j)} < n-(s+\tilde{s}) ; otherwise, we stop for 0+v_{2}^{(1)}+\dots +v_{2}^{(j)} and land in the second summand. Only y_{1}+\dots+y_{j}\leq k matters since A(z, m) = 0 whenever m < 0 . Also, observe that v_{2}^{(1)}+\dots+v_{2}^{(j)}\geq n-(s+\tilde{s}) can only happen if

    j\geq\frac{n-(s+\tilde{s})}{s+\tilde{s}+1},

    since we "gain" at most s+\tilde{s}+1 to the second component in each v^{(i)} .

    Let \alpha > 0 be large enough such that

    \begin{align*} \phi(\alpha): & = \sum\limits_{y\in\{0, 1\}}\sup\limits_{\substack{z\in{\mathbb{Z}}^{2}\\ z_{2} < n-(s+\tilde{s})} }\#D^{\infty}\cdot h(z, y)\cdot e^{-\alpha y}\leq\frac{1}{4}+\#D^{\infty}\cdot e^{-\alpha}\leq\frac{1}{2}. \end{align*}

    Then,

    \begin{align*} {\mathbb{P}}\big(A(o, k)\big)\leq\ \ & e^{\alpha k}\sum\limits_{\substack{y_{1}, \dots, y_{l}\\ y_{1}+\dots+y_{l}\leq k} }\Big(\#D^{\infty}\sup\limits_{\substack{v\in{\mathbb{Z}}^{2}\\ v_{2} < n-(s+\tilde{s})} }h(v, y_{m})\Big)^{l}\prod\limits_{i\leq l}e^{-\alpha y_{i}}\\ + & e^{\alpha k}\sum\limits_{\frac{n-(s+\tilde{s})}{s+\tilde{s}+1}\leq j\leq l}\sum\limits_{\substack{y_{1}, \dots, y_{j}\\ y_{1}+\dots+y_{j}\leq k} }\Big(\#D^{\infty}\sup\limits_{\substack{v\in{\mathbb{Z}}^{2}\\ v_{2} < n-(s+\tilde{s})} }h(v, y_{m})\Big)^{j}\prod\limits_{i\leq j}e^{-\alpha y_{i}}\\ \leq\ \ & e^{\alpha k}\phi(\alpha)^{l}+e^{\alpha k}\sum\limits_{j\geq\frac{n-(s+\tilde{s})}{s+\tilde{s}+1}}^{l}\phi(\alpha)^{j}\leq e^{\alpha k}\big[2^{-l}+2^{-\frac{n-(s+\tilde{s})}{s+\tilde{s}+1}+1}\big]. \end{align*}

    Since l\in{\mathbb{N}} was arbitrary, we get

    \begin{align*} {\mathbb{P}}\big(A(o, k)\big) & \leq e^{\alpha k}\cdot2^{-\frac{n-(s+\tilde{s})}{s+\tilde{s}+1}+1} = e^{\alpha k-\big[\frac{n-(s+\tilde{s})}{s+\tilde{s}+1}-1\big]\ln2} = C_{3}e^{\alpha k-C_{4}n}. \end{align*}

    Now we finally make use of k . Setting C_{5}: = \frac{C_{4}}{2\alpha} and k: = C_{5}n , the claim follows:

    {\mathbb{P}}\big(A(o, C_{5}n)\big)\leq C_{3}\exp\big(-C_{4}n/2\big).

    All results of the publication have resulted from joint discussions. They can therefore not be attributed separately to the authors, but are to be regarded as joint achievements.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors thank the two anonymous reviewers for their valuable comments that helped to improve the manuscript. BJ and ADV received support from the Leibniz Association within the Leibniz Junior Research Group on Probabilistic Methods for Dynamic Communication Networks as part of the Leibniz Competition (grant no. J105/2020).

    The authors declare there is no conflict of interest.



    [1] H. Sung, J. Ferlay, R. L. Siegel, M. Laversanne, I. Soerjomataram, A. Jemal, et al., Global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries, CA Cancer J. Clin., 71 (2021), 209–249. https://doi.org/10.3322/caac.21660 doi: 10.3322/caac.21660
    [2] A. Das, M. S. Nair, S. D. Peter, Computer-aided histopathological image analysis techniques for automated nuclear atypia scoring of breast cancer: a review, J. Dig. Imaging, 33 (2020), 1–31. https://doi.org/10.1007/s10278-019-00295-z doi: 10.1007/s10278-020-00318-0
    [3] L. He, R. Long, S. Antani, G. R. Thoma, Histology image analysis for carcinoma detection and grading, Comput. Methods Prog. Biomed., 107 (2012), 538–556. https://doi.org/10.1016/j.cmpb.2011.12.007 doi: 10.1016/j.cmpb.2011.12.007
    [4] L. Dalton, Artificial intelligence grading of breast cancer: A study of ensemble learning, Lab. Investig., 102 (2022), 107–108.
    [5] J. G. Elmore, G. M. Longton, P. A. Carney, B. M. Geller, T. Onega, A. N. A. Tosteson, et al., Diagnostic concordance among pathologists interpreting breast biopsy specimens, JAMA, 313 (2015), 1122–1132. https://doi.org/10.1001/jama.2015.1405 doi: 10.1001/jama.2015.1405
    [6] K. E. Lindquist, C. Ciornei, S. Westbom-Fremer, I. Gudinaviciene, A. Ehinger, N. Mylona, et al., Difficulties in diagnostics of lung tumours in biopsies: an interpathologist concordance study evaluating the international diagnostic guidelines, J. Clin. Pathol., 75 (2022), 302–309. http://dx.doi.org/10.1136/jclinpath-2020-207257 doi: 10.1136/jclinpath-2020-207257
    [7] C. Merkouri, D. Grapsa, I. Tourkantonis, I. Gkiozos, A. Charpidou, G. Tournas, et al. Cytology-histology concordance for diagnosis, histological subtyping and molecular profiling of lung cancer, J. Thorac. Oncol., 16 (2021), S463. https://doi.org/10.1016/j.jtho.2021.01.794 doi: 10.1016/j.jtho.2021.01.794
    [8] V. K. Singh, H. A. Rashwan, M. Abdel-Nasser, F. Akram, R. Haffar, N. Pandey, et al., A computer-aided diagnosis system for breast cancer molecular subtype prediction in mammographic images, State Art Neural Networks Their Appl., 1 (2021), 153–178. https://doi.org/10.1016/B978-0-12-819740-0.00008-5 doi: 10.1016/B978-0-12-819740-0.00008-5
    [9] M. Perumal, A research on computer aided detection system for women breast cancer diagnosis from digital mammographic images, Int. J. Recent Technol. Eng., 2021 (2021). https://doi.org/10.35940/ijrte.B1169.0982S1119 doi: 10.35940/ijrte.B1169.0982S1119
    [10] S. Naik, S. Doyle, S. Agner, A. Madabhushi, M. Feldman, J. Tomaszewski, Automated gland and nuclei segmentation for grading of prostate and breast cancer histopathology, in 2008 5th IEEE International Symposium on Biomedical Imaging: From Nano to Macro, (2008). https://doi.org/10.1109/ISBI.2008.4540988
    [11] C. Jung, C. Kim, Segmenting clustered nuclei using H-minima transform-based marker extraction and contour parameterization, IEEE Trans. Biomed. Eng., 57 (2010), 2600–2604. https://doi.org/10.1109/TBME.2010.2060336 doi: 10.1109/TBME.2010.2060336
    [12] M. Kowal, P. Filipczuk, A. Obuchowicz, J. Korbicz, R. Monczak, Computer-aided diagnosis of breast cancer based on fine needle biopsy microscopic images, Comput. Biol. Med., 43 (2013), 1563–1572. https://doi.org/10.1016/j.compbiomed.2013.08.003 doi: 10.1016/j.compbiomed.2013.08.003
    [13] A. D. Belsare, M. M. Mushrif, M. A. Pangarkar, N. Meshram, Breast histopathology image segmentation using spatio-colour-texture based graph partition method, J. Microscopy, 262 (2016), 260–273. https://doi.org/10.1111/jmi.12361 doi: 10.1111/jmi.12361
    [14] Z. Senousy, M. M. Abdelsamea, M. M. Mohamed, M. M. Gaber, 3E-Net: Entropy-based elastic ensemble of deep convolutional neural networks for grading of invasive breast carcinoma histopathological microscopic images, Entropy, 23 (2021), 620. https://doi.org/10.3390/e23050620 doi: 10.3390/e23050620
    [15] Y. Yari, T. V. Nguyen, H. T. Nguyen, Deep learning applied for histological diagnosis of breast cancer, IEEE Access, 8 (2020), 162463–162448. https://doi.org/10.1109/ACCESS.2020.3021557 doi: 10.1109/ACCESS.2020.3021557
    [16] D. C. Ciresan, A. Giusti, L. M. Gambardella, J. Schmidhuber, Mitosis detection in breast cancer histology images with deep neural networks, in International Conference on Medical Image Computing and Computer-Assisted Intervention, (2013). https://doi.org/10.1007/978-3-642-40763-5_51
    [17] G. Litjens, C. I. Sánchez, N. Timofeeva, M. Hermsen, I. Nagtegaal, I. Kovacs, et al., Deep learning as a tool for increased accuracy and efficiency of histopathological diagnosis, Sci. Rep., 6 (2016), 26286. https://doi.org/10.1038/srep26286 doi: 10.1038/srep26286
    [18] O. Ronneberger, P. Fischer, T. Brox, U-Net: Convolutional networks for biomedical image segmentation, in International Conference on Medical Image Computing and Computer-Assisted Intervention, (2015). https://doi.org/10.1007/978-3-319-24574-4_28
    [19] D. J. Ho, D. Yarlagadda, T. M. D'Alfonso, M. G. Hanna, A. Grabenstetter, P. Ntiamoah, et al., Deep multi-magnification networks for multi-class breast cancer image segmentation, Comput. Med. Imaging Graphics, 88 (2021), 101866. https://doi.org/10.1016/j.compmedimag.2021.101866 doi: 10.1016/j.compmedimag.2021.101866
    [20] I. Anand, H. Negi, D. Kumar, M. Mittal, T. Kim, S. Roy, Residual U-network for breast tumor segmentation from magnetic resonance images, Comput. Mater. Continua, 67 (2021), 3107–3127. http://doi.org/10.32604/cmc.2021.014229 doi: 10.32604/cmc.2021.014229
    [21] L. C. Chen, G. Papandreou, I. Kokkinos, K. Murphy, A. L. Yuille, DeepLab: Semantic image segmentation with deep convolutional nets, atrous convolution, and fully connected CRFs, IEEE Trans. Pattern Anal. Mach. Intell., 40 (2018), 834–848. https://doi.org/10.1109/TPAMI.2017.2699184 doi: 10.1109/TPAMI.2017.2699184
    [22] J. Wang, X. Liu, Medical image recognition and segmentation of pathological slices of gastric cancer based on Deeplab v3+ neural network, Comput. Methods Prog. Biomed., 207 (2021), 106210. https://doi.org/10.1016/j.cmpb.2021.106210 doi: 10.1016/j.cmpb.2021.106210
    [23] C. Yu, J. Wang, C. Peng, C. Gao, G. Yu, N. Sang, BiSeNet: Bilateral segmentation network for real-time semantic segmentation, in Proceedings of the European Conference on Computer Vision (ECCV), (2018).
    [24] G. Aresta, T. Araújo, S. Kwok, S. S. Chennamsetty, M. Safwan, V. Alex, et al., BACH: Grand challenge on breast cancer histology images, Med. Image Anal., 56 (2019), 122–139. https://doi.org/10.1016/j.media.2019.05.010 doi: 10.1016/j.media.2019.05.010
    [25] S. Krishnamurthy, K. Mathews, S. McClure, M. Murray, M. Gilcrease, C. Albarracin, et al., Multi-institutional comparison of Whole slide digital imaging and optical microscopy for interpretation of hematoxylin-eosin-stained breast tissue sections, Arch. Pathol. Lab. Med., 137 (2013), 1733–1739. https://doi.org/10.5858/arpa.2012-0437-OA doi: 10.5858/arpa.2012-0437-OA
    [26] A. Canziani, A. Paszke, E. Culurciello, An analysis of deep neural network models for practical applications, preprint, arXiv: 1605.07678.
    [27] S. Liu, D. Huang, Y. Wang, Receptive field block net for accurate and fast object detection, in Proceedings of the European Conference on Computer Vision (ECCV), (2018).
    [28] F. Chollet, Xception: Deep learning with depthwise separable convolutions, in 2017 IEEE Conference on Computer Vision and Pattern Recognition (CVPR), (2017).
    [29] J. Hu, L. Shen, G. Sun, Squeeze-and-excitation networks, in Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR), (2018).
    [30] A. Buslaev, V. I. Iglovikov, E. Khvedchenya, A. Parinov, M. Druzhinin, A. A. Kalinin, Albumentations: fast and flexible image augmentations, Information, 11 (2020), 125. https://doi.org/10.3390/info11020125 doi: 10.3390/info11020125
    [31] S. Mantrala, P. S. Ginter, A. Mitkar, S. Joshi, H. Prabhala, V. Ramachandra, Concordance in breast cancer grading by artificial intelligence on whole slide images compares with a multi-Institutional cohort of breast pathologists, Arch. Pathol. Lab. Med., 2022 (2022), forthcoming. https://doi.org/10.5858/arpa.2021-0299-OA
  • 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(2780) PDF downloads(149) Cited by(2)

Article outline

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog