Homogenisation for the Stokes equations in randomly perforated domains under almost minimal assumptions on the size of the holes (Q2334991)

From MaRDI portal
Revision as of 21:35, 29 July 2024 by Daniel (talk | contribs) (‎Created claim: Wikidata QID (P12): Q127459995, #quickstatements; #temporary_batch_1722281465132)
(diff) ← Older revision | Latest revision (diff) | Newer revision → (diff)
scientific article
Language Label Description Also known as
English
Homogenisation for the Stokes equations in randomly perforated domains under almost minimal assumptions on the size of the holes
scientific article

    Statements

    Homogenisation for the Stokes equations in randomly perforated domains under almost minimal assumptions on the size of the holes (English)
    0 references
    0 references
    0 references
    13 November 2019
    0 references
    The authors prove a homogenization result for the Stokes problem \[-\Delta u_{\varepsilon}+\nabla p_{\varepsilon}=f,\;\;\; \nabla \cdot u_{\varepsilon}=0,\] posed in \(D^{\varepsilon}=D\setminus H^{\varepsilon}\), where \(D\subset \mathbb{R}^{d}\), \(d>2\), is an open and bounded domain, which is star-shaped with respect to the origin, and \(H^{\varepsilon}\) is the union of a random number of small balls \(B_{\varepsilon ^{\frac{d}{d-2}}\rho _{i}}(\varepsilon z_{i})\) with \(z_{i}\in \Phi +\frac{1}{\varepsilon}D\), \(\Phi \) being a Poisson point process on \(\mathbb{R}^{d}\) of homogeneous intensity rate \(\lambda >0\), and \(\{\rho _{i}\}\) are identically and independently distributed unbounded random variables. The source term \(f\) of the Stokes equation is supposed to belong to \(H^{-1}(D;\mathbb{R}^{d})\) and the fluid is supposed to perfectly adhere to the balls and to the boundary of \(D\). The probability space associated to the joint process \((\Phi ,\mathcal{R})\) of the centres and radii is denoted as \((\Omega ,\mathcal{F},\mathbb{P})\). The main result of the paper proves that, for \(\mathbb{P}\)-almost every \(\omega \in \Omega \), \(u_{\varepsilon}(\omega ,\cdot )\) weakly converges in \(H_{0}^{1}(D;\mathbb{R}^{d})\) to the solution \(u_{h}\) of Brinkman's equation \[ -\Delta u_{h}+\nabla p_{h}+C_{d}\lambda \left\langle \rho^{d-2}\right\rangle u_{h}=f, \] in \(D\) with the divergence property \(\nabla\cdot u_{h}=0\) and the boundary condition \(u_{h}=0\) on \(\partial D\). In the extra term of this homogenized problem, \(\left\langle \cdot \right\rangle\) denotes the expectation under the probability measure on the radii \(\rho_{i}\) and the positive constant \(C_{d}\) only depends on the dimension \(d\) and is equal to \(6\pi \) in the case \(d=3\). The authors prove a similar result for the stationary Navier-Stokes equation. Concerning the pressure, the authors prove the existence of a set \(E^{\varepsilon}\subseteq \mathbb{R}^{d}\) satisfying \(E^{\varepsilon}\supset H^{\varepsilon}\) and \(Cap(E^{\varepsilon}\setminus H^{\varepsilon})\rightarrow 0\) when \(\varepsilon \) goes to 0 such that for every compact set \(K\Subset D\), the modification of the pressure \(\widetilde{p}_{\varepsilon}=p_{\varepsilon}-\frac{1}{\left\vert K\setminus E^{\varepsilon}\right\vert}\int_{K\setminus E^{\varepsilon}}p_{\varepsilon}\) in \(K\setminus E^{\varepsilon}\) and 0 outside weakly converges in \(L_{0}^{q}(K;\mathbb{R})\) to \(p_{h}\) for every \(q<\frac{d}{d-1}\). The main tool of the proof of the first main result is the construction of a linear operator \(R_{\varepsilon}:\{v\in C_{0}^{\infty}(D;\mathbb{R}^{d}):\nabla \cdot v=0\}\rightarrow H^{1}(D;\mathbb{R}^{d})\) such that \(\int \nabla R_{\varepsilon}v:\nabla u_{\varepsilon}dx\rightarrow\int \nabla v:\nabla udx+C_{d}\lambda \left\langle \rho ^{d-2}\right\rangle\int v\cdot udx\), for all \(u_{\varepsilon}\in H^{1}(D^{\varepsilon}; \mathbb{R}^{d})\) satisfying \(\nabla \cdot u_{\varepsilon}=0\) in \(D\) and \(u_{\varepsilon}\rightharpoonup u\) weakly in \(H_{0}^{1}(D;\mathbb{R}^{d})\), among other properties. This construction is obtained splitting the set \(H^{\varepsilon}\) into a ``good'' set \(H_{g}^{\varepsilon}\) which contains holes which are small and well-separated and a ``bad'' set \(H_{b}^{\varepsilon}\) which contains big and overlapping holes. The authors build \(R_{\varepsilon}v\) such that it vanishes on \(H_{g}^{\varepsilon}\) following ideas by \textit{G. Allaire} in [Arch. Ration. Mech. Anal. 113, No. 3, 209--259 (1991; Zbl 0724.76020)] and by \textit{L. Desvillettes} et al. in [J. Stat. Phys. 131, No. 5, 941--967 (2008; Zbl 1154.76018)]. On \(H_{b}^{\varepsilon}\), the authors introduce a suitable covering \(\overline{H}_{b}^{\varepsilon}\) of the set \(H_{b}^{\varepsilon}\) selecting some of the balls in \(H_{b}^{\varepsilon}\), dilating them by a uniformly bounded factor \(\lambda _{\varepsilon}\leq\Lambda \), solving different Stokes problems in disjoint annuli and iterating this procedure a finite number of steps. This requires the derivation of geometric properties of the holes and improves a procedure previously introduced by the same authors and \textit{J. J. L. Velazquez} in [Commun. Partial Differ. Equations 43, No. 9, 1377--1412 (2018; Zbl 1448.60136)]. The last part of the paper gives some probabilistic results on the random set \(H^{\varepsilon}\), in terms of the size of the clusters generated by overlapping balls of comparable size.
    0 references
    random homogenization
    0 references
    Brinkman equations
    0 references
    0 references
    0 references
    0 references
    0 references
    0 references
    0 references

    Identifiers

    0 references
    0 references
    0 references
    0 references
    0 references
    0 references
    0 references
    0 references