This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Convergence of the CEM-GMsFEM for compressible flow in highly heterogeneous media

Leonardo A. Poveda Department of Mathematics, The Chinese University of Hong Kong, Shatin, Hong Kong SAR, China    Shubin Fu Eastern Institute for Advanced Study, Ningbo, China    Eric T. Chung Department of Mathematics, The Chinese University of Hong Kong, Shatin, Hong Kong SAR, China    Lina Zhao Department of Mathematics, City University of Hong Kong, Hong Kong SAR, China
Abstract

This paper presents and analyses a Constraint Energy Minimization Generalized Multiscale Finite Element Method (CEM-GMsFEM) for solving single-phase non-linear compressible flows in highly heterogeneous media. The construction of CEM-GMsFEM hinges on two crucial steps: First, the auxiliary space is constructed by solving local spectral problems, where the basis functions corresponding to small eigenvalues are captured. Then the basis functions are obtained by solving local energy minimization problems over the oversampling domains using the auxiliary space. The basis functions have exponential decay outside the corresponding local oversampling regions. The convergence of the proposed method is provided, and we show that this convergence only depends on the coarse grid size and is independent of the heterogeneities. An online enrichment guided by a posteriori error estimator is developed to enhance computational efficiency. Several numerical experiments on a three-dimensional case to confirm the theoretical findings are presented, illustrating the performance of the method and giving efficient and accurate numerical

Keywords: Constraint energy minimization, multiscale finite element methods, compressible flow, highly heterogeneous, local spectral problems

1 Introduction

The numerical solution of non-linear partial differential equations defined on domains with multiscale and heterogeneous properties is an active research subject in the scientific community. The subject is related to several engineering applications, such as composite materials, porous media flow, and fluid mechanics. A common feature for all these applications is that they are very computationally challenging and often impossible to solve within an acceptable tolerance using standard fine-scale approximations due to the disparity between scales that need to be represented and the inherent nonlinearities. For this reason, coarse-grid computational models are often used. These approaches are usually referred to as multiscale methods in the literature, among which we may mention: Multiscale Finite Element Method [17], the Variational Multiscale Method [18], Mixed Multiscale Finite Element Method [6], Mixed Mortar Multiscale Finite Element Method [2], the two-scale Finite Element Method [24], and the Multiscale Finite Volume method [19, 20]. The aforementioned methods share model reduction techniques, using different structures to find multiscale solutions, especially in many practical applications such as fluid flow simulations, for instance, [1, 5, 25, 21, 16, 27, 15, 26]. In particular, we consider a family of an extended version of MsFEM, known as the Generalized Multiscale Finite Element Method (GMsFEM) that was first introduced by [12, 13, 8, 11]. The main idea of GMsFEM is to construct localized basis functions by solving local spectral problems that are used to approximate the solution on a coarse grid incorporating fine-scale features. Following this, we construct an auxiliary space associated with local spectral problems in the coarse grid. The first few eigenfunctions corresponding to small eigenvalues (the convergence depends on the decay of the spectral problems [12]) are considered as the multiscale basis functions. In this paper, we extend the Constraint Energy Minimization Generalized Multiscale Finite Element Method (CEM-GMsFEM) developed in [9, 22] to solve single-phase nonlinear compressible flow. The key ideas of the method can be summarized as follows: First, we construct the auxiliary basis functions by solving local spectral problems. Then, by using oversampling techniques and localization (cf. [23]), we solve an appropriate energy subject to some constrainted oversampling regions to find the required basis functions. Finally, the resulting basis functions are shown to have exponential decay away from the target coarse element, and therefore, they are localizable. Then we rigorously analyze the convergence error estimates for the proposed scheme. Our theories indicate that the convergence rate behaves as H/ΛH/\Lambda, where HH denotes the coarse-grid size and Λ\Lambda is proportional to the minimum (taken over all coarse regions) of the eigenvalues that the corresponding eigenvectors are not included in the coarse space. Since the problems under consideration are nonlinear, some novel methodologies shall be incorporated to overcome the difficulties present in the analysis. Several numerical experiments are carried out to demonstrate the capabilities and efficiency of the proposed method.

The outline of the paper is following. In Section 2, we briefly derive a mathematical model for compressible fluid flows in porous media. Section 3 is devoted to constructing the offline multiscale space and framework of CEM-GMsFEM. Section 4 presents our convergence analysis for the proposed method. Numerical experiments are presented in Section 5. Finally, concluding remarks and future perspectives are given in Section 6.

2 Mathematical model for compressible fluid flow

In this section, we consider the governing equations of the single-phase, non-linear compressible fluid flow processes in a porous medium that are defined by

{t(ϕρ)(κμρp)=q,in ΩT:=Ω×(0,T],T>0,κμρpn=0,on ΓN×(0,T],p=pD,on ΓD×(0,T],p=p0,on Ω×{t=0}.\left\{\begin{split}\partial_{t}(\phi\rho)-\nabla\cdot\left(\frac{\kappa}{\mu}\rho\nabla p\right)&=q,\quad\mbox{in }\Omega_{T}:=\Omega\times(0,T],\quad T>0,\\ \frac{\kappa}{\mu}\rho\nabla p\cdot n&=0,\quad\mbox{on }\Gamma_{N}\times(0,T],\\ p&=p^{D},\quad\mbox{on }\Gamma_{D}\times(0,T],\\ p&=p_{0},\quad\mbox{on }\Omega\times\{t=0\}.\end{split}\right. (2.1)

For simplicity of presentation, let Ωd\Omega\subset\mathbb{R}^{d} be the computational domain with a boundary defined by Ω=ΓDΓN\partial\Omega=\Gamma_{D}\cup\Gamma_{N}. We will henceforth neglect the gravity effects and capillary forces and assume that ϕ\phi, the porosity of the medium, is assumed to be a constant. We aim to seek the fluid pressure pp, κ\kappa denotes the permeability field that may be highly heterogeneous, such that κ0κκ1\kappa_{0}\leq\kappa\leq\kappa_{1}, where 0<κ0<κ1<0<\kappa_{0}<\kappa_{1}<\infty; and μ\mu is the constant fluid viscosity. The fluid density ρ\rho is a function of the fluid pressure pp defined as

ρ(p)=ρrefec(ppref),\rho(p)=\rho_{\mathrm{ref}}e^{c(p-p_{\mathrm{ref}})}, (2.2)

where ρref\rho_{\mathrm{ref}} is the given reference density and prefp_{\mathrm{ref}} is the reference pressure. Finally, nn denotes the outward unit-normal vector on Ω\partial\Omega.

For a sub-domain DΩD\subset\Omega, let V:=H1(D)\mathrm{V}:=\mathrm{H}^{1}(D) be the standard Sobolev spaces endowed with the norm 1,D\|\cdot\|_{1,D}. We further denote by (,)(\cdot,\cdot) and 0,D\|\cdot\|_{0,D} the inner product and norm, respectively in L2(D)\mathrm{L}^{2}(D). The subscript DD will be omitted whenever D=ΩD=\Omega. In addition, we use the space V0:=H01(D)\mathrm{V}_{0}:=\mathrm{H}^{1}_{0}(D), which is a subspace of V\mathrm{V} made of functions that vanish at ΓD\Gamma_{D}. Finally, let L2(0,T;L2(D))\mathrm{L}^{2}(0,T;\mathrm{L}^{2}(D)) denote the set of functions with norm

vL2(0,T;L2(D))=(0Tv(,t)0,D2𝑑t)1/2.\|v\|_{\mathrm{L}^{2}(0,T;\mathrm{L}^{2}(D))}=\left(\int_{0}^{T}\|v(\cdot,t)\|^{2}_{0,D}dt\right)^{1/2}.

Throughout the paper, aba\preceq b means there exists a positive constant CC independent of the mesh size such that aCba\leq Cb.

2.1 A finite element approximation

In this subsection, we introduce the notions of fine and coarse grids to discretize the problem (2.1). Let 𝒯H{\cal T}^{H} be a usual conforming partition of the computational domain Ω\Omega into coarse block K𝒯HK\in{\cal T}^{H} with diameter HH. Then, we denote this partition as the coarse grid and assume that each coarse element is partitioned into a connected union of fine-grid blocks. In this case, the fine grid partition will be denoted by 𝒯h{\cal T}^{h}, and is, by definition, a refinement of the coarse grid 𝒯H{\cal T}^{H}, such that hHh\ll H. We shall denote {xi}i=1Ncoarse\{x_{i}\}_{i=1}^{N_{\tiny{\rm{coarse}}}} as the vertices of the coarse grid 𝒯H{\cal T}^{H}, where NcoarseN_{\tiny{\rm{coarse}}} denotes the number of coarse nodes. We define the neighborhood of the node xix_{i} by

ωi={Kj𝒯H:xiK¯j}.\omega_{i}=\bigcup\{K_{j}\in{\cal T}^{H}:x_{i}\in\overline{K}_{j}\}.

In addition, for CEM-GMsFEM considered in this paper, we have that given a coarse block KiK_{i}, we represent the oversampling region Ki,mΩK_{i,m}\subset\Omega obtained by enlarging KiK_{i} with m1m\geq 1 coarse grid layers, see Fig. 1.

Ki,2K_{i,2}KiK_{i}fine elementxix_{i}ωi\omega_{i}Ω\Omega
Figure 1: Illustration of the 22D multiscale grid with a typical coarse element KiK_{i} and oversampling domain Ki,2K_{i,2}, the fine grid element and neighborhood ωi\omega_{i} of the node xix_{i}.

We consider the linear finite element space Vh\mathrm{V}^{h} associated with the grid 𝒯h{\cal T}^{h}, where the basis functions in this space are the standard Lagrange basis functions defined as {ηi}i=1Nfine\{\eta^{i}\}_{i=1}^{N_{\tiny{\rm{fine}}}}. Then, the semi-discrete finite element approximation to (2.1) on the fine grid is to find phVhp^{h}\in\mathrm{V}^{h} such that

(ϕtρ(ph),v)+(κμρ(ph)ph,v)=(q,v),for each vVh.(\phi\partial_{t}\rho(p^{h}),v)+\left(\tfrac{\kappa}{\mu}\rho(p^{h})\nabla p^{h},\nabla v\right)=(q,v),\quad\mbox{for each }v\in\mathrm{V}^{h}. (2.3)

We can now define the fully-discrete scheme for the discrete formulation (2.3). Let 0=t0<t1<<tNtime1<tNtime=T0=t_{0}<t_{1}<\cdots<t_{N_{\tiny{\rm{time}}}-1}<t_{N_{\tiny{\rm{time}}}}=T be a partition of the interval [0,T][0,T], with time-step size given by Δn=tntn1\Delta_{n}=t_{n}-t_{n-1}, for n=1,,Ntimen=1,\dots,N_{\tiny{\rm{time}}}, where NtimeN_{\tiny{\rm{time}}} is an integer. The backward Euler time integration leads to finding pnhp^{h}_{n} such that

(ϕρ(pnh),v)(ϕρ(pn1h),v)+Δn(κμρ(pnh)pnh,v)=Δn(q,v),for each vVh.(\phi\rho(p^{h}_{n}),v)-(\phi\rho(p^{h}_{n-1}),v)+\Delta_{n}\left(\tfrac{\kappa}{\mu}\rho(p^{h}_{n})\nabla p^{h}_{n},\nabla v\right)=\Delta_{n}(q,v),\quad\mbox{for each }v\in\mathrm{V}^{h}. (2.4)

Linearization of (2.4) via Newton-Raphson iteration yields a iterative linear matrix problem,

𝐉n,k𝜹pn,k=𝐅n,k,{\mathbf{J}}^{n,k}\boldsymbol{\delta}_{p^{n,k}}=-{\mathbf{F}}^{n,k},

where 𝐉n,k:=[Jjin,k]i,j=1Nfine{\mathbf{J}}^{n,k}:=[J_{ji}^{n,k}]_{i,j=1}^{N_{\tiny{\rm{fine}}}} denotes the Jacobi matrix, with entries

Jjin,k:=(ϕρ(pnh,k)ηi,ηj)+Δn(κμρ(pnh,k)ηi,ηj)+Δn(cκμηiρ(pnh,k)ipni,kηi,ηj),J_{ji}^{n,k}:=(\phi\rho(p^{h,k}_{n})\eta_{i},\eta_{j})+\Delta_{n}\left(\frac{\kappa}{\mu}\rho(p^{h,k}_{n})\nabla\eta_{i},\nabla\eta_{j}\right)+\Delta_{n}\left(c\frac{\kappa}{\mu}\eta_{i}\rho(p^{h,k}_{n})\sum_{i}p^{i,k}_{n}\nabla\eta_{i},\nabla\eta_{j}\right),

𝐅n,k:=[Fjn,k]j=1Nfine{\mathbf{F}}^{n,k}:=[F_{j}^{n,k}]_{j=1}^{N_{\tiny{\rm{fine}}}} is the residual with entries

Fjn,k\displaystyle F_{j}^{n,k} =(ϕρ(i=1Nfinepni,kηi),ηj)(ϕρ(i=1Nfinepn1iηi),ηj)+Δn(κμρ(i=1Nfinepni,kηi)i=1Nfinepni,kηi,ηj)\displaystyle=\left(\phi\rho\left(\sum_{i=1}^{N_{\tiny{\rm{fine}}}}p^{i,k}_{n}\eta_{i}\right),\eta_{j}\right)-\left(\phi\rho\left(\sum_{i=1}^{N_{\tiny{\rm{fine}}}}p^{i}_{n-1}\eta_{i}\right),\eta_{j}\right)+\Delta_{n}\left(\frac{\kappa}{\mu}\rho\left(\sum_{i=1}^{N_{\tiny{\rm{fine}}}}p^{i,k}_{n}\eta_{i}\right)\sum_{i=1}^{N_{\tiny{\rm{fine}}}}p^{i,k}_{n}\nabla\eta_{i},\nabla\eta_{j}\right)
Δn(q,ηj)=0,\displaystyle\quad-\Delta_{n}(q,\eta_{j})=0,

and pnk+1=pnk+𝜹pn,kp^{k+1}_{n}=p^{k}_{n}+\boldsymbol{\delta}_{p^{n,k}}, where pnh,k=ipni,kηip^{h,k}_{n}=\sum_{i}p^{i,k}_{n}\eta_{i} and pn1h=ipn1iηip^{h}_{n-1}=\sum_{i}p^{i}_{n-1}\eta_{i}, with kk and k1k-1 the new and old Newton iteration. Here {ηi}i=1Nfine\{\eta^{i}\}_{i=1}^{N_{\tiny{\rm{fine}}}} represents the finite element basis functions for Vh\mathrm{V}^{h}.

3 Construction of CEM-GMsFEM basis functions

This section will describe the construction of CEM-GMsFEM basis functions using the framework of [9] and [22]. This procedure can be divided into two stages. The first stage involves constructing the auxiliary spaces by solving a local spectral problem in each coarse element KK, see [12]. The second stage is to provide the multiscale basis functions by solving some local constraint energy minimization problems in oversampling regions.

3.1 Auxiliary basis function

In this subsection, we present the construction of the auxiliary multiscale basis functions by solving the local eigenvalue problem for each coarse element KiK_{i}. We consider V(Ki):=V|Ki\mathrm{V}(K_{i}):=\mathrm{V}\scriptstyle\big{|}_{K_{i}} the restriction of the space V\mathrm{V} to the coarse element KiK_{i}. We solve the following local eigenvalue problem: find {λj(i),φj(i)}\{\lambda^{(i)}_{j},\varphi^{(i)}_{j}\} such that

ai(φj(i),w)=λj(i)si(φj(i),w),for each wV(Ki),a_{i}(\varphi^{(i)}_{j},w)=\lambda^{(i)}_{j}s_{i}(\varphi^{(i)}_{j},w),\quad\mbox{for each }w\in\mathrm{V}(K_{i}), (3.1)

where

ai(v,w):=Kiκρ(p0)vwdx,si(v,w):=Kiκ~vw𝑑x.a_{i}(v,w):=\int_{K_{i}}\kappa\rho(p_{0})\nabla v\cdot\nabla wd\mathrm{x},\quad s_{i}(v,w):=\int_{K_{i}}\widetilde{\kappa}vwd\mathrm{x}.

Here κ~=ρ(p0)κi=1Ncoarse|χi|2\widetilde{\kappa}=\rho(p_{0})\kappa\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}|\nabla\chi_{i}|^{2}, where NcoarseN_{\tiny{\rm{coarse}}} is the total number of neighborhoods, p0p_{0} is the pressure pp at the initial time and {χi}\{\chi_{i}\} is a set of partition of unity functions for the coarse grid 𝒯H{\cal T}^{H}, see [3]. The problem defined above is solved on the fine grid in the actual computation. We assume that the eigenfunctions satisfy the normalized condition si(φj(i),φj(i))=1s_{i}(\varphi^{(i)}_{j},\varphi^{(i)}_{j})=1. We let λj(i)\lambda^{(i)}_{j} be the eigenvalues of (3.1) arranged in ascending order. We shall use the first LiL_{i} eigenfunctions to construct the local auxiliary multiscale space Vaux(i):={φj(i):1jLi}\mathrm{V}_{\mathrm{aux}}^{(i)}:=\{\varphi^{(i)}_{j}:1\leq j\leq L_{i}\}. We can define the global auxiliary multiscale space as Vaux:=i=1NcoarseVaux(i)\mathrm{V}_{\mathrm{aux}}:=\bigoplus_{i=1}^{N_{\tiny{\rm{coarse}}}}\mathrm{V}_{\mathrm{aux}}^{(i)}.

For the local auxiliary space Vaux(i)\mathrm{V}_{\mathrm{aux}}^{(i)}, the bilinear form sis_{i} given above defines an inner product with norm vs(Ki)=si(v,v)1/2\|v\|_{s(K_{i})}=s_{i}(v,v)^{1/2}. Then, we can define the inner product and norm for the global auxiliary multiscale space Vaux\mathrm{V}_{\mathrm{aux}}, which are defined by

s(v,w)=i=1Ncoarsesi(v,w),vs:=s(v,v)1/2,for each v,wVaux.s(v,w)=\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}s_{i}(v,w),\quad\|v\|_{s}:=s(v,v)^{1/2},\quad\mbox{for each }v,w\in\mathrm{V}_{\mathrm{aux}}.

To construct the CEM-GMsFEM basis functions, we use the following definition.

Definition 3.1 ([9]).

Given a function φj(i)Vaux\varphi^{(i)}_{j}\in\mathrm{V}_{\mathrm{aux}}, if a function ψV\psi\in\mathrm{V} satisfies

s(ψ,φj(i)):=1,s(ψ,φj(i))=0,if jj or ii,s(\psi,\varphi^{(i)}_{j}):=1,\quad s(\psi,\varphi_{j^{\prime}}^{(i^{\prime})})=0,\quad\mbox{if }j^{\prime}\neq j\mbox{ or }i^{\prime}\neq i,

then, we say that is φj(i)\varphi^{(i)}_{j}-orthogonal where s(v,w)=i=1Nsi(v,w)s(v,w)=\sum_{i=1}^{N}s_{i}(v,w).

Now, we define π:VVaux\pi:\mathrm{V}\to\mathrm{V}_{\mathrm{aux}} to be the projection with respect to the inner product s(v,w)s(v,w). So, π\pi is defined by

π(v):=i=1Ncoarseπi(v)=i=1Ncoarsej=1Lisi(v,φj(i))φj(i),for each vV,\pi(v):=\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\pi_{i}(v)=\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\sum_{j=1}^{L_{i}}s_{i}(v,\varphi^{(i)}_{j})\varphi^{(i)}_{j},\quad\mbox{for each }v\in\mathrm{V},

where πi:L2(Ki)Vaux(i)\pi_{i}:L^{2}(K_{i})\to\mathrm{V}_{\mathrm{aux}}^{(i)} denotes the projection with respect to inner product si(,)s_{i}(\cdot,\cdot). The null space of the operator π\pi is defined by V~={vV:π(v)=0}\widetilde{\mathrm{V}}=\{v\in\mathrm{V}:\pi(v)=0\}. Now, we will construct the multiscale basis functions. Given a coarse block KiK_{i}, we denote the oversampling region Ki,mΩK_{i,m}\subset\Omega obtained by enlarging KiK_{i} with an arbitrary number of coarse grid layers m1m\geq 1, see Figure 1. Let V0(Ki,m):=H01(Ki,m)\mathrm{V}_{0}(K_{i,m}):=\mathrm{H}^{1}_{0}(K_{i,m}). Then, we define the multiscale basis function

ψj,ms(i)=argmin{a(ψ,ψ):ψV0(Ki,m),ψ is φj(i)-orthogonal},\psi_{j,\mathrm{ms}}^{(i)}=\mathrm{argmin}\{a(\psi,\psi):\psi\in\mathrm{V}_{0}(K_{i,m}),\,\psi\mbox{ is $\varphi^{(i)}_{j}$-orthogonal}\}, (3.2)

where V(Ki,m)\mathrm{V}(K_{i,m}) is the restriction of V0\mathrm{V}_{0} in Ki,mK_{i,m} and V0(Ki,m)\mathrm{V}_{0}(K_{i,m}) is the subspace of V(Ki,m)\mathrm{V}(K_{i,m}) with zero trace on Ki,m\partial K_{i,m}. The multiscale finite element space Vms\mathrm{V}^{\mathrm{ms}} is defined by

Vms=span{ψj,ms(i):1jLi,1iNcoarse}.\mathrm{V}^{\mathrm{ms}}=\mathrm{span}\{\psi_{j,\mathrm{ms}}^{(i)}:1\leq j\leq L_{i},1\leq i\leq N_{\tiny{\rm{coarse}}}\}.

By introducing the Lagrange multiplier, the problem (3.2) is equivalent to the explicit form: find ψj,ms(i)V0(Ki,m)\psi_{j,\mathrm{ms}}^{(i)}\in\mathrm{V}_{0}(K_{i,m}), λVaux(i)(Ki)\lambda\in\mathrm{V}_{\mathrm{aux}}^{(i)}(K_{i}) such that

{a(ψj,ms(i),η)+s(η,λ)=0,for all ηV(Ki,m),s(ψj,ms(i)φj(i),ν)=0,for all νVaux(i)(Ki,m),\begin{cases}a(\psi_{j,\mathrm{ms}}^{(i)},\eta)+s(\eta,\lambda)&=0,\quad\mbox{for all }\eta\in\mathrm{V}(K_{i,m}),\\ s(\psi_{j,\mathrm{ms}}^{(i)}-\varphi^{(i)}_{j},\nu)&=0,\quad\mbox{for all }\nu\in\mathrm{V}_{\mathrm{aux}}^{(i)}(K_{i,m}),\end{cases}

where Vaux(i)(Ki,m)\mathrm{V}_{\mathrm{aux}}^{(i)}(K_{i,m}) is the union of all local auxiliary spaces for KiKi,mK_{i}\subset K_{i,m}. Thus, the semi-discrete multiscale approximation reads as follows: find pnmsVmsp^{\mathrm{ms}}_{n}\in\mathrm{V}^{\mathrm{ms}} such that

(ϕtρ(pnms),v)+(κμρ(pnms)pnms,v)=(q,v),for each vVms.(\phi\partial_{t}\rho(p^{\mathrm{ms}}_{n}),v)+\left(\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n},\nabla v\right)=(q,v),\quad\mbox{for each }v\in\mathrm{V}^{\mathrm{ms}}. (3.3)

Using the backward Euler time-stepping scheme, we have a full-discrete formulation: find pnmsVmsp^{\mathrm{ms}}_{n}\in\mathrm{V}^{\mathrm{ms}} such that

(ϕρ(pnms),v)(ϕρ(pn1ms),v)+Δn(κμρ(pnms)pnms,v)=Δn(q,v),for each vVms.(\phi\rho(p^{\mathrm{ms}}_{n}),v)-(\phi\rho(p^{\mathrm{ms}}_{n-1}),v)+\Delta_{n}\left(\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n},\nabla v\right)=\Delta_{n}(q,v),\quad\mbox{for each }v\in\mathrm{V}^{\mathrm{ms}}. (3.4)

4 Convergence analysis

In this section, we establish the estimates of the convergence order of the proposed method.

4.1 Error estimates

In this subsection, we will present the convergence error estimates for the semi-discrete scheme (3.3). The analysis consists of two main steps. First, we derive the error estimate for the difference between the exact solution and its corresponding elliptic projection. Second, we estimate the difference between the solution of (2.1) and solution of (3.3) by the difference between exact solution and the elliptic projection solution of problem (2.1).

To begin, we let p^Vms\hat{p}\in\mathrm{V}^{\mathrm{ms}} be the elliptic projection of the function pVp\in\mathrm{V} that is defined by

(κμρ(p0)(pp^),w)=0,for each wVms.(\tfrac{\kappa}{\mu}\rho(p_{0})\nabla(p-\hat{p}),\nabla w)=0,\quad\mbox{for each }w\in\mathrm{V}^{\mathrm{ms}}. (4.1)

The following lemma gives us the error bound of p^\hat{p} for the nonlinear parabolic problem.

Lemma 4.1.

Let pp be the solution of (2.3). For each t>0t>0, we define the elliptic projection p^Vms\hat{p}\in\mathrm{V}^{\mathrm{ms}} that satisfies (4.1). Then,

(κμ)1/2(pp^)(t)0Λ1/2H(κμ)1/2(qtρ(p))(t)0,\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})(t)\|_{0}\preceq\Lambda^{-1/2}H\|(\tfrac{\kappa}{\mu})^{1/2}(q-\partial_{t}\rho(p))(t)\|_{0},

where Λ=min1iNλLi+1(i)\Lambda=\min_{1\leq i\leq N}\lambda^{(i)}_{L_{i}+1}.

Proof.

Let p^Vms\hat{p}\in\mathrm{V}^{\mathrm{ms}} be the projection of pp. By boundedness of ρ\rho and orthogonality property, we can write

(κμ)1/2(pp^)02Ω(κμ)ρ(p0)|(pp^)|2𝑑x\displaystyle\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|^{2}_{0}\preceq\int_{\Omega}(\tfrac{\kappa}{\mu})\rho(p_{0})|\nabla(p-\hat{p})|^{2}d\mathrm{x} =(κμρ(p0)(pp^),(pp^))\displaystyle=(\tfrac{\kappa}{\mu}\rho(p_{0})\nabla(p-\hat{p}),\nabla(p-\hat{p}))
=(κμρ(p0)p,(pp^)).\displaystyle=(\tfrac{\kappa}{\mu}\rho(p_{0})\nabla p,\nabla(p-\hat{p})).

Invoking again the boundedness of ρ\rho, we have

|(κμρ(p0)p,(pp^))||(κμρ(p)p,(pp^))|.|(\tfrac{\kappa}{\mu}\rho(p_{0})\nabla p,\nabla(p-\hat{p}))|\preceq|(\tfrac{\kappa}{\mu}\rho(p)\nabla p,\nabla(p-\hat{p}))|.

Now, from problem (2.3), we get that

(κμρ(p)p,(pp^))=(qϕtρ(p),(pp^)),for all t>0.(\tfrac{\kappa}{\mu}\rho(p)\nabla p,\nabla(p-\hat{p}))=(q-\phi\partial_{t}\rho(p),\nabla(p-\hat{p})),\quad\mbox{for all }t>0.

Therefore, we arrive at

(κμ)1/2(pp^)02(qϕtρ(p),pp^)κ~1/2(qϕtρ(p))0pp^s,for all t>0.\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|^{2}_{0}\preceq(q-\phi\partial_{t}\rho(p),p-\hat{p})\leq\|\widetilde{\kappa}^{-1/2}(q-\phi\partial_{t}\rho(p))\|_{0}\|p-\hat{p}\|_{s},\quad\mbox{for all }t>0. (4.2)

Since pp^V~p-\hat{p}\in\widetilde{\mathrm{V}}, implies that π(pp^)=0\pi(p-\hat{p})=0. According to [9], the coarse blocks KiK_{i} with i=1,,Ncoarsei=1,\dots,N_{\tiny{\rm{coarse}}} are disjoint, so we obtain that πi(pp^)=0\pi_{i}(p-\hat{p})=0, for all i=1,2,,Ncoarsei=1,2,\dots,N_{\tiny{\rm{coarse}}}. Thus, the local spectral problem (3.1) yields that

pp^s2=i=1Ncoarsepp^si2=i=1Ncoarse(Iπi)(pp^)si21Λi=1Ncoarse(κμ)1/2(pp^)0,Ki2,\|p-\hat{p}\|^{2}_{s}=\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|p-\hat{p}\|^{2}_{s_{i}}=\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|(I-\pi_{i})(p-\hat{p})\|^{2}_{s_{i}}\preceq\frac{1}{\Lambda}\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|^{2}_{0,K_{i}}, (4.3)

where Λ=min1iNλLi+1(i)\Lambda=\min_{1\leq i\leq N}\lambda^{(i)}_{L_{i}+1}. Therefore, by combining (4.2) and (4.3), and using the fact |χi|=𝒪(H1)|\nabla\chi_{i}|={\cal O}(H^{-1}), we obtain

(κμ)1/2(pp^)0Λ1/2Hκ1/2(qϕtρ(p))0.\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|_{0}\preceq\Lambda^{-1/2}H\|\kappa^{-1/2}(q-\phi\partial_{t}\rho(p))\|_{0}.

This completes the proof. ∎

The above estimate is the essence of the following result.

Lemma 4.2.

Under Assumptions of Lemma 4.1, the following estimates hold

pp^0\displaystyle\|p-\hat{p}\|_{0} Λ1H2κ1/2(qϕtρ(p))0,\displaystyle\preceq\Lambda^{-1}H^{2}\|\kappa^{-1/2}(q-\phi\partial_{t}\rho(p))\|_{0},
t(pp^)0\displaystyle\|\partial_{t}(p-\hat{p})\|_{0} Λ1H2κ1/2t(qϕtρ(p))0.\displaystyle\preceq\Lambda^{-1}H^{2}\|\kappa^{-1/2}\partial_{t}(q-\phi\partial_{t}\rho(p))\|_{0}.
Proof.

First, we will invoke the duality argument. For each t>0t>0, we define wV0w\in\mathrm{V}_{0} by

Ωκρ(p0)wvdx=Ω(pp^)v𝑑x,for each vV0,\int_{\Omega}\kappa\rho(p_{0})\nabla w\cdot\nabla vd\mathrm{x}=\int_{\Omega}(p-\hat{p})vd\mathrm{x},\quad\mbox{for each }v\in\mathrm{V}_{0},

and consider w^\hat{w} as elliptic projection of ww in Vms\mathrm{V}^{\mathrm{ms}}. By Lemma 4.1, for v=pp^v=p-\hat{p}, we have

pp^02=Ωκρ(p0)w(pp^)dx\displaystyle\|p-\hat{p}\|^{2}_{0}=\int_{\Omega}\kappa\rho(p_{0})\nabla w\cdot\nabla(p-\hat{p})d\mathrm{x} =Ωκρ(p0)(ww^)(pp^)dx\displaystyle=\int_{\Omega}\kappa\rho(p_{0})\nabla(w-\hat{w})\cdot\nabla(p-\hat{p})d\mathrm{x}
Ωκμρ(p0)(ww^)(pp^)dx\displaystyle\preceq\int_{\Omega}\tfrac{\kappa}{\mu}\rho(p_{0})\nabla(w-\hat{w})\cdot\nabla(p-\hat{p})d\mathrm{x}
(κμ)1/2(ww^)0(κμ)1/2(pp^)0\displaystyle\leq\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(w-\hat{w})\|_{0}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|_{0}
(HΛ1/2max{κ1/2}pp^0)\displaystyle\preceq\left(H\Lambda^{-1/2}\max\{\kappa^{-1/2}\}\|p-\hat{p}\|_{0}\right)
×(HΛ1/2κ1/2(qϕtρ(p))0).\displaystyle\quad\times\left(H\Lambda^{-1/2}\|\kappa^{-1/2}(q-\phi\partial_{t}\rho(p))\|_{0}\right).

Hence, we have

pp^0Λ1H2κ1/2(qϕtρ(p))0.\|p-\hat{p}\|_{0}\preceq\Lambda^{-1}H^{2}\|\kappa^{-1/2}(q-\phi\partial_{t}\rho(p))\|_{0}.

By a similar computation, we can obtain the second estimate. This completes the proof. ∎

We will derive an error estimate for the difference between the solution of (2.1) and the CEM-GMsFEM solution of (3.3) using the framework of [15].

Theorem 4.3.

Let pp be the solution obtained from (2.3), pmsVmsp^{\mathrm{ms}}\in\mathrm{V}^{\mathrm{ms}} be the multiscale solution of (3.3) using CEM-GMsFEM and p^\hat{p} be an elliptic projection of pp in Vms\mathrm{V}^{\mathrm{ms}}. Then, the following error estimate holds

(ppms)(t)02+0T(κμ)1/2(ppms)02𝑑t\displaystyle\|(p-p^{\mathrm{ms}})(t)\|_{0}^{2}+\int_{0}^{T}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-p^{\mathrm{ms}})\|^{2}_{0}dt Λ1H2(κ1/2t(qϕtρ(p))(t)02\displaystyle\preceq\Lambda^{-1}H^{2}\left(\|\kappa^{-1/2}\partial_{t}(q-\phi\partial_{t}\rho(p))(t)\|_{0}^{2}\right.
+0Tκ1/2t(qϕtρ(p))02dt)\displaystyle\quad+\left.\int_{0}^{T}\|\kappa^{-1/2}\partial_{t}(q-\phi\partial_{t}\rho(p))\|^{2}_{0}dt\right)
+(p^pms)(0)02.\displaystyle\quad+\|(\hat{p}-p^{\mathrm{ms}})(0)\|^{2}_{0}.
Proof.

Subtracting (3.3) from (2.3), and using (4.7), we have that

(ϕtρ(p),v)+(κμρ(p)p,v)(ϕtρ(pms),v)(κμρ(pms)pms,v)=0,for each vVms.(\phi\partial_{t}\rho(p),v)+\left(\tfrac{\kappa}{\mu}\rho(p)\nabla p,\nabla v\right)-(\phi\partial_{t}\rho(p^{\mathrm{ms}}),v)-\left(\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}})\nabla p^{\mathrm{ms}},\nabla v\right)=0,\quad\mbox{for each }v\in\mathrm{V}^{\mathrm{ms}}.

Since p^Vms\hat{p}\in\mathrm{V}^{\mathrm{ms}}, we put v=p^pmsv=\hat{p}-p^{\mathrm{ms}}, then follows that

(ϕt(ρ(p)ρ(pms)),p^pms)I1+(κμ(ρ(p)pρ(pms)pms),(p^pms))I2=0.\underset{I_{1}}{\underbrace{(\phi\partial_{t}(\rho(p)-\rho(p^{\mathrm{ms}})),\hat{p}-p^{\mathrm{ms}})}}+\underset{I_{2}}{\underbrace{\left(\tfrac{\kappa}{\mu}(\rho(p)\nabla p-\rho(p^{\mathrm{ms}})\nabla p^{\mathrm{ms}}),\nabla(\hat{p}-p^{\mathrm{ms}})\right)}}=0. (4.4)

About I1I_{1}, we can rewrite

I1=(ϕt(ρ(p^)ρ(pms)),p^pms)I3+(ϕt(ρ(p)ρ(p^)),p^pms)I4.I_{1}=\underset{I_{3}}{\underbrace{(\phi\partial_{t}(\rho(\hat{p})-\rho(p^{\mathrm{ms}})),\hat{p}-p^{\mathrm{ms}})}}+\underset{I_{4}}{\underbrace{(\phi\partial_{t}(\rho(p)-\rho(\hat{p})),\hat{p}-p^{\mathrm{ms}})}}. (4.5)

For I3I_{3}, we obtain

I3\displaystyle I_{3} =ddtΩϕ0p^pmsρ(p^+ξ)ξ𝑑ξ𝑑xΩϕ0p^pmsρ′′(p^+ξ)tp^ξdξdxI5\displaystyle=\frac{d}{dt}\int_{\Omega}\phi\int_{0}^{\hat{p}-p^{\mathrm{ms}}}\rho^{\prime}(\hat{p}+\xi)\xi d\xi d\mathrm{x}-\underset{I_{5}}{\underbrace{\int_{\Omega}\phi\int_{0}^{\hat{p}-p^{\mathrm{ms}}}\rho^{\prime\prime}(\hat{p}+\xi)\partial_{t}\hat{p}\xi d\xi d\mathrm{x}}}
+ΩϕΩ(ρ(pms)ρ(p^))tp^(p^pms)dxI6.\displaystyle\quad+\underset{I_{6}}{\underbrace{\int_{\Omega}\phi\int_{\Omega}(\rho^{\prime}(p^{\mathrm{ms}})-\rho^{\prime}(\hat{p}))\partial_{t}\hat{p}(\hat{p}-p^{\mathrm{ms}})d\mathrm{x}}}.

Following [25, 21], we have that the terms I5I_{5} and I6I_{6} are bounded by p^pms02\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0}. We deduce that

I3ddtΩϕ0p^pmsρ(p^+ξ)ξ𝑑ξ𝑑xC1p^pms02,I_{3}\geq\frac{d}{dt}\int_{\Omega}\phi\int_{0}^{\hat{p}-p^{\mathrm{ms}}}\rho^{\prime}(\hat{p}+\xi)\xi d\xi d\mathrm{x}-C_{1}\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0},

where C1C_{1} is a positive constant independent of the mesh size. In virtue of ρ\rho^{\prime} being bounded below positively, we have

Ωϕ0p^pmsρ(p^+ξ)ξ𝑑ξ𝑑xC2p^pms02.\int_{\Omega}\phi\int_{0}^{\hat{p}-p^{\mathrm{ms}}}\rho^{\prime}(\hat{p}+\xi)\xi d\xi d\mathrm{x}\geq C_{2}\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0}.

Then, for I3I_{3}, we obtain

I3=(ϕt(ρ(p^)ρ(pms)),p^pms)ddtp^pms02C3p^pms02.I_{3}=(\phi\partial_{t}(\rho(\hat{p})-\rho(p^{\mathrm{ms}})),\hat{p}-p^{\mathrm{ms}})\geq\frac{d}{dt}\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0}-C_{3}\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0}. (4.6)

For I4I_{4}, by using the chain rule and Young’s inequality, one can get

I4\displaystyle I_{4} =(ϕ(ρ(p)ρ(p^))tp^,p^pms)+(ϕρ(p)(tptp^),p^pms)\displaystyle=(\phi(\rho^{\prime}(p)-\rho^{\prime}(\hat{p}))\partial_{t}\hat{p},\hat{p}-p^{\mathrm{ms}})+(\phi\rho^{\prime}(p)(\partial_{t}p-\partial_{t}\hat{p}),\hat{p}-p^{\mathrm{ms}})
pp^02+t(pp^)02+p^pms02.\displaystyle\preceq\|p-\hat{p}\|^{2}_{0}+\|\partial_{t}(p-\hat{p})\|^{2}_{0}+\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0}.

Now, for I2I_{2}, we get

I2=(κμ(ρ(pms)(p^pms),(p^pms))I7+(κμ(ρ(p)pρ(pms)p^),(p^pms))I8I_{2}=\underset{I_{7}}{\underbrace{\left(\tfrac{\kappa}{\mu}(\rho(p^{\mathrm{ms}})\nabla(\hat{p}-p^{\mathrm{ms}}),\nabla(\hat{p}-p^{\mathrm{ms}})\right)}}+\underset{I_{8}}{\underbrace{\left(\tfrac{\kappa}{\mu}\left(\rho(p)\nabla p-\rho(p^{\mathrm{ms}})\nabla\hat{p}\right),\nabla(\hat{p}-p^{\mathrm{ms}})\right)}}

Then, for I7I_{7} we have

I7C(κμ)1/2(p^pms)02,I_{7}\geq C\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(\hat{p}-p^{\mathrm{ms}})\|^{2}_{0},

and for I8I_{8} by invoking Young’s inequality, one obtains

I8\displaystyle I_{8} =(κμ(ρ(p)ρ(pms))p,(p^pms))+(κμ(ρ(pms)pρ(pms)p^),(p^pms))\displaystyle=\left(\tfrac{\kappa}{\mu}(\rho(p)-\rho(p^{\mathrm{ms}}))\nabla p,\nabla(\hat{p}-p^{\mathrm{ms}})\right)+\left(\tfrac{\kappa}{\mu}(\rho(p^{\mathrm{ms}})\nabla p-\rho(p^{\mathrm{ms}})\nabla\hat{p}),\nabla(\hat{p}-p^{\mathrm{ms}})\right)
C((κμ)1/2(pp^)02+ppms02)+ϵ(κμ)1/2(p^pms)02,\displaystyle\leq C\left(\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|^{2}_{0}+\|p-p^{\mathrm{ms}}\|^{2}_{0}\right)+\epsilon\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(\hat{p}-p^{\mathrm{ms}})\|^{2}_{0},

where in the last inequality we use the boundedness of ρ\rho and ρ\rho^{\prime}, and pW1,(Ω)p\in W^{1,\infty}(\Omega). Combining the above estimates and taking ϵ\epsilon small enough, we can obtain

ddtp^pms02+(κμ)1/2(pmsp^)02\displaystyle\frac{d}{dt}\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0}+\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p^{\mathrm{ms}}-\hat{p})\|^{2}_{0} (κμ)1/2(pp^)02+pp^02+t(pp^)02\displaystyle\preceq\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|^{2}_{0}+\|p-\hat{p}\|^{2}_{0}+\|\partial_{t}(p-\hat{p})\|^{2}_{0}
+p^pms02.\displaystyle\quad+\|\hat{p}-p^{\mathrm{ms}}\|^{2}_{0}.

Integrating with respect to time tt and invoking the continuous Gronwall’s inequality [4], we can infer that

(p^pms)(t)02+0T(κμ)1/2(pmsp^)02𝑑t\displaystyle\|(\hat{p}-p^{\mathrm{ms}})(t)\|^{2}_{0}+\int_{0}^{T}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p^{\mathrm{ms}}-\hat{p})\|^{2}_{0}dt (p^pms)(0)02+0T(κμ)1/2(pp^)02𝑑t\displaystyle\preceq\|(\hat{p}-p^{\mathrm{ms}})(0)\|^{2}_{0}+\int_{0}^{T}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|^{2}_{0}dt
+0T(pp^02+t(pp^)02)𝑑t.\displaystyle\quad+\int_{0}^{T}\left(\|p-\hat{p}\|^{2}_{0}+\|\partial_{t}(p-\hat{p})\|^{2}_{0}\right)dt.

Thus, we use the triangle inequality to get

(ppms)(t)02+0T(κμ)1/2(ppms)02𝑑t\displaystyle\|(p-p^{\mathrm{ms}})(t)\|^{2}_{0}+\int_{0}^{T}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-p^{\mathrm{ms}})\|^{2}_{0}dt (p^pms)(0)02+0T((κμ)1/2(pp^)02dt\displaystyle\preceq\|(\hat{p}-p^{\mathrm{ms}})(0)\|^{2}_{0}+\int_{0}^{T}(\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p-\hat{p})\|^{2}_{0}dt
+0T(pp^02+t(pp^)02)𝑑t+(pp^)(t)02.\displaystyle\quad+\int_{0}^{T}\left(\|p-\hat{p}\|^{2}_{0}+\|\partial_{t}(p-\hat{p})\|^{2}_{0}\right)dt+\|(p-\hat{p})(t)\|^{2}_{0}.

Finally, the proof is completed by using Lemmas 4.1 and 4.2. ∎

4.2 A posteriori error estimate

We shall give an a posteriori error estimate, which provides a computable error bound to access the quality of the numerical solution. To begin, notice that since VmsVh\mathrm{V}^{\mathrm{ms}}\subset\mathrm{V}^{h}, we can derive from the fully-discrete approximation a residual expression defined by

rnms(v):=(ϕρ(pnms),v)(ϕρ(pn1ms),v)+Δn(κμρ(pnms)pnms,v)Δn(q,v),for each vVms.r^{\mathrm{ms}}_{n}(v):=(\phi\rho(p^{\mathrm{ms}}_{n}),v)-(\phi\rho(p^{\mathrm{ms}}_{n-1}),v)+\Delta_{n}\left(\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n},\nabla v\right)-\Delta_{n}(q,v),\quad\mbox{for each }v\in\mathrm{V}^{\mathrm{ms}}. (4.7)

We also consider, the local residuals. For each coarse node xix_{i}, we define ωi\omega_{i} be the set of coarse blocks having the vertex xix_{i}. For each coarse neighborhood ωi\omega_{i}, we define the local residual functional ri:Vr_{i}:\mathrm{V}\to\mathbb{R} by

rn(i)(v)=r(χiv)=(ϕρ(pnms),χiv)(ϕρ(pn1ms),χiv)+Δn(κμρ(pnms)pnms,χiv)Δn(q,χiv),r_{n}^{(i)}(v)=r(\chi_{i}v)=(\phi\rho(p^{\mathrm{ms}}_{n}),\chi_{i}v)-(\phi\rho(p^{\mathrm{ms}}_{n-1}),\chi_{i}v)+\Delta_{n}\left(\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n},\nabla\chi_{i}v\right)-\Delta_{n}(q,\chi_{i}v),

for all vVv\in\mathrm{V}. The local residual rir_{i} gives a measure of the error ppnmsp-p^{\mathrm{ms}}_{n} in the coarse neighborhood ωi\omega_{i}.

Theorem 4.4.

Let pnp_{n} be the solution obtained from (2.1) at tnt_{n} and pnmsVmsp^{\mathrm{ms}}_{n}\in\mathrm{V}^{\mathrm{ms}} denote the CEM-GMsFEM solution of the fully discrete scheme of (3.4) at tnt_{n}. Then, there exists a positive constant CC independent of the mesh size such that

pNtimepNtimems02+n=1NtimeΔn(κμ)1/2(pnpnms)02(1+Λ1)n=1Ntimei=1Ncoarser~inVi2+p0p0ms02,\|p_{N_{\tiny{\rm{time}}}}-p^{\mathrm{ms}}_{N_{\tiny{\rm{time}}}}\|^{2}_{0}+\sum_{n=1}^{N_{\tiny{\rm{time}}}}\Delta_{n}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p_{n}-p^{\mathrm{ms}}_{n})\|^{2}_{0}\preceq(1+\Lambda^{-1})\sum_{n=1}^{N_{\tiny{\rm{time}}}}\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|\widetilde{r}^{n}_{i}\|^{2}_{\mathrm{V}^{*}_{i}}+\|p_{0}-p^{\mathrm{ms}}_{0}\|^{2}_{0},

where

r~n(i)=Δnωiqnv𝑑xωiϕ(ρ(pnms)ρ(pn1ms))v𝑑xΔnωiκμρ(pnms)pnmsvdx,\widetilde{r}_{n}^{(i)}=\Delta_{n}\int_{\omega_{i}}q_{n}vd\mathrm{x}-\int_{\omega_{i}}\phi(\rho(p^{\mathrm{ms}}_{n})-\rho(p^{\mathrm{ms}}_{n-1}))vd\mathrm{x}-\Delta_{n}\int_{\omega_{i}}\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n}\cdot\nabla vd\mathrm{x},

and the residual norm is defined by

r~n(i)Vi=supvL2(tn,tn+1;H01(ωi))r~n(i)vVi.\|\widetilde{r}_{n}^{(i)}\|_{\mathrm{V}^{*}_{i}}=\sup_{v\in\mathrm{L}^{2}(t_{n},t_{n+1};\mathrm{H}^{1}_{0}(\omega_{i}))}\frac{\widetilde{r}_{n}^{(i)}}{\|v\|_{\mathrm{V}_{i}}}.
Proof.

Subtracting (3.4) from (2.4), we get for pVp\in\mathrm{V} at tnt_{n}

(ϕ(ρ(pn)ρ(pnms)),v)(ϕ(ρ(pn1)ρ(pn1ms),v)+Δn(κμ(ρ(pn)pnρ(pnms)pnms,v)=0,(\phi(\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})),v)-(\phi(\rho(p_{n-1})-\rho(p^{\mathrm{ms}}_{n-1}),v)+\Delta_{n}\left(\tfrac{\kappa}{\mu}(\rho(p_{n})\nabla p_{n}-\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n},\nabla v\right)=0, (4.8)

for each vVmsv\in\mathrm{V}^{\mathrm{ms}} Putting v=pnpnmsv=p_{n}-p^{\mathrm{ms}}_{n} and using the fact that ρ\rho is bounded below positively, we easily obtain that

(ϕ(ρ(pn)ρ(pnms)),pnpnms)Cpnpnms02.(\phi(\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})),p_{n}-p^{\mathrm{ms}}_{n})\geq C\|p_{n}-p^{\mathrm{ms}}_{n}\|^{2}_{0}.

Similarly, for the second term of (4.8), we can use the boundedness of ρ\rho and the Young’s inequality to yield

(ϕ(ρ(pn1)ρ(pn1ms),pnpnms)Cpn1pn1ms02+ϵpnpnms02.(\phi(\rho(p_{n-1})-\rho(p^{\mathrm{ms}}_{n-1}),p_{n}-p^{\mathrm{ms}}_{n})\leq C\|p_{n-1}-p^{\mathrm{ms}}_{n-1}\|^{2}_{0}+\epsilon\|p_{n}-p^{\mathrm{ms}}_{n}\|^{2}_{0}.

Gathering the above inequalities and for ϵ\epsilon small enough, we arrive to

(ϕ(ρ(pn)ρ(pnms)),pnpnms)(ϕ(ρ(pn1)ρ(pn1ms),pnpnms)C(pnpnms02pn1pn1ms02).\begin{split}(\phi(\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})),p_{n}-p^{\mathrm{ms}}_{n})-(\phi(\rho(p_{n-1})-\rho(p^{\mathrm{ms}}_{n-1}),p_{n}-p^{\mathrm{ms}}_{n})&\geq C\left(\|p_{n}-p^{\mathrm{ms}}_{n}\|^{2}_{0}\right.\\ &\quad-\left.\|p_{n-1}-p^{\mathrm{ms}}_{n-1}\|^{2}_{0}\right).\end{split}

For third term of (4.8), we have that

(κμ(ρ(pn)pnρ(pnms)pnms),(pnpnms))C(κμ)1/2(pnpnms)02.\left(\tfrac{\kappa}{\mu}\left(\rho(p_{n})\nabla p_{n}-\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n}\right),\nabla(p_{n}-p^{\mathrm{ms}}_{n})\right)\geq C\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p_{n}-p^{\mathrm{ms}}_{n})\|_{0}^{2}.

Thus, these above inequalities drive us the expression

pnpnms02\displaystyle\|p_{n}-p^{\mathrm{ms}}_{n}\|^{2}_{0} pn1pn1ms02+Δn(κμ)1/2(pnpnms)02(ϕ(ρ(pn)ρ(pnms)),pnpnms)\displaystyle-\|p_{n-1}-p^{\mathrm{ms}}_{n-1}\|^{2}_{0}+\Delta_{n}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p_{n}-p^{\mathrm{ms}}_{n})\|^{2}_{0}\preceq(\phi(\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})),p_{n}-p^{\mathrm{ms}}_{n})
(ϕ(ρ(pn1)ρ(pn1ms)),pnpnms)\displaystyle\quad-(\phi(\rho(p_{n-1})-\rho(p^{\mathrm{ms}}_{n-1})),p_{n}-p^{\mathrm{ms}}_{n})
+Δn(κμ(ρ(pn)pnρ(pnms)pnms),(pnpnms))\displaystyle\quad+\Delta_{n}\left(\tfrac{\kappa}{\mu}\left(\rho(p_{n})\nabla p_{n}-\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n}\right),\nabla(p_{n}-p^{\mathrm{ms}}_{n})\right)
=(ϕ(ρ(pn)ρ(pnms)),pnpnms)(ϕ(ρ(pn1)ρ(pn1ms)),pnpnms)\displaystyle=(\phi(\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})),p_{n}-p^{\mathrm{ms}}_{n})-(\phi(\rho(p_{n-1})-\rho(p^{\mathrm{ms}}_{n-1})),p_{n}-p^{\mathrm{ms}}_{n})
+Δn(κμ(ρ(pn)pnρ(pn)pnms),(pnpnms))\displaystyle\quad+\Delta_{n}\left(\tfrac{\kappa}{\mu}\left(\rho(p_{n})\nabla p_{n}-\rho(p_{n})\nabla p^{\mathrm{ms}}_{n}\right),\nabla(p_{n}-p^{\mathrm{ms}}_{n})\right)
+Δn(κμ(ρ(pn)pnmsρ(pnms)pnms),(pnpnms)).\displaystyle\quad+\Delta_{n}\left(\tfrac{\kappa}{\mu}\left(\rho(p_{n})\nabla p^{\mathrm{ms}}_{n}-\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n}\right),\nabla(p_{n}-p^{\mathrm{ms}}_{n})\right).

Reorganizing the terms and using the definition of the weak formulation (2.1), we get that

pnpnms02pn1pn1ms02+Δn(κμ)1/2(pnpnms)02Δn(qn,pnpnms)(ϕ(ρ(pnms)ρ(pn1ms)),pnpnms)Δn(κμρ(pn)pnms,(pnpnms))+Δn(κμ(ρ(pn)ρ(pnms))pnms,(pnpnms)).\begin{split}&\|p_{n}-p^{\mathrm{ms}}_{n}\|^{2}_{0}-\|p_{n-1}-p^{\mathrm{ms}}_{n-1}\|^{2}_{0}+\Delta_{n}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p_{n}-p^{\mathrm{ms}}_{n})\|^{2}_{0}\preceq\Delta_{n}(q^{n},p_{n}-p^{\mathrm{ms}}_{n})\\ &\quad-(\phi(\rho(p^{\mathrm{ms}}_{n})-\rho(p^{\mathrm{ms}}_{n-1})),p_{n}-p^{\mathrm{ms}}_{n})-\Delta_{n}\left(\tfrac{\kappa}{\mu}\rho(p_{n})\nabla p^{\mathrm{ms}}_{n},\nabla(p_{n}-p^{\mathrm{ms}}_{n})\right)\\ &\quad+\Delta_{n}\left(\tfrac{\kappa}{\mu}\left(\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})\right)\nabla p^{\mathrm{ms}}_{n},\nabla(p_{n}-p^{\mathrm{ms}}_{n})\right).\end{split} (4.9)

We will limit the right-hand side of (4.9). In light of residual expression (4.7), we have that

rms(v)=0,for each vVms.r^{\mathrm{ms}}(v)=0,\quad\mbox{for each }v\in\mathrm{V}^{\mathrm{ms}}.

Denote w=pnmspnw=p^{\mathrm{ms}}_{n}-p_{n} and use w^Vms\hat{w}\in\mathrm{V}^{\mathrm{ms}}, the elliptic projection of ww. Thus,

rnms(w)=rnms(ww^)\displaystyle r^{\mathrm{ms}}_{n}(w)=r^{\mathrm{ms}}_{n}(w-\hat{w}) =Δn(qn,ww^)(ϕ(ρ(pnms)ρ(pn1ms)),ww^)\displaystyle=\Delta_{n}(q_{n},w-\hat{w})-(\phi(\rho(p^{\mathrm{ms}}_{n})-\rho(p^{\mathrm{ms}}_{n-1})),w-\hat{w})
Δn(κμρ(pnms)pnms,(ww^)).\displaystyle\quad-\Delta_{n}\left(\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n},\nabla(w-\hat{w})\right).

Let us rewrite rnms(ww^)=i=1Ncoarser~n(i)(χi(ww^))r^{\mathrm{ms}}_{n}(w-\hat{w})=\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\widetilde{r}_{n}^{(i)}(\chi_{i}(w-\hat{w})) [10, 22], then

i=1Ncoarser~n(i)(χi(ww^))=Δni=1Ncoarse(ωiqnχi(ww^)dxωiϕρ(pnms)ρ(pn1ms)Δnχi(ww^)dxωiκμρ(pnms)pnmsχi(ww^)dx).\begin{split}\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\widetilde{r}_{n}^{(i)}(\chi_{i}(w-\hat{w}))&=\Delta_{n}\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\left(\int_{\omega_{i}}q_{n}\chi_{i}(w-\hat{w})d\mathrm{x}-\int_{\omega_{i}}\phi\frac{\rho(p^{\mathrm{ms}}_{n})-\rho(p^{\mathrm{ms}}_{n-1})}{\Delta_{n}}\chi_{i}(w-\hat{w})d\mathrm{x}\right.\\ &\quad-\left.\int_{\omega_{i}}\tfrac{\kappa}{\mu}\rho(p^{\mathrm{ms}}_{n})\nabla p^{\mathrm{ms}}_{n}\cdot\nabla\chi_{i}(w-\hat{w})d\mathrm{x}\right).\end{split} (4.10)

Note that

i=1Ncoarser~in(χi(ww^))i=1Ncoarser~inViχi(ww^)Vi,\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\widetilde{r}^{n}_{i}(\chi_{i}(w-\hat{w}))\leq\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|\widetilde{r}^{n}_{i}\|_{\mathrm{V}^{*}_{i}}\|\chi_{i}(w-\hat{w})\|_{\mathrm{V}_{i}}, (4.11)

where vVi=v0,ωi2+Δn(κμ)1/2v0,ωi2\|v\|_{\mathrm{V}_{i}}=\|v\|^{2}_{0,\omega_{i}}+\Delta_{n}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla v\|^{2}_{0,\omega_{i}}, where 0,ωi\|\cdot\|_{0,\omega_{i}} denotes the L2\mathrm{L}^{2}-norm restricted to ωi\omega_{i}. Notice also that,

(κμ)1/2χi(ww^)0,ωi(ww^s(ωi)2+(κμ)1/2(ww^)0,ωi2)1/2,\|(\tfrac{\kappa}{\mu})^{1/2}\nabla\chi_{i}(w-\hat{w})\|_{0,\omega_{i}}\preceq\left(\|w-\hat{w}\|^{2}_{s(\omega_{i})}+\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(w-\hat{w})\|^{2}_{0,\omega_{i}}\right)^{1/2}, (4.12)

where s(ωi)\|\cdot\|_{s(\omega_{i})} represents the ss-norm restricted to ωi\omega_{i}. For the second term on the right-hand side, by using the orthogonality property, i.e., ((κμ)1/2ρ(p0)(ww^),v)=0((\tfrac{\kappa}{\mu})^{1/2}\rho(p_{0})\nabla(w-\hat{w}),\nabla v)=0, for all vVmsv\in\mathrm{V}^{\mathrm{ms}}, we get

(κμ)1/2(ww^)0,ωi2(κμ)1/2w0,ωi2.\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(w-\hat{w})\|^{2}_{0,\omega_{i}}\preceq\|(\tfrac{\kappa}{\mu})^{1/2}\nabla w\|^{2}_{0,\omega_{i}}. (4.13)

Now, for the first term on the right-hand side of (4.12), we shall use the duality argument. Let g=κ~(ww^)g=\widetilde{\kappa}(w-\hat{w}) and zV0z\in\mathrm{V}_{0} be the solution of problem below

ωiκρ(p0)zvdx=ωigv𝑑x,for each vV0.\int_{\omega_{i}}\kappa\rho(p_{0})\nabla z\cdot\nabla vd\mathrm{x}=\int_{\omega_{i}}gvd\mathrm{x},\quad\mbox{for each }v\in\mathrm{V}_{0}.

Putting v=ww^v=w-\hat{w}, using the Cauchy-Schwarz inequality and equation (4.13), we arrives to

ww^s(ωi)2\displaystyle\|w-\hat{w}\|^{2}_{s(\omega_{i})} =ωig(ww^)𝑑x=ωiκρ(p0)z(ww^)dx\displaystyle=\int_{\omega_{i}}g(w-\hat{w})d\mathrm{x}=\int_{\omega_{i}}\kappa\rho(p_{0})\nabla z\cdot\nabla(w-\hat{w})d\mathrm{x}
=ωiκρ(p0)(zz^)(ww^)dx\displaystyle=\int_{\omega_{i}}\kappa\rho(p_{0})\nabla(z-\hat{z})\cdot\nabla(w-\hat{w})d\mathrm{x}
(κρ(p0))1/2(zz^)0,ωi(κρ(p0))1/2(ww^)0,ωi\displaystyle\leq\|(\kappa\rho(p_{0}))^{1/2}\nabla(z-\hat{z})\|_{0,\omega_{i}}\|(\kappa\rho(p_{0}))^{1/2}\nabla(w-\hat{w})\|_{0,\omega_{i}}
Λ1/2κ~1/2gL2(ωi)(κμ)1/2(ww^)0,ωi\displaystyle\preceq\Lambda^{-1/2}\|\widetilde{\kappa}^{-1/2}g\|_{\mathrm{L}^{2}(\omega_{i})}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(w-\hat{w})\|_{0,\omega_{i}}
Λ1/2ww^s(ωi)(κμ)1/2w0,ωi.\displaystyle\preceq\Lambda^{-1/2}\|w-\hat{w}\|_{s(\omega_{i})}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla w\|_{0,\omega_{i}}.

So, we have

ww^s(ωi)Λ1/2(κμ)1/2w0,ωi.\|w-\hat{w}\|_{s(\omega_{i})}\preceq\Lambda^{-1/2}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla w\|_{0,\omega_{i}}. (4.14)

Gathering (4.12)-(4.14), we arrive to

(κμ)1/2χi(ww^)0,ωi(1+Λ1)1/2(κμ)1/2w0,ωi.\|(\tfrac{\kappa}{\mu})^{1/2}\nabla\chi_{i}(w-\hat{w})\|_{0,\omega_{i}}\preceq(1+\Lambda^{-1})^{1/2}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla w\|_{0,\omega_{i}}.

Analogously, we estimate χi(ww^)0,ωi(κμ)1/2w0,ωi\|\chi_{i}(w-\hat{w})\|_{0,\omega_{i}}\preceq\|(\tfrac{\kappa}{\mu})^{1/2}\nabla w\|_{0,\omega_{i}}. Therefore,

i=1Ncoarser~n(i)Viχi(ww^)Vi(1+Λ1)1/2i=1Ncoarser~n(i)Vi(κμ)1/2w0,ωi.\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|\widetilde{r}_{n}^{(i)}\|_{\mathrm{V}^{*}_{i}}\|\chi_{i}(w-\hat{w})\|_{\mathrm{V}_{i}}\preceq(1+\Lambda^{-1})^{1/2}\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|\widetilde{r}_{n}^{(i)}\|_{\mathrm{V}^{*}_{i}}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla w\|_{0,\omega_{i}}. (4.15)

For the last term of the right-hand side of (4.9), we have

Δn(κμ(ρ(pn)ρ(pnms))pnms,(pnpnms))Δnρ(pn)ρ(pnms)0(κμ)1/2(pnpnms)0.\Delta_{n}\left(\tfrac{\kappa}{\mu}\left(\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})\right)\nabla p^{\mathrm{ms}}_{n},\nabla(p_{n}-p^{\mathrm{ms}}_{n})\right)\preceq\Delta_{n}\|\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})\|_{0}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p_{n}-p^{\mathrm{ms}}_{n})\|_{0}. (4.16)

Combining (4.15) and (4.16), and using Young’s inequality, one can express for (4.9) by summing over all nn

pNtimepNtimems02+n=1NtimeΔn(κμ)1/2(pnpnms)02\displaystyle\|p_{N_{\tiny{\rm{time}}}}-p^{\mathrm{ms}}_{N_{\tiny{\rm{time}}}}\|^{2}_{0}+\sum_{n=1}^{N_{\tiny{\rm{time}}}}\Delta_{n}\|(\tfrac{\kappa}{\mu})^{1/2}\nabla(p_{n}-p^{\mathrm{ms}}_{n})\|^{2}_{0} (1+Λ1)n=1Ntimei=1Ncoarser~n(i)Vi2\displaystyle\preceq(1+\Lambda^{-1})\sum_{n=1}^{N_{\tiny{\rm{time}}}}\sum_{i=1}^{N_{\tiny{\rm{coarse}}}}\|\widetilde{r}_{n}^{(i)}\|^{2}_{\mathrm{V}^{*}_{i}}
+n=1NtimeΔnρ(pn)ρ(pnms)0+p0p0ms02.\displaystyle\quad+\sum_{n=1}^{N_{\tiny{\rm{time}}}}\Delta_{n}\|\rho(p_{n})-\rho(p^{\mathrm{ms}}_{n})\|_{0}+\|p_{0}-p^{\mathrm{ms}}_{0}\|^{2}_{0}.

The proof is completed by using the discrete Gronwall inequality. ∎

5 Numerical results

We now present numerical results of the nonlinear single-phase compressible flow in highly heterogeneous porous media with the performance of the CEM-GMsFEM that are summarized in the following three separate experiments. All parameters on the flow model, boundary, and initial conditions in each numerical experiment are described in detail to allow their proper reproduction of them. The main aim of the simulation is to demonstrate the viability of the proposed numerical approximation and improve the convergence rate shown in Section 4. We implement the CEM-GMsFEM in Matlab language and use the numerical experiments presented in [27, 15] as a reference guide to our three-dimensional experiments. We will use the Euler-backward for the time discretization and a Newton-Raphson method with a tolerance of 10610^{-6} for the non-linear problem. Only 242-4 Newton’s iterations are needed in the computations presented below.

Refer to caption
(a) Perm1
Refer to caption
(b) perm1
Refer to caption
(c) Perm3
Refer to caption
(d) Perm4
Figure 2: Different permeability fields.

We consider three high-contrast permeability fields that are the disjoint union of a background region with 10510^{5} millidarcys and other regions of 10910^{9} millidarcys (see Figure 2). We also consider a fractured porous medium. In this case, the permeability value in fractures is much larger than in the surrounding medium. Finally, we employ the first 3030 layers of SPE10 33D dataset from [7], which is widely used in the reservoir simulation community to test multiscale approaches. All experiments employ parameters of viscosity μ=5\mu=5cP, porosity ϕ=500\phi=500, fluid compressibility c=1.0×108Pa1c=1.0\times 10^{-8}\text{Pa}^{-1}, the reference pressure pref=2.00×107p_{\mathrm{ref}}=2.00\times 10^{7}Pa, and the reference density ρref=850\rho_{\mathrm{ref}}=850kg/m3\mbox{m}^{3}.

Example 5.1.

For our first example, we set a fine grid resolution of 64364^{3}, with a size of h=20h=20m, and a coarse grid resolution of 838^{3} of size H=8hH=8h. The coarsening factor is chosen due this coarse grid provides the most computationally efficient performance for the method. For the CEM-GMsFEM, we use 44 basis functions and 44 oversampling layers. We know well that the number of bases is sufficient to improve the accuracy of CEM-GMsFEM; see [9]. Then, we have a coarse system with dimension 4916(=729×number of basis functions)4916\,(=729\times\mbox{number of basis functions}), and the fine-scale system has a dimension of 274625274625. The permeability field κ1\kappa_{1} used in this experiment is depicted in Figure 2a. We define a model configuration as follows: four vertical injectors in each corner and a unit sink in the center of the domain to drive the flow, and employ the full zero Neumann boundary condition and an initial pressure field p0=2.16×107p_{0}=2.16\times 10^{7}Pa. We consider a fine grid resolution of 64364^{3}, whose fine grid size is given by h=20h=20m, meanwhile the coarse grid resolution of 838^{3}, whose coarse size is given by H=8hH=8h. The time step Δn\Delta_{n} is 77 days, and the total time simulation will be T=25Δn(=175 days)T=25\Delta_{n}(=175\mbox{ days}). Figure 3 shows the pressure profiles with the sink term and zero Neumann boundary conditions in Fig. 3 for different instants at day t=77t=77 and t=147t=147. In this case, we obtain a relative L2L^{2}- and H1H^{1}-error of 2.1138E-03 and 3.8058E-02 respectively.

Refer to caption
(a) example1
Refer to caption
(b) example1
Figure 3: Numerical solution of Example 5.1 using full-zero Neumann boundary conditions and high-contrast permeability field κ1\kappa_{1}, see Figure 2a. The fine-scale reference solution (left) and CEM-GMsFEM solution (right) with 44 basis function and 44 oversampling layers at (a) t=77t=77 and (b) t=147t=147.
Example 5.2.

We consider a combination of zero Neumann and nonzero Dirichlet boundary conditions as in [27, 15]. We set a fine grid resolution of 64364^{3}, with a size of h=20h=20m, and different coarse grid resolutions of 43,834^{3},8^{3} and 16316^{3}. The time step Δn\Delta_{n} and total time simulation are the same as Example 5.1. We impose zero Neumann condition on boundaries of planes xyxy and xzxz and let p=2.16×107p=2.16\times 10^{7}Pa in the first yzyz plane and p=2.00×107p=2.00\times 10^{7}Pa in the last yzyz plane for all time instants, no additional source is imposed. The permeability field used is κ2\kappa_{2} (Figure 2b). The pressure difference will drive the flow, and the initial field p0p_{0} linearly decreases along the xx axis and is fixed in the yzyz plane. Table 1 shows that numerical results use 44 basis functions on each coarse block with different coarse grid sizes (H=4h,8hH=4h,8h and 16h16h), where ε0\varepsilon_{0} and ε1\varepsilon_{1} denote the relative L2L^{2} and energy error estimate between the reference solution and CEM-GMsFEM solution defined by

ε0=(i=1Ntime(pihpims)2i=1Ntime(pih)2)1/2,ε1=(i=1Ntime(κμ)1/2ρ(p0)(pihpims)2i=1Ntime(pih)2)1/2,\varepsilon_{0}=\left(\frac{\sum_{i=1}^{N_{\tiny{\rm{time}}}}(p^{h}_{i}-p^{\mathrm{ms}}_{i})^{2}}{\sum_{i=1}^{N_{\tiny{\rm{time}}}}(p^{h}_{i})^{2}}\right)^{1/2},\quad\varepsilon_{1}=\left(\frac{\sum_{i=1}^{N_{\tiny{\rm{time}}}}(\frac{\kappa}{\mu})^{1/2}\rho(p_{0})\nabla(p^{h}_{i}-p^{\mathrm{ms}}_{i})^{2}}{\sum_{i=1}^{N_{\tiny{\rm{time}}}}(p^{h}_{i})^{2}}\right)^{1/2},

where pihp^{h}_{i} denotes the references solution and pimsp^{\mathrm{ms}}_{i} is the CEM-GMsFEM approximation for i=1,,Ntimei=1,\dots,N_{\tiny{\rm{time}}}. For instance, for a coarse grid size of H=8hH=8h, we obtain the relative errors ε0=\varepsilon_{0}= 4.5581E-04 and ε1=\varepsilon_{1}= 2.5249E-01. In Figure 4, we depict the numerical solution profiles with a fine grid resolution of 64364^{3} and coarse grid resolution of 838^{3} at day t=105t=105 and t=140t=140, which is hard to find any difference between the reference solution and CEM-GMsFEM solution. Therefore, we have a good agreement.

Number of basis HH Number of oversampling layers mm ε0\varepsilon_{0} ε1\varepsilon_{1}
4 4h4h 33 2.4004E-03 4.4227E-01
4 8h8h 44 4.5581E-04 2.5249E-01
4 16h16h 55 1.4257E-04 1.258E-01
Table 1: Convergence rate for Example 5.2 with different numbers of oversampling layers (mm) with a combination of zero Neumann and nonzero Dirichlet boundary conditions.
Refer to caption
(a) example2
Refer to caption
(b) example2
Figure 4: Numerical solution of Example 5.2 combining a zero Neumann boundary condition and nonzero Dirichlet boundary condition. High-contrast permeability field κ2\kappa_{2}, fine-scale reference solution (left), and CEM-GMsFEM solution (right) with 44 basis function and 44 number of oversampling layers at (a) t=105t=105 and (b) t=140t=140.
Example 5.3.

For the third experiment, we consider the combination of zero Neumann and nonzero Dirichlet boundary conditions as in Example 5.2. We set a fine grid resolution of 32332^{3} (fine-scale system with dimension 3593735937), with a size of h=20h=20m, and different coarse grid resolutions of 43,834^{3},8^{3} and 16316^{3} (coarse-scale system with dimension 500500, 29162916 and 1965219652 respectively). The time step Δn\Delta_{n} and total time simulation are the same as Example 5.1. The fractured medium κ3\kappa_{3} used is depicted in Figure 2c. For this experiment, we employ the framework from [14] and apply the CEM-GMsFEM to the 33D model. The domain Ω\Omega can be represented by

Ω=Ω0(iΩfrac,i),\Omega=\Omega_{0}\cup(\cup_{i}\Omega_{\mathrm{frac},i}),

where Ω0\Omega_{0} represents the matrix and subscript frac\mathrm{frac} denotes the fracture regions. Then, we can write the finite element discretization corresponding to equation (2.3)

(ϕtρ(ph),v)+(κμρ(ph)ph,v)\displaystyle(\phi\partial_{t}\rho(p^{h}),v)+\left(\tfrac{\kappa}{\mu}\rho(p^{h})\nabla p^{h},\nabla v\right) =(ϕtρ(ph),v)Ω0+i(ϕtρ(ph),v)Ωfrac,i\displaystyle=(\phi\partial_{t}\rho(p^{h}),v)_{\Omega_{0}}+\sum_{i}(\phi\partial_{t}\rho(p^{h}),v)_{\Omega_{\mathrm{frac},i}}
+(κμρ(ph)ph,v)Ω0+i(κμρ(ph)ph,v)Ωfrac,i\displaystyle\quad+\left(\tfrac{\kappa}{\mu}\rho(p^{h})\nabla p^{h},\nabla v\right)_{\Omega_{0}}+\sum_{i}\left(\tfrac{\kappa}{\mu}\rho(p^{h})\nabla p^{h},\nabla v\right)_{\Omega_{\mathrm{frac},i}}
=(q,v),for each vVh,\displaystyle=(q,v),\quad\mbox{for each }v\in\mathrm{V}^{h},

In Table 2, we give the convergence rate T=25Δn(=175 days)T=25\Delta_{n}(=175\text{ days}) with different coarse-grid sizes HH. We notice that the error significantly decreases as the size of the coarse grid is finer. Then, we have that the CEM-GMsFEM gives a good approximation of the solution for the case of the fractured medium. Figure 5 shows the numerical solutions at t=70t=70 and t=140t=140.

Number of basis HH Number of oversampling layers mm ε0\varepsilon_{0} ε1\varepsilon_{1}
4 4h4h 33 1.6111E-03 2.9441E-01
4 8h8h 44 1.9532E-04 1.1653E-01
4 16h16h 55 1.0160E-04 5.1492E-02
Table 2: Convergence rate of Example 5.3 with different numbers of oversampling layers (mm) with a combination of zero Neumann and nonzero Dirichlet boundary conditions.
Refer to caption
(a) example3
Refer to caption
(b) example2
Figure 5: Numerical solution of Example 5.3 combining a zero Neumann boundary condition and nonzero Dirichlet boundary condition. High-contrast permeability field κ3\kappa_{3}. The fine-scale reference solution (left) and CEM-GMsFEM solution (right) with 44 basis function and 44 number of oversampling layers at (a) t=70t=70 and (b) t=140t=140.
Example 5.4.

For the last experiment, we consider the same boundary conditions as Example 5.2. The permeability field used is κ4\kappa_{4} (see Figure LABEL:fig:perm4) and the time step Δn\Delta_{n} is 77 days, with total time simulation will be T=26Δn(=186 days)T=26\Delta_{n}(=186\mbox{ days}). We set a fine grid resolution of 220×60×30220\times 60\times 30 (fine-scale system with a dimension of 417911417911), with a size of h=20h=20m, and coarse grid resolution of 10310^{3} (coarse-scale system with a dimension 53245324). We show the pressure profiles comparison in Figure 6. This experiment obtains an error estimate of ε0=\varepsilon_{0}=1.8377E-03 and ε1=\varepsilon_{1}=3.4547E-01, respectively.

Refer to caption
(a) example4
Refer to caption
(b) example4
Figure 6: Example 4. Mixed boundary conditions (a) t=84t=84 and (b) t=126t=126.

6 Concluding remarks

This paper studies the convergence of the numerical approximations to the highly heterogeneous nonlinear single-phase compressible flow by CEM-GMsFEM. We first build an auxiliary for the proposed method by solving spectral problems. Then, we construct a multiscale basis function by solving some constraint energy minimization problems in the oversampling local regions. So, we obtain multiscale basis functions for the pressure. This work defines the elliptic projection in the multiscale space spanned by CEM-GMsFEM basis functions for convergence analysis. Thus, we present the convergence of the semi-discrete formulation. The convergence depends on the coarse mesh size and the decay of eigenvalues of local spectral problems; an a posteriori error estimate is derived underlining discretization. Some numerical examples have been presented to verify the feasibility of the proposed method concerning convergence and stability. We observe that the CEM-GMsFEM is shown to have a second-order convergence rate in the L2L^{2}-norm and a first-order convergence rate in the energy norm concerning the coarse grid size.

A foreseeable result in ongoing research is to boost the performance of the coarse-grid simulation, mainly where the source term is singular; one may need to further improve the accuracy of the approximation without additional refinement in the grid. We can enrich the multiscale space for such a goal by adding additional basis functions in the online stage [10]. These new multiscale basis functions are constructed on the oversampling technique and the information on local residuals. Consequently, we could present an adaptive enrichment algorithm to reduce error in some regions with large residuals.

Acknowledgement

The research of Eric Chung is partially supported by the Hong Kong RGC General Research Fund (Projects: 14305222 and 14304021).

References

  • [1] Manal Alotaibi, Huangxin Chen, and Shuyu Sun. Generalized multiscale finite element methods for the reduced model of darcy flow in fractured porous media. Journal of Computational and Applied Mathematics, 413:114305, 2022.
  • [2] T. Arbogast, G. Pencheva, M. F. Wheeler, and I. Yotov. A multiscale mortar mixed finite element method. Multiscale Model. Simul., 6(1):319–346, 2007.
  • [3] I. Babuška and J. M. Melenk. The partition of unity method. Internat. J. Numer. Methods Engrg., 40(4):727–758, 1997.
  • [4] John R. Cannon, Richard E. Ewing, Yinnian He, and Yanping Lin. A modified nonlinear Galerkin method for the viscoelastic fluid motion equations. Internat. J. Engrg. Sci., 37(13):1643–1662, 1999.
  • [5] Jie Chen, Eric T Chung, Zhengkang He, and Shuyu Sun. Generalized multiscale approximation of mixed finite elements with velocity elimination for subsurface flow. Journal of Computational Physics, 404:109133, 2020.
  • [6] Zhiming Chen and Thomas Y. Hou. A mixed multiscale finite element method for elliptic problems with oscillating coefficients. Math. Comp., 72(242):541–576, 2003.
  • [7] M. A. Christie and M. J. Blunt. Tenth SPE Comparative Solution Project: A Comparison of Upscaling Techniques. SPE Reservoir Evaluation & Engineering, 4(04):308–317, August 2001. _eprint: https://onepetro.org/REE/article-pdf/4/04/308/2586053/spe-72469-pa.pdf.
  • [8] Eric T. Chung, Yalchin Efendiev, and Shubin Fu. Generalized multiscale finite element method for elasticity equations. GEM Int. J. Geomath., 5(2):225–254, 2014.
  • [9] Eric T. Chung, Yalchin Efendiev, and Wing Tat Leung. Constraint energy minimizing generalized multiscale finite element method. Comput. Methods Appl. Mech. Engrg., 339:298–319, 2018.
  • [10] Eric T. Chung, Yalchin Efendiev, and Wing Tat Leung. Fast online generalized multiscale finite element method using constraint energy minimization. J. Comput. Phys., 355:450–463, 2018.
  • [11] Eric T. Chung, Yalchin Efendiev, and Guanglian Li. An adaptive GMsFEM for high-contrast flow problems. J. Comput. Phys., 273:54–76, 2014.
  • [12] Yalchin Efendiev, Juan Galvis, and Thomas Y. Hou. Generalized multiscale finite element methods (GMsFEM). J. Comput. Phys., 251:116–135, 2013.
  • [13] Yalchin Efendiev, Juan Galvis, Guanglian Li, and Michael Presho. Generalized multiscale finite element methods: Oversampling strategies. International Journal for Multiscale Computational Engineering, 12(6), 2014.
  • [14] Yalchin Efendiev, Seong Lee, Guanglian Li, Jun Yao, and Na Zhang. Hierarchical multiscale modeling for flows in fractured media using generalized multiscale finite element method. GEM Int. J. Geomath., 6(2):141–162, 2015.
  • [15] Shubin Fu, Eric Chung, and Lina Zhao. Generalized Multiscale Finite Element Method for Highly Heterogeneous Compressible Flow. Multiscale Model. Simul., 20(4):1437–1467, 2022.
  • [16] Hadi Hajibeygi and Patrick Jenny. Multiscale finite-volume method for parabolic problems arising from compressible multiphase flow in porous media. J. Comput. Phys., 228(14):5129–5147, 2009.
  • [17] Thomas Y. Hou and Xiao-Hui Wu. A multiscale finite element method for elliptic problems in composite materials and porous media. J. Comput. Phys., 134(1):169–189, 1997.
  • [18] Thomas J. R. Hughes, Gonzalo R. Feijóo, Luca Mazzei, and Jean-Baptiste Quincy. The variational multiscale method—a paradigm for computational mechanics. Comput. Methods Appl. Mech. Engrg., 166(1-2):3–24, 1998.
  • [19] Patrick Jenny, SH Lee, and Hamdi A Tchelepi. Multi-scale finite-volume method for elliptic problems in subsurface flow simulation. Journal of computational physics, 187(1):47–67, 2003.
  • [20] Patrick Jenny, SH Lee, and Hamdi A Tchelepi. Adaptive multiscale finite-volume method for multiphase flow and transport in porous media. Multiscale Model. Simul., 3(1):50–64, 2004.
  • [21] Mi-Young Kim, Eun-Jae Park, Sunil G. Thomas, and Mary F. Wheeler. A multiscale mortar mixed finite element method for slightly compressible flows in porous media. J. Korean Math. Soc., 44(5):1103–1119, 2007.
  • [22] Mengnan Li, Eric Chung, and Lijian Jiang. A constraint energy minimizing generalized multiscale finite element method for parabolic equations. Multiscale Model. Simul., 17(3):996–1018, 2019.
  • [23] Axel Må lqvist and Daniel Peterseim. Localization of elliptic multiscale problems. Math. Comp., 83(290):2583–2603, 2014.
  • [24] Ana M. Matache and Christoph Schwab. Homogenization via pp-FEM for problems with microstructure. In Proceedings of the Fourth International Conference on Spectral and High Order Methods (ICOSAHOM 1998) (Herzliya), volume 33, pages 43–59, 2000.
  • [25] Eun-Jae Park. Mixed finite element methods for generalized Forchheimer flow in porous media. Numer. Methods Partial Differential Equations, 21(2):213–228, 2005.
  • [26] Maria Vasilyeva, Eric T Chung, Siu Wun Cheung, Yating Wang, and Georgy Prokopev. Nonlocal multicontinua upscaling for multicontinua flow problems in fractured porous media. Journal of Computational and Applied Mathematics, 355:258–267, 2019.
  • [27] Yixuan Wang, Hadi Hajibeygi, and Hamdi A Tchelepi. Algebraic multiscale solver for flow in heterogeneous porous media. Journal of Computational Physics, 259:284–303, 2014.