
Citation: Paola F. Antonietti, Chiara Facciolà, Marco Verani. Unified analysis of discontinuous Galerkin approximations of flows in fractured porous media on polygonal and polyhedral grids[J]. Mathematics in Engineering, 2020, 2(2): 340-385. doi: 10.3934/mine.2020017
[1] | Luca Formaggia, Alessio Fumagalli, Anna Scotti . A multi-layer reactive transport model for fractured porous media. Mathematics in Engineering, 2022, 4(1): 1-32. doi: 10.3934/mine.2022008 |
[2] | Thomas J. Radley, Paul Houston, Matthew E. Hubbard . Quadrature-free polytopic discontinuous Galerkin methods for transport problems. Mathematics in Engineering, 2024, 6(1): 192-220. doi: 10.3934/mine.2024009 |
[3] | Raimondo Penta, Ariel Ramírez-Torres, José Merodio, Reinaldo Rodríguez-Ramos . Effective governing equations for heterogenous porous media subject to inhomogeneous body forces. Mathematics in Engineering, 2021, 3(4): 1-17. doi: 10.3934/mine.2021033 |
[4] | Caterina Ida Zeppieri . Homogenisation of high-contrast brittle materials. Mathematics in Engineering, 2020, 2(1): 174-202. doi: 10.3934/mine.2020009 |
[5] | Chiara Gavioli, Pavel Krejčí . Deformable porous media with degenerate hysteresis in gravity field. Mathematics in Engineering, 2025, 7(1): 35-60. doi: 10.3934/mine.2025003 |
[6] | Federica Botta, Matteo Calafà, Pasquale C. Africa, Christian Vergara, Paola F. Antonietti . High-order discontinuous Galerkin methods for the monodomain and bidomain models. Mathematics in Engineering, 2024, 6(6): 726-741. doi: 10.3934/mine.2024028 |
[7] | Paola F. Antonietti, Ilario Mazzieri, Laura Melas, Roberto Paolucci, Alfio Quarteroni, Chiara Smerzini, Marco Stupazzini . Three-dimensional physics-based earthquake ground motion simulations for seismic risk assessment in densely populated urban areas. Mathematics in Engineering, 2021, 3(2): 1-31. doi: 10.3934/mine.2021012 |
[8] | Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082 |
[9] | M. M. Bhatti, Efstathios E. Michaelides . Oldroyd 6-constant Electro-magneto-hydrodynamic fluid flow through parallel micro-plates with heat transfer using Darcy-Brinkman-Forchheimer model: A parametric investigation. Mathematics in Engineering, 2023, 5(3): 1-19. doi: 10.3934/mine.2023051 |
[10] | Lina Zhao, Eun-Jae Park . A locally conservative staggered least squares method on polygonal meshes. Mathematics in Engineering, 2024, 6(2): 339-362. doi: 10.3934/mine.2024014 |
Many geophysical and engineering applications, including, for example, fluid-structure interaction, crack and wave propagation problems, and flow in fractured porous media, are characterized by a strong complexity of the physical domain, possibly involving thousands of fault/fractures, heterogeneous media, moving geometries/interfaces and complex topographies. Whenever classical Finite-Element-based approaches are employed to discretize the underlying differential model, the process of mesh generation can be the bottleneck of the whole simulation, as classical finite elements only support computational grids composed by tetrahedral/hexahedral/prismatic elements. To overcome this limitation, in the last decade a wide strand of literature focused on the design of numerical methods that support computational meshes composed of general polygonal and polyhedral (polytopic, for short) elements. In the conforming setting, we mention for example the composite finite element method [1,2], the mimetic finite difference method [3,4,5,6,7,8], the polygonal finite element method [9], the extended finite element method [10,11,12], the virtual element method [13,14,15,16,17] and the hybrid high-order method [18,19,20,21]. In the setting of non-conforming/discontinuos polygonal methods, we mention, for example, hybridizable discontinuous Galerkin methods [22,23,24,25], composite discontinuous finite element methods [26,27], non-conforming VEM [28,29,30], gradient schemes [31] and the polytopic discontinuous Galerkin method [32,33,34,35,36,37,38,39,40,41,42,43,44].
Within this framework, we focus our attention on the problem of modelling the flow in a fractured porous medium, which is fundamental in many energy or environmental engineering applications, such as tracing oil migration, isolation of radioactive waste, groundwater contamination, etc. Fractures are regions of the porous medium that are typically characterized both by a different porous structure and by a very small width. The first feature implies that fractures have a very strong impact on the flow, since they can possibly act as barriers or as preferential paths for the fluid. The second feature entails the need for a very large number of elements for the discretization of the fracture layer and, consequently, a high computational cost. For this reason, one popular modelling choice consists in a reduction strategy, so that fractures are treated as (d−1)-dimensional interfaces between d-dimensional porous matrices, d=2,3. In particular, we refer to the model for single-phase flow developed in [45,46,47,48]. Here, the flow in the porous medium (bulk) is assumed to be governed by Darcy's law and a suitable reduced version of the law is formulated on the surface modelling the single fracture. The two problems are then coupled through physically consistent conditions to account for the exchange of fluid between them. We remark that this model is able to handle both fractures with low and large permeability. Moreover, its extension to the case of two-phase flows has been addressed in [49,50], while the case of a totally immersed fracture has been considered e.g., in [51].
Even if the use of this kind of dimensionally reduced models avoids the need for extremely refined grids inside the fracture domains, in realistic cases, the construction of a computational grid aligned with the fractures is still a major issue. For example, a fractured oil reservoir can be cut by several thousands of fractures, which often intersect, create small angles or are nearly coincident [52]. In line with the previous discussion, in order to avoid the limitations imposed by standard finite element methods, various numerical methods supporting polytopic elements have been employed in the literature for the approximation of the coupled bulk-fracture problem. For example, in [8,52] a mixed approximation based on the use of mimetic finite difference method has been explored; in [53,54] a framework for treating flows in discrete fracture networks based on the virtual element method has been introduced, and in [55] the hybrid high-order method has been employed. We also mention that an alternative strategy consists in the use of non-conforming discretizations, where fractures are allowed to arbitrarily cut the bulk grid, which can then be chosen fairly regular. In particular, we refer to [49,56,57], where an approximation employing extended finite element method (XFEM) has been proposed and to the recent work [58], where the use of the cut finite element method (CUTFEM) has been explored.
Recently in [59], in the setting of conforming discretizations, we developed a numerical approximation of the coupled bulk-fracture model based on polytopic discountinuous Galerkin (PolyDG) methods. In particular, the intrinsic "discontinuous" nature of DG methods allows very general polytopic elements because of the freedom in representing the underlying (local) polynomial space. Indeed the degrees of freedom are not "attached" to any geometric quantity, (vertexes, edges, etcc), so that mesh elements with edges/faces that may be in arbitrary number and whose measure may be arbitrarily small compared to the diameter of the corresponding element are naturally supported with a solid theoretical background. This approach is then very well suited to tame the geometrical complexity featured by most of applications in the computational geoscience field. Moreover, since the interface conditions between the bulk and fracture problem can be naturally formulated using jump and average operators, the coupling of the two problems can be naturally embedded in the variational formulation.
The goal of this paper is to extend the results obtained in [59], where the proposed discretization based on PolyDG was in a primal-primal setting. Indeed, when dealing with the approximation of Darcy's flow there are two possible approaches: primal and mixed. The primal approach considers a single-field formulation with the pressure field of the fluid as only unknown. The mixed approach describes the flow not only through the pressure field, but also through an additional unknown representing Darcy's velocity, which is of primary interest in many engineering applications. The primal setting has the advantage of featuring less degrees of freedom and leads to a symmetric positive definite algebraic system of equations that can be efficiently solved based on employing, for example, multigrid techniques [39,44,60]. In this case, Darcy's velocity is afterwards reconstructed taking the gradient of the computed pressure and multiplying it by the permeability tensor. However, this process usually entails a loss of accuracy and does not guarantee mass conservation, see for example [61,62]. For this reason, the mixed setting is sometimes preferred. In this case Darcy's velocity is directly computed, so that a higher degree of accuracy is achieved, together with local and global mass conservation. However, the drawback of this approach is the complexity of the resulting scheme, which leads to a generalized saddle point algebraic system of equations. From the above discussion we may infer that, according to the desired approximation properties of the model, one may resort to either a primal or mixed approximation for the problem in the bulk, as well as to a primal or mixed approximation for the problem in the fracture. Our aim is then to design and analyze, in the unified framework of [63] based on the flux-formulation, all the possible combinations of primal-primal, mixed-primal, primal-mixed and mixed-mixed formulations for the bulk and fracture problems, respectively. In particular, the primal discretizations are obtained using the symmetric interior penalty discontinuous Galerkin method [64,65], whereas the mixed discretizations are based on employing the local DG (LDG) method of [66], both in their generalization to polytopic grids [36,37,38,40,41]. Moreover, the coupling conditions between bulk and fracture are imposed through a suitable definition of the numerical fluxes on the fracture faces. Such an abstract setting allows to analyse theoretically at the same time all the possible formulations. We perform a unified analysis of all the derived combinations of DG discretizations for the bulk-fracture problem. We prove their well-posedness and derive a priori hp-version error estimates in a suitable (mesh-dependent) energy norm. Finally, we present preliminary numerical experiments assessing the validity of the theoretical error estimates and comparing the accuracy of the proposed formulations on simplified geometries with manufactured solutions.
The rest of the paper is organized as follows. In Section 2 we introduce the model problem; its weak formulation is discussed in Section 3. The discretization based on employing PolyDG methods is presented, in the unified setting of [63], in Section 4. In Section 5, we address the problem of stability and prove that all formulations, namely primal-primal (PP), mixed-primal (MP), primal-mixed (PM) and mixed-mixed (MM) are well-posed. The corresponding error analysis is presented in Section 6. Several numerical tests, focusing, for the sake of brevity, on the primal-primal (PP) and mixed-primal (MP) cases, are presented in Section 7 to confirm the theoretical bounds. Moreover, we assess the capability of the method of handling more complicated geometries, presenting some test cases featuring networks of partially immersed fractures.
For simplicity, we consider the case where the porous medium is cut by a single, non immersed fracture. The extension to the case of a network of disjoint fractures can be treated analogously. The case where the fracture is partially or totally immersed in the domain is more complex to analyze, and we refer to [51,52] for its discussion. Nevertheless, the capability of our method to deal with networks of partially immersed fractures will be explored via numerical experiments in Section 7.4 in the mixed-primal setting (MP). Finally, the case of a network of interecting fractures will be the object of a future work and we refer to [59] for preliminary numerical results (in the primal-primal setting) showing that our method is able to handle also such cases.
The porous matrix is represented by the domain Ω⊂Rd, d=2,3, which we assume to be open, bounded, convex and polygonal/polyhedral. Moreover, following the strategy of [47], we suppose that the fracture may be described by the (d−1)-dimensional C∞ manifold (with no curvature) Γ⊂Rd−1, d=2,3. This approach is justified by the fact that the thickness of the fracture domain is typically some orders of magnitude smaller than the size of the domain. Since we are assuming that Γ is not immersed, it separates Ω into two connected disjoint subdomains, Ω∖Γ=Ω1∪Ω2 with Ω1∩Ω2=∅. We decompose the boundary of Ω into two disjoint subsets ∂ΩD and ∂ΩN, i.e., ∂Ω=∂ΩD∪∂ΩN, with ∂ΩD∩∂ΩN=∅, and we define ∂ΩD,i=∂ΩD∩∂Ωi and ∂ΩN,i=∂ΩN∩∂Ωi, for i=1,2. Finally, we denote by ni, i=1,2 the unit normal vector to Γ pointing outwards from Ωi and, for a (regular enough) scalar-valued function q and a (regular enough) vector-valued function v, we define the classical jump and average operators across the fracture Γ as
{q}=12(q1+q2)[[q]]=q1n1+q2n2,{v}=12(v1+v2)[[v]]=v1⋅n1+v2⋅n2, | (2.1) |
where the subscript i=1,2 denotes the restriction to the subdomain Ωi. Note that, since we are assuming that the fracture is continuously differentiable, it holds n1=−n2. We also denote by nΓ the normal unit vector on Γ with a fixed orientation from Ω1 to Ω2, so that we have nΓ=n1=−n2. In Figure 1 we report an example of domain cut by a single fracture.
We can now introduce the governing equations for our model. In the bulk, we suppose that the flow is governed by Darcy's law. The motion of an incompressible fluid in each domain Ωi, i=1,2, with pressure pi and velocity ui may then be described by:
{ui=νi∇piinΩi,−∇⋅ui=fiinΩi,pi=gD,ion∂ΩD,i,ui⋅ni=0on∂ΩN,i, | (2.2) |
where f∈L2(Ω) represents a source term, gD∈H1/2(∂ΩD) is the Dirichlet boundary datum and ν=ν(x)∈Rd×d is the bulk permeability tensor, which we assume to be symmetric, positive definite, uniformly bounded from below and above and with entries that are bounded, piecewise continuous real-valued functions.
On the manifold Γ representing the fracture, we formulate a reduced version of Darcy's law in the tangential direction (we refer to [47] for a rigorous derivation of the model). To this aim we assume that the fracture permeability tensor νΓ, has a block-diagonal structure of the form
νΓ=[νnΓ00ντΓ], | (2.3) |
when written in its normal and tangential components. Here, ντΓ∈R(d−1)×(d−1) is a positive definite, uniformly bounded tensor (it reduces to a positive number for d=2). Moreover, we assume that νΓ satisfies the same regularity assumptions as those satisfied by the bulk permeability ν. Setting ∂Γ=Γ∩∂Ω, ∂Γ=∂ΓD∪∂ΓN, introducing the fracture thickness ℓΓ>0 and denoting by pΓ and uΓ the fracture pressure and velocity, the governing equations for the fracture flow are
{uΓ=ντΓℓΓ∇τpΓinΓ,−∇τ⋅uΓ=fΓ−[[u]]inΓ,pΓ=gΓon∂ΓD,uΓ⋅τ=0on∂ΓN, | (2.4) |
where fΓ∈L2(Γ), gΓ∈H1/2(∂Γ), τ is vector in the tangent plane of Γ normal to ∂Γ and ∇τ and ∇τ⋅ denote the tangential gradient and divergence operators, respectively.
Finally, following [47], we close the model providing the interface conditions to couple problems (2.2) and (2.4) along their interface. Given a positive real number ξ≠12 that will be chosen later on, the coupling conditions read as follows
−{u}⋅nΓ=βΓ[[p]]⋅nΓonΓ, | (2.5a) |
−[[u]]=αΓ({p}−pΓ)onΓ, | (2.5b) |
where βΓ=12ηΓ and αΓ=2ηΓ(2ξ−1) and ηΓ=ℓΓνnΓ, νnΓ being the normal component of the fracture permeability tensor, see (2.3).
In conclusion, the coupled bulk-fracture model problem is the following:
{ui=νi∇piinΩi,−∇⋅ui=fiinΩi,pi=gD,ionγD,i,ui⋅ni=0onγN,iuΓ=ντΓℓΓ∇τpΓinΓ,−∇τ⋅uΓ=fΓ−[[u]]inΓ,pΓ=gΓon∂ΓD,uΓ⋅τ=0on∂ΓN,−{u}⋅nΓ=βΓ[[p]]⋅nΓonΓ,−[[u]]=αΓ({p}−pΓ)onΓ. | (2.6) |
In this section we introduce the weak formulation of our model problem (2.6) and prove its well-posedness. We start with the introduction of the functional setting.
We will employ the following notation. For an open, bounded domain D⊂Rd, d=2,3, we denote by Hs(D) the standard Sobolev space of order s, for a real number s≥0. For s=0, we write L2(D) in place of H0(D). The usual norm on Hs(D) is denoted by ||⋅||s,D and the usual seminorm by |⋅|s,D. We also introduce the standard space Hdiv(D)={v:D→Rd:||v||0,D+||∇⋅v||0,D<∞}. Given a decomposition of the domain into elements Th, we will denote by Hs(Th) the standard broken Sobolev space, equipped with the broken norm ||⋅||s,Th. Furthermore, we will denote by Pk(D) the space of polynomials of total degree less than or equal to k≥1 on D. The symbol ≲ (and ≳) will signify that the inequalities hold up to multiplicative constants which are independent of the discretization parameters, but might depend on the physical parameters.
Next, we introduce the functional spaces for our weak formulation. For the bulk pressure and velocity, we introduce the spaces Mb=L2(Ω) and Vb={v∈Hdiv(Ω):[[v]]|Γ∈L2(Γ),{v}|Γ∈[L2(Γ)]d,v⋅n|∂ΩN=0}, and equip the space Vb with the norm ||v||2Vb=||v||20,Ω+||∇⋅v||20,Ω+||[[v]]||20,Γ+||{v}||20,Γ.
Similarly, for the fracture pressure and velocity we define the spaces MΓ=L2(Γ) and VΓ={vΓ∈Hdiv,τ(Γ):vΓ⋅τ|∂Γ=0}. The norm on VΓ is given by ||vΓ||2VΓ=||vΓ||20,Γ+||∇τ⋅vΓ||20,Γ. Finally, we define the global spaces for the pressure and the velocity as M=Mb×MΓ and W=Vb×VΓ, equipped with the canonical norms for product spaces. In order to deal with non-homogeneous boundary conditions, we also introduce the affine spaces Vbg=Lg+Vb and VΓg=LgΓ+VΓ, where Lg∈Hdiv(Ω) and LgΓ∈Hdiv,τ(Γ) are liftings of the boundary data g and gΓ, respectively. We can then define the global space Wg=Vbg×VΓg.
We can now formulate problem (2.6) in weak form as follows: Find (u,uΓ)∈Wg and (p,pΓ)∈M such that
{A((u,uΓ),(v,vΓ))+B((v,vΓ),(p,pΓ))=Fu(v,vΓ)−B((u,uΓ),(q,qΓ))=Fp(q,qΓ) | (3.1) |
where the bilinear form A(⋅,⋅):Wg×Wg→R is defined as A((u,uΓ),(v,vΓ))=a(u,v)+aΓ(uΓ,vΓ) with
a(u,v)=∫Ων−1u⋅v+∫Γ1αΓ[[u]][[v]]+∫Γ1βΓ{u}⋅{v},aΓ(uΓ,vΓ)=∫Γ(ντΓℓΓ)−1uΓ⋅vΓ, |
and the bilinear form B(⋅,⋅):Wg×M→R is defined as B((v,vΓ),(q,qΓ))=b(v,q)+bΓ(vΓ,qΓ)+d(v,qΓ), with
b(v,q)=∫Ω∇⋅vq,bΓ(vΓ,qΓ)=∫Γ∇τ⋅vΓqΓ,d(v,qΓ)=−∫Γ[[v]]qΓ. |
Finally the linear operators Fu(⋅):Wg→R and Fp(⋅):M→R are defined as
Fu(v,vΓ)=∫∂Ωgv⋅n+∫∂ΓgΓvΓ⋅τ,Fp(q,qΓ)=∫Ωfq+∫ΓfΓqΓ. |
Next, we prove that formulation (3.1) is well-posed. For the sake of simplicity, we will assume that homogeneous Dirichlet boundary conditions are imposed for both the bulk and fracture problems, i.e., gD,i=0, i=1;2, and gΓ=0 and that the domain and fracture are smooth enough. The extension to the general non-homogeneous case is straightforward. Note that the existence and uniqueness of the problem can be proven only under the condition that the parameter ξ>1/2.
Theorem 3.1. Suppose that ξ>1/2. Then problem (3.1) admits a unique solution.
Proof. For the proof we follow the technique of [47]. First, we define the subspace of W, ˆW={(v,vΓ)∈W:B((v,vΓ),(q,qΓ))=0∀(q,qΓ)∈M}. To show existence and uniqueness of the solution of (3.1), we only need to show that A(⋅,⋅) is ˆW-elliptic and that B(⋅,⋅) satisfies the inf-sup condition, that is
inf(v,vΓ)∈ˆWA((v,vΓ),(v,vΓ))||(v,vΓ)||2W≳1,inf(q,qΓ)∈Msup(v,vΓ)∈WB((v,vΓ),(q,qΓ))||(q,qΓ)||M||(v,vΓ)||W≳1. |
First, we prove that A(⋅,⋅) is ˆW-elliptic. Since for elements in (v,vΓ)∈ˆW we have ∇⋅v=0 in L2(Ω) and ∇τ⋅vΓ=[[v]]|Γ in L2(Γ), the norm ||(v,vΓ)||W is equivalent to ||v||20,Ω+||vΓ||20,Γ+||[[v]]||20,Γ+||{v}||20,Γ. Owing to the regularity properties of the permeability tensors ν and νΓ, this implies that
A((v,vΓ),(v,vΓ))≳||(v,vΓ)||2W. |
Note that the hidden constant also depends on the parameter αΓ, and that we need to assume αΓ>0, or, equivalently, ξ>1/2, for the inequality to hold true.
To show that B satisfies the inf-sup condition, given (q,qΓ)∈M, we construct, exploiting the adjoint problem, (v,vΓ)∈W such that B((v,vΓ),(q,qΓ))=||(q,qΓ)||2M and ||(v,vΓ)||W≲||(q,qΓ)||M. Given (q,qΓ)∈M, let (ϕ,ϕΓ) be the solution of
{−Δϕ=q,on Ωϕ=0,on ∂Ω and {−ΔτϕΓ=qΓ,on ΓϕΓ=0,on ∂Γ. |
If we set v=(v1,v2) with vi=−∇ϕ|Ωi, i=1,2, and vΓ=−∇τϕΓ, we obtain ∇⋅v=q, ∇τ⋅vΓ=qΓ and [[v]]|Γ=0, since v∈H1(Ω). This implies that (v,vΓ)∈W and B((v,vΓ),(q,qΓ))=||q||20,Ω+||qΓ||20,Γ=||(q,qΓ)||2M. Finally, from elliptic regularity, we have ||(v,vΓ)||2W=||∇ϕ||20,Ω+||∇τϕΓ||20,Γ+||q||20,Ω+||qΓ||20,Γ+||{∇ϕ}||20,Γ≲||q||20,Ω+||qΓ||20,Γ, and this concludes the proof.
In this section we present a family of discrete formulations for the coupled bulk-fracture problem (3.1), which are based on discontinuous Galerkin methods on polytopic grids. In particular, since we can choose to discretize the problem in the bulk and the one in the fracture either in their mixed or in their primal form, we derive four formulations that embrace all the possible combinations of primal-primal, mixed-primal, primal-mixed and mixed-mixed discretizations. The mixed discretizations will be based on the local discontinuous Galerkin method (LDG) [66,67,68], while the primal discretizations on the Symmetric Interior Penalty method (SIPDG) [64,65], all supporting polytopic grids [36,37,40,41]. The derivation of our discrete formulations will be carried out following the same strategy as in [63], so that it will be based on the introduction of the numerical fluxes, which approximate the trace of the solutions on the boundary of each mesh element. In particular, the imposition of the coupling conditions (2.5a)-(2.5b) will be achieved through a proper definition of the numerical fluxes on the faces belonging to the fracture.
First, we introduce the notation related to the discretization of the domains by means of polytopic meshes. For the problem in the bulk, we consider a family of meshes Th made of disjoint open polygonal/polyhedral elements. Following [36,37,40], we introduce the concept of mesh interface, defined as the intersection of the (d−1)-dimensional facets of two neighbouring elements. We need now to distinguish between the case when d=3 and d=2:
● when d=3, each interface consists of a general polygon, which we assume may be decomposed into a set of co-planar triangles. We assume that a sub-triangulation of each interface is provided and we denote the set of all these triangles by Fh. We then use the terminology face to refer to one of the triangular elements in Fh;
● when d=2, each interface simply consists of a line segment, so that the concepts of face and interface are in this case coincident. We still denote by Fh the set of all faces.
Note that Fh is always defined as a set of (d−1)-dimensional simplices (triangles or line segments). As in [36,37,40], no limitation on either the number of faces of each polygon E∈Th or on the relative size of the faces compared to the diameter of the element is imposed.
We consider meshes Th that are aligned with the fracture Γ, so that any element E∈Th cannot be cut by Γ and it belongs exactly to one of the two disjoint subdomains Ω1 or Ω2. This implies that each mesh Th induces a subdivision of the fracture Γ into faces, which we denote by Γh. It follows that we can write
Fh=FIh∪FBh∪Γh, |
where FBh is the set of faces lying on the boundary of the domain ∂Ω and FIh is the set of interior faces not belonging to the fracture. In addition, we write FBh=FDh∪FNh, where FDh and FNh are the boundary faces contained in ∂ΩD and ∂ΩN, respectively (we assume the decomposition to be matching with the partition of ∂Ω into ∂ΩD and ∂ΩN).
The induced discretization of the fracture Γh consists of the faces of the elements of Th that share part of their boundary with the fracture, so that Γh is made up of line segments when d=2 and of triangles when d=3. Note that, in the 3D case, the triangles are not necessarily shape-regular and they may present hanging nodes, due to the fact that the sub-triangulations of each elemental interface is chosen independently from the others. For this reason, we extend the concept of interface also to the (d−2)-dimensional facets of elements in Γh, defined again as intersection of boundaries of two neighbouring elements. When d=2, the interfaces reduce to points, while when d=3 they consists of line segments. We denote by EΓ,h the set of all the interfaces (that we will also call edges) of the elements in Γh, and we write, accordingly to the previous notation, EΓ,h=EIΓ,h∪EBΓ,h, with EBΓ,h=EDΓ,h∪ENΓ,h.
For each element E∈Th, we denote by |E| its measure, by hE its diameter and we set h=maxE∈ThhE. Given an element E∈Th, for any face F⊂∂E, with F∈Fh, we define nF as the unit normal vector on F that points outward of E. We can then define the standard jump and average operators across a face F∈Fh∖FBh for (regular enough) scalar and vector-valued functions similarly to (2.1). We also recall a well-known identity [65] for scalar and vector-valued functions q and v that are piecewise smooth on Th:
∑E∈Th∫∂Eqv⋅nE=∫Fh{v}⋅[[q]]+∫Fh∖FBh[[v]]{q}, | (4.1) |
where we have used the compact notation ∫Fh⋅=∑F∈Fh∫F⋅ and jump and average operators on a boundary face F∈FBh are defined as [[q]]=qnF and {v}=v.
Analogous definitions may be also set up on the fracture. In particular, given an element F∈Γh, with measure |F| and diameter hF, for any edge e⊂∂F, with e∈EΓ,h, we define ne as the unit normal vector on e pointing outward of F (it reduces to ±1 when d=2). Finally, standard jump and average operators across every edge e can be defined for (regular enough) scalar and vector-valued functions and an analogous version of formula (4.1) can be stated for piecewise smooth function on the fracture mesh Γh.
We have now all the ingredients to introduce the discrete formulation of model problem (3.1).
For simplicity in the forthcoming analysis, we will suppose that the permeability tensors ν and νΓ are piecewise constant on mesh elements, i.e., ν|E∈[P0(E)]d×d for all E∈Th, and νΓ|F∈[P0(F)](d−1)×(d−1) for all F∈Γh. First, we introduce the finite-dimensional spaces where we will set our discrete problem. We set
Qbh={q∈L2(Ω):q|E∈PkE(E)∀E∈Th}Wbh={v∈[L2(Ω)]d:v|E∈[PkE(E)]d∀E∈Th}QΓh={qΓ∈L2(Γ):qΓ|F∈PkF(F)∀F∈Γh}WΓh={vΓ∈[L2(Γ)]d−1:vΓ|F∈[PkF(F)]d−1∀F∈Γh}. |
Note that, to each element E∈Th is associated the polynomial degree kE≥1, as well as to each face F∈Γh is associated the degree kF≥1. We remark that the polynomial degrees in the bulk and fracture discrete spaces just defined are chosen independently of each other.
We first focus on the problem in the bulk. Multiplying the first and second equations in (2.2) by test functions v∈Wbh and q∈Qbh, respectively, and integrating by parts over an element E∈Th, we obtain
∫Eν−1u⋅v=−∫Ep∇⋅v+∫∂Epv⋅nE,∫Eu⋅∇q=∫∂Equ⋅nE+∫Efq. |
In the spirit of [63], we start the derivation of our DG discretization from these equations. Adding over the elements E∈Th, the general discrete formulation for the problem in the bulk will then be: Find ph∈Qbh and uh∈Wbh, such that for all E∈Th we have
∑E∈Th∫Eν−1uh⋅v=−∑E∈Th∫Eph∇⋅v+∑E∈Th∫∂EˆpEv⋅nE∑E∈Th∫Euh⋅∇q=∑E∈Th∫∂EqˆuE⋅nE+∑E∈Th∫Efq, |
where the numerical fluxes ˆpE and ˆuE are approximations to the exact solutions u and p, respectively, on the boundary of E. The definition of the numerical fluxes in terms of ph, uh, of the boundary data and of the coupling conditions (2.5a)-(2.5b) will determine the method. Using identity (4.1), we get
∫Thν−1uh⋅v=−∫Thph∇⋅v+∫FIh∪Γh{ˆp}[[v]]+∫FIh∪FBh∪Γh[[ˆp]]⋅{v}, | (4.2) |
∫Thuh⋅∇q−∫FIh∪FBh∪Γh{ˆu}⋅[[q]]−∫FIh∪Γh[[ˆu]]{q}=∫Thfq, | (4.3) |
where we have introduced ˆp=(ˆpE)E∈Th and ˆu=(ˆuE)E∈Th. The numerical fluxes ˆp and ˆu must be interpreted as linear functionals taking values in the spaces ΠE∈ThL2(∂E) and [ΠE∈ThL2(∂E)]d, respectively. In particular, this means that they are, in general, double-valued on FIh∪Γh and single-valued on FBh. We also observe for future use that, after integrating by parts and using again identity (4.1), Eq. (4.2) may also be rewritten as
∫Thν−1uh⋅v=∫Th∇ph⋅v+∫FIh∪Γh{ˆp−ph}[[v]]+∫FIh∪FBh∪Γh[[ˆp−ph]]⋅{v}. | (4.4) |
We now reason analogously for the fracture. Multiplying the first and second equations in (2.4) by test functions vΓ and qΓ, respectively, integrating by parts over an element F∈Γh and summing over all the elements in Γh we obtain the following problem: Find pΓ,h∈QΓh and uΓ,h∈WΓh such that for all F∈Γh we have
∑F∈Γh∫F(ντΓℓΓ)−1uΓ,h⋅vΓ=−∑F∈Γh∫FpΓ,h∇⋅vΓ+∑F∈Γh∫∂FˆpΓ,Fv⋅nF,∑F∈Γh∫FuΓ,h⋅∇qΓ−∑F∈Γh∫∂FqΓˆuΓ,F⋅nF=∑F∈Γh∫FfΓqΓ−∑F∈Γh∫F[[ˆu]]qΓ. |
Here, we have introduced the numerical fluxes ˆpΓ,F and ˆuΓ,F. Again, the idea is that they represent approximations on the boundary of the fracture face F of the exact solutions pΓ and uΓ, respectively. Note also that here ˆu is the numerical flux approximating the bulk velocity on Γh. Using identity (4.1), we get
∫Γh(ντΓℓΓ)−1uΓ,h⋅vΓ=−∑F∈Γh∫FpΓ,h∇⋅vΓ+∫EIΓ,h{ˆpΓ}[[vΓ]]+∫EΓ,h{vΓ}⋅[[ˆpΓ]] | (4.5) |
∫ΓhuΓ,h⋅∇qΓ−∫EIΓ,h{qΓ}[[ˆuΓ]]−∫EΓ,h{ˆuΓ}⋅[[qΓ]]=∫ΓhfΓqΓ−∫Γh[[ˆu]]qΓ | (4.6) |
We point out that, in all previous equations, the gradient and divergence operators are actually tangent operators. Here, we have dropped the subscript τ in order to simplify the notation.
In the following, we explore all possible combinations of primal-primal, mixed-primal, primal mixed and mixed-mixed formulations for the bulk and fracture, respectively.
In order to obtain the primal-primal formulation, we need to eliminate the velocities uh and uΓ,h from equations (4.2)-(4.3) and (4.5)-(4.6). To do so, we need to express uh solely in terms of ph (and pΓ,h), and uΓ,h solely in terms of pΓ,h. As in [63] this will be achieved via the definition of proper lifting operators.
We start by focusing on the problem in the bulk. In order to complete the specification of the DG method that we want to use for the approximation, we need to give an expression to the numerical fluxes. We choose the classic symmetric interior penalty method (SIPDG). Moreover, coupling conditions (2.5a)-(2.5b) are imposed through a suitable definition of the numerical fluxes on the fracture faces. Since we want a primal formulation, the definition of ˆp and ˆu will not contain uh. The numerical fluxes are defined as follows:
ˆp=ˆp(ph)={{ph}onFIhgDonFDhphonFNhphonΓh |
ˆu=ˆu(ph,pΓ,h)={{ν∇ph}−σF[[ph]]onFIhν∇ph−σF(ph−gD)nFonFDh0onFNh−[αΓ({ph}−pΓ,h)nF2+βΓ[[ph]]]onΓh | (4.7) |
Here, we have introduced the discontinuity penalization parameter σ. In particular, σ is a non-negative bounded function, i.e., σ∈L∞(FIh∪FDh) and its precise definition will be given in Definition 5.2 below. Moreover, we have used the notation σF=σ|F, for F∈FIh∪FDh. We remark that, with this choice, the numerical flux ˆp is doubled valued on Γh and single valued on FIh∪FBh.
Using the definition of the numerical fluxes, it follows that
{ˆp−ph}=0,[[ˆu]]=0onFIh,{ˆp−ph}=0,[[ˆu]]=−αΓ({ph}−pΓ,h)onΓh,[[ˆp−ph]]=−[[ph]],{ˆu}={ν∇ph}−σF[[ph]]onFIh,[[ˆp−ph]]=(gD−ph)nF,{ˆu}=ν∇ph−σF(ph−gD)nFonFDh,[[ˆp−ph]]=0,{ˆu}=0onFNh,[[ˆp−ph]]=0,{ˆu}=−βΓ[[ph]]onΓh, |
so we rewrite (4.4) as
∫Thν−1uh⋅v=∫Th∇ph⋅v−∫FIh∪FDh[[ph]]⋅{v}+∫FDhgDv⋅n. | (4.8) |
At this point, we proceed with the elimination of the auxiliary variable uh from our equations. To this aim, we introduce the lifting operator LSIPb:[L1(FIh∪FDh)]d→Wbh defined by
∫ThLSIPb(ξ)⋅v=−∫FIh∪FDh{v}⋅ξ∀v∈Wbh. | (4.9) |
Similarly, we define the lifting Gb(gD)∈Wbh of the Dirichlet boundary datum gD as
∫ThGb⋅v=∫FDhgDv⋅n∀v∈Wbh. | (4.10) |
Thanks to the introduction of the lifting operators, Eq. (4.8) may be rewritten as
∫Th(uh−ν[∇ph+LSIPb([[ph]])+Gb])⋅v=0. |
Since ∇Qbh⊆Wbh, we can write
uh=ν[∇ph+LSIPb([[ph]])+Gb], | (4.11) |
where ∇ph+LSIPb([[ph]])+Gb can be seen as a discrete approximation of the gradient ∇p.
We can then rewrite Eq. (4.3) as
∫Thν[∇ph+LSIPb([[ph]])+Gb]⋅∇q−∫FIh∪FBh∪Γh{ˆu}⋅[[q]]−∫FIh∪Γh[[ˆu]]{q}=∫Thfq. |
Using the definition of the discrete gradient (4.11), of the lifting operators (4.9) and (4.10) and of the numerical flux ˆu (4.7), we have
∫Thν∇ph⋅∇q+∫ThνLSIPb([[ph]])⋅∇q+∫ThνLSIPb([[q]])⋅∇ph+∫FIh∪FDhσF[[ph]]⋅[[q]]+∫ΓhβΓ[[ph]]⋅[[q]]+∫ΓhαΓ({ph}−pΓ,h){q}=∫Thfq+∫FDhgDσFq−∫ThνGb⋅∇q. | (4.12) |
Now we move our attention to the problem in the fracture. We define the numerical fluxes ˆpΓ and ˆuΓ in order to obtain a symmetric interior penalty approximation as follows:
ˆpΓ=ˆpΓ(pΓ,h)={{pΓ,h}onEIΓ,hgΓonEDΓ,hpΓ,honENΓ,h, |
ˆuΓ=ˆuΓ(pΓ,h)={{ντΓℓΓ∇pΓ,h}−σe[[pΓ,h]]onEIΓ,hντΓℓΓ∇pΓ,h−σe(pΓ,h−gΓ)neonEDΓ,h0onENΓ,h. | (4.13) |
Again, we have introduced the discontinuity penalization parameter σΓ∈L∞(EIΓ,h∪EDΓ,h) and we set σe=σΓ|e for e∈EIΓ,h∪EDΓ,h. Its precise definition will be given in Definition 5.3 below. Next, as before, we introduce the lifting operator LSIPΓ:[L1(EIΓ,h∪EDΓ,h)]d−1→WΓh and the lifting of the boundary datum GΓ(gΓ,D)∈WΓh defined by
∫ΓhLSIPΓ(ξΓ)⋅vΓ=−∫EIΓ,h∪EDΓ,hξΓ⋅{vΓ}∀vΓ∈WΓh, | (4.14) |
∫ΓhGΓ⋅vΓ=∫EDΓ,hgΓ,DvΓ⋅nτ∀vΓ∈WΓh. | (4.15) |
Integrating by parts and using (4.1), we can rewrite Eq. (4.5) as
∫Γh(uΓ,h−ντΓℓΓ[∇pΓ,h+LSIPΓ([[pΓ,h]])+GΓ])⋅vΓ=0. |
Again, since ∇QΓh⊆WΓh elementwise, we can write
uΓ,h=ντΓℓΓ[∇pΓ,h+LSIPΓ([[pΓ,h]])+GΓ]. |
Plugging this last identity and the definition of the numerical fluxes ˆu (see (4.7)) and ˆuΓ (see (4.13)) into Eq. (4.6), we obtain
∫ΓhντΓℓΓ∇pΓ,h⋅∇qΓ+∫ΓhντΓℓΓLSIPΓ([[pΓ,h]])⋅∇qΓ+∫ΓhντΓℓΓLSIPΓ([[qΓ]])⋅∇pΓ,h+∫EIΓ,h∪EDΓ,hσe[[pΓ,h]]⋅[[qΓ]]+∫ΓhαΓpΓ,hqΓ−∫ΓhαΓ{ph}qΓ=∫ΓhfΓqΓ+∫EDΓ,hgΓσeqΓ−∫ΓhντΓℓΓGΓ⋅∇qΓ. | (4.16) |
In conclusion, summing Eqs. (4.12) and (4.16) we obtain the following discrete formulation: Find (ph,pΓh)∈Qbh×QΓh such that
APPh((ph,pΓh),(q,qΓ))=LPPh(q,qΓ)∀(q,qΓ)∈Qbh×QΓh, | (4.17) |
where PP stands for primal-primal and where Lh:Qbh×QΓh→R is defined as LPPh(q,qΓ)=LPb(q)+LPΓ(qΓ) and APPh:(Qbh×QΓh)×(Qbh×QΓh)→R is defined as
APPh((ph,pΓh),(q,qΓ))=APb(ph,q)+APΓ(pΓ,h,qΓ)+I((ph,pΓ,h),(q,qΓ)), |
with
APb(ph,q)=∫Thν∇ph⋅∇q+∫ThνLSIPb([[ph]])⋅∇q |
+∫ThνLSIPb([[q]])⋅∇ph+∫FIh∪FDhσF[[ph]]⋅[[q]], | (4.18) |
APΓ(pΓ,h,qΓ)=∫ΓhντΓℓΓ∇pΓ,h⋅∇qΓ+∫ΓhντΓℓΓLPΓ([[pΓ,h]])⋅∇qΓ |
+∫ΓhντΓℓΓLSIPΓ([[qΓ]])⋅∇pΓ,h+∫EIΓ,h∪EDΓ,hσe[[pΓ,h]]⋅[[qΓ]], | (4.19) |
I((ph,pΓ,h),(q,qΓ))=∫ΓhβΓ[[ph]]⋅[[q]]+∫ΓhαΓ({ph}−pΓ,h)({q}−qΓ,h), | (4.20) |
and
LPb(q)=∫Thfq+∫FDhgDσFq−∫ThνGb⋅∇q, | (4.21) |
LPΓ(qΓ)=∫ΓhfΓqΓ+∫EDΓ,hgΓσeqΓ−∫ΓhντΓℓΓGΓ⋅∇qΓ. | (4.22) |
We remark that we have recovered the formulation already obtained in [59] (in its not strongly consistent version), the only difference being that the bilinear form for the problem in the fracture is in the shape of SIPDG method, instead of classical conforming finite elements.
In this section, we discretize the problem in the bulk in its mixed form. To this aim, we use the local discontinuous Galerkin (LDG) method [66,67,68,69]. The LDG method is a particular DG method that can be included in the class of mixed finite element methods. However, the variable uh can be locally solved in terms of ph and then eliminated from the equations, giving rise to a primal formulation with ph as only unknown.
In what follows, we first derive the formulation of our method in a mixed setting. After that, we recast it in a primal setting, in order to perform the analysis in the framework of [63,69]. However, we remark that the mixed formulation is the one that will actually be implemented for the numerical experiments of Section 7. As far as the problem in the fracture is concerned, we work again in a primal setting, using the SIPDG method for the discretization.
In the bulk, we define the numerical fluxes as
ˆp=ˆp(ph)={{ph}+b⋅[[ph]]onFIhgDonFDhphonFNhphonΓh |
ˆu=ˆu(uh,ph,pΓ,h)={{uh}−b[[uh]]−σF[[ph]]onFIhuh−σF(phnF−gDnF)onFDh0onFNh−[αΓ({ph}−pΓ,h)nF2+βΓ[[ph]]]onΓh |
Here, b∈[L∞(FIh)]d is a (possibly null) vector-valued function which is constant on each face. It is chosen such that
||b||∞,FIh≤B, | (4.23) |
with B≥0 independent if the discretization parameters. Moreover, σ is the penalization parameter introduced in (4.7), whose precise definition will be given in (5.2) below. Note that the numerical flux ˆp does not depend on uh. This will allow for an element-by-element elimination of the variable uh, generating a primal formulation of the problem. We also point out that the definition of the numerical fluxes on the fracture faces is the same as in the primal SIPDG setting.
With this definition of the numerical fluxes, and after integration by parts as in (4.4), Eq. (4.2) becomes
∫Thν−1uh⋅v−∫Th∇ph⋅v+∫FIh[[ph]]⋅({v}−b[[v]])+∫FDhphv⋅nF=∫FDhgDv⋅nF, | (4.24) |
while Eq. (4.3) turns into
∫Thuh⋅∇q−∫FIh({uh}−b[[uh]])⋅[[q]]+∫FIh∪FDhσF[[ph]]⋅[[q]]−∫FDhquh⋅nF+∫ΓhβΓ[[ph]]⋅[[q]]+∫ΓhαΓ({ph}−pΓ,h){q}=∫Thfq+∫FDhσFgDq. | (4.25) |
If we discretize the problem in the fracture with the SIPDG method, we obtain the following discrete mixed problem: Find ((ph,uh),pΓh)∈Qbh×Wbh×QΓh such that
Mb(uh,v)+Bb(ph,v)=Fb(v)∀v∈Wbh, |
−Bb(q,uh)+Sb(ph,q)+I1(ph,q,pΓ,h)=Gb(q)∀q∈Qbh, | (4.26) |
APΓ(pΓ,h,qΓ)+I2(ph,pΓ,h,qΓ)=LPΓ(qΓ)∀qΓ∈QΓh, |
where
Mb(uh,v)=∫Thν−1uh⋅v, | (4.27) |
Bb(ph,v)=−∫Th∇ph⋅v+∫FIh[[ph]]⋅({v}−b[[v]])+∫FDhphv⋅nF, |
Sb(ph,q)=∫FIh∪FDhσF[[ph]]⋅[[q]], |
I1(ph,q,pΓ,h)=∫ΓhβΓ[[ph]]⋅[[q]]+∫ΓhαΓ({ph}−pΓ,h){q}, |
I2(ph,pΓ,h,qΓ)=∫ΓhαΓ(pΓ,h−{ph})qΓ, |
Fb(v)=∫FDhgDv⋅nF, |
Gb(q)=∫Thfq+∫FDhσFgDq, |
and APΓ(⋅,⋅) and LPΓ(⋅) are defined as in (4.19) and (4.22), respectively. Also note that we have I((ph,pΓ,h),(q,qΓ))=I1(ph,q,pΓ,h)+I2(ph,pΓ,h,qΓ).
We now focus on rewriting the problem in the bulk in a primal form, taking advantage of the local solvability of LDG method. We proceed as in the SIPDG case and introduce an appropriate lifting operator, LLDGb:[L1(FIh∪FDh)]d→Wbh, defined by
∫ThLLDGb(ξ)⋅v=−∫FIh({v}−b[[v]])⋅ξ−∫FDhξ⋅v∀v∈Wbh | (4.28) |
From Eq. (4.24) we obtain
uh=ν(∇ph+LLDGb([[ph]])+Gb), | (4.29) |
where Gb is the lifting of the Dirichlet boundary datum defined in (4.10). Eq. (4.25) now becomes
∫Thν∇ph⋅∇q+∫ThνLLDGb([[ph]])⋅∇qh+∫ThνGb⋅∇q−∫FIh({uh}+b[[uh]])⋅[[q]]−∫FDhquh⋅nF+∫FIh∪FDhσF[[ph]]⋅[[q]]+∫ΓhβΓ[[ph]]⋅[[q]]+∫ΓhαΓ({ph}−pΓ,h){q}=∫Thfq+∫FDhσFqgD. |
Using again the definition of the lifting LLDGb and the identity (4.29), we obtain
∫Thν(∇ph+LLDGb([[ph]]))⋅(∇q+LLDGb([[q]]))+∫FIh∪FDhσF[[ph]]⋅[[q]]+∫ΓhβΓ[[ph]]⋅[[q]]+∫ΓhαΓ({ph}−pΓ){q}=∫Thfq+∫FDhσFqgD−∫ThνGb⋅(∇q+LLDGb([[q]])). | (4.30) |
Summing Eqs. (4.30) and (4.16) we obtain the following discrete formulation: Find (ph,pΓh)∈Qbh×QΓh such that
AMPh((ph,pΓh),(q,qΓ))=LMPh(q,qΓ)∀(q,qΓ)∈Qbh×QΓh, | (4.31) |
where MP stands for mixed-primal and where AMPh:(Qbh×QΓh)×(Qbh×QΓh)→R is defined as
AMPh((ph,pΓh),(q,qΓ))=AMb(ph,q)+APΓ(pΓ,h,qΓ)+I((ph,pΓ,h),(q,qΓ)), |
and LMPh:Qbh×QΓh→R is defined as
LMPh(q,qΓ)=LMb(q)+LPΓ(qΓ) |
with
AMb(ph,q)=∫Thν(∇ph+LLDGb([[ph]]))⋅(∇q+LLDGb([[q]]))+∫FIh∪FDhσF[[ph]]⋅[[q]] |
+∫ΓhβΓ[[ph]]⋅[[q]]+∫ΓhαΓ({ph}−pΓ){q}, | (4.32) |
LMb(q)=∫Thfq+∫FDhσFqgD−∫ThνGb⋅(∇q+LLDGb([[q]])). |
Note that the mixed formulation (4.26) is equivalent to the primal formulation (4.31) together with the definition of the lifting operator (4.28) and Eq. (4.29).
We now want to approximate the problem in the fracture in mixed form, employing the LDG method and the problem in the bulk using the SIPDG method. We define the numerical fluxes as follows
ˆpΓ=ˆpΓ(pΓ,h)={{pΓ,h}+bΓ⋅[[pΓ,h]]onEIΓ,hgΓonEDΓ,hpΓ,honENΓ,h |
ˆuΓ=ˆuΓ(uΓ,h,pΓ,h)={{uΓ,h}−bΓ[[uΓ,h]]−σe[[pΓ,h]]onEIΓ,huΓ,h−σe(pΓ,hne−gΓne)onEDΓ,h0onENΓ,h |
Here, bΓ∈[L∞(EIΓ,h)]d−1 is a vector-valued function that is constant on each edge and it is chosen such that ||bΓ||∞,EIΓ,h≤BΓ, with BΓ≤0 independent of the discretization parameters. Eqs. (4.5) and (4.6) now become
∫Γh(ντΓℓΓ)−1uΓ,h⋅vΓ−∫ΓhvΓ⋅∇pΓ,h+∫EIΓ,h[[pΓ,h]]⋅({vΓ}−bΓ[[vΓ]])+∫EDΓ,hpΓ,hvΓ⋅ne=∫EDΓ,hgΓvΓ⋅ne | (4.33) |
∫ΓhuΓ,h⋅∇qΓ−∫EIΓ,h[[qΓ]]⋅({uΓ,h}−bΓ[[uΓ,h]])+∫EΓ,hσe[[pΓ,h]]⋅[[qΓ]]−∫EDΓ,hqΓuΓ,h⋅ne=∫ΓhfΓqΓ+∫ΓhαΓ({ph−pΓ,h})qΓ+∫EΓ,hσegΓqΓ, | (4.34) |
where we have also used the definition of the numerical flux ˆu on Γh (see (4.7)) to rewrite −[[ˆu]]=αΓ({ph}−pΓ,h). For the bulk we proceed as in the primal-primal section using for the discretization the SIPDG method. We then obtain the following primal-mixed problem: Find (ph,(pΓh,uΓ,h))∈Qbh×QΓh×WΓh such that
3APb(ph,q)+I1((ph,q),pΓ,h)=LPb(q)∀q∈Qbh |
MΓ(uΓ,h,vΓ)+BΓ(pΓ,h,vΓ)=FΓ(vΓ)∀vΓ∈WΓh, | (4.35) |
−BΓ(qΓ,uΓ,h)+SΓ(pΓ,h,qΓ)+I2(ph,(pΓ,h,qΓ))=GΓ(qΓ)∀qΓ∈QΓh, |
where
MΓ(uΓ,h,vΓ)=∫Γh(ντΓℓΓ)−1uΓ,h⋅vΓ,BΓ(pΓ,h,vΓ)=−∫ΓhvΓ⋅∇pΓ,h+∫EIΓ,h[[pΓ,h]]⋅({vΓ}−bΓ[[vΓ]])+∫EDΓ,hpΓ,hvΓ⋅ne,Sb(pΓ,h,qΓ)=∫EΓ,hσe[[pΓ,h]]⋅[[qΓ]],FΓ(vΓ)=∫EDΓ,hgΓvΓ⋅ne,GΓ(qΓ)=∫ΓhfΓqΓ+∫EΓ,hσegΓqΓ, |
and APb(ph,q) and LPb(q) are defined as in (4.18) and (4.21), respectively.
Aiming at rewriting the problem in the fracture in primal form, we introduce the lifting operator, LLDGΓ:[L1(EIh∪EDh)]d→WΓh, defined by
∫ΓhLLDGΓ(ξΓ)⋅vΓ=−∫EIΓ,h({vΓ}−bΓ[[vΓ]])⋅ξΓ−∫EDΓ,hξΓ⋅vΓ∀vΓ∈WΓh | (4.36) |
From Eq. (4.33) we obtain
uΓ,h=ντΓℓΓ[∇pΓ,h+LLDGΓ([[pΓ,h]])+GΓ] | (4.37) |
where GΓ is the lifting of the Dirichlet boundary datum defined in (4.15). Eq. (4.34) now becomes
∫ΓhντΓℓΓ(∇pΓ,h+LLDGΓ([[pΓ,h]]))⋅(∇qΓ+LLDGΓ([[qΓ]]))+∫EIΓ,h∪EDΓ,hσe[[pΓ,h]]⋅[[qΓ]]+∫ΓhαΓ(pΓ,h)−{ph})=∫ΓhfΓqΓ+∫EDΓ,hσeqΓgΓ−∫ΓhντΓℓΓGΓ⋅(∇qΓ+LLDGΓ([[qΓ]])). |
We obtain the following primal formulation: Find (ph,pΓh)∈Qbh×QΓh such that
APMh((ph,pΓh),(q,qΓ))=LPMh(q,qΓ)∀(q,qΓ)∈Qbh×QΓh, | (4.38) |
where PM stands for primal-mixed and where APMh:(Qbh×QΓh)×(Qbh×QΓh)→R is defined as
APMh((ph,pΓh),(q,qΓ))=APb(ph,q)+AMΓ(pΓ,h,qΓ)+I((ph,pΓ,h),(q,qΓ)), |
and LPMh:Qbh×QΓh→R is defined as LPMh(q,qΓ)=LPb(q)+LMΓ(qΓ) with
AMΓ(pΓ,h,qΓ)=∫ΓhντΓℓΓ(∇pΓ,h+LLDGΓ([[pΓ,h]]))⋅(∇qΓ+LLDGΓ([[qΓ]])) |
+∫EIΓ,h∪EDΓ,hσe[[pΓ,h]]⋅[[qΓ]], | (4.39) |
LMΓ(qΓ)=∫ΓhfΓqΓ+∫EDΓ,hσeqΓgΓ−∫ΓhντΓℓΓGΓ⋅(∇qΓ+LLDGΓ([[qΓ]])). |
Finally, if we approximate both the problem in the bulk and in the fracture with the LDG method, we obtain the following formulation: Find (ph,pΓ,h)∈Qbh×QΓh and (uh,uΓ,h)∈Wbh×WΓh such that
3Mb(uh,v)+Bb(ph,v)=Fb(v)∀v∈Wbh,−Bb(q,uh)+Sb(ph,q)+I1(ph,q,pΓ,h)=Gb(q)∀q∈Qbh,MΓ(uΓ,h,vΓ)+BΓ(pΓ,h,vΓ)=FΓ(vΓ)∀vΓ∈WΓh,−BΓ(qΓ,uΓ,h)+SΓ(pΓ,h,qΓ)+I2(ph,(pΓ,h,qΓ))=GΓ(qΓ)∀qΓ∈QΓh, | (4.40) |
This formulation, together with the definition of the lifting operators (4.28) and (4.36) and of the discrete gradients (4.29) and (4.37) is equivalent to the following: Find (ph,pΓ,h)∈Qbh×QΓh such that
AMMh((ph,pΓh),(q,qΓ))=LMMh(q,qΓ)∀(q,qΓ)∈Qbh×QΓh, | (4.41) |
where MM stands for mixed-mixed and where AMMh:(Qbh×QΓh)×(Qbh×QΓh)→R is defined as
AMMh((ph,pΓh),(q,qΓ))=AMb(ph,q)+AMΓ(pΓ,h,qΓ)+I((ph,pΓ,h),(q,qΓ)), |
and LMMh:Qbh×QΓh→R is defined as LMMh(q,qΓ)=LMb(q)+LMΓ(qΓ).
Next, we perform a unified analysis of all of the derived DG discretizations for the fully-coupled bulk-fracture problem. We remark that the analysis will be performed considering the mixed LDG discretizations recast in their primal form, following [69]. For clarity, in Table 1 we summarize the bilinear forms for all the four approaches.
Method | Primal bilinear form |
Primal-Primal (PP) | APb(p,q)+APΓ(pΓ,qΓ)+I((p,q),(pΓ,qΓ)) |
Mixed-Primal (MP) | AMb(p,q)+APΓ(pΓ,qΓ)+I((p,q),(pΓ,qΓ)) |
Primal-Mixed (PM) | APb(p,q)+AMΓ(pΓ,qΓ)+I((p,q),(pΓ,qΓ)) |
Mixed-Mixed (MM) | AMb(p,q)+AMΓ(pΓ,qΓ)+I((p,q),(pΓ,qΓ)) |
The bulk, fracture and interface bilinear forms are defined in:
APb(p,q):(4.18) APΓ(pΓ,qΓ):(4.19) I((p,q),(pΓ,qΓ)):(4.20)AMΓ(pΓ,qΓ):(4.32) AMb(p,q):(4.39) |
In this section, we address the problem of stability. We prove that the primal-primal (PP) (4.17), mixed-primal (MP) (4.31), primal-mixed (PM) (4.38) and mixed-mixed (MM)(4.41) formulations are well-posed. We remark that all these formulations are not strongly consistent, due to the presence of the lifting operators. This implies that the analysis will be based on Strang's second Lemma, [70].
We recall that, for simplicity in the analysis, we are assuming the permeability tensors ν and ντΓ to be piecewise constant. We will employ the following notation ˉνE=|√ν|E|22 and ˉντF=|√ντΓ|F|22, where |⋅|2 denotes the l2-norm.
To consider the boundedness and stability of our primal bilinear forms, we introduce the spaces Qb(h)=Qbh+˜Qb and QΓ(h)=QΓh+˜QΓ where ˜Qb={q=(q1,q2)∈H1(Ω1)×H1(Ω2)}∩H2(Th) and ˜QΓ=H1(Γ)∩H2(Γh). We remark that all the bilinear forms A∗∗h(⋅,⋅) are also well-defined on the extended space Qb(h)×QΓ(h).
Further, we introduce the following energy norm on the discrete space Qbh×QΓh
|||(q,qΓ)|||2=||q||2DG+||qΓ||2DG+||(q,qΓ)||2I |
where
||q||2DG=||ν1/2∇q||20,Th+||σ1/2F[[q]]||20,FIh∪FDh||qΓ||2DG=||(ντΓℓΓ)1/2∇qΓ||20,Γh+||σ1/2e[[qΓ]]||20,EIΓ,h∪EDΓ,h||(q,qΓ)||2I=||β1/2Γ[[q]]||20,Γh+||α1/2Γ({q}−qΓ)||20,Γh |
Note that |||⋅||| is also well defined on the extended space Qb(h)×QΓ(h).
Since our discretization employ general polytopic grids, we need introduce some technical instruments to work in this framework [36,37,38,40,41]. In particular, we need trace inverse estimates to bound the norm of a polynomial on a polytope's face/edge by the norm on the element itself. To this aim, we give the following
Definition 5.1. A mesh Th is said to be polytopic-regular if, for any E∈Th, there exists a set of non-overlapping (not necessarily shape-regular) d-dimensional simplices {SiE}nEi=1 contained in E, such that ˉF=∂ˉE∩¯SiE, for any face F⊆∂E, and
hE≲d|SiE||F|,i=1,…,nE, |
with the hidden constant independent of the discretization parameters, the number of faces of the element nE, and the face measure.
We remark that this definition does not give any restriction on the number of faces per element, nor on their measure relative to the diameter of the element the face belongs to.
Assumption 5.1. We assume that Th and Γh are polytopic-regular meshes.
With this hypothesis, we can state the following inverse-trace estimate that is valid for polytopic elements [38,41].
Lemma 5.2. Let E be a polygon/polyhedron belonging to a mesh satisfying Definition 5.1 and let v∈PkE(E). Then, we have
||v||2L2(∂E)≲k2EhE||v||2L2(E), | (5.1) |
where the hidden constant depends on the dimension d, but it is independent of the discretization parameters, of the number of faces of the element and of the relative size of the face compared to the diameter kE of E.
The second fundamental tool to deal with polytopic discretizations, is an appropriate definition of the discontinuity penalization parameter, which allows for the use of elements with arbitrarily small faces. Taking as a reference [36,37,38,40,41], we give the following two definitions for the bulk and fracture penalty functions:
Definition 5.2. The discontinuity-penalization parameter σ:Fh∖Γh→R+ for the bulk problem is defined facewise as
σ(x)=σ0{maxE∈{E+,E−}ˉνEk2EhEif x⊂F∈FIh,ˉF=∂ˉE+∩∂ˉE−,ˉνEk2EhEif x⊂F∈FDh,ˉF=∂ˉE∩∂ˉΩ, | (5.2) |
with σ0>0 independent of kE, |E| and |F|.
Definition 5.3. The discontinuity-penalization parameter σΓ:EΓ,h→R+ for the fracture problem is defined edgewise as
σΓ(x)=σ0,Γ{maxF∈{F+,F−}ˉντFk2FhFif x⊂e∈EIΓ,h,ˉe=∂ˉF+∩∂ˉF−,ˉντFk2FhF,if x⊂e∈EDΓ,h,ˉe=∂ˉF∩∂ˉΓ, | (5.3) |
with σ0,Γ>0 independent of kF, |F| and |e|.
Now we have all the technical tools to work in a polytopic framework. Next, we will state and prove some estimates that will be instrumental for the proof of the well-posedness of our discrete formulations. We start deriving some bounds on the lifting operators, with arguments similar to those of [68,69,71]. Note that all the results hold true on the extended spaces Qb(h) and QΓ(h).
Lemma 5.3. Let LSIPb(⋅) be the lifting operator defined in (4.9). Then, for every q∈Qb(h) it holds
||ν1/2LSIPb([[q]])||0,Ω≲1σ1/20||σ1/2F[[q]]||0,FIh∪FDh. | (5.4) |
Proof. Denoting by ΠWbh the L2-projection onto Wbh, by definition of the lifting operator LSIPb and Cauchy-Schwarz inequality, we have
||ν1/2LSIPb([[q]])||0,Ω=supz∈[L2(Ω)]d∫Ων1/2LSIPb([[q]])⋅z||z||0,Ω=supz∈[L2(Ω)]d∫ΩLSIPb([[q]])⋅ΠWbh(ν1/2z)||z||0,Ω=−supz∈[L2(Ω)]d∫FIh∪FDhσ1/2F[[q]]⋅σ−1/2F{ΠWbh(ν1/2z)}||z||0,Ω≤supz∈[L2(Ω)]d||σ1/2F[[q]]||0,FIh∪FDh||σ−1/2F{ΠWbh(ν1/2z)}||0,FIh∪FDh||z||0,Ω. |
Using the triangular inequality, the definition of the penalization coefficient σF (5.2), the inverse inequality (5.1), the assumptions on the permeability tensor ν and the continuity property of the L2-projector we have
||σ−1/2F{ΠWbh(ν1/2z)}||20,FIh∪FDh≲∑E∈Th1σ0hEˉνEk2E||ΠWbh(ν1/2z)||20,∂E≲∑E∈Th1ˉνE1σ0||ΠWbh(ν1/2z)||20,E | (5.5) |
≲1σ0∑E∈Th||z||20,E=1σ0||z||20,Ω. |
This proves the desired estimate.
Lemma 5.4. Let LSIPΓ(⋅) be the lifting operator defined in (4.14). Then, for every qΓ∈QΓ(h) it holds
||(ντΓℓΓ)1/2LSIPΓ([[qΓ]])||0,Γ≲1σ1/20,Γ||σ1/2e[[qΓ]]||0,EIΓ,h∪EDΓ,h. |
Proof. Same arguments as in in the proof of Lemma 5.3.
Lemma 5.5. Let LLDGb(⋅) be the lifting operator defined in (4.28). Then, for every q∈Qb(h) it holds
||ν1/2LLDGb([[q]])||0,Ω≲1+Bσ1/20||σ1/2F[[q]]||0,FIh∪FDh. | (5.6) |
Proof. We proceed as in the proof of Lemma 5.3. By definition of the lifting operator LLDGb and Cauchy-Schwarz inequality, we have
||ν1/2LLDGb([[q]])||0,Ω=supz∈[L2(Ω)]d∫Ων1/2LLDGb([[q]])⋅z||z||0,Ω=supz∈[L2(Ω)]d∫ΩLLDGb([[q]])⋅ΠWbh(ν1/2z)||z||0,Ω≤supz∈[L2(Ω)]d|−∫FIhσ1/2F[[q]]⋅σ−1/2F({ΠWbh(ν1/2z)}−b[[ΠWbh(ν1/2z)]])|||z||0,Ω+supz∈[L2(Ω)]d|−∫FDhσ1/2F[[q]]⋅σ−1/2FΠWbh(ν1/2z)|||z||0,Ω≤supz∈[L2(Ω)]d||σ1/2F[[q]]||0,FIh∪FDh||σ−1/2F{ΠWbh(ν1/2z)}||0,FIh∪FDh||z||0,Ω+supz∈[L2(Ω)]d||σ1/2F[[q]]||0,FIh||σ−1/2Fb[[ΠWbh(ν1/2z)]]||0,FIh||z||0,Ω=(a)+(b) |
From (5.5) we know that (a)≲1σ1/20||σ1/2F[[q]]||0,FIh∪FDh, while using similar arguments and bound (4.23) on b, we can prove that
||σ−1/2Fb[[ΠWbh(ν1/2z)]]||20,FIh≲B2σ0||z||20,Ω, |
so that (b)≲Bσ1/20||σ1/2F[[q]]||0,FIh∪FDh. This concludes the proof.
Lemma 5.6. Let LLDGΓ(⋅) be the lifting operator defined in (4.36). Then, For every qΓ∈QΓ(h) it holds
||(ντΓℓΓ)1/2LLDGΓ([[qΓ]])||0,Γ≲1+BΓσ1/20,Γ||σ1/2e[[qΓ]]||0,EIΓ,h∪EDΓ,h. |
Proof. Same arguments as in in the proof of Lemma 5.5.
Using these results, we can now prove that the bilinear forms for the bulk problem APb(⋅,⋅) and AMb(⋅,⋅) are continuous on Qb(h) and coercive on Qbh, as well as the fracture bilinear forms APΓ(⋅,⋅) and AMΓ(⋅,⋅) are continuous on QΓ(h) and coercive on QΓh.
Lemma 5.7. APb(⋅,⋅) is coercive on Qbh×Qbh and continuous on Qb(h)×Qb(h), that is
APb(q,q)≳||q||2DG∀q∈Qbh, |
APb(p,q)≲||p||DG||q||DG∀p,q∈Qb(h), |
provided that σ0 is chosen big enough.
Proof. We start with coercivity. Taking p=q∈Qbh, we have
APb(q,q)=∑E∈Th[||ν1/2∇q||20,E+2∫EνLSIPb([[q]])⋅∇q]+∑F∈FIh∪FDh||σ1/2F[[q]]||20,F |
From Young inequality we have
2∫EνLSIPb([[q]])⋅∇q≥−2||ν1/2LSIPb([[q]])||0,E||ν1/2∇q||0,E≥−ε||ν1/2LSIPb([[q]])||20,E−1ε||ν1/2∇q||20,E, |
so that, using the bound on the lifting (5.4), we obtain
APb(q,q)≥∑E∈Th[(1−ε)‖ν1/2∇q‖20,E−1ε‖ν1/2LSIPb([[q]])‖20,E]+∑F∈FIh∪FDh||σ1/2F[[q]]||20,F≳(1−ε)∑E∈Th‖ν1/2∇q‖20,E+(1−1σ0ε)∑F∈FIh∪FDh||σ1/2F[[q]]||20,F≳||q||2DG, |
for σ0 big enough.
Continuity directly follows from Cauchy Schwarz inequality and the bound on the lifting (5.4).
Lemma 5.8. APΓ(⋅,⋅) is coercive on QΓh×QΓh and continuous on QΓ(h)×QΓ(h), that is
APΓ(qΓ,qΓ)≳||qΓ||2DG∀qΓ∈QΓh, |
APΓ(pΓ,qΓ)≲||pΓ||DG||qΓ||DG∀pΓ,qΓ∈QΓ(h), |
provided that σ0,Γ is chosen big enough.
Proof. Analogous to the proof of Lemma 5.7.
Lemma 5.9. AMb(⋅,⋅) is coercive on Qbh×Qbh and continuous on Qb(h)×Qb(h), that is
AMb(q,q)≳||q||2DG∀q∈Qbh, |
AMb(p,q)≲||p||DG||q||DG∀p,q∈Qb(h). |
Proof. We start with coercivity. From Young's inequality and the bound on the lifting (5.10) we have, for every 0<ε<1,
AMb(q,q)=∑E∈Th[||ν1/2∇q||20,E+||ν1/2LLDGb([[q]])||20,E+2∫EνLLDGb([[q]])⋅∇q]+∑F∈FIh∪FDh||σ1/2F[[q]]||20,F≥∑E∈Th[(1−ε)‖ν1/2∇q‖20,E+(1−1ε)||ν1/2LLDGb([[q]])||20,E]+∑F∈FIh∪FDh||σ1/2F[[q]]||20,F≳(1−ε)∑E∈Th‖ν1/2∇q‖20,E+(1+C)∑F∈FIh∪FDh||σ1/2F[[q]]||20,F |
with C=(1+B)σ0ε(1−1ε), so that AMb(⋅,⋅) is coercive for every choice of the parameters σ0>0 and B>0*. Continuity is again a direct consequence of Cauchy Schwarz's inequality and the bound on the lifting (5.6).
* More in detail: we need 1+C>0, with 0<ε<1. We obtain 1+(1−1ε)(1+B)2σ0>0, that is ε>11+σ0(1+B)2=˜C, being 0<˜C<1 for every possible choice of σ0>0 and B>0.
Lemma 5.10. AMΓ(⋅,⋅) is coercive on QΓh×QΓh and continuous on QΓ(h)×QΓ(h), that is
AMΓ(qΓ,qΓ)≳||qΓ||2DG∀qΓ∈QΓh, |
AMΓ(pΓ,qΓ)≲||pΓ||DG||qΓ||DG∀pΓ,qΓ∈QΓ(h). |
Proof. Analogous to the proof of Lemma 5.9.
Employing Lemmas 5.7, 5.9, 5.8 and 5.10, we can easily prove the well-posedness of all of our discrete problems, as stated in the following stability result.
Proposition 5.11. Let the penalization parameters σ for the problem in the bulk and in the fracture be defined as in (5.2) and (5.3), respectively. Then, the fully-coupled discrete problems PP (4.22), MP (4.31), PM (4.38) and MM (4.41) are well-posed provided that σ0 and σ0,Γ are chosen big enough for the primal formulations.
Proof. In order to use Lax-Milgram Lemma, we prove that the bilinear forms APPh(⋅,⋅), AMPh(⋅,⋅), APMh(⋅,⋅) and AMMh(⋅,⋅) are continuous on Qb(h)×QΓ(h) and coercive on Qbh×QΓh. We have, from Cauchy-Schwarz inequality
I((q,qΓ),(q,qΓ))=||(q,qΓ)||2II((q,qΓ),(w,wΓ))≤∑F∈Γh||β1/2Γ[[q]]||2L2(F)||β1/2Γ[[w]]||2L2(F)+∑F∈Γh||α1/2Γ({q}−qΓ)||2L2(F)||α1/2Γ({w}−wΓ)||2L2(F)≤|||(q,qΓ)|||⋅|||(w,wΓ)|||, |
so that coercivity and continuity are a direct consequence of the definition of the norm |||⋅||| and of Lemmas 5.7, 5.9, 5.8 and 5.10. The continuity of the linear operators LPPh(⋅), LMPh(⋅), LPMh(⋅) and LMMh(⋅) on Qb(h)×QΓ(h) can be easily proved by using Cauchy-Schwarz's inequality, thanks to the regularity assumptions on the forcing terms f and fΓ and on the boundary data gD and gΓ.
In this section we derive error estimates for our discrete problems.
The tool at the base of DG-method error analysis are hp-interpolation estimates. Here, we summarize the results contained in [36,37,38,40,41], where standard estimates on simplices are extended to arbitrary polytopic elements.
First, we give the following definitions.
Definition 6.1. A covering T#={TE} related to the polytopic mesh Th is a set of shape-regular d-dimensional simplices TE, such that for each E∈Th, there exists a TE∈T# such that E⊊TE.
Assumption 6.1. [36,37,38,40,41] There exists a covering T# of Th (see Definition 6.1) and a positive constant OΩ, independent of the mesh parameters, such that
maxE∈Thcard{E′∈Th:E′∩TE≠∅,TE∈T#s.t.E⊂TE}≤OΩ, |
and hTE≲hE for each pair E∈Th and TE∈T#, with E⊂TE.
Moreover, there exists a covering F# of Γh and a positive constant OΓ, independent of the mesh parameters, such that
maxF∈Γhcard{F′∈Γh:F′∩TF≠∅,TF∈F#s.t.F⊂TF}≤OΓ, |
and hTF≲hF for each pair F∈Γh and TF∈F#, with F⊂TF.
We can now state the following approximation result:
Lemma 6.2. [36,37,38,40,41] Let E∈Th and TE∈T# denote the corresponding simplex such that E⊂TE (see Definition 6.1). Suppose that v∈L2(Ω) is such that Ev|TE∈HrE(TE), for some rE≥0. Then, if Assumption 5.1 and 6.1 are satisfied, there exists ˜Πv, such that ˜Πv|E∈PkE(E), and the following bound holds
||v−˜Πv||Hq(E)≲hsE−qEkrE−qE||Ev||HrE(TE),0≤q≤rE. | (6.1) |
Moreover, if rE>1/2,
||v−˜Πv||L2(∂E)≲hsE−1/2EkrE−1/2E||Ev||HrE(TE). | (6.2) |
Here, sE=min(kE+1,rE) and the hidden constants depend on the shape-regularity of TE, but are independent of v, hE, kE and the number of faces per element and E is the continuous extension operator as defined in [72].
Proof. See [36] for a detailed proof of (6.1) and [38] for the proof of (6.2). Clearly, analogous approximation results can be stated on the fracture spaces, since Assumptions 5.1 and 6.1 are both satisfied.
For each subdomain Ωi, i=1,2, we denote by Ei the classical continuous extension operator (cf. [72], see also [59]) Ei:Hs(Ωi)→Hs(Rd), for s∈N0. Similarly, we denote by EΓ the continuous extension operator EΓ:Hs(Γ)→Hs(Rd−1), for s∈N0. We then make the following regularity assumptions for the exact solution (p,pΓ) of problem (3.1):
Assumption 6.3. Let T#={TE} and F#={TF} denote the associated coverings of Ω and Γ, respectively, of Definition 6.1. We assume that the exact solution (p,pΓ) is such that:
A1. for every E∈Th, if E⊂Ωi, it holds Eipi|TE∈HrE(TE), with rE≥1+d/2 and TE∈T# with E⊂TE;
A2. for every F∈Γh, it holds EΓpΓ|TF∈HrF(TF), with rF≥1+(d−1)/2 and TF∈F# with F⊂TF.
Assumption 6.4. We assume that the normal components of the exact fluxes ν∇p and ℓΓντΓ∇pΓ are continuous across mesh interfaces, that is [[ν∇p]]=0 on FIh and [[ℓΓντΓ∇pΓ]]=0 on EIΓ,h.
From Proposition 5.11 and Strang's second Lemma the following abstract error bound directly follows.
Lemma 6.5. Assuming that the hypotheses of Proposition 5.11 are satisfied, for all the discrete problems PP (4.17), MP (4.31), MM (4.41) and PM (4.41) the following abstract error bound holds
|||(p,pΓ)−(ph,pΓ,h)|||≲inf(q,qΓ)∈Qbh×QΓh|||(p,pΓ)−(q,qΓ)|||+sup(w,wΓ)∈Qbh×QΓh|R∗∗h((p,pΓ),(w,wΓ))||||(w,wΓ)|||, |
where the residual R∗∗h is defined as
R∗∗h((p,pΓ),(w,wΓ))=A∗∗h((p,pΓ),(w,wΓ))−L∗∗h(w,wΓ), |
with ∗∗∈{PP,MP,MM,PM}.
Note that, irrespective of the numerical method chosen for the discretization (PP, MP, PM or MM), the residual can always be split into two contributions, one deriving from the approximation of the problem in the bulk and one deriving from the approximation of the problem in the fracture, i.e.,
R∗∗h((p,pΓ),(w,wΓ))=R∗b(p,w)+R∗Γ(pΓ,wΓ) | (6.3) |
It follows that, to derive a bound for the global residual, we can bound each of the two contributions separately. With this in mind, we state and prove the next two lemmas.
Lemma 6.6. Let (p,pΓ) be the exact solution of problem (3.1) satisfying the regularity Assumptions 6.4 and 6.3. Then, for every w∈Qb(h) and wΓ∈QΓ(h), it holds
|RPb(p,w)|2≲∑E∈Thh2(sE−1)Ek2(rE−1)E||Ep||2HrE(TE)[ˉν2EmaxF⊂∂E∖Γσ−1F(kEhE+k2EhE)]⋅||w||2DG, | (6.4) |
|RPΓ(pΓ,wΓ)|2≲∑F∈Γhh2(sF−1)Fk2(rF−1)F||EpΓ||2HRF(TF)[(ˉντFℓΓ)2maxe⊆∂Fσ−1e(kFhF+k2FhF)]⋅||wΓ||2DG. | (6.5) |
Proof. First, we prove (6.4). Let ΠWbh be the L2-orthogonal projector onto Wbh, then, integrating by parts elementwise, using the fact that p satisfies (2.2) and recalling that, from Assumption 6.4, [[ν∇p]] vanishes on FIh, we obtain the following expression for the residual RPb:
RPb(p,w)=∑F∈FIh∪FDh∫F{ν(∇p−ΠWbh(∇p))}⋅[[w]],∀w∈Qb(h). |
Employing the Cauchy-Schwarz's inequality and the definition of the norm |||⋅|||, we then obtain
|RPb(p,w)|2≲(∑F∈FIh∪FDhσ−1F∫F|{ν(∇p−ΠWbh(∇p))}|2)⋅||w||2DG,∀w∈Qb(h). |
If we still denote by ˜Π the vector-valued generalization of the projection operator ˜Π defined in Lemma 6.2, we observe that
∑F∈FIh∪FDhσ−1F∫F|{ν(∇p−ΠWbh(∇p))}|2≤∑F∈FIh∪FDhσ−1F∫F|{ν(∇p−˜Π(∇p))}|2+∑F∈FIh∪FDhσ−1F∫F|{νΠWbh(∇p−˜Π(∇p))}|2≡(1)+(2). |
To bound the term (1), we employ the approximation result stated in Lemma 6.2. We obtain
(1)≲∑E∈Thh2(sE−1)Ek2(rE−1)E(ˉν2EmaxF⊂∂E∖Γσ−1Fh−1Ek−1E)||Ep||2HrE(TE). |
Exploiting, the boundedness of the permeability tensor ν, the inverse inequality (5.1), the L2-stability of the projector ΠWbh and the approximation results stated in Lemma 6.2, we can bound term (2) as follows:
(2)≲∑E∈ThmaxF⊂∂E∖Γσ−1Fˉν2E||ΠWbh(˜Π(∇p)−∇p)||2L2(∂E∖Γ)≲∑E∈ThmaxF⊂∂E∖Γσ−1Fˉν2Ek2EhE||˜Π(∇p)−∇p||2L2(E)≲∑E∈Thh2(sE−1)Ek2(rE−1)E||Ep||2HrE(TE)(ˉν2Ek2EhEmaxF⊂∂E∖Γσ−1F), |
which concludes the proof of (6.4).
Proceeding as above we obtain the following expression for the residual RPΓ:
RPΓ(pΓ,wΓ)=∑e∈EIΓ,h∪EDΓ,h∫e{ντΓℓΓ(∇pΓ−ΠWΓh(∇pΓ))}⋅[[wΓ]]. |
Estimate (6.5) can then be proven with analogous arguments.
Lemma 6.7. Let (p,pΓ) be the exact solution of problem (3.1) satisfying the regularity Assumptions 6.4 and 6.3. Then, for every w∈Qb(h) and wΓ∈QΓ(h), it holds
|RMb(p,w)|2≲∑E∈Thh2(sE−1)Ek2(rE−1)E||Ep||2HrE(TE)[(1+B)ˉν2EmaxF⊂∂E∖Γσ−1F(kEhE+k2EhE)]⋅||w||2DG, | (6.6) |
|RMΓ(pΓ,wΓ)|2≲∑F∈Γhh2(sF−1)Fk2(rF−1)F||EpΓ||2HRF(TF)[(1+BΓ)(ˉντFℓΓ)2maxe⊆∂Fσ−1e(kFhF+k2FhF)]⋅||wΓ||2DG. | (6.7) |
Proof. We focus on the proof of (6.6), estimate (6.7) can be obtained likewise. Recalling that ΠWbh denotes the L2-orthogonal projector onto Wbh, the residual RMb has the following expression:
RMb(p,w)=∑F∈FIh∪FDh∫F({ν(∇p−ΠWbh(∇p))}−b[[ν(∇p−ΠWbh(∇p))]])⋅[[w]]+∑F∈FDh∫Fwν(∇p−ΠWbh(∇p))⋅nF, |
where we have used the identity LLDGb([[p]])=−Gb and the continuity of ν∇p across internal faces (Assumption 6.4). From Cauchy-Schwarz and triangular inequalities and the bound on the coefficient b (4.23), we have
|RMb(p,w)|2≲(∑F∈FIh∪FDhσ−1F[∫F|{ν(∇p−˜Π(∇p))}|2+∫F|{νΠWbh(∇p−˜Π(∇p))}|2]+B∑F∈FIh∪FDhσ−1F[∫F|[[ν(∇p−˜Π(∇p))]]|2+∫F|[[νΠWbh(∇p−˜Π(∇p))]]|2])⋅||w||2DG, |
where we recall that, with a slight abuse of notation, ˜Π still denotes the vector-valued generalization of the projection operator ˜Π defined in Lemma 6.2. The thesis now follows from the boundedness of the permeability tensor ν, the inverse inequality (5.1), the L2-stability of the projector ΠWbh and the approximation results stated in Lemma 6.2.
Theorem 6.8. Let T#={TE} and F#={TF} denote the associated coverings of Ω and Γ, respectively, consisting of shape-regular simplexes as in Definition 4, satisfying Assumptions 6.1. Let (p,pΓ) be the solution of problem (3.1) and (ph,pΓ,h)∈Qbh×QΓh be its approximation obtained with the method PP, MP, MM or PM, with the penalization parameters given by (5.2) and (5.3) and σ0 and σ0,Γ sufficiently large for the primal formulations. Moreover, suppose that the exact solution (p,pΓ) satisfies the regularity Assumptions 6.4 and 6.3. Then, the following error bound holds:
|||(p,pΓ)−(ph,pΓ,h)|||2≲∑E∈Thh2(sE−1)Ek2(rE−1)EG∗E(hE,kE,ˉνE)||Ep||2HrE(TE)+∑F∈Γhh2(sF−1)Fk2(rF−1)FG∗F(hF,kF,ˉντF)||EΓpΓ||2HrF(TF), |
where the Ep is to be interpreted as E1p1 when E⊂Ω1 or as E2p2 when E⊂Ω2. Here, sE=min(kE+1,rE) and sF=min(kF+1,rF), and the constants are defined according to the chosen approximation method as follows:
GPE(hE,kE,ˉνE)=ˉνE+hEk−1EmaxF⊂∂E∖ΓσF+(αΓ+βΓ)hEk−1E+ˉν2Eh−1EkEmaxF⊂∂E∖Γσ−1F+ˉν2Eh−1Ek2EmaxF⊂∂E∖Γσ−1F,GPF(hF,kF,ˉντF)=ˉντFℓΓ+hFk−1Fmaxe⊆∂Fσe+αΓh2Fk−2F+(ˉντFℓΓ)2h−1FkFmaxe⊆∂Fσ−1e+(ˉντFℓΓ)2h−1Fk2Fmaxe⊆∂Fσ−1e,GME(hE,kE,ˉνE)=ˉνE+hEk−1EmaxF⊂∂E∖ΓσF+(αΓ+βΓ)hEk−1E+(1+B)ˉν2Eh−1EkEmaxF⊂∂E∖Γσ−1F+(1+B)ˉν2Eh−1Ek2EmaxF⊂∂E∖Γσ−1F,GMF(hF,kF,ˉντF)=ˉντFℓΓ+hFk−1Fmaxe⊆∂Fσe+αΓh2Fk−2F+(1+BΓ)(ˉντFℓΓ)2h−1FkFmaxe⊆∂Fσ−1e+(1+BΓ)(ˉντFℓΓ)2h−1Fk2Fmaxe⊆∂Fσ−1e. |
Proof. From Lemma 6.5 we know that the error satisfies the following bound
|||(p,pΓ)−(ph,pΓ,h)|||≲inf(q,qΓ)∈Qbh×QΓh|||(p,pΓ)−(q,qΓ)|||⏟I+sup(w,wΓ)∈Qbh×QΓh|Rh((p,pΓ),(w,wΓ))||||(w,wΓ)|||⏟II. | (6.8) |
We estimate the two terms on the right-hand side of (6.8) separately. We can rewrite term I as
I=inf(q,qΓ)∈Qbh×QΓh(||p−q||2DG+||pΓ−qΓ||2DG+||(p−q,pΓ−qΓ||2I)≤||p−˜Πp||2DG⏟(a)+||pΓ−˜ΠpΓ||2DG⏟(b)+||(p−˜Πp,pΓ−˜ΠpΓ)||2I⏟(c). |
Again we consider each of the three terms separately. To bound term (a), we exploit the two approximation results stated in Lemma 6.2 and obtain
(a)≤||p−˜Πp||2DG=∑E∈Th||ν1/2∇(p−˜Πp)||2L2(E)+∑F∈FIh∪FDhσF||[[p−˜Πp]]||2L2(F)≲∑E∈Th[ˉνE|p−˜Πp|2H1(E)+(maxF⊂∂E∖ΓσF)||p−˜Πp||2L2(∂E∖Γ)]≲∑E∈Th[h2(sE−1)Ek2(rE−1)EˉνE||Ep||2HrE(TE)+∑F⊂∂E∖Γh2(sE−1/2)Ek2(rE−1/2)E(maxF⊂∂E∖ΓσF)||Ep||2HrE(TE)]=∑E∈Thh2(sE−1)Ek2(rE−1)E||Ep||2HrE(TE)(ˉνE+hEkE(maxF⊂∂E∖ΓσF)). |
Using analogous interpolation estimates on the fracture we can bound term (b) as follows:
(b)≤||pΓ−˜ΠpΓ||2DG≲∑F∈Γh||ντΓℓΓ∇(pΓ−˜ΠpΓ)||2L2(F)+∑e∈EIΓ,h∪EDΓ,hσe||[[pΓ−˜ΠpΓ]]||2L2(e)≲∑F∈Γhh2(sF−1)Fk2(rF−1)F||EpΓ||2HRF(TF)(ˉντFℓΓ+hFkFmaxe⊆∂Fσe) |
Finally, for term (c), we have
(c)≤||(p−˜Πp,pΓ−˜ΠpΓ)||2I≤βΓ∑F∈Γh||[[p−˜Πp]]||2L2(F)+αΓ∑F∈Γh||{p−˜Πp}||2L2(F)+αΓ∑F∈Γh||pΓ−˜ΠpΓ||2L2(F). |
Exploiting the approximation result (6.2), we obtain
βΓ∑F∈Γh||[[p−˜Πp]]||2L2(F)≤βΓ∑E∈Th∂E∩Γ≠∅||p−˜Πp||2L2(∂E)≲βΓ∑E∈Th∂E∩Γ≠∅h2(sE−12)Ek2(rE−12)E||Ep||2HrE(TE)=βΓ∑E∈Th∂E∩Γ≠∅h2(sE−1)Ek2(rE−1)E||Ep||2HrE(TE)hEkE. |
Similarly, we have
αΓ∑F∈Γh||{p−˜Πp}||2L2(F)≲αΓ∑E∈Th∂E∩Γ≠∅h2(sE−1)Ek2(rE−1)E||Ep||2HrE(TE)hEkE. |
Finally, using the interpolation estimates for the fracture, we deduce that
αΓ∑F∈Γh||pΓ−˜ΠpΓ||2L2(F)≲αΓ∑F∈Γhh2sFFk2rF||EpΓ||2HrF(TF)=αΓ∑F∈Γhh2(sF−1)Fk2(rF−1)F||EpΓ||2HRF(TF)h2Fk2F. |
In conclusion, combining all the previous estimates, we can bound the term I on the right-hand side of (6.8) as follows:
I≲∑E∈Thh2(sE−1)Ek2(rE−1)E||Ep||2HrE(TE)[ˉνE+hEkEmaxF⊂∂E∖ΓσF+(αΓ+βΓ)hEkE]+∑F∈Γhh2(sF−1)Fk2(rF−1)F||EpΓ||2HRF(TF)[ˉντFℓΓ+hFkFmaxe⊆∂Fσe+αΓh2Fk2F]. | (6.9) |
Finally, the desired estimates follow from the combination of (6.9), together with the bound on Term II deriving from what observed in (6.3) and Lemmas 6.6 and 6.7.
Finally, from the above result we can derive some error estimates also for the velocities u and uΓ.
Theorem 6.9. Let all the hypotheses of Theorem 6.8 hold. Let (u,uΓ)∈Wg and (p,pΓ)∈M be the solution of problem (3.1). Then:
● if ((ph,uh),pΓ,h)∈Qbh×Wbh×QΓh is its approximation obtained with the MP method (4.26), it holds
||u−uh||20,Th≲∑E∈Thh2(sE−1)Ek2(rE−1)EGME||Ep||2HrE(TE)+∑F∈Γhh2(sF−1)Fk2(rF−1)FGPF||EΓpΓ||2HrF(TF); |
● if (ph,(pΓ,h,uΓ,h))∈Qbh×QΓh×WΓh is its approximation obtained with the PM method (4.35), it holds
||uΓ−uΓ,h||20,Γh≲∑E∈Thh2(sE−1)Ek2(rE−1)EGPE||Ep||2HrE(TE)+∑F∈Γhh2(sF−1)Fk2(rF−1)FGMF||EΓpΓ||2HrF(TF); |
● if ((ph,uh),(pΓ,h,uΓ,h))∈Qbh×Wbh×QΓh×WΓh is its approximation obtained with the MM method (4.40), it holds
||u−uh||20,Th+||uΓ−uΓ,h||20,Γh≲∑E∈Thh2(sE−1)Ek2(rE−1)EGME||Ep||2HrE(TE)+∑F∈Γhh2(sF−1)Fk2(rF−1)FGMF||EΓpΓ||2HrF(TF), |
where the constants GME, GPF, GPE and GMF are defined as in Theorem 6.8.
Proof. Let ((ph,uh),pΓ,h) and ((ph,pΓ,h),(uh,uΓ,h)) be the discrete solutions obtained with the MP method and with the MM method, respectively. Then, using identity (4.29) and the fact that LLDGb([[p]])=−Gb, we can rewrite
uh−u=ν∇ph+νLLDGb([[ph]])+νGb−ν∇p=ν(∇ph−∇p)+νLLDGb([[ph−p]]). |
From the uniform boundedness of ν, the triangular inequality, the bound on the lifting (5.6) and the definition of the ||⋅||DG norm it follows that
||u−uh||0,Th≲||ν1/2∇(ph−p)||0,Th+||ν1/2LLDGb([[ph−p]])||0,Th≲||ph−p||DG+1+Bσ1/20||σ1/2F[[ph−p]]||0,FIh∪FDh≲||ph−p||DG. |
In particular, this implies that ||u−uh||0,Th≲|||(p,pΓ)−(ph,pΓ,h)|||. Similarly, one can prove that, if (ph,(pΓ,h,uΓ,h)) and ((ph,pΓ,h),(uh,uΓ,h)) are the discrete solutions obtained with the PM method and with the MM method, respectively, it holds ||uΓ−uΓ,h||0,Γh≲|||(p,pΓ)−(ph,pΓ,h)|||. The thesis is now a direct consequence of Theorem 6.8.
In this section we present some two-dimensional numerical experiments with the aim of validating the obtained theoretical convergence results and of comparing the accuracy of the proposed formulations. In particular, we will focus on the paradigmatic primal-primal and mixed-primal settings. This means that, for the approximation of the problem in the bulk, we will employ the SIPDG method in the first setting and the LDG method in the second setting, while, for the problem in the fracture, we will employ the SIPDG method in both settings. All the numerical tests have been implemented in \texttt{Matlab}^{®} . For the generation of polygonal meshes conforming to the fractures we have suitably modified the code \texttt{PolyMesher} [73].
In particular, we present three sets of numerical experiments. The first set (Sections 7.1 and 7.2) is obtained assuming that analytical solutions are known and aims at verifying the a-priori error estimates obtained in Theorems 6.8 and 6.9. The second set (Section 7.3) is derived from physical considerations and aims at testing how different values of the fracture permeability may influence the flow in the bulk. Finally, the last set of experiments (Section 7.4) aims at showing how the method is capable of handling more complicated geometries, specifically networks of partially immersed fractures.
In this first test case we take \Omega = (0, 1)^2 and choose as exact solutions in the bulk and in the fracture \Gamma = \{(x, y) \in \Omega: \; \; x+y = 1\}
\begin{equation*} p = \begin{cases} e^{x+y} & \mbox{in }\Omega_1, \\ e^{x+y}+ \frac{4 \eta_{\Gamma}}{\sqrt{2}}e & \mbox{in } \Omega_2, \end{cases} \;\;\; \quad \quad \;\;\; \textbf{u} = \begin{cases} -e^{x+y} & \mbox{in }\Omega_1, \\ -e^{x+y} & \mbox{in } \Omega_2, \end{cases} \;\;\; \quad \quad \;\;\; p_{\Gamma} = e+ \frac{2 \eta_{\Gamma}}{\sqrt{2}}e. \end{equation*} |
It is easy to prove that \textbf{u} , p and p_{\Gamma} satisfy the coupling conditions (2.5a)-(2.5b) with \xi = 1 , \ell_\Gamma = 0.001 and \boldsymbol{\nu} = \boldsymbol{\nu}_\Gamma = \textbf{I} . Note that in this case f_{\Gamma} = 0 since the solution in the fracture is constant and \left[\kern-0.15em\left[ { \textbf{u} } \right]\kern-0.15em\right] = 0 .
Figure 2 shows three successive levels of refinements for the polygonal mesh employed in this set of experiments. In order to test the behaviour of the energy norm of the error, thus validating the h -convergence properties of our methods proved in Theorem 6.8, we compute the quantity {||p-p_h||_{1, \mathcal{T}_h} + ||p_{\Gamma}- p_{\Gamma, h}||_{1, \Gamma_h}} (Figure 3(a) and 3(b)). We also want to validate the results of Theorem 6.9 for the velocity, computing ||\textbf{u}-\textbf{u}_h||_{L^2(\Omega)} (Figure 3(e)). In addition, we test the behaviour of the L^2 -norm of the error for the primal unknowns, i.e., ||p-p_h||_{L^2(\Omega)} + ||p_{\Gamma}- p_{\Gamma, h}||_{L^2(\Gamma)} (Figure 3(c) and 3(d)). All the plots in Figure 3 show the computed errors as a function of the inverse of the mesh size h (loglog scale). Each plot consists of four lines: Every line shows the behaviour of the computed error for a different polynomial degree in the bulk (we consider uniform degree k_E = k = 1, 2, 3, 4 for all E \in \mathcal{T}_h ). For the fracture problem we always choose uniform quadratic polynomial degree, i.e., k_F = k_\Gamma = 2 for all F \in \Gamma_h . On the left we report the results obtained with the (MP) approximation scheme and on the right with the (PP) scheme. We observe that, for both methods, the convergence rates are superoptimal with respect to the expected rate of min(k, k_\Gamma) (they coincide with the bulk polynomial degree k ). This is probably due to the constant nature of the solution in the fracture. Indeed this behaviour will not be observed in the next set of experiments, cf. Section 7.2, where the solution in the fracture is trigonometric. Moreover, Figure 3(c) and 3(d)) show that one order of convergence is gained for the L^2 -norm. Finally, one can clearly see that the level of accuracy achieved by the (PP) and (MP) schemes is the same.
Next, we consider the domain \Omega = (0, 1)^2 and the fracture \Gamma = \{(x, y) \in \Omega: \; \; x = 0.5\} . We reproduce the numerical test already proposed in [59] for the primal-primal setting, choosing the exact solutions in the bulk and in the fracture as follows
\begin{equation*} p = \begin{cases} \sin(4x)\cos(\pi y)& \mbox{if } x \lt 0.5, \\ \cos(4x)\cos(\pi y) & \mbox{if } x \gt 0.5, \end{cases} \quad \quad \quad p_{\Gamma} = \xi[\cos(2) + \sin(2)] \cos(\pi y). \end{equation*} |
We impose Dirichlet boundary conditions on the whole \partial \Omega and also on \partial \Gamma . Notice that coupling conditions (2.5a)-(2.5b) are satisfied provided that we take \boldsymbol{\nu} = \textbf{I} , and \beta_\Gamma = 2 , that is \boldsymbol{\nu}^n_\Gamma / \ell_\Gamma = 4 . The source terms are then chosen accordingly as
\begin{equation*} f = \begin{cases} \sin(4x)\cos(\pi y)(16 + \pi^2) & \mbox{if }x \lt 0.5, \\\cos(4x)\cos(\pi y)(16+\pi^2) & \mbox{if } x \gt 0.5, \end{cases} \quad \quad f_{\Gamma} = \cos(\pi y)[\cos(2) + \sin(2)](\xi \boldsymbol{\nu}_{\Gamma}^{\tau} \pi^2 + \frac{4}{\ell_\Gamma}). \end{equation*} |
In the experiments, we set the components of the fracture permeability tensor \boldsymbol{\nu}_{\Gamma}^{\tau} = 10^2 and \nu_{\Gamma}^n = 4 \cdot 10^{-2} , we set the fracture thickness \ell_\Gamma = 10^{-2} and the closure parameter \xi = \frac{3}{4} . As before, in order to test the h -convergence properties of the primal-primal and mixed-primal schemes, we compute the quantity {||p-p_h||_{1, \mathcal{T}_h} + ||p_{\Gamma}- p_{\Gamma, h}||_{1, \Gamma_h}} . In Figure 4 we show the computed errors as a function of the inverse of the mesh size h (loglog scale), together with the expected convergence rates. As before, we fix the polynomial degree for the fracture problem k_\Gamma = 2 , and we vary the polynomial degree for the problem in the bulk taking k = 1, 2, 3, 4 . Figure 4(a) encloses the results obtained with the mixed-primal approximation, while Figure 4(b) with the primal-primal approximation. In each plot, the four lines describe the behaviour of the energy norm of the error for a different polynomial degree in the bulk. Notice that in this case, for both the (PP) and (MP) method, the theoretical convergence rates are clearly obtained, coinciding with \min(k, k_\Gamma) (no superconvergence is observed). In particular, the convergence rate is equal to 1 when the approximation in the bulk is linear, i.e., when k = 1 and it is equal to 2 in all the other cases, since we always choose a quadratic approximation for the problem in the fracture. Moreover, we remark that also in this case the (PP) and (MP) methods achieve the same level of accuracy.
Next, we reproduce some numerical experiments first presented in [47]. We examine two test cases with bulk domain \Omega = (0, 2) \times (0, 1) and fracture domain \Gamma = \{(x, y) \in \mathbb{R}^2: x = 1, \, 0\leq y \leq 1\} . In the first case, we consider a fracture with constant permeability, while in the second case we consider a fracture with lower permeability in its middle part, thus presenting a discontinuity. In particular:
(a) Case 1: constant permeability: The permeability tensor in the fracture is given by \boldsymbol{\nu}_{\Gamma}^n = \boldsymbol{\nu}_{\Gamma}^\tau = 100 . The bulk permeability \boldsymbol{\nu} is chosen to be constant and isotropic, i.e., \boldsymbol{\nu} = \textbf{I} . We impose Dirichlet boundary conditions on the left and right side of the bulk domain and homogeneous Neumann conditions on the top and bottom sides. On the fracture boundaries we impose Dirichlet boundary conditions.
(b) Case 2: discontinous permeabilty: The fracture \Gamma is subdivided into two areas having different values for the permeability tensor: In the initial and ending part of the fracture \Gamma_1 = \{ (x, y) \in \Gamma, \; 0\leq y \leq 0.25 \; \mbox{and} \; 0.75 \leq y \leq 1\} the permeability tensor \boldsymbol{\nu}_{\Gamma_1} is defined as \boldsymbol{\nu}_{\Gamma_1}^n = \boldsymbol{\nu}_{\Gamma_1}^\tau = 1 , while in the middle part \Gamma_2 = \{ (x, y) \in \Gamma, 0.25\leq y \leq 0.75 \} the permeability is low and is defined as \boldsymbol{\nu}_{\Gamma_2}^n = \boldsymbol{\nu}_{\Gamma_2}^\tau = 0.002 . The bulk permeability tensor is chosen again equal to the identity matrix, i.e., \boldsymbol{\nu} = \textbf{I} . In the bulk, we impose the same boundary conditions as in the previous test case, while at the fracture extremities we impose homogeneous Neumann conditions.
The two geometrical configurations are shown in Figures 5(a)-5(b), together with the boundary conditions. For both test cases we take the fracture thickness \ell_\Gamma = 0.01 and the model parameter \xi = 2/3 . Moreover, we discretize the problem in the bulk taking as polynomial degree k = 1 and the problem in the fracture taking k_\Gamma = 2 .
The obtained results are shown in Figure 6. For both cases (constant at the top, discontinuous at the bottom) we report the pressure field and Darcy velocity in the bulk (here the grid is very coarse only for visualization purposes) and the value of the pressure along the fracture. In the first case, since we have taken \boldsymbol{\nu}_{\Gamma}^n = \boldsymbol{\nu}_{\Gamma}^\tau = 100 > 1 , we can observe that the fluid has the tendency to flow along the fracture. In the second case, one can see that the part of the fracture with low (normal) permeability acts as a geological barrier, so that the fluid tends to avoid it and we can observe a jump of the bulk pressure across it. Our results are in agreement with those obtained in [47].
With this last set of numerical experiments we investigate the capability of our discretization method to deal with more complicated geometrical configurations, considering a network of partially immersed fractures. Our reference is, in this case, [51], where the mathematical model developed in [47] has been extended to fully immersed fractures. In [59] we showed that our method in a primal-primal setting is capable of efficiently handling the configuration. Here, we reproduce the same numerical experiments to demonstrate that this holds true also in a mixed-primal setting.
In order to deal with immersed fractures, we need to supplement our model (2.6) with an equation describing the behaviour of the fracture pressure at the immersed tips. Following [51], we impose a homogeneous Neumann condition, thus assuming that the mass transfer across the immersed tips can be neglected, i.e., \boldsymbol{\nu}_\Gamma^\tau \nabla_\tau p_\Gamma \cdot \boldsymbol{\tau} = 0 on \partial \Gamma . At the extremities of the fractures that are non-immersed, i.e., \partial \Gamma \cap \partial \Omega , we impose boundary conditions that are consistent with those imposed on \partial\Omega in that point.
We consider the bulk domain \Omega = [0, 1]^2 cut by a network made of four partially immersed fractures: \Gamma_1 = \{(x, y) \in [0, 1]^2: x \geq 0.3, y = 0.2 \} , \Gamma_2 = \{(x, y) \in [0, 1]^2: x\leq 0.7, y = 0.4 \} , \Gamma_3 = \{(x, y) \in [0, 1]^2: x\geq 0.3, y = 0.6 \} and \Gamma_4 = \{(x, y) \in [0, 1]^2: x\leq 0.7, y = 0.8 \} . We perform two numerical experiments. In both of them, the fractures \Gamma_2 and \Gamma_4 are impermeable ( \boldsymbol{\nu}_\Gamma^\tau = \nu_\Gamma^n = 10^{-2} ), while \Gamma_1 and \Gamma_3 are partially permeable. In the first configuration, we consider for \Gamma_1 and \Gamma_3 the permeabilities \nu_\Gamma^n = 10^{-2} and \boldsymbol{\nu}_\Gamma^\tau = 100 , while in the second, we consider \nu_\Gamma^n = 10^{-2} and \boldsymbol{\nu}_\Gamma^\tau = 1 . Moreover, we vary the imposed boundary conditions as illustrated in Figure 7.
In both the experiments we consider an isotropic bulk permeability tensor i.e., \boldsymbol{\nu} = \textbf{I} and we assume that all the fractures have aperture \ell_\Gamma = 0.01 . The flow is only generated by boundary conditions, since we take all the forcing terms f = f_\Gamma = 0 . Finally, we choose as model parameter \xi = 0.55 .
To obtain our results, we employed cartesian grids featuring approximately the same number of elements as those employed in [51] and such that the immersed tips of the fractures coincide with one of the mesh vertices. For the approximation of the problem in the bulk and in the fracture we chose the polynomial degrees k = k_\Gamma = 2 . In Figure 8, we show the results obtained for the two test cases with a mesh of 26 051 elements. In particular, we report the pressure field in the bulk with the streamlines of the velocity (left), the value of the bulk pressure along the line x = 0.65 (middle) and the pressure field inside the four fractures (right). Our results are in perfect agreement with those obtained in [51] and in [59], thus showing that, also in a mixed-primal setting, our method is able to efficiently handle this configuration.
In this paper we have proposed a formulation based on discontinuous Galerkin methods on polygonal/polyhedral grids for the simulation of flows in fractured porous media. In particular, we have designed and analysed, in the unified framework of [63] based on the flux-formulation, a polyDG approximation for all the possible combinations of primal-primal, mixed-primal, primal-mixed and mixed-mixed formulations for the bulk and fracture problems, respectively. The novelty of our method relies on the imposition of coupling conditions between bulk and fracture through a suitable definition of the numerical fluxes on the fracture faces. We have proved in an unified setting the well-posedness of all the formulations and we have derived a priori hp -version error estimates in a suitable (mesh-dependent) energy norm, whose validity has been assessed performing some preliminary numerical experiments, focusing on the paradigmatic primal-primal and mixed-primal methods. In our test cases we have also compared, in a simplified setting, the performance of our approximation schemes. In particular, we have shown that the same level of accuracy may be obtained irrespective of the chosen method. The other the factors that should be taken into account when choosing which one between the (PP), (MP), (PM), and (MM) setting to employ, are summarized in the following.
● The desired accuracy in the approximation of the physical quantities (pressure and velocity) according to the application at hand. The primal approach considers a single-field formulation with the pressure field of the fluid as only unknown, so that velocity may only be afterwards reconstructed taking the gradient of the computed pressure and multiplying it by the permeability tensor. This process usually entails a loss of accuracy and does not guarantee mass conservation, see for example [61,62], so it may not be suitable for those Engineering applications (such as oil recovery or groundwater pollution modeling) where the simulation of the phenomenon requires very accurate approximation of the velocities of the involved fluids in order to be effective. In such cases, the mixed approach should be preferred, so that Darcy's velocity can be directly computed and a higher degree of accuracy can be achieved, together with local and global mass conservation.
● The numerical linear algebra resulting from the discretization. The primal setting has the advantage of featuring less degrees of freedom and leads to a symmetric positive definite algebraic system of equations, which can be efficiently solved based on employing, for example, multigrid techniques [39,44,60]. The mixed approach leads to a so-called generalized saddle point system of equations. For example, in the MP setting, problem (4.26) translates into an algebraic system of the form
\begin{equation} \begin{bmatrix} M & B & 0\\ -B^T & S +C_1 & C_2\\ 0 &C_2^T &A_\Gamma + A_\Gamma^R \end{bmatrix}, \end{equation} |
where, referring to Eq. (4.27):
- M is the mass matrix related to the bilinear form \mathcal{M}_b ;
- B is related to the bilinear form \mathcal{B}_b ;
- S is related to the bilinear form \mathcal{S}_b ;
- C_1 is related to the terms involving only bulk unknowns contained in the interface bilinear forms \mathcal{I}_1 and \mathcal{I}_2 ;
- C_2 is related to the terms involving both bulk and fracture unknowns in \mathcal{I}_1 and \mathcal{I}_2 ;
- A_\Gamma is related to the fracture primal bilinear form \mathcal{A}_{\Gamma}^{ P} ;
- A_\Gamma^R is related to the terms involving only fracture unknowns contained in the interface bilinear forms \mathcal{I}_1 and \mathcal{I}_2 , which result in a reaction term for the fracture problem.
In this case the Schur complement of M , i.e. the matrix D+B^TM^{-1}B , with D = \begin{bmatrix} S +C_1 & C_2\\ C_2^T & A_\Gamma + A_\Gamma^R \end{bmatrix} , can be computed explicitly, thanks to the fact that, in the DG setting, M either has a block-diagonal or a diagonal structure, depending on whether a modal or nodal expansion is chosen to span the discrete space. We also point out that the mixed case involves a (three-time) larger number of unknowns, so that efficient memory handling may be needed.
● The conditioning of the resulting systems. The behaviour of the condition number and the possibility of building efficient preconditioners is certainly of fundamental importance in the choice between the primal or mixed approach, especially for three-dimensional problems. We did not address this issue in the present work. However, one may refer to [76], where preconditioners for the system stemming from the discretization of problem (2.6) in mixed-mixed form, employing mimetic finite differences, are constructed. In the DG setting, for the hp -preconditioning of systems stemming from the discretization of diffusion problems, the reader may refer to [71] for standard grids and to [39,44,78] for the extension to the polytopic setting.
Finally, we mention that our method may be extended in order to deal with network of intersecting fractures. To this aim, the mathematical model needs to be complemented with some suitable physical conditions at the intersections, prescribing the behaviour of the fluid. One possible choice is to impose pressure continuity and flux conservation as in [52,74]. From the DG-discretization point of view, the key point to deal with this case is the generalization of the concepts of jump and average at the intersections. We refer to [75] for a rigorous analysis of the method in the primal-primal setting and to [59] for a preliminary numerical test case involving a totally immersed network of fractures.
Paola F. Antonietti and Chiara Facciolà have been partially supported by SIR Starting grant n. RBSI14VT0S "PolyPDEs: Non-conforming polyhedral finite element methods for the approximation of partial differential equations" funded by MIUR. Paola F. Antonietti and Marco Verani have been partially funded by the Italian research project PRIN n. 201744KLJL. All the authors have been partially supported by INdAM-GNCS.
The authors declare no conflict of interest.
[1] |
Hackbusch W, Sauter SA (1997) Composite finite elements for the approximation of PDEs on domains with complicated micro-structures. Numer Math 75: 447-472. doi: 10.1007/s002110050248
![]() |
[2] | Hackbusch W, Sauter SA (1997) Composite finite elements for problems containing small geometric details. Part II: Implementation and numerical results. Comput Visual Sci 1: 15-25. |
[3] |
Hyman J, Shashkov M, Steinberg S (1997) The numerical solution of diffusion problems in strongly heterogeneous non-isotropic materials. J Comput Phys 132: 130-148. doi: 10.1006/jcph.1996.5633
![]() |
[4] |
Brezzi F, Lipnikov K, Simoncini V (2005) A family of Mimetic finite difference methods on polygonal and polyhedral meshes. Math Mod Meth Appl S 15: 1533-1551. doi: 10.1142/S0218202505000832
![]() |
[5] |
Brezzi F, Lipnikov K, Shashkov M (2005) Convergence of the mimetic finite difference method for diffusion problems on polyhedral meshes. SIAM J Numer Anal 43: 1872-1896. doi: 10.1137/040613950
![]() |
[6] |
Brezzi F, Lipnikov K, Shashkov M (2006) Convergence of mimetic finite difference method for diffusion problems on polyhedral meshes with curved faces. Math Mod Meth Appl S 16: 275-297. doi: 10.1142/S0218202506001157
![]() |
[7] | Beirão da Veiga L, Lipnikov K, Manzini G (2014) The Mimetic Finite Difference Method for Elliptic Problems, Cham: Springer. |
[8] |
Antonietti PF, Formaggia L, Scotti A, et al. (2016) Mimetic finite difference approximation of flows in fractured porous media. ESAIM Math Model Numer Anal 50: 809-832. doi: 10.1051/m2an/2015087
![]() |
[9] |
Sukumar N, Tabarraei A (2004) Conforming Polygonal Finite Elements. Int J Numer Meth Eng 61: 2045-2066. doi: 10.1002/nme.1141
![]() |
[10] |
Dolbow J, Moës N, Belytschko T (2001) An extended finite element method for modeling crack growth with frictional contact. Comput Method Appl M 190: 6825-6846. doi: 10.1016/S0045-7825(01)00260-2
![]() |
[11] |
Tabarraei A, Sukumar N (2008) Extended finite element method on polygonal and quadtree meshes. Comput Method Appl M 197: 425-438. doi: 10.1016/j.cma.2007.08.013
![]() |
[12] | Fries TP, Belytschko T (2010) The extended/generalized finite element method: An overview of the method and its applications. Int J Numer Meth Eng 84: 253-304. |
[13] |
Beirão da Veiga L, Brezzi F, Cangiani A, et al. (2013) Basic principles of virtual element methods. Math Mod Meth Appl S 23: 199-214. doi: 10.1142/S0218202512500492
![]() |
[14] |
Beirão da Veiga L, Brezzi F, Marini L, et al. (2016) Virtual element method for general second-order elliptic problems on polygonal meshes. Math Mod Meth Appl S 26: 729-750. doi: 10.1142/S0218202516500160
![]() |
[15] |
Antonietti PF, da Veiga LB, Mora D, et al. (2014) A stream virtual element formulation of the Stokes problem on polygonal meshes. SIAM J Numer Anal 52: 386-404. doi: 10.1137/13091141X
![]() |
[16] |
Beirão da Veiga L, Brezzi F, Marini LD, et al. (2016) Mixed virtual element methods for general second order elliptic problems on polygonal meshes. ESAIM Math Model Numer Anal 50: 727-747. doi: 10.1051/m2an/2015067
![]() |
[17] |
Antonietti PF, da Veiga LB, Scacchi S, et al. (2016) A C1 virtual element method for the CahnHilliard equation with polygonal meshes. SIAM J Numer Anal 54: 34-56. doi: 10.1137/15M1008117
![]() |
[18] | Di Pietro DA, Ern A, Lemaire S (2014) An arbitrary-order and compact-stencil discretization of diffusion on general meshes based on local reconstruction operators. Comput Meth Appl Math 14: 461-472. |
[19] |
Di Pietro DA, Ern A (2015) A hybrid high-order locking-free method for linear elasticity on general meshes. Comput Meth Appl Mech Eng 283: 1-21. doi: 10.1016/j.cma.2014.09.009
![]() |
[20] | Di Pietro DA, Ern A (2015) Hybrid high-order methods for variable-diffusion problems on general meshes. CR Math 353: 31-34. |
[21] | Di Pietro DA, Ern A, Lemaire S (2016) A review of hybrid high-order methods: Formulations, computational aspects, comparison with other methods. In: Barrenechea GR, Brezzi F, Cangiani A, et al. Editors, Building Bridges: Connections and Challenges in Modern Approaches to Numerical Partial Differential Equations, Cham: Springer. |
[22] |
Cockburn B, Dond B, Guzmán J (2008) A superconvergent LDG-hybridizable Galerkin method for second-order elliptic problems. Math Comput 77: 1887-1916. doi: 10.1090/S0025-5718-08-02123-6
![]() |
[23] |
Cockburn B, Gopalakrishnan J, Lazarov R (2009) Unified hybridization of discontinuous Galerkin, mixed, and continuous Galerkin methods for second order elliptic problems. SIAM J Numer Anal 47: 1319-1365. doi: 10.1137/070706616
![]() |
[24] |
Cockburn B, Gopalakrishnan J, Sayas FJ (2010) A projection-based error analysis of HDG methods. Math Comput 79: 1351-1367. doi: 10.1090/S0025-5718-10-02334-3
![]() |
[25] |
Cockburn B, Guzmán J, Wang H (2009) Superconvergent discontinuous Galerkin methods for second-order elliptic problems. Math Comput 78: 1-24. doi: 10.1090/S0025-5718-08-02146-7
![]() |
[26] |
Antonietti PF, Giani S, Houston P (2013) hp-version composite discontinuous Galerkin methods for elliptic problems on complicated domains. SIAM J Sci Comput 35: A1417-A1439. doi: 10.1137/120877246
![]() |
[27] |
Antonietti PF, Giani S, Houston P (2014) Domain decomposition preconditioners for discontinuous Galerkin methods for elliptic problems on complicated domains. J Sci Comput 60: 203-227. doi: 10.1007/s10915-013-9792-y
![]() |
[28] |
Antonietti PF, Manzini G, Verani M (2018) The fully nonconforming virtual element method for biharmonic problems. Math Mod Meth Appl S 28: 387-407. doi: 10.1142/S0218202518500100
![]() |
[29] |
Ayuso de Dios B, Lipnikov K, Manzini G (2016) The nonconforming virtual element method. ESAIM Math Model Numer Anal 50: 879-904. doi: 10.1051/m2an/2015090
![]() |
[30] | Cangiani A, Manzini G, Sutton OJ (2017) Conforming and nonconforming virtual element methods for elliptic problems. IMA J Numer Anal 37: 1317-1354. |
[31] |
Droniou J, Eymard R, Herbin R (2016) Gradient schemes: Generic tools for the numerical analysis of diffusion equations. ESAIM Math Model Numer Anal 50: 749-781. doi: 10.1051/m2an/2015079
![]() |
[32] |
Antonietti PF, Brezzi F, Marini LD (2009) Bubble stabilization of discontinuous Galerkin methods. Comput Meth Appl Mech Eng 198: 1651-1659. doi: 10.1016/j.cma.2008.12.033
![]() |
[33] |
Bassi F, Botti L, Colombo A, et al. (2012) On the flexibility of agglomeration based physical space discontinuous Galerkin discretizations. J Comput Phys 231: 45-65. doi: 10.1016/j.jcp.2011.08.018
![]() |
[34] |
Bassi F, Botti L, Colombo A, et al. (2012) Agglomeration based discontinuous Galerkin discretization of the Euler and Navier-Stokes equations. Comput Fluids 61: 77-85. doi: 10.1016/j.compfluid.2011.11.002
![]() |
[35] |
Bassi F, Botti L, Colombo A (2014) Agglomeration-based physical frame dG discretizations: An attempt to be mesh free. Math Mod Meth Appl S 24: 1495-1539. doi: 10.1142/S0218202514400028
![]() |
[36] |
Cangiani A, Georgoulis EH, Houston P (2014) hp-version discontinuous Galerkin methods on polygonal and polyhedral meshes. Math Mod Meth Appl S 24: 2009-2041. doi: 10.1142/S0218202514500146
![]() |
[37] |
Cangiani A, Dong Z, Georgoulis EH, et al. (2016) hp-version discontinuous Galerkin methods for advection-diffusion-reaction problems on polytopic meshes. ESAIM Math Model Numer Anal 50: 699-725. doi: 10.1051/m2an/2015059
![]() |
[38] |
Cangiani A, Dong Z, Georgoulis EH (2017) hp-version space-time discontinuous Galerkin methods for parabolic problems on prismatic meshes. SIAM J Sci Comput 39: A1251-A1279. doi: 10.1137/16M1073285
![]() |
[39] |
Antonietti PF, Houston P, Hu X, et al. (2017) Multigrid algorithms for hp-version interior penalty discontinuous Galerkin methods on polygonal and polyhedral meshes. Calcolo 54: 1169-1198. doi: 10.1007/s10092-017-0223-6
![]() |
[40] | Antonietti PF, Cangiani A, Collis J, et al. (2016) Review of discontinuous Galerkin finite element methods for partial differential equations on complicated domains, In: Lecture Notes in Computational Science and Engineering, Springer, 281-310. |
[41] | Cangiani A, Dong Z, Georgoulis EH, et al. (2017) hp-Version Discontinuous Galerkin Methods on Polytopic Meshes, Springer International Publishing. |
[42] |
Antonietti PF, Mazzieri I (2018) High-order discontinuous Galerkin methods for the elastodynamics equation on polygonal and polyhedral meshes. Comput Meth Appl Mech Eng 342: 414-437. doi: 10.1016/j.cma.2018.08.012
![]() |
[43] | Antonietti PF, Bonaldi F, Mazzieri I (2018) A high-order discontinuous galerkin approach to the elasto-acoustic problem. arXiv preprint arXiv:180301351. |
[44] |
Antonietti PF, Pennesi G (2019) V-cycle multigrid algorithms for discontinuous Galerkin methods on non-nested polytopic meshes. J Sci Comput 78: 625-652. doi: 10.1007/s10915-018-0783-x
![]() |
[45] | Alboin C, Jaffré J, Roberts JE, et al. (2000) Domain decomposition for some transmission problems in flow in porous media. In: Numerical Treatment of Multiphase Flows in Porous Media, Berlin: Springer, 22-34. |
[46] | Alboin C, Jaffré J, Roberts JE, et al. (2002) Modeling fractures as interfaces for flow and transport in porous media. In: Fluid Flow and Transport in Porous Media: Mathematical and Numerical Treatment (South Hadley, MA, 2001), Providence: Amer. Math. Soc., 13-24. |
[47] |
Martin V, Jaffré J, Roberts JE (2005) Modeling fractures and barriers as interfaces for flow in porous media. SIAM J Sci Comput 26: 1667-1691. doi: 10.1137/S1064827503429363
![]() |
[48] |
Frih N, Roberts JE, Saada A (2008) Modeling fractures as interfaces: A model for Forchheimer fractures. Comput Geosci 12: 91-104. doi: 10.1007/s10596-007-9062-x
![]() |
[49] |
Fumagalli A, Scotti A (2013) A numerical method for two-phase flow in fractured porous media with non-matching grids. Adv Water Resour 62: 454-464. doi: 10.1016/j.advwatres.2013.04.001
![]() |
[50] |
Jaffré J, Mnejja M, Roberts J (2011) A discrete fracture model for two-phase flow with matrix-fracture interaction. Procedia Comput Sci 4: 967-973. doi: 10.1016/j.procs.2011.04.102
![]() |
[51] |
Angot P, Boyer F, Hubert F (2009) Asymptotic and numerical modelling of flows in fractured porous media. M2AN Math Model Numer Anal 43: 239-275. doi: 10.1051/m2an/2008052
![]() |
[52] |
Formaggia L, Scotti A, Sottocasa F (2018) Analysis of a mimetic finite difference approximation of flows in fractured porous media. ESAIM Math Model Numer Anal 52: 595-630. doi: 10.1051/m2an/2017028
![]() |
[53] |
Benedetto MF, Berrone S, Pieraccini S, et al. (2014) The virtual element method for discrete fracture network simulations. Comput Meths Appl Mech Eng 280: 135-156. doi: 10.1016/j.cma.2014.07.016
![]() |
[54] |
Benedetto MF, Berrone S, Scialò S (2016) A globally conforming method for solving flow in discrete fracture networks using the virtual element method. Finite Elem Anal Des 109: 23-36. doi: 10.1016/j.finel.2015.10.003
![]() |
[55] |
Chave FA, Di Pietro D, Formaggia L (2018) A hybrid high-order method for Darcy flows in fractured porous media. SIAM J Sci Comput 40: A1063-A1094. doi: 10.1137/17M1119500
![]() |
[56] |
D'Angelo C, Scotti A (2012) A mixed finite element method for darcy flow in fractured porous media with non-matching grids. ESAIM Math Mod Numer Anal 46: 465-489. doi: 10.1051/m2an/2011148
![]() |
[57] | Flemisch B, Fumagalli A, Scotti A (2016) A review of the XFEM-based approximation of flow in fractured porous media. In: Advances in Discretization Methods, Springer, 47-76. |
[58] |
Burman E, Hansbo P, Larson MG, et al. (2019) Cut finite elements for convection in fractured domains. Comput Fluids 179: 726-734. doi: 10.1016/j.compfluid.2018.07.022
![]() |
[59] |
Antonietti PF, Facciolà C, Russo A, et al. (2019) Discontinuous Galerkin approximation of flows in fractured porous media on polytopic grids. SIAM J Sci Comput 41: A109-A138. doi: 10.1137/17M1138194
![]() |
[60] |
Antonietti PF, Sarti M, Verani M (2015) Multigrid algorithms for hp-discontinuous Galerkin discretizations of elliptic problems. SIAM J Numer Anal 53: 598-618. doi: 10.1137/130947015
![]() |
[61] |
Masud A, Hughes TJ (2002) A stabilized mixed finite element method for darcy flow. Comput Meth Appl Mech Eng 191: 4341-4370. doi: 10.1016/S0045-7825(02)00371-7
![]() |
[62] | Brezzi F, Hughes TJ, Marini LD, et al. (2005) Mixed discontinuous Galerkin methods for darcy flow. J Sci Comput 22: 119-145. |
[63] | Arnold DN, Brezzi F, Cockburn B, et al. (2001/02) Unified analysis of discontinuous Galerkin methods for elliptic problems. SIAM J Numer Anal 39: 1749-1779. |
[64] |
Wheeler MF (1978) An elliptic collocation-finite element method with interior penalties. SIAM J Numer Anal 15: 152-161. doi: 10.1137/0715010
![]() |
[65] |
Arnold DN (1982) An interior penalty finite element method with discontinuous elements. SIAM J Numer Anal 19: 742-760. doi: 10.1137/0719052
![]() |
[66] |
Cockburn B, Shu CW (1998) The local discontinuous Galerkin method for time-dependent convection-diffusion systems. SIAM J Numer Anal 35: 2440-2463. doi: 10.1137/S0036142997316712
![]() |
[67] |
Castillo P, Cockburn B, Perugia I, et al. (2000) An a priori error analysis of the local discontinuous Galerkin method for elliptic problems. SIAM J Numer Anal 38: 1676-1706. doi: 10.1137/S0036142900371003
![]() |
[68] | Perugia I, Schötzau D (2003) The hp-local discontinuous galerkin method for low-frequency time-harmonic maxwell equations. Math Comput 72: 1179-1214. |
[69] |
Perugia I, Schötzau D (2002) An hp-analysis of the local discontinuous galerkin method for diffusion problems. J Sci Comput 17: 561-571. doi: 10.1023/A:1015118613130
![]() |
[70] | Strang G, Fix GJ (1973) An analysis of the finite element method, Prentice-Hall Englewood Cliffs. |
[71] |
Antonietti PF, Houston P (2011) A class of domain decomposition preconditioners for hpdiscontinuous Galerkin finite element methods. J Sci Comput 46: 124-149. doi: 10.1007/s10915-010-9390-1
![]() |
[72] | Stein EM (1970) Singular integrals and differentiability properties of functions, Princeton university press. |
[73] |
Talischi C, Paulino GH, Pereira A, et al. (2012) Polymesher: A general-purpose mesh generator for polygonal elements written in matlab. Struct Multidiscip O 45: 309-328. doi: 10.1007/s00158-011-0706-z
![]() |
[74] | Brenner K, Hennicker J, Masson R, et al. (2016) Gradient discretization of hybrid-dimensional Darcy flow in fractured porous media with discontinuous pressures at matrix-fracture interfaces. IMA J Numer Anal 37: 1551-1585. |
[75] | Antonietti PF, Facciolà C, Verani M (2020) Polytopic discontinuous Galerkin methods for the numerical modelling of flow in porous media with networks of intersecting fractures. MOX Rep 08, arXiv:2002.06420. |
[76] | Antonietti PF, De Ponti J, Formaggia L, et al. (2019) Preconditioning techniques for the numerical solution of flow in fractured porous media. MOX Rep 17. |
[77] | Antonietti PF, Houston P (2013) Preconditioning high-order discontinuous galerkin discretizations of elliptic problems. In: Domain Decomposition Methods in Science and Engineering XX, Springer, 231-238. |
[78] | Antonietti PF, Houston P, Pennesi G, et al. (2020) An agglomeration-based massively parallel non-overlapping additive Schwarz preconditioner for high-order discontinuous Galerkin methods on polytopic grids. Math Comput DOI: https://doi.org/10.1090/mcom/3510. |
1. | Paola F. Antonietti, Jacopo De Ponti, Luca Formaggia, Anna Scotti, Preconditioning Techniques for the Numerical Solution of Flow in Fractured Porous Media, 2021, 86, 0885-7474, 10.1007/s10915-020-01372-0 | |
2. | Chongbin Zhao, Bruce Hobbs, Alison Ord, An accurate porosity‐velocity‐concentration approach for solving reactive mass transport problems involving chemical dissolution in fluid‐saturated porous media with arbitrarily initial porosity distributions, 2021, 122, 0029-5981, 7354, 10.1002/nme.6833 | |
3. | Igor Mozolevski, Marcio A. Murad, Luciane A. Schuh, High order discontinuous Galerkin method for reduced flow models in fractured porous media, 2021, 190, 03784754, 1317, 10.1016/j.matcom.2021.07.012 | |
4. | Paola F. Antonietti, Michele Botti, Ilario Mazzieri, On Mathematical and Numerical Modelling of Multiphysics Wave Propagation with Polytopal Discontinuous Galerkin Methods: a Review, 2022, 50, 2305-221X, 997, 10.1007/s10013-022-00566-3 | |
5. | Paola F. Antonietti, Chiara Facciolà, Paul Houston, Ilario Mazzieri, Giorgio Pennesi, Marco Verani, 2021, Chapter 5, 978-3-030-69362-6, 159, 10.1007/978-3-030-69363-3_5 | |
6. | Jiajing Li, Guang Fu, Zhe Liu, Quantitative Characterization of the Lower Limit of the Physical Properties of Tight Oil Reservoirs in Nano-Material Porous Media, 2022, 227, 1058-4587, 288, 10.1080/10584587.2022.2072118 | |
7. | Paola F. Antonietti, Chiara Facciolà, Marco Verani, Polytopic discontinuous Galerkin methods for the numerical modelling of flow in porous media with networks of intersecting fractures, 2022, 116, 08981221, 116, 10.1016/j.camwa.2021.08.015 | |
8. | Rui Li, Yali Gao, Zhangxin Chen, Adaptive discontinuous Galerkin finite element methods for the Allen-Cahn equation on polygonal meshes, 2024, 95, 1017-1398, 1981, 10.1007/s11075-023-01635-5 | |
9. | Shuangshuang Chen, Discontinuous Galerkin method for hybrid-dimensional fracture models of two-phase flow, 2023, 488, 00219991, 112244, 10.1016/j.jcp.2023.112244 | |
10. | Marcio A. Murad, Luciane A. Schuh, Igor Mozolevski, Josue Barroso, An extended two-parameter mixed-dimensional model of fractured porous media incorporating entrance flow and boundary-layer transition effects, 2024, 193, 03091708, 104838, 10.1016/j.advwatres.2024.104838 |
Method | Primal bilinear form |
Primal-Primal (PP) | \mathcal{A}_b^P(p, q) + \mathcal{A}_\Gamma^P(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |
Mixed-Primal (MP) | \mathcal{A}_b^M(p, q) + \mathcal{A}_\Gamma^P(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |
Primal-Mixed (PM) | \mathcal{A}_b^P(p, q) + \mathcal{A}_\Gamma^M(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |
Mixed-Mixed (MM) | \mathcal{A}_b^M(p, q) + \mathcal{A}_\Gamma^M(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |
Method | Primal bilinear form |
Primal-Primal (PP) | \mathcal{A}_b^P(p, q) + \mathcal{A}_\Gamma^P(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |
Mixed-Primal (MP) | \mathcal{A}_b^M(p, q) + \mathcal{A}_\Gamma^P(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |
Primal-Mixed (PM) | \mathcal{A}_b^P(p, q) + \mathcal{A}_\Gamma^M(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |
Mixed-Mixed (MM) | \mathcal{A}_b^M(p, q) + \mathcal{A}_\Gamma^M(p_\Gamma, q_\Gamma) + \mathcal{I}((p, q), (p_\Gamma, q_\Gamma)) |