
We consider single-phase flow in a fractured porous medium governed by Darcy's law with spatially varying hydraulic conductivity matrices in both bulk and fractures. The width-to-length ratio of a fracture is of the order of a small parameter ε and the ratio Kf⋆/Kb⋆ of the characteristic hydraulic conductivities in the fracture and bulk domains is assumed to scale with εα for a parameter α∈R. The fracture geometry is parameterized by aperture functions on a submanifold of codimension one. Given a fracture, we derive the limit models as ε→0. Depending on the value of α, we obtain five different limit models as ε→0, for which we present rigorous convergence results.
Citation: Maximilian Hörl, Christian Rohde. Rigorous derivation of discrete fracture models for Darcy flow in the limit of vanishing aperture[J]. Networks and Heterogeneous Media, 2024, 19(1): 114-156. doi: 10.3934/nhm.2024006
[1] | Alexei Heintz, Andrey Piatnitski . Osmosis for non-electrolyte solvents in permeable periodic porous media. Networks and Heterogeneous Media, 2016, 11(3): 471-499. doi: 10.3934/nhm.2016005 |
[2] | Mohamed Belhadj, Eric Cancès, Jean-Frédéric Gerbeau, Andro Mikelić . Homogenization approach to filtration through a fibrous medium. Networks and Heterogeneous Media, 2007, 2(3): 529-550. doi: 10.3934/nhm.2007.2.529 |
[3] | Giuseppe Maria Coclite, Lorenzo di Ruvo, Jan Ernest, Siddhartha Mishra . Convergence of vanishing capillarity approximations for scalar conservation laws with discontinuous fluxes. Networks and Heterogeneous Media, 2013, 8(4): 969-984. doi: 10.3934/nhm.2013.8.969 |
[4] | Karoline Disser, Matthias Liero . On gradient structures for Markov chains and the passage to Wasserstein gradient flows. Networks and Heterogeneous Media, 2015, 10(2): 233-253. doi: 10.3934/nhm.2015.10.233 |
[5] | Clément Cancès . On the effects of discontinuous capillarities for immiscible two-phase flows in porous media made of several rock-types. Networks and Heterogeneous Media, 2010, 5(3): 635-647. doi: 10.3934/nhm.2010.5.635 |
[6] | Jinhu Zhao . Natural convection flow and heat transfer of generalized Maxwell fluid with distributed order time fractional derivatives embedded in the porous medium. Networks and Heterogeneous Media, 2024, 19(2): 753-770. doi: 10.3934/nhm.2024034 |
[7] | Pedro Aceves-Sanchez, Benjamin Aymard, Diane Peurichard, Pol Kennel, Anne Lorsignol, Franck Plouraboué, Louis Casteilla, Pierre Degond . A new model for the emergence of blood capillary networks. Networks and Heterogeneous Media, 2021, 16(1): 91-138. doi: 10.3934/nhm.2021001 |
[8] | Leda Bucciantini, Angiolo Farina, Antonio Fasano . Flows in porous media with erosion of the solid matrix. Networks and Heterogeneous Media, 2010, 5(1): 63-95. doi: 10.3934/nhm.2010.5.63 |
[9] | Yangyang Qiao, Huanyao Wen, Steinar Evje . Compressible and viscous two-phase flow in porous media based on mixture theory formulation. Networks and Heterogeneous Media, 2019, 14(3): 489-536. doi: 10.3934/nhm.2019020 |
[10] | Brahim Amaziane, Leonid Pankratov, Andrey Piatnitski . An improved homogenization result for immiscible compressible two-phase flow in porous media. Networks and Heterogeneous Media, 2017, 12(1): 147-171. doi: 10.3934/nhm.2017006 |
We consider single-phase flow in a fractured porous medium governed by Darcy's law with spatially varying hydraulic conductivity matrices in both bulk and fractures. The width-to-length ratio of a fracture is of the order of a small parameter ε and the ratio Kf⋆/Kb⋆ of the characteristic hydraulic conductivities in the fracture and bulk domains is assumed to scale with εα for a parameter α∈R. The fracture geometry is parameterized by aperture functions on a submanifold of codimension one. Given a fracture, we derive the limit models as ε→0. Depending on the value of α, we obtain five different limit models as ε→0, for which we present rigorous convergence results.
Porous media with fractures or other thin heterogeneities, such as membranes, occur in a wide range of applications in nature and industry including carbon sequestration, groundwater flow, geothermal engineering, oil recovery, and biomedicine. Fractures are characterized by an extreme geometry with a small aperture but a significantly larger longitudinal extent, typically by several orders of magnitude. Therefore, it is often computationally unfeasible to represent fractures explicitly in full-dimensional numerical methods, especially in the case of fracture networks, as this results in thin equi-dimensional domains that require a high resolution. However, the presence of fractures can have a crucial impact on the flow profile in a porous medium with the fractures acting either as major conduits or as barriers. Moreover, in order to obtain accurate predictions for the flow profile, generally, one also has to take into account the geometry of fractures, i.e., curvature and spatially varying aperture [1,2].
In the following paragraph, we provide a brief overview on modeling approaches for flow in fractured porous media with a focus on discrete fracture models. For details on modeling and discretization strategies, we refer to the review article [3] and the references therein. Conceptually, one can distinguish between models with an explicit representation of fractures and models that represent fractures implicitly by an effective continuum. For the latter category, there is a distinction between equivalent porous medium models [4,5], where fractures are modeled by modifying the permeability of the underlying porous medium, and multi-continuum models [6,7], where the fractured porous medium is represented by multiple superimposed interacting continua---in the simplest case by a fracture continuum and a matrix continuum. In contrast, discrete fracture models represent fractures explicitly as interfaces of codimension one within a porous medium. In comparison with implicit models, there is an increase in geometrical complexity but no upscaled description based on effective quantities. Besides, there are also hybrid approaches for fracture networks, where only dominant fractures are represented explicitly [8,9]. The most popular method for the derivation of discrete fracture models is vertical averaging [10,11,12,13,14,15,16,17], where the governing equations inside the fracture are integrated in normal direction. This leads to a description based on averaged fracture quantities on an interface of codimension one. However, the integration in normal direction provides no relation between the resulting interfacial model and the bulk flow model. Thus, using this approach, the resulting mixed-dimensional model is typically closed by making assumptions on the flow profile inside the fracture, which eventually renders the model derivation formal. Most commonly, averaged discrete fracture models are based on the conception of a planar fracture geometry with constant aperture. However, there are also works that consider curved fractures and fractures with spatially varying aperture [1,18]. Moreover, there are papers that take a fully mathematically rigorous approach for the derivation of discrete fracture models by applying weak compactness arguments to prove (weak) convergence towards a mixed-dimensional model in the limit of vanishing aperture [19,20,21,22,23,24,25]. This is also the approach that we follow here. In this case, in contrast to the method of vertical averaging, the width-to-length ratio of a fracture serves as a scaling parameter ε and one has to specify how the model parameters, such as the hydraulic conductivity, scale with respect to ε in the limit ε→0. Depending on their scaling, one can identify different regimes with fundamentally different limit problems as ε→0. Similar to the idea of homogenization theory, in the first place, this approach provides insight on the behavior of the system in the limit of vanishing width-to-length ratio ε→0 but the resulting limit models can be also be viewed as a computationally efficient approximation for real fractures with positive width-to-length ratio 0<ε≪1. Further, we mention [26,27], where formal asymptotic expansions are employed to obtain limit models for the Richards equation and two-phase Darcy flow in the limit of vanishing aperture, and [28], where a rigorous asymptotic approximation is presented for a convection-diffusion problem in a thin graph-like network. Besides, rigorous error estimates for classical solutions of discrete fracture models are obtained in [29,30]. In particular, in [30], an asymptotic expansion based on a Fourier transform is used to obtain the reduced model for one particular scaling of the fracture hydraulic conductivity with respect to the fracture aperture. Further, the authors in [31] have developed a mixed-dimensional functional analysis, which is utilized in [32] to obtain a poromechanical discrete fracture model using a formal "top-down" approach. In addition, we also mention phase-field models [33], which are convenient to track the propagation of fractures and can be combined with discrete fracture models [34].
In this paper, we consider single-phase fluid flow in a porous medium with an isolated fracture. Here, the term fracture refers to a thin heterogeneity inside the bulk porous medium which may itself be described as another porous medium with a distinctly different permeability, e.g., a debris- or sediment-filled crack inside a porous rock. We assume that the flow is governed by Darcy's law in both bulk and fracture. Further, we introduce the characteristic width-to-length ratio ε>0 of the fracture as a scaling parameter. Given that the ratio Kf⋆/Kb⋆ of characteristic hydraulic conductivities in the fracture and bulk domain scales with εα, we obtain five different limit models as ε→0 depending on the value of the parameter α∈R. As the mathematical structure of the limit models is different in each case and reaches from a simple boundary condition to a PDE on the interfacial limit fracture, the different cases require different analytical approaches. Aside from delicate weak compactness arguments, the convergence proofs rely on tailored parameterizations and a novel coordinate transformation with controllable behavior with respect to the scaling parameter ε. Besides, we show the wellposedness of the limit models and strong convergence.
The limit of vanishing width-to-length ratio ε→0 is also considered in some of the works mentioned above for systems with simple geometries and constant hydraulic conductivities. In particular, for more simple systems, this is discussed in [20,22] for the case α=−1 and in [25] for the case α=1. Moreover, our approach is related to the approach in [21], where Richards equation is considered for α<1. However, while their focus is on dealing with the nonlinearity and time-dependency of unsaturated flow, our focus is on the derivation of limit models for general fracture geometries and spatially varying tensor-valued hydraulic conductivities for the whole range of parameters α∈R. This aspect is not considered in [21]. In particular, in our case, the presence of off-diagonal elements in the hydraulic conductivity matrix inside the fracture complicates the analysis in the cases α=−1 and α=1. Moreover, one of the limit models (α=−1) will explicitly depend on these off-diagonal components.
The structure of this paper is as follows. In Section 2, we define the full-dimensional model problem of Darcy flow in a porous medium with an isolated fracture and introduce the characteristic width-to-length ratio ε of the fracture as a scaling parameter. Section 3 deals with the derivation of a-priori estimates for the family of full-dimensional solutions parameterized by ε>0. Further, in Section 4, depending on the choice of parameters, we identify the limit models as ε→0 and provide rigorous proofs of convergence. A short summary of the geometric background is given in Appendix A.
First, in Section 2.1, we define the geometric setting and introduce the full-dimensional model problem of single-phase Darcy flow in a porous medium with an isolated fracture in dimensional form. Then, in Section 2.2, dimensional quantities are rescaled by characteristic reference quantities to obtain a non-dimensional problem. Section 2.3 discusses the dependence of the domains and parameters on the width-to-length ratio ε of the fracture, which is introduced as a scaling parameter. Further, given an atlas for the surface that represents the fracture in the limit ε→0, Section 2.4 introduces suitable local parameterizations for the bulk and fracture domains, which, in Section 2.5, allow us to transform the weak formulation of the non-dimensional problem from Section 2.2 into a problem with ε-independent domains.
In the following, dimensional quantities are denoted with a tilde to distinguish them from the non-dimensional quantities that are introduced in Section 2.2. Constant reference quantities are marked by a star.
Let n∈N with n≥2 denote the spatial dimension of a porous medium. Of pratical interest are the cases n∈{2,3} but we also allow n>3. First, we introduce a technical domain ˜G⊂Rn, which we suppose to be bounded with ∂˜G∈C2 (see Figure 2). We write N∈C1(∂˜G;Rn) for the outer unit normal field on ∂˜G. Subsequently, we will consider the limit of vanishing width-to-length ratio for an isolated fracture in a porous medium such that ¯˜γ represents the closure of the interfacial fracture in the limit model. It has to satisfy ∅≠¯˜γ⊂∂˜G as a compact and connected C0,1-submanifold with boundary ∂¯˜γ and dimension n−1. The interior of ¯˜γ is denoted by ˜γ. We remark that ˜γ⊂∂˜G is in fact a C2-submanifold without boundary, while ¯˜γ as a submanifold with boundary is only required to be of class C0,1 (i.e., ¯˜γ can have corners for example). The domain ˜G plays a purely technical role: It induces an orientation on ˜γ. Besides, the domain ˜G (or rather its boundary ∂˜G) allows to us to directly apply geometrical results for (compact) manifolds without boundary without worrying about ¯˜γ as a manifold with boundary. In particular, ∂˜G is endowed with a signed distance function d∂˜G↔. Moreover, ∂˜G has bounded curvature. Thus, there exists a neighborhood of ∂˜G where the orthogonal projection P∂˜G and the signed distance function d∂˜G↔ are well-defined and differentiable. We refer to Appendix A.1 for the relevant geometric background.
In the following, we define the geometry of the full-dimensional model. Given aperture functions ˜ai∈C0,1(¯˜γ) for i∈{+,−} such that the total aperture ˜a:=˜a++˜a−≥0 is non-negative, we define the fracture domain ˜Ωf and its boundary segments ˜γ± by
˜Ωf:={˜π+˜sN(˜π)∈Rn | ˜π∈˜γ,−˜a−(˜π)<˜s<˜a+(˜π)}, | (2.1a) |
˜γ±:={˜π±˜a±(˜π)N(˜π)∈Rn | ˜π∈˜γ}. | (2.1b) |
Here and subsequently, we use the index ± as an abbreviation to simultaneously refer to two different quantities or domains on the inside (−) and outside (+) of the domain ˜G. Further, we distinguish between the parts of the fracture interface ˜γ and the boundary segments ˜γ± with non-zero and zero aperture ˜a, i.e., ˜γ=˜Γ˙∪˜Γ00 and ˜γ±=˜Γ±˙∪˜Γ0, where
2˜Γ:={˜π∈˜γ | ˜a(˜π)>0},˜Γ00:=˜γ∖˜Γ, | (2.2a) |
˜Γ0:=˜γ+∩˜γ−,˜Γ±:=˜γ±∖˜Γ0. | (2.2b) |
We assume that ˜Ωf is connected with λn(˜Ωf)>0, where λn denotes the n-dimensional Lebesgue measure. In addition, we assume that the aperture functions ˜a± are sufficiently small such that ˜Ωf⊂unpp(∂˜G) with unpp(∂˜G)⊂Rn as defined in Definition A.2. Besides, we denote by ˜Ω±⊂Rn two disjoint and bounded Lipschitz domains such that ˜Ω±∩˜Ωf=∅ and ∂˜Ω±∩∂˜Ωf=¯˜γ±. ˜Ω+ and ˜Ω− are bulk domains adjacent to the fracture domain ˜Ωf. Further, we define the total domain
˜Ω:=˜Ω+∪˜Ω−∪˜Ωf∪˜γ+∪˜γ−, | (2.3) |
which we assume to be a Lipschitz domain. Moreover, we write
˜ϱ±:=∂˜Ω±∖˜γ±=˜ϱ±,D˙∪˜ϱ±,N, | (2.4a) |
˜ϱf:=∂˜Ω∖(˜ϱ+∪˜ϱ−)=˜ϱf,D˙∪˜ϱf,N⊂∂˜Ωf | (2.4b) |
for the external boundaries of the bulk domains ˜Ωi⊂˜Ω, i∈{+,−,f}, which are composed of disjoint Dirichlet and Neumann segments ˜ϱi,D and ˜ϱi,N. The resulting geometric configuration is sketched in Figure 1. Besides, the position of the technical domain ˜G is sketched in Figure 2.
Now, let ˜K±∈L∞(˜Ω±;Rn×n) and ˜Kf∈L∞(˜Ωf;Rn×n) be symmetric and uniformly elliptic hydraulic conductivity matrices. Further, for i∈{+,−,f}, let ˜pi denote the pressure head in ˜Ωi. Then, given the source terms ˜q±∈L2(˜Ω±) and ˜qf∈L2(˜Ωf), we consider the following problem of Darcy flow in ˜Ω.
Find ˜p±:˜Ω±→R and ˜pf:˜Ωf→R such that
3−˜∇⋅(˜Ki˜∇˜pi)=˜qiin ˜Ωi,i∈{+,−,f}, | (2.5a) |
˜p±=˜pfon ˜Γ±, | (2.5b) |
˜K±˜∇˜p±⋅n±=˜Kf˜∇˜pf⋅n±on ˜Γ±, | (2.5c) |
˜p+=˜p−on ˜Γ0, | (2.5d) |
˜K+˜∇˜p+⋅n+=−˜K−˜∇˜p−⋅n−on ˜Γ0, | (2.5e) |
˜pi=0on ˜ϱi,D,i∈{+,−,f}, | (2.5f) |
˜Ki˜∇˜pi⋅n=0on ˜ϱi,N,i∈{+,−,f}. | (2.5g) |
Here, n is the outer unit normal on ∂˜Ω and n± denotes the unit normal on ˜γ± pointing into ˜Ω±. We remark that the choice of homogeneous boundary conditions in Eq (2.5) is only made for the sake of simplicity. The extension to the inhomogeneous case is straightforward.
We write L⋆[m] and a⋆[m] for the characteristic values of the length and aperture of the fracture given by
L⋆:=λ∂˜G(˜Γ)1n−1anda⋆:=1λ∂˜G(˜Γ)∫˜Γ˜adλ∂˜G. | (2.6) |
Here, λ∂˜G denotes the Riemannian measure on the submanifold ∂˜G⊂Rn (cf. Appendix A.3). Then, we define ε:=a⋆/L⋆>0 as the characteristic width-to-length ratio of the fracture. Subsequently, in Sections 3 and 4, we will treat ε as scaling parameter and analyze the limit behavior as ε→0.
Next, let K⋆b[m/s] and K⋆f[m/s] be characteristic values of the hydraulic conductivities ˜K± and ˜Kf in the bulk and fracture. In addition, we define the non-dimensional position vector x:=˜x/L⋆. The non-dimensionalization of the position vector x results in a rescaling of spatial derivative operators, e.g., ∇=L⋆˜∇. Besides, it necessitates the definition of non-dimensional domains (and boundary interfaces), which will be denoted without tilde, e.g., Ω:=˜Ω/L⋆. If a domain or interface depends on the width-to-length ε of the fracture, this is indicated by an additional index, e.g., Ωε+:=˜Ω+/L⋆. Moreover, we define
ϱεb,D:=ϱε+,D∪ϱε−,D,ϱεD:=ϱε+,D∪ϱε−,D∪ϱεf,D. | (2.7) |
We require λ∂Ω(ϱεb,D)>0. Besides, we sometimes require the stronger assumption
λ∂Ω(ϱε+,D)>0andλ∂Ω(ϱε−,D)>0,(A) |
i.e., both bulk domains Ωε+ and Ωε− have a boundary part with Dirichlet conditions (and not possibly only one of them). This is subsequently referred to as "assumption (A)". Further, we define the non-dimensional quantities
pε±:=˜p±p⋆b,Kε±:=˜K±K⋆b,qε±:=˜q±q⋆b,a±:=˜a±a⋆,a:=˜aa⋆,pεf:=˜pfp⋆f,Kεf:=˜KfK⋆f,qεf:=˜qfq⋆f, | (2.8) |
where p⋆b:=L⋆, p⋆f:=L⋆, and q⋆b:=K⋆b/L⋆. We assume that there exist parameters α∈R and β≥−1 such that the characteristic fracture quantities K⋆f and q⋆f scale like
K⋆f=εαK⋆bandq⋆f=εβq⋆b. | (2.9) |
The dimensional Darcy system in Eq (2.5) now corresponds to the following non-dimensional problem.
Find pε±:Ωε±→R and pεf:Ωεf→R such that
3−∇⋅(Kε±∇pε±)=qε±in Ωε±, | (2.10a) |
−∇⋅(εαKεf∇pεf)=εβqεfin Ωεf, | (2.10b) |
pε±=pεfon Γε±, | (2.10c) |
Kε±∇pε±⋅nε±=εαKεf∇pεf⋅nε±on Γε±, | (2.10d) |
pε+=pε−on Γε0, | (2.10e) |
Kε+∇pε+⋅nε+=−Kε−∇pε−⋅nε−on Γε0, | (2.10f) |
pεi=0on ϱεi,D,i∈{+,−,f}, | (2.10g) |
Kεi∇pεi⋅n=0on ϱεi,N,i∈{+,−,f}. | (2.10h) |
In Eq (2.10), n is the outer unit normal on ∂Ω and nε± denotes the unit normal on γε± pointing into Ωε±. The geometry of the non-dimensional problem (2.10) with full-dimensional fracture Ωεf, as well as the limit geometry as ε→0, are sketched in Figure 3.
Next, we define the space
Φε:={(φε+,φε−,φεf)∈×i∈{+,−,f}H10,ϱεi,D(Ωεi) |φε+|Γε0=φε−|Γε0,φε±|Γε±=φεf|Γε±}≅H10,ϱεD(Ω), | (2.11) |
where the equalities on Γε0 and Γε± are to be understood in the sense of traces. Then, a weak formulation of the system in Eq (2.10) is given by the following problem.
Find (pε+,pε−,pεf)∈Φε such that, for all (φε+,φε−,φεf)∈Φε,
∑i=±(Kεi∇pεi,∇φεi)L2(Ωεi)+εα(Kεf∇pεf,∇φεf)L2(Ωεf)=∑i=±(qεi,φεi)L2(Ωεi)+εβ(qεf,φεf)L2(Ωεf). | (2.12) |
As a consequence of the Lax-Milgram theorem, the Darcy problem (2.12) admits a unique solution (pε+,pε−,pεf)∈Φε.
Let κk∈C0(∂G), k∈{1,…,n−1}, denote the principal curvatures on ∂G and set
κmax:=maxπ∈¯γmaxk∈{1,…,n−1}|κk(π)|. | (2.13) |
Then, we have κmax<∞ due to the compactness of ¯γ. Further, we define
ˆˆε:=min{1,13κmax,reach(∂G)}>0,ˆε:=ˆˆε2[maxi=±{‖ai‖L∞(γ)}]−1>0 | (2.14) |
with reach(∂G) as defined in Definition A.2. In the following, we require ε∈(0,ˆε]. In Eq (2.14), the condition ˆˆε≤1 ensures that ˆˆε is finite and the condition ˆˆε<[3κmax]−1 guarantees the invertibility of certain ε-perturbed identity operators on the tangent space TπΓ (cf. Lemma 2.2 below). Besides, the condition ˆˆε<reach(∂G) allows us to use the results from Appendix A.1 on the regularity and wellposedness of the orthogonal projection P∂G and the signed distance function d∂G↔.
The dependence of the non-dimensional domains and quantities on the width-to-length ratio ε of the fracture is made explicit in the notation. For the non-dimensional fracture domain Ωεf, the ε-dependence is evident. Specifically, we have
Ωεf={π+sN(π)∈Rn | π∈γ, −εa−(π)<s<εa+(π)}. | (2.15) |
Accordingly, the hydraulic conductivity Kεf and the source term qεf scale like
Kεf(x)=Kˆεf(Tεf(x)),qεf(x)=qˆεf(Tεf(x)), | (2.16) |
where the transformation Tεf:Ωεf→Ωˆεf is given by
Tεf(x)=P∂G(x)−ˆεεd∂G↔(x)N(P∂G(x)). | (2.17) |
Further, we define
Ωε±,in:=Ωε±∩{π±sN(π) | π∈γ,εa±(π)<s<ˆˆε}, | (2.18a) |
Ω±,out:=Ωε±∖Ωε±,in. | (2.18b) |
Note that only the inner region Ωε±,in of the bulk domain Ωε± depends on the scaling parameter ε, while the outer region Ω±,out does not. For the inner region Ωε±,in, we impose a linear deformation in normal direction with decreasing ε, i.e., the hydraulic conductivity Kε± and the source term qε± satisfy
Kε±(x)=K0±(Tε±(x)),qε±(x)=q0±(Tε±(x)) | (2.19) |
for x∈Ωε±,in, where the transformation Tε±:Ωε±,in→Ω0±,in is given by
Tε±(x):=P∂G(x)+tε±(P∂G(x),−d∂G↔(x))N(P∂G(x)), | (2.20a) |
tε±(π,λ):=ˆˆεˆˆε−εa±(π)[λ∓εa±(π)]. | (2.20b) |
It is now easy to see that the following lemma holds.
Lemma 2.1. Let ε∈(0,ˆε]. Then, Tεf:Ωεf→Ωˆεf is a C1-diffeomorphism. Besides, Tε±:Ωε±,in→Ω0±,in is bi-Lipschitz. The inverses T_εf:=(Tεf)−1 and T_ε±:=(Tε±)−1 are given by
T_εf(x)=P∂G(x)−εˆεd∂G↔(x)N(P∂G(x)), | (2.21) |
T_ε±(x)=P∂G(x)+t_ε±(P∂G(x),−d∂G↔(x))N(P∂G(x)), | (2.22a) |
t_ε±(π,λ)=ˆˆε−εa±(π)ˆˆελ±εa±(π). | (2.22b) |
Subsequently, beginning with an atlas for the fracture interface Γ, we develop a suitable local parameterization of the fracture domain Ωεf and the interior bulk domains Ωε±,in. Further, we will introduce transformations onto ε-independent domains and characterize how they depend on the scaling parameter ε. Eventually, in Section 2.5, this will allow us to reformulate the Darcy problem (2.10) in terms of ε-independent domains. In the following, we use the definitions and notations from Appendix A.3.
We observe that Γ⊂∂G is open so that Γ⊂Rn is itself a C2-submanifold of dimension n−1. Besides, ¯Γ⊂Rn is a C0,1-submanifold with boundary. Now, let {(Uj,ψj,Vj)}j∈J be a C2-atlas for Γ consisting of charts ψj:Uj→Vj, where Uj⊂Γ and Vj⊂Rn−1 are open. Then, for j∈J and ε∈(0,ˆε], we write ψ_j:=ψ−1j for the inverse charts and define
Uεf,j:={π+sN(π) | π∈Uj,−εa−(π)<s<εa+(π)}, | (2.23a) |
Vf,j:={(ϑ′,ϑn)∈Rn | ϑ′∈Vj, −a−(ψ_j(ϑ′))<ϑn<a+(ψ_j(ϑ′))} | (2.23b) |
for ε∈(0,ˆε], as well as
Uε±,j:={π±sN(π) | π∈Uj, εa±(π)<s<ˆˆε}, | (2.24a) |
V±,j:={(ϑ′,±ϑn)∈Rn | ϑ′∈Vj, 0<ϑn<ˆˆε} | (2.24b) |
for ε∈[0,ˆε]. In the following, we will also think of the subdomains Ωεf, Ωε±,in⊂Rn as n-dimensional C0,1-submanifolds. With the given atlas for Γ, we can construct a C0,1-atlas {(Uεf,j,ψεf,j,Vf,j)}j∈J for Ωεf for ε∈(0,ˆε], as well as C0,1-atlases {(Uε±,j,ψε±,j,V±,j)}j∈J for Ωε±,in for ε∈[0,ˆε]. For j∈J, the charts ψεf,j and ψε±,j, as well as their inverses ψ_εf,j and ψ_ε±,j, are given by
2ψεf,j:Uεf,j→Vf,j,x↦(ψj(P∂G(x)), −ε−1d∂G↔(x)), | (2.25a) |
ψ_εf,j:Vf,j→Uεf,j,(ϑ′,ϑn)↦ψ_j(ϑ′)+εϑnN(ψ_j(ϑ′)), | (2.25b) |
ψε±,j:Uε±,j→V±,j,x↦(ψj(P∂G(x)), tε±(P∂G(x),−d∂G↔(x))), | (2.26a) |
ψ_ε±,j:V±,j→Uε±,j,(ϑ′,ϑn)↦ψ_j(ϑ′)+t_ε±(ψ_j(ϑ′),ϑn)N(ψ_j(ϑ′)). | (2.26b) |
Further, we introduce the product-like n-dimensional C2-submanifold
Γa:={(π,ϑn) | π∈Γ, −a−(π)<ϑn<a+(π)}⊂Rn×R. | (2.27) |
Then, Γa is the interior of the following C0,1-manifolds with boundary.
¯Γpaerp:={(π,ϑn) | π∈Γ, −a−(π)≤ϑn≤a+(π)}⊂Rn×R, | (2.28a) |
¯Γpaar:={(π,ϑn) | π∈¯Γ, −a−(π)<ϑn<a+(π)}⊂Rn×R. | (2.28b) |
Besides, we write
ϱa,D:={(π,ϑn)∈¯Γpaar | π+ˆεϑnN(π)∈ϱˆεf,D}⊂∂¯Γpaar | (2.29) |
for the external boundary segment of ¯Γpaar with Dirichlet conditions. A C2-atlas of Γa is given by {(Uaj,ψaj,Vf,j)}j∈J, where
Uaj:={(π,ϑn) | π∈Uj,−a−(π)<ϑn<a+(π)}, | (2.30a) |
ψaj:Uaj→Vf,j,(π,ϑn)↦(ψj(π),ϑn). | (2.30b) |
Further, for f∈H1(Γa), we decompose the gradient ∇Γaf into a tangential and a normal component, i.e.,
∇Γaf:=∇Γf+∇Nf,∇Nf(π,ϑn):=∂f(π,ϑn)∂ϑnN(π). | (2.31) |
Next, we write Sψj(ϑ′)∈R(n−1)×(n−1) for the matrix representation of the shape operator Sψ_j(ϑ′) of Γ at ψ_j(ϑ′) with respect to the basis
{∂ψ_j(ϑ′)∂ϑ1,…,∂ψ_j(ϑ′)∂ϑn−1}⊂Tψ_j(ϑ′)Γ. | (2.32) |
Details on the shape operator Sψ_j(ϑ′) can be found in Appendix A.2. In addition, for j∈J and ϑ=(ϑ′,ϑn)∈Vf,j or V±,j, we introduce the abbreviations
Rεf,j(ϑ):=In−1−εϑnSψj(ϑ′), | (2.33a) |
Rε±,j(ϑ):=In−1−t_ε±(ψ_j(ϑ′),ϑn)Sψj(ϑ′), | (2.33b) |
where In−1∈R(n−1)×(n−1) is the identity matrix. Besides, we define the operators
R_εf|(π,ϑn):TπΓ→TπΓ, | (2.34a) |
R_ε±|x:TP∂G(x)Γ→TP∂G(x)Γ, | (2.34b) |
R0±|x:TP∂G(x)Γ→TP∂G(x)Γ | (2.34c) |
for all (π,ϑn)∈Γa and x∈Ω0±,in by
R_εf|(π,ϑn):=(idTπΓ−εϑnSπ)−1, | (2.35a) |
R_ε±|x:=(idTP∂G(x)Γ−t_ε±(P∂G(x),−d∂G↔(x))SP∂G(x))−1, | (2.35b) |
R0±|x:=idTP∂G(x)Γ+d∂G↔(x)SP∂G(x). | (2.35c) |
The operators in Eq (2.34) will appear when considering gradients of yet to be introduced transformations "Ωε±,in→Ω0±,in" and "Ωεf→Γa" onto ε-independent domains (cf. Eq (2.45) and Lemma 2.4 (iii) and (iv) below). Moreover, the operators in Eq (2.34) have the following properties. In particular, we can characterize their behavior as ε→0.
Lemma 2.2. (i) The operators R_εf and R_ε± exist for all ε∈(0,ˆε].
(ii) For all (π,ϑn)∈Γa and x∈Ω0±,in, the operators
R_εf|(π,ϑn),R_ε±|x,andR0±|x |
are self-adjoint for ε∈(0,ˆε]. In particular, for i∈{+,−,f}, it is
g|ψj(ϑ′)Rεi,j(ϑ)=[Rεi,j(ϑ)]tg|ψj(ϑ′). | (2.36) |
(iii) For j∈J and ε∈(0,ˆε], the matrix representations of the operators
R_εf|ψ_aj(ϑ),R_ε±|ψ_0±,j(ϑ),andR0±|ψ_0±,j(ϑ) |
with respect to the basis (2.32) are given by [Rεf,j(ϑ)]−1, [Rε±,j(ϑ)]−1, and R0±,j(ϑ).
(iv) As ε→0, we have
(a)sup(π,ϑn)∈Γa‖idTπΓ−R_εf|(π,ϑn)‖=O(ε), | (2.37a) |
(b)supx∈Ω0±,in‖idTP∂G(x)Γ−R_ε±|x∘R0±|x‖=O(ε) | (2.37b) |
for (π,ϑn)∈Γa and x∈Ω0±,in.
Proof. (i) Using Eq (2.14) and the self-adjointness of Sπ, we find
‖εϑnSπ‖≤εκmaxmaxi=±{‖ai‖L∞(γ)}≤ˆˆε2κmax≤16<1. |
Thus, the operator R_εf|(π,ϑn) exists for all (π,ϑn)∈Γa and ε∈(0,ˆε].
Further, with Eq (2.14) and the self-adjointness of Sπ, we have
‖d∂G↔(x)SP∂G(x)‖≤ˆˆεκmax≤13<1 |
for all x∈Ω0±,in so that R0±|x is invertible with
‖[R0±|x]−1‖≤11−‖d∂G↔(x)SP∂G(x)‖≤32. |
Besides, it is
‖εa±(P∂G(x))[ˆˆε−1d∂G↔(x)±1]SP∂G(x)‖≤32ε‖a±‖L∞(γ)κmax≤14<‖[R0±|x]−1‖−1, |
where we have used that 0≤|ˆˆε−1d∂G↔(x)±1|≤32. Consequently, the operator
R_ε±|x=[R0±|x−εa±(P∂G(x))[ˆˆε−1d∂G↔(x)±1]SP∂G(x)]−1 |
exists for all x∈Ω0±,in and ε∈(0,ˆε].
(ii) The result follows directly from the self-adjointness of the shape operator.
(iii) We have
∂ψ_j(ϑ′)∂ϑi=Dψ_j(ϑ′)Rεf,j(ϑ)[Rεf,j(ϑ)]−1ei=(idTψ_j(ϑ′)Γ−εϑnSψ_j(ϑ′))(Dψ_j(ϑ′)[Rεf,j(ϑ)]−1ei) |
for i∈{1,…,n−1}, where ei∈Rn−1 denotes the ith unit vector, and hence
Dψ_j(ϑ′)[Rεf,j(ϑ)]−1ei=R_εf|ψ_aj(ϑ)(∂ψ_j(ϑ′)∂ϑi). |
The result for R_ε± follows analogously. The result for R0± is trivial.
(iv-a) Using (ii), we find
sup(π,ϑn)∈Γa‖idTπΓ−R_εf|(π,ϑn)‖=sup(π,ϑn)∈Γamaxk∈{1,…,n−1}|1−11−εϑnκk(π)|=O(ε). |
Here, κk∈C0(∂G), k∈{1,…,n−1}, denote the principal curvatures on ∂G, which are bounded due to the compactness of ∂G.
(iv-b) Using Eq (2.14) and the self-adjointness of SP∂G(x), we find
supx∈Ω0±,in|t_ε±(P∂G(x),−d∂G↔(x))|‖SP∂G(x)‖≤[ˆˆε+32ε‖a±‖L∞(γ)]κmax≤712<1. |
Thus, we can express R_ε±|x as a Neumann series and obtain
R_ε±|x∘R0±|x=[∞∑k=0t_ε±(P∂G(x),−d∂G↔(x))kSkP∂G(x)]∘R0±|x=idTP∂G(x)Γ+[t_ε±(P∂G(x),−d∂G↔(x))+d∂G↔(x)]R_ε±|x∘SP∂G(x), |
where t_ε±(P∂G(x),−d∂G↔(x))+d∂G↔(x)=O(ε).
Further, for j∈J, the Jacobians of the inverse charts ψ_εf,j, ψ_ε±,j are given by
Dψ_εf,j(ϑ)=[Dψ_j(ϑ′)Rεf,j(ϑ)εN(ψ_j(ϑ′))], | (2.38a) |
Dψ_ε±,j(ϑ)=Aε±,j(ϑ)+εN(ψ_j(ϑ′))[v±,j(ϑ)]t, | (2.38b) |
where
Aε±,j(ϑ):=[Dψ_j(ϑ′)Rε±,j(ϑ)N(ψ_j(ϑ′))], | (2.39a) |
v±,j(ϑ):=[[±1−ˆˆε−1ϑn][Dψ_j(ϑ′)]t∇Γa±(ψ_j(ϑ′))−ˆˆε−1a±(ψ_j(ϑ′))]. | (2.39b) |
Consequently, with [Dψ_j(ϑ′)]tN(ψ_j(ϑ′))=0, we find that the metric tensors of Ωεf and Ωε±,in in coordinates of the charts ψεf,j and ψε±,j, j∈J, are given by
g|ψεf,j(ϑ)=[[Rεf,j(ϑ)]tg|ψj(ϑ′)Rεf,j(ϑ)00ε2], | (2.40a) |
g|ψε±,j(ϑ)=[[Rε±,j(ϑ)]tg|ψj(ϑ′)Rε±,j(ϑ)001]+(εv±,j(ϑ)+en)(εv±,j(ϑ)+en)t−enetn, | (2.40b) |
where en∈Rn is the nth unit vector and g|ψj denotes the metric tensor on Γ in coordinates of the chart ψj. Subsequently, for j∈J, we will use the notation
μj:=√detg|ψj,με±,j:=√detg|ψε±,j,μεf,j:=√detg|ψεf,j. | (2.41) |
Moreover, we have the following result on the metric tensors.
Lemma 2.3. Let ε∈(0,ˆε] and j∈J. As ε→0, we have
μεf,j(ϑ)=ε[1+O(ε)]μj(ϑ′), | (2.42a) |
με±,j(ϑ)=[1+O(ε)]μ0±,j(ϑ). | (2.42b) |
The prefactors on the right-hand side of Eq (2.42) do not depend on j∈J.
Proof. Given the principal curvatures κk∈C0(∂G) on ∂G, k∈{1,…,n−1}, which are bounded on the compact submanifold ∂G, we have
detRεf,j(ϑ)=n−1∏k=1[1−εϑnκk(ψ_j(ϑ′))],detRε±,j(ϑ)=n−1∏k=1[1−t_ε±(ψ_j(ϑ′),ϑn)κk(ψ_j(ϑ′))] |
for j∈J, where t_ε±(ψ_j(ϑ′),ϑn)=t_0±(ψ_j(ϑ′),ϑn)+O(ε). This yields
detRεf,j(ϑ)=1+O(ε), | (2.43a) |
detRε±,j(ϑ)=[1+O(ε)]detR0±,j(ϑ) | (2.43b) |
so that Eq (2.42a) follows. Moreover, as a consequence of Sylvester's determinant theorem, the relation
det(A+cdt+eft)=det(A)[(dtA−1c+1)(ftA−1e+1)−dtA−1eftA−1c]. | (2.44) |
holds true for any invertible matrix A∈Rn×n and c,d,e,f∈Rn. Thus, with Eq (2.43b), we find
detg|ψε±,j(ϑ)=(1−εˆˆε−1a±(ψ_j(ϑ′)))2(detRε±,j(ϑ))2detg|ψj(ϑ′)=[1+O(ε)](detR0±,j(ϑ))2detg|ψj(ϑ′)=[1+O(ε)]detg|ψ0±,j(ϑ). |
Next, given a partition of unity {χj}j∈J of Γ that is subordinate to the covering {Uj}j∈J, we define the partitions of unity
● {χε±,j}j∈J on Ωε±,in subordinate to {Uε±,j}j∈J by χε±,j:=χj∘P∂G|Ωε±,in,
● {χεf,j}j∈J on Ωεf subordinate to {Uεf,j}j∈J by χεf,j:=χj∘P∂G|Ωεf,
● {χaj}j∈J on Γa subordinate to {Uaj}j∈J by χaj(π,ϑn):=χj(π).
Further, for ε∈(0,ˆε], we define the transformations Yε±:L2(Ω0±)→L2(Ωε±) and Yεf:L2(Γa)→L2(Ωεf) by
(Yε±φ±)(x):={φ±(Tε±(x)),if x∈Ωε±,in,φ±(x)if x∈Ω±,out, | (2.45a) |
(Yεfφf)(x)↦φf(P∂G(x),−ε−1d∂G↔(x)). | (2.45b) |
The inverse maps Y_ε±:=(Yε±)−1 and Y_εf:=(Yεf)−1 are given by
(Y_ε±φε±)(x):={φε±(T_ε±(x)),if x∈Ω0±,in,φε±(x)if x∈Ω±,out, | (2.46a) |
(Y_εfφεf)(π,ϑn):=φεf(π+εϑnN(π)). | (2.46b) |
Moreover, we define the product map
Yε:L2(Ω0+)×L2(Ω0−)×L2(Γa)→L2(Ωε+)×L2(Ωε−)×L2(Ωεf),(φ+,φ−,φf)↦(Yε+φ+,Yε−φ−,Yεfφf) | (2.47) |
and write Y_ε:=(Yε)−1 for its inverse. Then, the following result for the asymptotics of the transformations Yε±, Yεf between the final domains Ω0±, Γa and the ε-dependent original domains Ωε±, Ωεf holds true as ε→0.
Lemma 2.4. There is an ˉε>0 such that the following results hold for all ε∈(0,ˉε].
(i) Yεf:L2(Γa)→L2(Ωεf) defines an isomorphism with
‖Yεfφf‖2L2(Ωεf)=ε[1+O(ε)]‖φf‖2L2(Γa) | (2.48) |
for all φf∈L2(Γa).
(ii) Yε±:L2(Ω0±)→L2(Ωε±) defines an isomorphism with
‖Yε±φ±‖L2(Ωε±)=[1+O(ε)]‖φ±‖L2(Ω0±) | (2.49) |
for all φ±∈L2(Ω0±).
(iii) Yεf|H1(Γa):H1(Γa)→H1(Ωεf) is an isomorphism. In particular, we have
∇(Yεfφf)(π+εϑnN(π))=(R_εf∇Γφf)(π,ϑn)+ε−1∇Nφf(π,ϑn) | (2.50) |
for φf∈H1(Γa) and a.a. (π,ϑn)∈Γa and hence
‖∇(Yεfφf)‖2L2(Ωεf)=[1+O(ε)](ε‖∇Γφf‖2L2(Γa)+ε−1‖∇Nφf‖2L2(Γa)). | (2.51) |
(iv) Yε±|H1(Ω0±):H1(Ω0±)→H1(Ωε±) is an isomorphism. In particular, we have
∇(Yε±φ±)(T_ε±(x))=Mε±(x)∇φ±(x) | (2.52) |
for φ±∈H1(Ω0±) and a.a. x∈Ω0±,in, where
Mε±(x):=[DTε±(T_ε±(x))]t=[DT_ε±(x)]−t. | (2.53) |
Besides, it is
supx∈Ω0±,in‖Mε±(x)−In‖=O(ε), | (2.54) |
where In∈Rn×n denotes the identity matrix. Thus, we obtain
‖∇(Yε±φ±)‖L2(Ωε±)=[1+O(ε)]‖∇φ±‖L2(Ω0±). | (2.55) |
Proof. (i) It is easy to see that Yεf is linear and bijective with inverse Y_εf. Moreover, with Lemma 2.3, we have
‖Yεfφf‖2L2(Ωεf)=∑j∈J∫Vf,j[χεf,j(Yεfφf)2]|ψ_εf,j(ϑ)μεf,j(ϑ)dλn(ϑ)=ε[1+O(ε)]∑j∈J∫Vf,j[χajφ2f]|ψ_aj(ϑ)μj(ϑ′)dϑndλn−1(ϑ′)=ε[1+O(ε)]‖φf‖2L2(Γa). |
(ii) Yε± is clearly linear and bijective with inverse Y_ε±. Further, we have
‖Yε±φ±‖2L2(Ωε±)=‖φ±‖2L2(Ω±,out)+‖φ±∘Tε±‖2L2(Ωε±,in). |
Then, by using Lemma 2.3 and Tε±∘ψ_ε±,j=ψ_0±,j, we find
‖φ±∘Tε±‖2L2(Ωε±,in)=∑j∈J∫V±,j[χε±,j(φ±∘Tε±)2]|ψ_ε±,j(ϑ)με±,j(ϑ)dλn(ϑ)=[1+O(ε)]∑j∈J∫V±,j[χ0±,jφ2±]|ψ_0±,j(ϑ)μ0±,j(ϑ)dλn(ϑ)=[1+O(ε)]‖φ±‖2L2(Ω0±,in). |
(iii) Let φf∈H1(Γa) and φεf:=Yεfφf. Then, by using the Eqs (2.38a), (2.40a) and Lemma 2.2 (ii) and (iii), we find
∇φεf(ψ_εf,j(ϑ))=Dψ_εf,j(ϑ)g−1|ψεf,j(ϑ)∇(φεf∘ψ_εf,j)(ϑ)=Dψ_j(ϑ′)g−1|ψj(ϑ′)[Rεf,j(ϑ)]−t∇′(φεf∘ψ_εf,j)(ϑ)+ε−1∇N(φεf∘ψ_εf,j)(ϑ)=Dψ_j(ϑ′)[Rεf,j(ϑ)]−1g−1|ψj(ϑ′)∇′(φεf∘ψ_εf,j)(ϑ)+ε−1∇N(φεf∘ψ_εf,j)(ϑ)=R_εf|ψ_aj(ϑ)(∇Γφf)(ψ_aj(ϑ))+ε−1∇Nφf(ψ_aj(ϑ)). |
for j∈J and a.a. ϑ=(ϑ′,ϑn)∈Vf,j. Thus, with Lemma 2.2 (iv), we have
|∇φεf(π+εϑnN(π))|2=[1+O(ε)]|∇Γφf(π,ϑn)|2+ε−2|∇Nφf(π,ϑn)|2 |
for a.a. (π,ϑn)∈Γa so that Eq (2.51) follows with Lemma 2.3.
(iv) Equation (2.52) follows by applying the chain rule. Now, let φ±∈H1(Ω0±) and φε±:=Yε±φ±. Then, by using that ψ_ε±,j=T_ε±∘ψ_0±,j, the chain rule yields
Mε±(ψ_0±,j(ϑ))=[DT_ε±(ψ_0±,j(ϑ))]−t=[Dψ_ε±,j(ϑ)]−t[Dψ_0±,j(ϑ)]t. |
With Eq (2.38b) and the Sherman-Morrison formula, we obtain
[Dψ_ε±,j(ϑ)]−1=(In−εen[v±,j(ϑ)]t1−εˆˆε−1a±(ψ_j(ϑ′)))[Aε±,j(ϑ)]−1, |
where In∈Rn×n is the identity matrix and
[Aε±,j(ϑ)]−1=[[Rε±,j(ϑ)]−1g−1|ψj(ϑ′)001][Dψ_j(ϑ′)N(ψ_j(ϑ′))]t. |
Consequently, with Lemma 2.2 (ii), we find
Mε±(ψ_0±,j(ϑ))=[Aε±,j(ϑ)]−t(In−εv±,j(ϑ)etn1−εˆˆε−1a±(ψ_j(ϑ′)))[A0±,j(ϑ)]t=Dψ_j(ϑ′)[Rε±,j(ϑ)]−1R0±,j(ϑ)g−1|ψj(ϑ′)[Dψ_j(ϑ′)]t+wε±,j(ϑ)[N(ψ_j(ϑ′))]t, |
where we have used the abbreviation
wε±,j(ϑ):=ˆˆεN(ψ_j(ϑ′))−ε[±ˆˆε−ϑn][R_ε±|ψ_0±,j(ϑ)∇Γa±(ψ_j(ϑ′))]ˆˆε−εa±(ψ_j(ϑ′)). |
Thus, using that
In=Dψ_j(ϑ′)g−1|ψj(ϑ′)[Dψ_j(ϑ′)]t+N(ψ_j(ϑ′))[N(ψ_j(ϑ′))]t, |
we find
[Mε±(ψ_0±,j(ϑ))−In]z=[R_ε±|ψ_0±,j(ϑ)∘R0±|ψ_0±,j(ϑ)−idTψ_j(ϑ′)Γ](Π|ψ_j(ϑ′)z)+[wε±,j(ϑ)−N(ψ_j(ϑ′))][N(ψ_j(ϑ′))⋅z] |
for all z∈Rn, where
Π|ψ_j(ϑ′):Rn→Rn,u↦Dψ_j(ϑ′)g−1|ψj(ϑ′)[Dψ_j(ϑ′)]tu |
denotes the orthogonal projection onto Tψ_j(ϑ′)Γ. Further, it is easy to see that
supj∈Jsupϑ∈Vf,j|wε±,j(ϑ)−N(ψ_j(ϑ′))|=O(ε). |
Thus, the result follows with Lemma 2.2 (iv).
Subsequently, we will rewrite the integrals in the weak formulation (2.12) on Ωε± and Ωεf as integrals on Ω0± and Γa by using the results on the transformations Yε± and Yεf from Lemma 2.4. In this way, we avoid working with ε-dependent domains and can more easily identify the dominant behavior for vanishing ε.
Let ˉε>0 be as in Lemma 2.4. Then, for ε∈(0,ˉε], we define the solution and test function space
Φ:=Y_ε(Φε)⊂[×i=±H1(Ω0i)]×H1(Γa). | (2.56) |
As a consequence of Lemma 2.4 (iii) and (iv), the space Φ does not depend on ε (cf. Lemma 3.2). In addition, we define
ˆKf:=Y_εfKεf=Y_ˆεfKˆεf∈L∞(Γa;Rn×n), | (2.57a) |
ˆqf:=Y_εfqεf=Y_ˆεfqˆεf∈L2(Γa). | (2.57b) |
Next, for ε∈(0,ˉε], let (φ+,φ−,φf)∈Φ and set
(φε+,φε−,φεf):=Yε(φ+,φ−,φf)∈Φε. | (2.58) |
Further, given the unique solution (pε+,pε−,pεf)∈Φε of Eq (2.12), we define
(ˆpε+,ˆpε−,ˆpεf):=Y_ε(pε+,pε−,pεf)∈Φ. | (2.59) |
Then, with Lemma 2.3, we have
∫Ωεfqεfφεfdλn=∑j∈J∫Vf,j[χεf,jqεfφεf]|ψ_εf,j(ϑ)μεf,j(ϑ)dλn(ϑ)=ε[1+O(ε)]∑j∈J∫Vf,j[χajˆqfφf]|ψ_aj(ϑ)μj(ϑ′)dϑndλn−1(ϑ′)=ε[1+O(ε)]∫ΓaˆqfφfdλΓa. | (2.60) |
In the same way, by additionally using Lemma 2.4 (iii), we obtain
[1+O(ε)]∫ΩεfKεf∇pεf⋅∇φεfdλn=ε∫ΓaˆKfR_εf∇Γˆpεf⋅R_εf∇ΓφfdλΓa+∫ΓaˆKf∇Nˆpεf⋅R_εf∇ΓφfdλΓa+∫ΓaˆKfR_εf∇Γˆpεf⋅∇NφfdλΓa+ε−1∫ΓaˆKf∇Nˆpεf⋅∇NφfdλΓa. | (2.61) |
Moreover, it is
∫Ωε±qε±φε±dλn=∫Ω±,outq0±φ±dλn+∫Ωε±,inqε±φε±dλn=[1+O(ε)]∫Ω0±q0±φ±dλn, | (2.62) |
where we have used that Tε±∘ψ_ε±,j=ψ_0±,j for j∈J and hence, with Lemma 2.3,
∫Ωε±,inqε±φε±dλn=∑j∈J∫V±,j[χε±,jqε±φε±]|ψ_ε±,j(ϑ)με±,j(ϑ)dλn(ϑ)=[1+O(ε)]∑j∈J∫V±,j[χ0±,jq0±φ±]|ψ_0±,j(ϑ)μ0±,j(ϑ)dλn(ϑ)=[1+O(ε)]∫Ω0±,inq0±φ±dλn. | (2.63) |
Analogously, by additionally using Lemma 2.4 (iv), we obtain
∫Ωε±Kε±∇pε±⋅∇φε±dλn=∫Ω±,outK0±∇ˆpε±⋅∇φ±dλn+[1+O(ε)]∫Ω0±,inK0±Mε±∇ˆpε±⋅Mε±∇φ±dλn. | (2.64) |
Thus, by combining the Eqs (2.60)–(2.64), we find that, if (pε+,pε−,pεf)∈Φε solves the weak formulation (2.12), then (ˆpε+,ˆpε−,ˆpεf)∈Φ satisfies
∑i=±,fAεi(ˆpεi,φi)=[1+O(ε)][∑i=±(q0i,φi)L2(Ω0i)+εβ+1(ˆqf,φf)L2(Γa)] | (2.65) |
for all φ=(φ+,φ−,φf)∈Φ as ε→0. The bilinear forms A±:Ω0±×Ω0±→R and Aεf:Γa×Γa→R are given by
Aε±(ˆpε±,φ±):=(K0±∇ˆpε±,∇φ±)L2(Ω±,out)+(K0±Mε±∇ˆpε±,Mε±∇φ±)L2(Ω0±,in), | (2.66) |
Aεf(ˆpεf,φf):=εα+1(ˆKf[R_εf∇Γˆpεf+1ε∇Nˆpεf],[R_εf∇Γφf+1ε∇Nφf])L2(Γa)=εα+1(ˆKfR_εf∇Γˆpεf,R_εf∇Γφf)L2(Γa)+εα(ˆKf∇Nˆpεf,R_εf∇Γφf)L2(Γa)+εα(ˆKfR_εf∇Γˆpεf,∇Nφf)L2(Γa)+εα−1(ˆKf∇Nˆpεf,∇Nφf)L2(Γa). | (2.67) |
In this section, we obtain a-priori estimates for the solution (ˆpε+,ˆpε−,ˆpεf)∈Φ of the transformed weak formulation (2.65) and, consequently, can identify a weakly convergent subsequence as ε→0. The main results are developed in Section 3.3. They build on trace inequalities from Section 3.1 and Poincaré-type inequalities from Section 3.2.
First, we introduce useful functions spaces on Γ and Γa, as well as averaging operators on Γa. Given a λΓ-measurable, non-negative weight function , we define the weighted Lebesgue space as the -space on with measure . Further, we define the weighted Sobolev space as the completion of
(3.1) |
with respect to the norm . Besides, we define the space as the closure of the space
(3.2) |
with respect to the norm . Moreover, we introduce the averaging operators
(3.3a) |
(3.3b) |
We begin by introducing a trace operator on for the lateral boundaries of .
Lemma 3.1. There exists a uniquely defined bounded linear operator
(3.4) |
such that, for all , we have
(3.5) |
Proof. W.l.o.g., we consider . The operator can be treated analogously.
Let . Then, for all , we have
An integration over yields
By applying Hölder's inequality, we obtain
The result now follows from the fact that is dense in .
Besides, we obtain the following characterization of the space .
Lemma 3.2. We have
(3.6) |
In particular, for , it is
(3.7) |
Proof. Using that and are dense, we find
almost everywhere for any for all and . Thus, it is easy to see that
where denotes the space on the right-hand side of Eq (3.6). Besides, Eq (3.7) is a consequence of the trace inequality in .
Further, it is easy to see that the following lemma holds, which introduces a trace operator on the weighted Sobolev space .
Lemma 3.3. Let denote the trace operator on from Lemma A.3. Further, we introduce the constant extension operator
(3.8) |
Then, the trace operator defined by
(3.9) |
is bounded and satisfies
(3.10) |
We obtain two Poincaré-type inequalities for functions in .
Lemma 3.4. Let and . Then, we have
(3.11) |
Proof. We prove the inequality (3.11) for and . The case is analogous. The general case follows from a density argument. We now have
Lemma 3.5. Let and . Then, we have
(3.12) |
Proof. Subsequently, we prove the inequality (3.12) for and . Then, the desired inequality is obtained from a density argument. The case follows by analogy. Now, let . Then, we have
and hence, by using Hölder's inequality,
Consequently, we obtain
An additional integration on yields
(3.13) |
Further, we have so that the result follows by applying the reverse triangle inequality in Eq (3.13).
We can now combine Poincaré's inequality and the Lemmas 3.2 and 3.5 to obtain the following estimate for function triples , which fits the setting of the coupled Darcy problem in Eq (2.65).
Lemma 3.6. Let .
(i) There exists an such that, for all and , we have
(3.14) |
(ii) Let and . Given additionally the assumption , we have
(3.15) |
Proof. (i) Let and, for , define by
Then, with Lemma 2.4 (ii) and Poincaré's inequality, we have
if is sufficiently small. Moreover, Lemma 2.4 (iii) and (iv) yield
By using Poincaré's inequality and the Lemmas 3.2 and 3.5, we obtain
(ii) Follows directly from Poincaré's inequality and the Lemmas 3.2 and 3.5.
Using Lemma 3.6, we can obtain the following a-priori estimates for the solution of the transformed Darcy problem (2.65).
Proposition 3.7. Let . Besides, let either or, given the assumption , let . Further, let . Then, there exists such that, for all , the solution of the transformed Darcy problem (2.65) satisfies
(3.16a) |
(3.16b) |
Proof. We use the solution as a test function in the transformed weak formulation Eq (2.65). The uniform ellipticity of the hydraulic conductivity yields
Here, we have used that, as a consequence of Lemma 2.3 and Lemma 2.4 (iv), it is
Besides, by using Lemma 2.2 (iv) and the uniform ellipticity of , we obtain
By applying Hölder's inequality on the right-hand side of Eq (2.65), we find
(3.17) |
if is sufficiently small. Thus, the inequality (3.16a) follows after applying Lemma 3.6 on the right-hand side of Eq (3.17). Then, the inequality in Eq (3.16b) follows from Eq (3.16a) and Lemma 3.6.
As a consequence of Proposition 3.7, the solution families , , have weakly convergent subsequences in the following sense as .
Proposition 3.8. Let . Besides, let either or, given the assumption , let . Then, there exists a sequence with as such that
(3.18a) |
(3.18b) |
(3.18c) |
(3.18d) |
(3.18e) |
In particular, we have if and if , where denotes the closure of in .
Proof. The weak convergence statements (3.18a), (3.18b), (3.18c), and (3.18d) are a direct consequence of the estimates in Proposition 3.7 and the Rellich-Kondrachov theorem. Further, the weak convergence (3.18e) follows from Proposition 3.7 and
Besides, we have if since is convex and closed in . Further, is convex and closed in and hence if .
Using Proposition 3.7, we can conclude that the limit solution in is constant in -direction if and completely constant if .
Proposition 3.9. Let . Besides, let either or, given the assumption , let .
(i) Let . Then, for a.a. , the limit function from Proposition 3.8 satisfies
(3.19) |
(ii) Let . Then, for a.a. , the limit function from Proposition 3.8 satisfies
(3.20) |
Proof. The results follow from the Propositions 3.7 and 3.8.
If , we obtain continuity of the limit solution across the interface .
Proposition 3.10. Let . Besides, let either or, given the assumption , let . Then, if , the limit functions and from Proposition 3.8 satisfy
(3.21) |
Proof. Let . Then, we have
(3.22) |
Using a version of the Sobolev trace inequality [35, Thm. II.4.1], we obtain
where, with Proposition 3.8, the first term vanishes as and the second term is bounded. Besides, by using the Lemmas 3.2 and 3.4 and Proposition 3.7, we find
Further, the last term on the right-hand side of Eq (3.22) vanishes due to the weak convergence (3.18e) in Proposition 3.8 as .
In the following, we present the convergence proofs and resulting limit models for vanishing . Depending on the value of the parameter , we obtain five different limit models. We distinguish between the following cases that are discussed in separate subsections.
● Section 4.1 discusses the case of a highly conductive fracture, where the limit pressure head inside the fracture becomes completely constant.
● In Section 4.2, we discuss the case of a conductive fracture, where the fracture pressure head in the limit model solves a PDE of effective Darcy flow on the interface .
● In Section 4.3, we examine the case , where the fracture disappears in the limit model, i.e., we have both the continuity of pressure and normal velocity across the interface without any effect of the fracture conductivity.
● Section 4.4 is concerned with the case , where the fracture turns into a permeable barrier with a pressure jump across the interface that scales with an effective conductivity.
● Section 4.5 discusses the case , where the fracture acts like a solid wall in the limit model.
The parameter determines the conductivity of the fracture in the limit . In particular, in accordance with Proposition 3.10, and in agreement with the models in [21,26,27], the pressure will be continuous across the fracture interface for , which is indicative for a conductive fracture. Besides, for , the normal velocity will be continuous across , which is indicative for an obstructing fracture. Further, the parameter controls the presence of a source term for the fracture in the limit model. For a fracture source term will remain in the limit , while, for , the source term will vanish. The role of the parameters and and the corresponding limit model regimes are briefly summarized in Table 1.
parameter | limit model |
fracture source term | |
no fracture source term | |
fracture = major conduit | |
fracture = conduit | |
fracture disappears | |
fracture = permeable barrier | |
fracture = impermeable barrier |
Each subsection is structured as follows. First, we state the strong formulation of the respective limit model and introduce a corresponding weak formulation. Then, we prove weak convergence towards the limit model for the subsequence as and express the limit solution in terms of the limit functions from Proposition 3.8. In a second step, we show strong convergence for the whole sequence as and discuss the wellposedness of the limit model.
Further, for and with well-defined (normal) trace on , we define the jump operators
(4.1) |
Besides, regarding the convergence of the bulk solution, we obtain the following result that will be useful for all cases.
Lemma 4.1. Let . Besides, let either or, given the assumption , let . Then, for all , we have
(4.2) |
as , where denote the limit functions from Proposition 3.8.
Proof. For all , we find
As , the first two terms on the right-hand side vanish due to Lemma 2.4 (iv), the third term due to Proposition 3.8. Thus, the result follows with Proposition 3.8.
As a consequence of Lemma 4.1, the bulk part of the limit problem as will have the following structure in all of the cases.
Find such that
(4.3a) |
(4.3b) |
(4.3c) |
(4.3d) |
(4.3e) |
Here, the functions can be identified as the limit functions from Proposition 3.8. The bulk problem (4.3) is incomplete and has to be supplemented with a fracture problem or suitable conditions on the fracture interface , which will depend on the choice of the parameter .
If , the fracture conductivity is much larger than the bulk conductivity. As a result, the pressure head inside the fracture becomes constant as , i.e., pressure fluctuations in the fracture are instantaneously equilibrated. This matches with the models obtained in [21,27] for Richards equation for . The range of achievable constants for the fracture pressure head in the limit model may be constrained by the choice of Dirichlet conditions at the external fracture boundary. For this reason, we define the set
(4.4) |
of admissible constants for the limit pressure head in the fracture. Then, the set can be characterized as follows.
Remark 4.2. (i) It is either or .
(ii) If , then we have .
(iii) If and for a constant , then we have .
The strong formulation of the limit problem for and as now reads as follows.
Find and such that
(4.5a) |
and the bulk problem (4.3) is satisfied. Moreover, if , the model is closed by the condition
(4.5b) |
Here, and are defined by
(4.6) |
A weak formulation of the system in the Eqs (4.3) and (4.5) is given by the following problem.
Find such that, for all ,
(4.7) |
Here, the space is given by
(4.8) |
Further, we obtain the following weak convergence result.
Theorem 4.3. Let and . Then, is a weak solution of problem (4.7), where , denote the limit functions from Proposition 3.8. Moreover, we have for a.a. .
Proof. Take a test function triple with . By inserting into the transformed weak formulation (2.65), we obtain
Thus, by letting and using Lemma 4.1, we find that the limit solution satisfies Eq (4.7). Besides, with the Propositions 3.8 and 3.9, it is with and hence .
Moreover, we obtain strong convergence in the following sense.
Theorem 4.4. Let and . Then, for the whole sequence , , we have strong convergence
(4.9a) |
(4.9b) |
as . Further, is the unique weak solution of problem (4.7).
Proof. The solution of Eq (4.7) is unique as a consequence of the Lax-Milgram theorem. Thus, the weak convergence (3.18a) and (3.18c) in Proposition 3.8 hold for the whole sequence , . This follows from Proposition 3.7 and the fact that every weakly convergent subsequence has the same limit.
Now, in order to show the strong convergence (4.9), we define the norm on by
Then, with Lemma 3.6, it is easy to see that the norm on the space is equivalent to the natural product norm of . Moreover, with analogous arguments as in Lemma 4.1, we find
(4.10) |
The uniform ellipticity of and Proposition 3.7 yield
Thus, with and Eq (2.65), we have
With Proposition 3.8 and Theorem 4.3, we find
Consequently, with the weak lower semicontinuity of the norm, we obtain
For and , the fracture pressure head in the limit models fulfills a Darcy-like PDE on the interface with an effective hydraulic conductivity matrix . The inflow from the bulk domains into the fracture is modeled by an additional source term on the right-hand side of the interfacial PDE. The bulk and interface solution are coupled by the continuity of the pressure heads across the interface , which corresponds to the case of a conductive fracture in accordance with the choice of the parameter . We remark that the effective conductivity matrix for the limit fracture in Eq (4.13) below explicitly depends on the off-diagonal entries of the full-dimensional conductivity matrix , which is not accounted for in previous works with equivalent scaling of bulk and fracture conductivities [20,21,22].
The resulting limit model for resembles discrete fracture models for Darcy flow that are derived by averaging methods [1,15]. The averaging approach leads to a Darcy-like PDE on the fracture interface as in Eq (4.11a) below. However, the choice of coupling conditions between bulk and interface solution does not occur naturally in this case, especially if the averaged model aspires to describe both conductive and blocking fractures. Therefore, coupling conditions in averaged models are typically obtained by making formal assumptions on the flow profile inside the fracture and usually include a jump of pressure across the fracture interface. Here, only the conductive case corresponding to is considered. In particular, as a consequence of Proposition 3.10, the pressure is continuous across the fracture interface in the limit model.
The strong formulation of the limit problem for and now reads as follows.
Find and such that
(4.11a) |
(4.11b) |
(4.11c) |
(4.11d) |
and the bulk problem (4.3) is satisfied. Here, and in Eq (4.11a) are given by
(4.12) |
(4.13) |
In Eq (4.13), the application of the operator is to be understood componentwise. We remark that agrees with on if is constant and is an eigenvector of , which is in agreement with the models in [21,22]. The boundary parts , in the Eqs (4.11d) and (4.11c) are given by
(4.14a) |
(4.14b) |
Generally, in particular, we have , i.e., we also have a homogeneous Neumann condition at closing points of the fracture inside the domain.
A weak formulation of the system in the Eqs (4.3) and (4.11) is given by the following problem.
Find such that, for all ,
(4.15) |
Here, the space is defined by
(4.16) |
We now have the following weak convergence result.
Theorem 4.5. Let and . Then, is a weak solution of problem (4.15), where , denote the limit functions from Proposition 3.8. Further, for a.a. , we have .
Proof. According to Proposition 3.7, we have . Thus, there exists such that
(4.17) |
as . By multiplying the transformed weak formulation (2.65) by and taking the limit , we find
(4.18) |
for any test function triple , where we have used Lemma 2.2 (iv). A solution for is now clearly given by
(4.19) |
Moreover, suppose that is another solution of Eq (4.18). Then, with Eq (4.18), we find
Thus, by choosing as
we obtain a.e. in , i.e., is uniquely determined by Eq (4.19).
Next, we define the space
and take a test function triple with . Then, there is a function with a.e. in . With Proposition 3.8, Lemma 4.1, and Eq (4.17), we obtain
as . Here, we have used that
for all , where the first two terms on the right-hand side vanish according to Lemma 2.2 (iv) as and the third terms tends to zero with Proposition 3.8. Moreover, with Eq (4.19) and Proposition 3.9 (ii), we have
where is defined by Eq (4.13). Thus, by inserting into the transformed weak formulation (2.65) and letting , it follows that the limit solution satisfies Eq (4.15). Besides, with Lemma 3.3 and Proposition 3.10, we have .
The effective hydraulic conductivity matrix has the following properties.
Lemma 4.6. (i) The effective hydraulic conductivity matrix from Eq (4.13) is symmetric and positive semidefinite. In addition, for all and , we have .
(ii) If , then is uniformly elliptic on , i.e., for all and , we have .
Proof. (i) is symmetric by definition. Moreover, for , we have
With the Cauchy-Schwarz inequality, we obtain
with strict inequality if .
(ii) Suppose that, for all , there exist and such that
W.l.o.g., we assume for all . Then, with the Bolzano-Weierstraß theorem, there exists a subsequence such that
as . In particular, we have
as , which is a contradiction to (i).
Further, the following strong convergence result holds true.
Theorem 4.7. Let and . Then, we have strong convergence
(4.20a) |
(4.20b) |
(4.20c) |
as , where is given by Eq (4.19). Moreover, if is uniformly elliptic on , is the unique weak solution of the problem in Eq (4.15) and the strong convergence in Eq (4.20) holds for the whole sequence , .
Proof. First, we define the norm
on . Then, with Lemma 3.6, it is easy to see that the norm is equivalent to the product norm on . With Lemma 2.2 (iv), Proposition 3.7, and the Eqs (2.65) and (4.10), we find
Thus, with the Proposition 3.8 and Theorem 4.5, we obtain
Additionally, with the Eqs (4.13) and (4.19) and Proposition 3.9, it is
Thus, with Proposition 3.9, we have
Now, let be uniformly elliptic on and . Then, we have
Hence, we obtain coercivity on by applying Lemma 3.6. Thus, as a consequence of the Lax-Milgram theorem, is the unique weak solution of the problem in Eq (4.15). Further, this implies the convergence of the whole sequence , , as since every convergent subsequence has the same limit.
For and , the hydraulic conductivities in bulk and fracture are of similar magnitude such that the fracture disappears in the limit . No effect of the fracture conductivity is visible in the limit model and pressure and normal velocity are continuous across the interface (except for source terms if ). This fits the models derived in [21,27] for , where Richards equation is considered. The strong formulation of the limit problem reads as follows.
Find such that
(4.21a) |
(4.21b) |
and the bulk problem (4.3) is satisfied, where is defined as in Eq (4.12).
A weak formulation of the system in the Eqs (4.3) and (4.21) is given by the following problem.
Find such that, for all with ,
(4.22) |
Here, the space is given by
(4.23) |
We now obtain the following convergence results.
Theorem 4.8. Let and . Besides, let either or assume that holds. Then, given the limit functions and from Proposition 3.8, we find that is a weak solution of Eq (4.22). Moreover, we have on and for a.a. .
Proof. Take a test function triple such that a.e. in . Then, by inserting into the transformed weak formulation (2.65), we obtain
(4.24) |
Further, with Lemma 2.2 (iv) and Proposition 3.7, we have
if is sufficiently small. Thus, by using Proposition 3.8 and Lemma 4.1 and letting in Eq (4.24), it follows that the limit solution pair satisfies the weak formulation (4.22). Besides, with Proposition 3.10, we have .
Theorem 4.9. Let and . Then, given the assumption , we have strong convergence
(4.25a) |
(4.25b) |
as for the whole sequence , . Besides, is the unique weak solution of Eq (4.22).
Proof. As consequence of the Lax-Milgram theorem, the problem in Eq (4.22) has a unique weak solution. Thus, the weak convergence (3.18a) holds for the whole sequence . This follows from Proposition 3.7 and the fact that every weakly convergent subsequent has the same limit. Besides, with the Propositions 3.7, 3.9, and 3.10, the weak convergence (3.18d) is satisfied for the whole sequence .
Next, we equip the space with the norm defined by
(4.26) |
which, as a consequence of Lemma 3.6, is equivalent to the usual product norm on . Besides, with Proposition 3.7, we have
Thus, using the Eqs (2.65) and (4.10) and , we find
Further, Theorem 4.8 yields
With the weak lower semicontinuity of the norm, we now have
For and , the fracture becomes a permeable barrier in limit with a jump of pressure heads across the interface but continuous normal velocity (except for source terms).
In the following, we will derive two different limit models for and . First, in Section 4.4.1, we obtain a coupled limit problem, where the pressure head in the fracture satisfies a parameter-dependent Darcy-type ODE inside the full-dimensional fracture domain . The ODE is formulated with respect to the normal coordinate , while the tangential coordinate acts as a parameter. This resembles the limit problem in [27] for Richards equation with the respective scaling of hydraulic conductivities. However, in Section 4.4.2, it then turns out that the bulk problem can be solved independently from the fracture problem. This is akin to the limit model in [25], where the Laplace equation is considered. In the decoupled bulk limit problem, the jump of pressure heads across the interface scales with an effective hydraulic conductivity, that is defined as a non-trivial mean value of the fracture conductivity in normal direction and reminds of a result from homogenization theory. In particular, if one is still interested in the fracture solution, it is possible to first solve the decoupled bulk limit problem in Section 4.4.2, which will then provide the boundary conditions to solve the ODE for the fracture pressure head in Section 4.4.1.
The strong formulation of the coupled limit problem for and reads as follows.
Find and such that
(4.27a) |
(4.27b) |
(4.27c) |
and the bulk problem (4.3) is satisfied. Here, and are defined by
(4.28) |
(4.29) |
A weak formulation of the system in the Eqs (4.3) and (4.27) is given by the following problem.
Find such that, for all ,
(4.30) |
We obtain the following convergence results.
Theorem 4.10. Let and . Then, given the assumption , the triple is a weak solution of problem (4.30), where and denote the limit functions from Proposition 3.8.
Proof. According to Proposition 3.7, we have
and hence
(4.31) |
as . As a result, we have
where, as , the first term vanishes with Lemma 2.2 (iv) and the second term with Eq (4.31). Thus, with the Propositions 3.7 and 3.8 and the Lemmas 2.2 (iv) and 4.1, we conclude that solves Eq (4.30) by taking the limit in the transformed weak formulation (2.65).
Theorem 4.11. Let and . Then, given the assumption , we have strong convergence
(4.32a) |
(4.32b) |
(4.32c) |
as for the whole sequence , . Besides, we find that is the unique weak solution of the problem in Eq (4.30).
Proof. Clearly, the bilinear form of the weak formulation (4.30) is continuous and coercive with respect to the norm defined by Eq (4.26). Thus, with the Lax-Milgram theorem, we obtain that is the unique solution of Eq (4.30). As a result, every weakly convergent subsequence has the same limit and hence, with Proposition 3.7, the weak convergence statements (3.18a) and (3.18d) in Proposition 3.8 hold for the whole sequence , .
Further, we define the space
and equip the product space with the norm
Then, with Lemma 3.6, it is easy to see that the norm is equivalent to the standard product norm on . Moreover, with Lemma 2.2 (iv) and the Eq (2.65) and (4.10), we have
Thus, with Proposition 3.8 and Theorem 4.10, we find
Starting from the coupled limit problem (4.30), we will subsequently derive a decoupled limit problem for the bulk solution only. The strong formulation of the decoupled bulk limit problem reads as follows.
Find such that
(4.33a) |
(4.33b) |
and the bulk problem (4.3) is satisfied, where is given by Eq (4.12). and the effective hydraulic conductivity with are defined by
(4.34a) |
(4.34b) |
(4.35) |
A weak formulation of the system in the Eqs (4.3) and (4.33) is given by the following problem.
Find such that, for all ,
(4.36) |
Here, the space is given by
(4.37) |
We require the following auxiliary result.
Lemma 4.12. The map
(4.38) |
defines a continuous embedding .
Proof. With Lemma 3.2, we have
for a.a. . Thus, an additional integration on yields
We now obtain the following convergence result.
Theorem 4.13. Let and . Then, given that the assumption holds true, is the unique solution of problem (4.36), where denote the limit functions from Proposition 3.8.
Proof. Let . We define by
where is given by Eq (4.35). It is easy to check that . In particular, we have
Thus, by inserting the test function triple into the weak formulation (4.30) and by using that
we find that satisfies Eq (4.36). With Lemma 4.12, we have . The uniqueness of the solution follows from the Lax-Milgram theorem.
For and , the fracture becomes a solid wall as , i.e., the interface is an impermeable barrier with zero flux across . This matches the formally derived limit model in [27] in the case , where the Richards equation is considered. The strong formulation of the limit problem reads as follows.
Find such that
(4.39) |
and the bulk problem (4.3) is satisfied. A weak formulation of the system in the Eqs (4.3) and (4.39) is given by the following problem.
Find such that, for all ,
(4.40) |
Here, the space is given by
(4.41) |
We now have the following convergence results.
Theorem 4.14. Let and . Then, given the assumption , is a weak solution of problem (4.40), where denote the limit functions from Proposition 3.8.
Proof. With Proposition 3.7, we have
Thus, with the Lemmas 2.2 (iv) and 4.1, the result follows by letting in the transformed weak formulation (2.65).
Theorem 4.15. Let and . Then, given the assumption , we have strong convergence
(4.42) |
as for the whole sequence . Moreover, is the unique weak solution of the problem in Eq (4.40).
Proof. The result follows with analogous arguments as in the cases above.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)–project number 327154368–SFB 1313, and under Germany's Excellence Strategy–EXC 2075 – 390740016. The authors also were supported by the Stuttgart Center for Simulation Science (SimTech).
The authors declare that there is no conflict of interest.
In the following, we summarize useful definitions and results related to the geometry of Euclidean submanifolds.
Definition A.1. Let and with . Besides, let and . Then, is called an -dimensional submanifold of class if, for all , there exists open with and a -diffeomorphism , where open, such that
(A.1) |
Here, denotes the zero vector.
We introduce the orthogonal projection and (signed) distance function of a set and state selected properties and regularity results. For details, we refer to [36].
Definition A.2. Let .
(i) We write , for the distance function of . If for a set , we can define the signed distance function of by
(A.2) |
(ii) A set is said to have the unique nearest point property with respect to if, for all , there exists a unique such that . We write for the maximal set with this property.
(iii) We define the orthogonal projection onto by
(A.3) |
(iv) Let . Then, we define the -neighborhood of by
(A.4) |
For , we also write .
(v) We define the reach of by
(A.5) |
Let be a -submanifold, . Then, the orthogonal projection is -differentiable on [36, Thm. 2]. If , we have for and with [36, Prop. 2]. Besides, if is compact and , we have [36, Prop. 6]. Moreover, if for a set of class , , the signed distance function is -differentiable on (cf. [37, Thm. 7.8.2] and [36, Thm. 2]).
Let and be an -dimensional -submanifold with a global unit normal vector field . We define the shape operator of at for each as the negative directional derivative . Then, for each , the shape operator is a self-adjoint linear operator . The eigenvalues of the shape operator are called the principal curvatures of at . In particular, we have .
Let be an -dimensional -submanifold with boundary . We denote charts for as triples , i.e., and (or for charts with boundary) are open and is bi-Lipschitz. For the inverse chart , we also use the symbol . Besides, we write for the metric tensor in coordinates of the chart , i.e., . For , we write for the Lebesgue space on with respect to the Riemannian measure . Moreover, we define . Following [38], we define the first-order Sobolev space as the completion of
(A.6) |
with respect to the norm , where denotes the gradient of . In local coordinates, we have
(A.7) |
Besides, is a reflexive Hilbert space. For the more general case of Sobolev spaces of arbitrary order and on Riemannian manifolds, we refer to [38]. Further, if is compact, we can alternatively define the Sobolev space by using local coordinates [39]. Given a finite atlas of and a subordinate partition of unity , we define the space
(A.8) |
with the norm . It is easy to check that the two definitions for are equivalent. Consequently, it is , where denotes the interior of . Moreover, with analogous arguments as in [40,§11], one can prove the following trace theorem.
Lemma A.3. Let be compact. Then, there exists a unique bounded linear operator such that for all .
[1] |
S. Burbulla, M. Hörl, C. Rohde, Flow in porous media with fractures of varying aperture, SIAM J. Sci. Comput., 45 (2023), A1519–A1544. https://doi.org/10.1137/22M1510406 doi: 10.1137/22M1510406
![]() |
[2] |
L. Y. Wang, Z. Y. Yin, Fluid flow and mass transport in fractured media with curved fractures and varying apertures: A 3d modeling approach, Int. J. Numer. Methods Eng., 124(2023), 4311–4338. https://doi.org/10.1002/nme.7314 doi: 10.1002/nme.7314
![]() |
[3] |
I. Berre, F. Doster, E. Keilegavlen, Flow in fractured porous media: A review of conceptual models and discretization approaches, Transp. Porous Med., 130 (2019), 215–236. https://doi.org/10.1007/s11242-018-1171-6 doi: 10.1007/s11242-018-1171-6
![]() |
[4] |
R. C. Liu, B. Li, Y. J. Jiang, N. Huang, Review: Mathematical expressions for estimating equivalent permeability of rock fracture networks, Hydrogeol. J., 24 (2016), 1623–1649. https://doi.org/10.1007/s10040-016-1441-8 doi: 10.1007/s10040-016-1441-8
![]() |
[5] |
M. Oda, Permeability tensor for discontinuous rock masses, Géotechnique, 35 (1985), 483–495. https://doi.org/10.1680/geot.1985.35.4.483 doi: 10.1680/geot.1985.35.4.483
![]() |
[6] |
T. Arbogast, J. Douglas, Jr., U. Hornung, Derivation of the double porosity model of single phase flow via homogenization theory, SIAM J. Numer. Anal., 21 (1990), 823–836. https://doi.org/10.1137/0521046 doi: 10.1137/0521046
![]() |
[7] |
G. I. Barenblatt, I. P. Zheltov, I. N. Kochina, Basic concepts in the theory of seepage of homogeneous liquids in fissured rocks, J. Appl. Math. Mech., 24 (1960), 1286–1303. https://doi.org/10.1016/0021-8928(60)90107-6 doi: 10.1016/0021-8928(60)90107-6
![]() |
[8] |
M. Chen, S. A. Masum, H. R. Thomas, 3d hybrid coupled dual continuum and discrete fracture model for simulation of CO2 injection into stimulated coal reservoirs with parallel implementation, Int. J. Coal Geol., 262 (2022), 104103. https://doi.org/10.1016/j.coal.2022.104103 doi: 10.1016/j.coal.2022.104103
![]() |
[9] |
D. C. Karvounis, P. Jenny, Adaptive hierarchical fracture model for enhanced geothermal systems, Multiscale Model Simul., 14 (2016), 207–231. https://doi.org/10.1137/140983987 doi: 10.1137/140983987
![]() |
[10] |
E. Ahmed, J. Jaffré, J. E. Roberts, A reduced fracture model for two-phase flow with different rock types, Math. Comput. Simul., 137 (2017), 49–70. https://doi.org/10.1016/j.matcom.2016.10.005 doi: 10.1016/j.matcom.2016.10.005
![]() |
[11] |
K. Brenner, J. Hennicker, R. Masson, P. Samier, Hybrid-dimensional modelling of two-phase flow through fractured porous media with enhanced matrix fracture transmission conditions, J. Comput. Phys., 357 (2018), 100–124. https://doi.org/10.1016/j.jcp.2017.12.003 doi: 10.1016/j.jcp.2017.12.003
![]() |
[12] | M. Bukac, I. Yotov, P. Zunino, Dimensional model reduction for flow through fractures in poroelastic media, ESAIM Math. Model. Numer. Anal., 51 (2017), 1429–1471. |
[13] |
S. Burbulla, C. Rohde, A finite-volume moving-mesh method for two-phase flow in dynamically fracturing porous media, J. Comput. Phys., 458 (2022), 111031. https://doi.org/10.1016/j.jcp.2022.111031 doi: 10.1016/j.jcp.2022.111031
![]() |
[14] |
M. Lesinigo, C. D'Angelo, A. Quarteroni, A multiscale Darcy–Brinkman model for fluid flow in fractured porous media, Numer. Math., 117 (2011), 717–752. https://doi.org/10.1007/s00211-010-0343-2 doi: 10.1007/s00211-010-0343-2
![]() |
[15] |
V. Martin, J. Jaffré, J. E. Roberts, Modeling fractures and barriers as interfaces for flow in porous media, SIAM J. Sci. Comput., 26 (2005), 1667–1691. https://doi.org/10.1137/S1064827503429363 doi: 10.1137/S1064827503429363
![]() |
[16] |
I. Rybak, S. Metzger, A dimensionally reduced Stokes–Darcy model for fluid flow in fractured porous media, Appl. Math. Comput., 384 (2020), 125260. https://doi.org/10.1016/j.amc.2020.125260 doi: 10.1016/j.amc.2020.125260
![]() |
[17] | M. Starnoni, I. Berre, E. Keilegavlen, J. M. Nordbotten, Modeling and discretization of flow in porous media with thin, full-tensor permeability inclusions, Internat. J. Numer. Methods Engrg., 122 (2021), 4730–4750. |
[18] |
P. Paranamana, E. Aulisa, M. Toda, Geometric model of the fracture as a manifold immersed in porous media, J. Math. Phys., 62 (2021), 051508. https://doi.org/10.1063/1.5109730 doi: 10.1063/1.5109730
![]() |
[19] |
A. Armiti-Juber, On the limit of a two-phase flow problem in thin porous media domains of Brinkman type, Math. Methods Appl. Sci., 45 (2022), 2563–2581. https://doi.org/10.1002/mma.7940 doi: 10.1002/mma.7940
![]() |
[20] |
H. P. Huy, E. Sanchez-Palencia, Phénomènes de transmission à travers des couches minces de conductivité élevée, J. Math. Anal. Appl., 47 (1974), 284–309. https://doi.org/10.1016/0022-247X(74)90023-7 doi: 10.1016/0022-247X(74)90023-7
![]() |
[21] |
F. List, K. Kumar, I. S. Pop, F. A. Radu, Rigorous upscaling of unsaturated flow in fractured porous media, SIAM J. Math. Anal., 52 (2020), 239–276. https://doi.org/10.1137/18M1203754 doi: 10.1137/18M1203754
![]() |
[22] | F. A. Morales, R. E. Showalter, The narrow fracture approximation by channeled flow, J. Math. Anal. Appl., 365 (2010), 320–331. |
[23] |
F. A. Morales, R. E. Showalter, A Darcy–Brinkman model of fractures in porous media, J. Math. Anal. Appl., 452 (2017), 1332–1358. https://doi.org/10.1016/j.jmaa.2017.03.063 doi: 10.1016/j.jmaa.2017.03.063
![]() |
[24] |
F. A. Morales, R. E. Showalter, Interface approximation of darcy flow in a narrow channel, Math. Methods Appl. Sci., 35 (2012), 182–195. https://doi.org/10.1002/mma.1555 doi: 10.1002/mma.1555
![]() |
[25] | E. Sanchez-Palencia, Problèmes de perturbations liés aux phénomènes de conduction à travers des couches minces de grande résistivité, J. Math. Pures Appl., 53 (1974), 251–269. |
[26] |
M. Dugstad, K. Kumar, Dimensional reduction of a fractured medium for a two-phase flow, Adv. Water Resour., 162 (2022), 104140. https://doi.org/10.1016/j.advwatres.2022.104140 doi: 10.1016/j.advwatres.2022.104140
![]() |
[27] |
K. Kumar, F. List, I. S. Pop, F. A. Radu, Formal upscaling and numerical validation of unsaturated flow models in fractured porous media, J. Comput. Phys., 407 (2020), 109138. https://doi.org/10.1016/j.jcp.2019.109138 doi: 10.1016/j.jcp.2019.109138
![]() |
[28] |
T. Mel'nyk, C. Rohde, Asymptotic approximations for semilinear parabolic convection-dominated transport problems in thin graph-like networks, J. Math. Anal. Appl., 529 (2024), 127587. https://doi.org/10.1016/j.jmaa.2023.127587 doi: 10.1016/j.jmaa.2023.127587
![]() |
[29] | Jan Březina, Jan Stebel, Analysis of model error for a continuum-fracture model of porous media flow, In T. Kozubek, R. Blaheta, J. Šístek, M. Rozložník, M. Čermák, editors, High Performance Computing in Science and Engineering, Cham: Springer, 2016,152–160. |
[30] |
M. J. Gander, J. Hennicker, R. Masson, Modeling and analysis of the coupling in discrete fracture matrix models, SIAM J. Numer. Anal., 59 (2021), 195–218. https://doi.org/10.1137/20M1312125 doi: 10.1137/20M1312125
![]() |
[31] |
W. M. Boon, J. M. Nordbotten, J. E. Vatne, Functional analysis and exterior calculus on mixed-dimensional geometries, Ann. Mat. Pura Appl., 200 (2021), 757–789. https://doi.org/10.1007/s10231-020-01013-1 doi: 10.1007/s10231-020-01013-1
![]() |
[32] |
W. M. Boon, J. M. Nordbotten, Mixed-dimensional poromechanical models of fractured porous media, Acta Mech., 234 (2023), 1121–1168. https://doi.org/10.1007/s00707-022-03378-1 doi: 10.1007/s00707-022-03378-1
![]() |
[33] |
A. Mikelić, M. F. Wheeler, T. Wick, Phase-field modeling of a fluid-driven fracture in a poroelastic medium, Comput. Geosci., 19 (2015), 1171–1195. https://doi.org/10.1007/s10596-015-9532-5 doi: 10.1007/s10596-015-9532-5
![]() |
[34] |
S. Burbulla, L. Formaggia, C. Rohde, A. Scotti, Modeling fracture propagation in poro-elastic media combining phase-field and discrete fracture models, Comput. Methods Appl. Mech. Eng., 403 (2023), 115699. https://doi.org/10.1016/j.cma.2022.115699 doi: 10.1016/j.cma.2022.115699
![]() |
[35] | G. P. Galdi, An Introduction to the Mathematical Theory of the Navier-Stokes Equations, New York: Springer, 2011. |
[36] |
G. Leobacher, A. Steinicke, Existence, uniqueness and regularity of the projection onto differentiable manifolds, Ann. Glob. Anal. Geom., 60 (2021), 559–587. https://doi.org/10.1007/s10455-021-09788-z doi: 10.1007/s10455-021-09788-z
![]() |
[37] | M. C. Delfour, J. P. Zolésio, Shapes and Geometries, Philadelphia: SIAM, 2011. |
[38] | E. Hebey, Nonlinear Analysis on Manifolds, Providence: American Mathematical Society, 2000. |
[39] | J. Wloka, Partial Differential Equations, Cambridge: Cambridge University Press, 1987. |
[40] | B. Booß-Bavnbek, K. P. Wojciechowski, Elliptic Boundary Problems for Dirac Operators, Boston: Birkhäuser, 1993. |
1. | Wei Liu, Kai Li, Decoupled bound-preserving algorithms for compressible Darcy-Brinkman flow with advection-diffusion transport problem in fractured media, 2025, 495, 00963003, 129298, 10.1016/j.amc.2025.129298 |
parameter | limit model |
fracture source term | |
no fracture source term | |
fracture = major conduit | |
fracture = conduit | |
fracture disappears | |
fracture = permeable barrier | |
fracture = impermeable barrier |