
This study investigates the global well-posedness of a coupled Navier–Stokes–Darcy model incorporating the Beavers–Joseph–Saffman–Jones interface boundary condition in two-dimensional Euclidean space. We establish the existence of global strong solutions for the system in both linear and nonlinear cases where porosity depends on pressure. When dealing with the time-dependent porous media, the primary challenge in obtaining closed prior estimates arises from the presence of complex, sharp interfaces. To address this issue, we employ the classical Trace Theorem. Such space-time variable coupled systems are crucial for understanding underground fluid flow.
Citation: Linlin Tan, Bianru Cheng. Global well-posedness of 2D incompressible Navier–Stokes–Darcy flow in a type of generalized time-dependent porosity media[J]. Electronic Research Archive, 2024, 32(10): 5649-5681. doi: 10.3934/era.2024262
[1] | Jianxia He, Qingyan Li . On the global well-posedness and exponential stability of 3D heat conducting incompressible Navier-Stokes equations with temperature-dependent coefficients and vacuum. Electronic Research Archive, 2024, 32(9): 5451-5477. doi: 10.3934/era.2024253 |
[2] | Nisachon Kumankat, Kanognudge Wuttanachamsri . Well-posedness of generalized Stokes-Brinkman equations modeling moving solid phases. Electronic Research Archive, 2023, 31(3): 1641-1661. doi: 10.3934/era.2023085 |
[3] | Jiwei Jia, Young-Ju Lee, Yue Feng, Zichan Wang, Zhongshu Zhao . Hybridized weak Galerkin finite element methods for Brinkman equations. Electronic Research Archive, 2021, 29(3): 2489-2516. doi: 10.3934/era.2020126 |
[4] | Jie Zhang, Gaoli Huang, Fan Wu . Energy equality in the isentropic compressible Navier-Stokes-Maxwell equations. Electronic Research Archive, 2023, 31(10): 6412-6424. doi: 10.3934/era.2023324 |
[5] | Keqin Su, Rong Yang . Pullback dynamics and robustness for the 3D Navier-Stokes-Voigt equations with memory. Electronic Research Archive, 2023, 31(2): 928-946. doi: 10.3934/era.2023046 |
[6] | Jingjing Zhang, Ting Zhang . Local well-posedness of perturbed Navier-Stokes system around Landau solutions. Electronic Research Archive, 2021, 29(4): 2719-2739. doi: 10.3934/era.2021010 |
[7] | Jiangshan Wang, Lingxiong Meng, Hongen Jia . Numerical analysis of modular grad-div stability methods for the time-dependent Navier-Stokes/Darcy model. Electronic Research Archive, 2020, 28(3): 1191-1205. doi: 10.3934/era.2020065 |
[8] | Jie Qi, Weike Wang . Global solutions to the Cauchy problem of BNSP equations in some classes of large data. Electronic Research Archive, 2024, 32(9): 5496-5541. doi: 10.3934/era.2024255 |
[9] | Wei Shi, Xinguang Yang, Xingjie Yan . Determination of the 3D Navier-Stokes equations with damping. Electronic Research Archive, 2022, 30(10): 3872-3886. doi: 10.3934/era.2022197 |
[10] | Chengchun Hao, Tao Luo . Some results on free boundary problems of incompressible ideal magnetohydrodynamics equations. Electronic Research Archive, 2022, 30(2): 404-424. doi: 10.3934/era.2022021 |
This study investigates the global well-posedness of a coupled Navier–Stokes–Darcy model incorporating the Beavers–Joseph–Saffman–Jones interface boundary condition in two-dimensional Euclidean space. We establish the existence of global strong solutions for the system in both linear and nonlinear cases where porosity depends on pressure. When dealing with the time-dependent porous media, the primary challenge in obtaining closed prior estimates arises from the presence of complex, sharp interfaces. To address this issue, we employ the classical Trace Theorem. Such space-time variable coupled systems are crucial for understanding underground fluid flow.
We begin by providing a comprehensive explanation of the simplified Navier-Stokes-Darcy system within the open strip domain Ω⊂R2, as illustrated in Figure 1. The incompressible free fluid flows within the confined conduit region Ωf⊂R2, which is interconnected with a time-dependent porous medium region Ωm⊂R2. These two regions are separated by an interface denoted as Γi, where the two types of fluid flow interact. The boundaries ΓU=∂Ωf∖Γi and ΓL=∂Ωm∖Γi represent the upper and lower boundaries, respectively.
The porous medium flow in the conduit Ωm is governed by the continuity equation and Darcy's law, as stated in [1,2,3,4,5,6]:
{∂tϕ(p)−divv=0,inΩm,v=−Kμ2∇p,inΩm, | (1.1) |
where p=p(x,t) denotes the pressure, μ2 is the constant viscosity of the fluid, and the real constant symmetric matrix K=kI represents the permeability of the porous medium. Here, k is a positive constant satisfying 0<λ≤k≤Λ, where λ and Λ are known constants, and I is the identity matrix. Additionally, ϕ=ϕ(p) denotes the pressure-dependent porosity. Combining the two equations in (1.1), we obtain the heat equation for the porous medium flow:
∂tϕ(p)−div(Kμ2∇p)=0,inΩm. | (1.2) |
The flow in the conduit Ωf is described by the Navier–Stokes equations [7]:
{∂tu−div(T(u,pf))+u⋅∇u=0,inΩf,divu=0,inΩf, | (1.3) |
where u=u(x,t) and pf=pf(x,t) denote the velocity and pressure of the free flow, respectively, and μ1 is the constant viscosity. The stress tensor is given by T(u,pf)=2μ1D(u)−pfI, where D(u)=12(∇u+∇uT) represents the rate of strain tensor. The Lions condition in (1.4)1 (see [8,9]) describes the balance of forces in the normal direction, while the Beavers–Joseph–Saffman–Jones interface condition in (1.4)2 (see [10,11]) explains the relationship between shear stress and tangential velocity. Furthermore, we require the continuity of velocity at the interface as follows:
{−n⋅T(u,pf)n+12|u|2=p,onΓi,−τ⋅T(u,pf)n=μ1Gτ⋅u,onΓi,u⋅n=v⋅n,onΓi, | (1.4) |
where n denotes the unit normal vector to Ωf, τ is the tangential unit vector to Γi, and the constant G relates to the trace of the matrix K and experimental data used in the numerical analysis. Additionally, we impose a no-slip condition at the top boundary ΓU and an impermeable condition at the bottom boundary ΓL:
{u=0,onΓU,v⋅n=0,onΓL. | (1.5) |
The initial conditions are as follows:
{u(x,0)=u0(x),inΩf×{t=0},p(x,0)=p0(x),inΩm×{t=0}. | (1.6) |
Due to the extensive application of the Navier–Stokes–Darcy equation in modeling porous media coupled with free fluid flow in industrial contexts, its mathematical analysis has garnered significant attention from scholars, particularly in the field of numerical analysis. This research has yielded a wealth of results [12,13,14,15,16,17,18,19,20,21] for the case of steady porosity. When the system incorporates temperature variability, results for the Stokes–Darcy–Boussinesq equation can be found in the literature [22,23]. Recently, a series of numerical results [11,24,25,26,27] have been obtained for the Cahn–Hilliard–Navier–Stokes–Darcy equation, which addresses a two-phase miscibility phenomenon at the interface.
Although there is an abundance of numerical analysis on the time-dependent Navier-Stokes-Darcy equation, results on fundamental theories, particularly the well-posedness of strong solutions, remain limited. The strong coupling of the interface involving convection phenomena is evidently more complex than that in a single fluid (see [28,29,30]), making high-order estimates more challenging to obtain.
A comprehensive review of the selection of interface conditions for this system of equations is provided by M. McCurdy et al. [31]. This review indicates that the relationship between shear stress and tangential velocity in (1.4)1 can be replaced by the Beavers–Joseph (BJ) condition [32] or the Beavers–Joseph–Jones (BJJ) condition [33] on Γi. Additionally, the balance of forces in the normal direction of the interface in (1.4)2 can be replaced by Rankine–Hugoniot interface conditions:
−n⋅T(u,pf)n=p, in Γi. | (1.7) |
Here, we highlight several results and corresponding references related to the mathematical analysis of the coupled Navier–Stokes and Darcy equations. For the flow system involving non-deformable porous media with constant porosity ϕ(p)≡C and a permeability tensor K that is either a constant matrix or satisfies ellipticity conditions, typical mathematical analyses include those by Layton [18] and Discacciati [19]. The analysis of underground mixed-phase displacement problems has been addressed by H. Alt and S. Luckhaus [34], as well as P. Fabrie and M. Langlais [35], P. Fabrie and T. Gallouët [36], and F. Marpeau and M. Saad [37]. P. Liu and W. Liu [38] obtained global well-posedness results for the strong solution of the porous medium fluid coupled with free flow in the two-dimensional case with the BJSJ–Lions interface condition. Subsequently, M. Cui et al. [39] achieved global well-posedness results for the strong solution of the porous media fluid coupled with free flow system under BJSJ–Rankine–Hugoniot interface conditions with periodic boundary conditions, and demonstrated the attenuation of the solution. L. Tan et al. [40] extended this result to the three-dimensional case with flat domains, removing the assumption of periodic boundary conditions.
The paper is structured as follows: in Section 2, we provide the necessary definitions, hypotheses, main theorems, and preparatory lemmas. Section 3 introduces the a priori estimates and their proofs. Finally, Section 4 presents the proofs of the main theorems.
For the convenience of notation in this article, we make the following conventions:
1) The function B(p)∈L∞(0,T;L1(Ωm)) satisfying:
B(p)≜∫p0rϕ′(r) dr. | (2.1) |
2) For any T∈(0,+∞], we first define a function space X(0,T) as
X(0,T)={(u,p)|u∈L∞(0,T;H2(Ωf))∩L∞(τ,T;H4(Ωf)),ut∈L∞(0,T;L2(Ωf))∩L2(0,T;H1(Ωf))∩L∞(τ,T;H2(Ωf)),utt∈L∞(τ,T;L2(Ωf))∩L2(τ,T;H1(Ωf)),p∈L∞(0,T;H2(Ωm))∩L∞(τ,T;H4(Ωm)),pt∈L2(0,T;H1(Ωm))∩L∞(τ,T;H2(Ωm)),ptt∈L∞(τ,T;L2(Ωm))∩L2(τ,T;H1(Ωm)),∀τ∈(0,T)}, |
then (u,p)∈X(0,T) is called the strong solution of (1.2)–(1.6), if it satisfies systems (1.2) and (1.3), and a.e. in Ω×(0,T), and fulfills the conditions (1.4)–(1.6).
3) Let ‖⋅‖p,f≜‖⋅‖Lp(Ωf), ‖⋅‖p,m≜‖⋅‖Lp(Ωm), ‖⋅‖p,i≜‖⋅‖Lp(Γi), for any p, 1≤p≤+∞.
Assumption 2.1. Suppose that ϕ∈C4(R), and
sup0≤t≤T‖ϕ(n)(p)ϕ′(p)m‖∞,m≪ϵ0,1≤m<n≤4,n∈N+, | (2.2) |
where ϵ0 is small enough depends only on λ, μ2, Ωm, ‖u0‖2H2(Ωf), ‖∇p0‖2H1(Ωm) and ‖B(p0)‖L1(Ωm).
Assumption 2.2. At least one of the following two equations is true:
lim supp→∞ϕ′(p)<∞. | (2.3) |
lim infp→∞ϕ′(p)>0. | (2.4) |
We are now ready to present the main results of this paper.
Theorem 2.1 (Global existence). For u0∈H2(Ωf), B(p0)∈L1(Ωm), ∇p0∈H1(Ωm), if Assumptions 1 and 2 hold, then there exists a sufficiently small constant ϵ0, depending only on μ1, μ2, λ, Ωf, and Ωm, such that if
‖u0‖2H2(Ωf)+‖∇p0‖2H1(Ωm)+‖B(p0)‖1,m≤ϵ0, |
the coupled time-dependent Navier–Stokes–Darcy flow system (1.2)–(1.6) has a unique global strong solution (u,p) in X(0,+∞).
A few remarks are in order:
Remark 1. This paper studies two cases of ϕ(p): both linear and nonlinear dependencies. For the linear dependency, as addressed in [41,42,43,44], we consider
ϕ(p)=ˉC1p+ˉC2,(linear) | (2.5) |
where ˉC1 and ˉC2 are constants. As is well known, if (2.5) holds, the Eq (1.2) is equivalent to
∂tp−div(Kμ2∇p)=0,inΩm, | (2.6) |
which simplifies to a typical parabolic equation. According to [41,42,43,44], research on this model has thus far been limited to global existence results for weak solutions.
The nonlinear relationship, as described in [45], is given by
ϕ(p)=ϕrexp{CR(p−pr)},(nonlinear) | (2.7) |
where ϕr is the critical value of porosity when the pressure p reaches the pressure limit pr, and CR is a constant representing the deformability of the rock components, if (2.7) holds, the Eq (1.2) is equivalent to
∂tϕ(p)−div(Kμ2∇p)=0,inΩm. | (2.8) |
It appears that this is the first exploration of the theoretical results for the incompressible Navier–Stokes flow coupled with the time-dependent Darcy flow (2.8), particularly concerning the global well-posedness of strong solutions. It is noteworthy that the constraint (2.2) is applicable to both (2.6) and (2.8). When the condition n≥2 is satisfied, ϕ(n)=0 becomes sufficiently small. Consequently, it is reasonable to disregard the term involving ‖ϕ(n)(ϕ′)m‖∞,m due to its insignificance, thereby ensuring that our estimation process effectively encompasses this scenario.
Remark 2. Note that the Assumption 2.2 is consistent with the system that concerns the physical significance of the rock compressibility ratio CR, as described in (2.7). Thus, we have
ϕ(p)(n)ϕ′(p)m=ϕ1−mrCn−mRexp{(1−m)CR(p−pr)}≪ϵ0,1≤m<n≤4, |
if and only if CR≪ϵ0, which is consistent with the slight compressibility of actual underground rocks.
As is well known, the global strong solution to nonlinear partial differential equations can be obtained by combining local solutions with global a priori estimates. The local solution can be derived through higher-order regularity estimates, which are omitted here. Instead, we present the necessary a priori estimates crucial for establishing the global well-posedness of the coupled problems (1.2)–(1.6).
Let (u,p) be the strong solution to (1.2)–(1.6) satisfying u0∈H2(Ωf), B(p0)∈L1(Ωm), and ∇p0∈H1(Ωm). The basic energy estimate is stated in the following lemma.
Lemma 3.1. Under the conditions of Theorem 2.1, it holds that
sup0≤t≤T(‖u‖22,f+2‖B(p)‖1,m)+∫T0(4μ1‖D(u)‖22,f+2λμ2‖∇p‖22,m)dt≤‖u0‖22,f+2‖B(p0)‖1,m. | (3.1) |
Proof. Multiplying (1.3)1 by u and integrating the resulting equation with respect to x over Ωf, we obtain
12ddt‖u‖22,f+2μ1‖D(u)‖22,f+μ1G‖u⋅τ‖22,i+∫Γip(u⋅n)dS≤0, | (3.2) |
where we have used the interface-boundary conditions and the vector decomposition of free flow velocity at the interface Γi:
u=(u⋅ n)n+(u⋅ τ)τ. | (3.3) |
Simultaneously, we multiply (1.2) by p and integrate the resulting equation with respect to x over Ωm. Using (2.1), we obtain
ddt∫ΩmB(p) dx−∫Γip(u⋅n)dS+λμ2‖∇p‖22,m≤0. | (3.4) |
Adding (3.2) and (3.4), and then integrating the resulting equation over (0,t) gives:
‖u‖22,f+2‖B(p)‖1,m+∫t0(4μ1‖D(u)‖22,f+2μ1G‖u⋅τ‖22,i+2λμ2‖∇p‖22,m)ds≤‖u0‖22,f+2‖B(p0)‖1,m. | (3.5) |
The proof of Lemma 3.1 is complete.
The second-order estimate is obtained under the a priori assumptions. We have the following proposition.
Proposition.
There exist positive constants ϵ0 depending only on λ, μ2, and Ωm, such that if (u,p) is a smooth solution of (1.2)–(1.6) on R2×(0,T] satisfying:
sup0≤t≤T(‖∇p‖H1(Ωm)+‖p‖∞,m)≤2M, | (3.6) |
the following estimates hold:
sup0≤t≤T(‖∇p‖H1(Ωm)+‖p‖∞,m)≤M, | (3.7) |
and
sup0≤t≤T(‖u‖2H2(Ωf)+‖ut‖22,f+‖∇p‖2H1(Ωm)+‖√|ϕ′|px‖22,m+‖√|ϕ′|pt‖22,m+‖√|ϕ′|pxx‖22,m)+∫T0(‖ut‖2H1(Ωf)+‖∇u‖2H1(Ωf)+‖∇p‖2H1(Ωm)+‖√|ϕ′|pt‖22,m+‖∇pt‖22,m+‖√|ϕ′|ptx‖22,m+‖D(uxx)‖22,f+‖∇pxx‖22,m)dt≤Cϵ0, | (3.8) |
provided
‖u0‖H2(Ωf)+‖∇p0‖H1(Ωm)+‖B(p0)‖1,m≤ϵ0, |
where M=max{1,Cϵ0}, and C depends only on μ1,μ2,λ,G,Ωf, andΩm.
Proposition 3.1 is a direct consequence of the following lemmas, from Lemma 3.2 to Lemma 3.5.
Lemma 3.2. Under the conditions of Proposition 3.1, it holds that
sup0≤t≤T(‖D(u)‖22,f+‖∇p‖22,m+‖√|ϕ′|px‖22,m)+∫T0(‖ut‖22,f+‖√|ϕ′|pt‖22,m+‖∇px‖22,m+‖D(ux)‖22,f)dt≤Cϵ0, | (3.9) |
where C depends only on μ1,μ2,λ,G,Ωf,Ωm.
Proof. To estimate the second derivative term of the fluid velocity with respect to space, one must consider the fact that
‖∇2u‖2,f≤C(‖D(ux)‖2,f+‖uyy‖2,f). | (3.10) |
Next, we focus on ‖uyy‖2,f. From the Eq (1.3)1, we know
‖uyy‖2,f≤C(‖ut‖2,f+‖u⋅∇u‖2,f+‖∇pf‖2,f+‖uxx‖2,f). | (3.11) |
It is straightforward to derive, using Hölder's inequality, Young's inequality, and the Gagliardo–Nirenberg inequality [46], that
‖u⋅∇u‖2,f≤C‖u‖4,f‖∇u‖4,f≤C‖u‖122,f‖∇u‖122,f‖∇u‖122,f‖∇u‖12H1(Ωf)≤ε‖∇2u‖2,f+Cε‖u‖2,f‖∇u‖22,f, | (3.12) |
where ε is sufficiently small. For ‖∇pf‖2,f, taking the divergence of (1.3)1 and using (1.2), we can derive
−Δpf=div(u⋅∇u),inΩf×(0,T), | (3.13) |
due to (1.4)1, we have
{pf=0, on ΓU×(0,T),pf=p−12|u|2−∂xu1, on Γi×(0,T). | (3.14) |
According to the standard results (Lemma 2.5, [39]) for the Dirichlet–Neumann problem, and using Young's inequality with ε (where ε is sufficiently small), the Trace theorem, and Poincaré's inequality [47], we have
‖∇pf‖2,f≤ε‖∇2u‖2,f+Cε‖u‖2,f‖∇u‖22,f+C(‖∇p‖2,f+‖D(ux)‖2,f). | (3.15) |
Combining (3.11), (3.12), and (3.13), we obtain
‖∇2u‖2,f≤C(‖ut‖2,f+‖u‖2,f‖∇u‖22,f+‖∇p‖2,f+‖D(ux)‖2,f). | (3.16) |
Next, we will estimate it sequentially.
Step 1. (L2 estimate of D(u))
Multiply (1.3)1 by ut and integrate the resulting equation with respect to x over Ωf. Similarly, multiply (1.2) by pt and integrate the resulting equation with respect to x over Ωm. Adding the two resulting equations and applying Green's formula yields
‖ut‖22,f+‖√|ϕ′|pt‖22,m+ddt(μ1‖D(u)‖22,f+μ12G‖u⋅τ‖22,i+λ2μ2‖∇p‖22,m)≤∫Ωf|∇|u|22−u⋅∇u|⋅utdx−∫Γip(ut⋅n)dS+∫Γipt(u⋅n)dS. | (3.17) |
Now, integrating (3.17) over (0,t) and applying (3.19), Young's inequality, and Grönwall's inequality, we obtain
μ1‖D(u)‖22,f+λ4μ2‖∇p‖22,m+12∫t0(‖ut‖22,f+‖√|ϕ′|pt‖22,m)ds≤μ12∫t0‖D(ux)‖22,fds+Cϵ0+Csup0≤t≤T‖u‖22,f‖D(u0)‖22,fexp{∫t0‖D(u0)‖22,fds}≤μ12∫t0‖D(ux)‖22,fds+Cϵ0, | (3.18) |
where we have used the following estimates (3.19), derived from Hölder's inequality, Korn's inequality, Young's inequality, the Gagliardo–Nirenberg inequality, (3.1), and (3.16):
∫Ωf|∇|u|22−u⋅∇u|⋅utdx≤‖ut‖2,f‖u‖4,f‖∇u‖4,f≤18‖ut‖22,f+C‖u‖2,f‖∇u‖22,f‖∇2u‖2,f≤18‖ut‖22,f+C‖u‖2,f‖∇u‖22,f(‖ut‖2,f+‖u‖2,f‖∇u‖22,f+‖∇p‖2,f+‖D(ux)‖2,f)≤C(‖u‖22,f‖D(u)‖42,f+‖∇p‖22,f)+μ12‖D(ux)‖22,f+14‖ut‖22,f, | (3.19) |
and the following estimation obtained by Green's formula and Lemma 3.1:
∫t0(ddt∫Γip(u⋅n)dS−2∫Γip(ut⋅n)dS)ds≤−2∫t0∫Ωf∇p⋅utdxds+∫Ωf∇p⋅udx−∫Ωf∇p⋅udx|t=0≤14∫t0‖ut‖22,fds+λ4μ2‖∇p‖22,m+Cϵ0. |
Step 2. (L2 estimate of D(ux)).
Differentiate (1.3)1 with respect to x, multiply by ux, and then integrate the resulting equation with respect to x over Ωf. Similarly, differentiate (1.2) with respect to x, multiply by px, and integrate the resulting equation with respect to x over Ωm. Adding the two equations, we obtain
12ddt(‖ux‖22,f+‖√|ϕ′|px‖22,m)+2μ1‖D(ux)‖22,f+μ1G‖u⋅τ‖22,i+λμ2‖∇px‖22,m≤∫Ωf∂x(∇|u|22−u⋅∇u)⋅uxdx−12∫Ωmϕ″ptp2xdx. | (3.20) |
Thus, applying Hölder's inequality, Young's inequality, the Gagliardo–Nirenberg inequality, and Korn's inequality [48], we derive that
∫Ωf∂x(∇|u|22−u⋅∇u)⋅uxdx≤‖u‖4,f‖ux‖4,f‖∇ux‖2,f+‖∇u‖2,f‖ux‖24,f≤C(‖u‖122,f‖∇u‖122,f‖ux‖122,f‖∇ux‖322,f+‖∇u‖2,f‖ux‖2,f‖∇ux‖2,f)≤μ12‖D(ux)‖22,f+C(‖u‖22,f+1)‖D(u)‖42,f. | (3.21) |
We know from (2.2), (3.6), Hölder's inequality, Young's inequality, and the Gagliardo–Nirenberg inequality that
−12∫Ωmϕ″ptp2xdx≤12‖ϕ″|ϕ′|12‖∞,m‖√|ϕ′|pt‖2,m‖px‖24,m≤Cϵ02‖√|ϕ′|pt‖2,m‖px‖2,m‖px‖H1(Ωm)≤CMϵ0‖√|ϕ′|pt‖2,m(‖∇px‖2,m+‖px‖2,m)≤λ2μ2‖∇px‖22,m+12‖√|ϕ′|pt‖22,m+C‖∇p‖22,m, | (3.22) |
where ϵ0 is sufficiently small. We then substitute the estimates from (3.21) and (3.22) into (3.20) to
12ddt(‖ux‖22,f+‖√|ϕ′|px‖22,m)+3μ12μ1‖D(ux)‖22,f+μ1G‖u⋅τ‖22,i+λ2μ2‖∇px‖22,m≤C(‖u‖22,f+1)‖D(u)‖42,f+12‖√|ϕ′|pt‖22,m+C‖∇p‖22,m, | (3.23) |
summing it up to (3.17), and integral the resulting equation over (0,t), using Grönwall's inequality and (3.18), we obtain
μ1‖D(u)‖22,f+λ4μ2‖∇p‖22,m+12(‖ux‖22,f+‖√|ϕ′|px‖22,m)+∫t0(12‖ut‖22,f+12‖√|ϕ′|pt‖22,m+λ2μ2‖∇px‖22,m+μ1‖D(ux)‖22,f)ds≤Cϵ0. |
Therefore, the proof of Lemma 3.2 is complete.
Lemma 3.3. Under the conditions of Proposition 3.1, it holds that
sup0≤t≤T(‖ut‖22,f+‖√|ϕ′|pt‖22,m)+∫T0(‖D(ut)‖22,f+‖∇pt‖22,m)dt≤Cϵ0, | (3.24) |
where C depends only on μ1,μ2,λ,G,Ωf,Ωm.
Proof. We have, given that ϕ(p)∈C4(R) and (3.6):
‖√|ϕ′|‖∞,m≤C‖p‖∞,m≤2CM. | (3.25) |
Differentiating (1.3)1 with respect to t, multiplying by ut, and integrating the result with respect to x over Ωf, and differentiating (1.2) with respect to t, then multiplying by pt, and integrating with respect to x over Ωm, we obtain:
12ddt(‖ut‖22,f+‖√|ϕ′|pt‖22,m)+2μ1‖D(ut)‖22,f+μ1G‖ut⋅τ‖22,i+λμ2‖∇pt‖22,m≤∫Ωf(∇(∂t|u|22)−∂t(u⋅∇u))utdx+∫Ωmϕ″2p3tdx. | (3.26) |
We deduce from Hölder's inequality, Young's inequality, Gagliardo–Nirenberg inequality, and Korn's inequality [48] that:
∫Ωf(∇∂t|u|22−∂t(u⋅∇u))utdx≤‖ut‖24,f‖∇u‖2,f+‖u‖4,f‖∇ut‖2,f‖ut‖4,f≤C(‖∇u‖2,f‖ut‖2,f‖∇ut‖2,f+‖u‖122,f‖∇u‖122,f‖ut‖122,f‖∇ut‖322,f)≤μ1‖D(ut)‖22,f+Cϵ0‖ut‖22,f. | (3.27) |
Based on Young's inequality, we obtain
∫Ωmϕ″2p3tdx≤‖ϕ″2|ϕ′|32‖∞,m‖√|ϕ′|pt‖2,m‖√|ϕ′|pt‖24,m≤Cϵ02‖√|ϕ′|pt‖2,m‖√|ϕ′|pt‖24,m≤C(ϵ0‖√|ϕ′|pt‖32,m‖∇p‖H1(Ωm)+Mϵ0‖√|ϕ′|pt‖22,m‖∇pt‖2,m+ϵ0‖√|ϕ′|pt‖32,m)≤Cϵ0‖√|ϕ′|pt‖22,m(‖√|ϕ′|pt‖22,m+‖∇p‖2H1(Ωm)+1)+λ2μ2‖∇pt‖22,m. | (3.28) |
For ‖√|ϕ′|pt‖24,m, applying Hölder's inequality, Young's inequality, Gagliardo–Nirenberg inequality [46], and Korn's inequality [48], we have:
‖√|ϕ′|pt‖24,m≤C‖√|ϕ′|pt‖2,m‖∇(√|ϕ′|pt)‖2,m+‖√|ϕ′|pt‖22,m≤C‖√|ϕ′|pt‖2,m(ϵ02‖∇p‖4,m‖√|ϕ′|pt‖4,m+M‖∇pt‖2,m)+‖√|ϕ′|pt‖22,m≤C(ϵ0‖√|ϕ′|pt‖22,m‖∇p‖H1(Ωm)+M‖√|ϕ′|pt‖2,m‖∇pt‖2,m)+‖√|ϕ′|pt‖22,m+12‖√|ϕ′|pt‖24,m. | (3.29) |
From ∇2p={pxx,pxy,pyy}, we can derive, using (1.2) and (3.25), that
‖∇2p‖22,m≤C(‖∇px‖22,m+‖ϕ′pt‖22,m)≤C(‖∇px‖22,m+M‖√|ϕ′|pt‖22,m). | (3.30) |
Therefore, recalling (3.1) and (3.9), we obtain
∫t0‖∇p‖2H1(Ωm)ds≤C∫t0‖∇2p‖22,m+‖∇p‖22,m)ds≤C∫t0(‖∇px‖22,m+M‖√|ϕ′|pt‖22,m+‖∇p‖22,m)ds≤Cϵ0(M+1). | (3.31) |
Then, by integrating (3.26) over (0,t) and utilizing (3.31), (3.27), and Grönwall's inequality,
12(‖ut‖22,f+‖√|ϕ′|pt‖22,m)+∫t0(μ1‖D(ut)‖22,f+λ2μ2‖∇pt‖22,m)ds≤Cϵ0‖ut|t=0‖22,fexp{∫t0‖D(u)‖22,fds}+Cϵ0∫t0‖√|ϕ′|pt‖22,mds+Cϵ0‖√|ϕ′|pt|t=0‖22,mexp{∫t0‖√|ϕ′|pt‖22,mds}+∫t0‖∇p‖2H1(Ωm)ds}≤Cϵ0. | (3.32) |
The proof of Lemma 3.3 is complete.
Lemma 3.4. Under the conditions of Proposition 3.1, it holds that
sup0≤t≤T(‖∇2u‖22,f+‖∇2p‖22,m)+∫T0(‖utx‖22,f+‖√|ϕ′|ptx‖22,m)dt≤Cϵ0, | (3.33) |
where C depends only on μ1,μ2,λ,G,Ωf,Ωm.
Proof. Step 1. (L2 estimate of D(ux))
Differentiating (1.3)1 with respect to x, multiplying by utx, and then integrating the resulting equation with respect to x over Ωf, we apply Hölder's inequality, Young's inequality, the extension theorem [7], Gagliardo–Nirenberg inequality, and Korn's inequality [48] to obtain
‖utx‖22,f+ddt(μ1‖D(ux)‖22,f+μ12G‖ux⋅τ‖22,i)≤−∫Γipx(utx⋅n)dS+∫Ωf∂x(∇|u|22−u⋅∇u)⋅utxdx≤‖∇px‖2,m‖utx‖2,f+‖ux‖4,f‖∇u‖4,f‖utx‖2,f+‖u‖∞,f‖∇ux‖2,f‖utx‖2,f≤12‖utx‖22,f+C(‖∇px‖22,m+‖D(u)‖42,f‖D(ux)‖22,f+‖u‖22,f‖D(ux)‖42,f+‖∇2u‖22,f)≤12‖utx‖22,f+C(‖∇px‖22,m+‖D(u)‖42,f‖D(ux)‖22,f+‖u‖22,f‖D(ux)‖42,f+‖ut‖22,f+‖u‖22,f‖∇u‖42,f+‖∇p‖22,f+‖D(ux)‖22,f). | (3.34) |
Next, integrating (3.34) over (0,t), and using (3.1), (3.9) and (3.24), along with Grönwall's inequality, we obtain
∫t012‖utx‖22,fds+μ1‖D(ux)‖22,f≤Cϵ0(‖u0‖2H2(Ωf)+1). | (3.35) |
By substituting (3.1) and (3.9) into (3.16), for any t∈[0,T], we obtain:
sup0≤t≤T‖∇2u‖22,f≤C(‖ut‖22,f+‖u‖22,f‖∇u‖42,f+‖∇p‖22,f+‖D(ux)‖22,f)≤Cϵ0. | (3.36) |
Step 2. (L2 estimate of ∇px)
Differentiating (1.2) with respect to x, multiplying by ptx, and then integrating the resulting equation with respect to x over Ωm, with Hölder's inequality, Gagliardo–Nirenberg inequality, Young's inequality and the extension theorem [7] to obtain:
‖√|ϕ′|ptx‖22,m+λ2μ2ddt‖∇px‖22,m≤∫Ωmϕ″ptpxptxdx+∫Γi(ux⋅n)ptxdS≤Cϵ0‖√|ϕ′|pt‖4,m‖px‖122,m‖px‖12H1(Ωm)‖√|ϕ′|ptx‖2,m+∫Γi(ux⋅n)ptxdS≤12‖√|ϕ′|ptx‖22,m+Cϵ0‖√|ϕ′|pt‖24,m‖px‖2,m‖px‖H1(Ωm)+ddt∫Γipx(ux⋅n)dS+C(‖∇px‖22,m+‖utx‖2L2(Ωf)) | (3.37) |
with (3.9), (3.24) and (3.29), we have
‖√|ϕ′|pt‖24,m‖px‖2,m‖px‖H1(Ωm)≤CMϵ0‖√|ϕ′|pt‖2,m‖∇pt‖2,m(‖∇px‖L2(Ωm)+‖px‖L2(Ωm))+Cϵ0‖∇p‖2H1(Ωm)≤C‖√|ϕ′|pt‖22,m+CM2ϵ0‖∇pt‖22,m‖∇px‖2L2(Ωm)+CM2ϵ0‖∇pt‖22,m+Cϵ0‖∇p‖2H1(Ωm). |
Then, integrating (3.37) over (0,t), and utilizing (3.9), (3.35), the extension theorem [7], Young's inequality, and Grönwall's inequality, we obtain:
12∫t0‖√|ϕ′|ptx‖22,mds+λ4μ2‖∇px‖22,m≤C(1+M2ϵ0+‖∇2p0‖22,m+‖∇u0‖22,f)≤Cϵ0. | (3.38) |
From (3.30), (3.32), and (3.38), we have
sup0≤t≤T‖∇2p‖22,m≤C(‖∇px‖22,m+M‖√|ϕ′|pt‖22,m)≤Cϵ0. |
Thus, the proof of Lemma 3.4 is complete.
We are now in a position to prove (3.7). If (2.3) in Assumption 2.2 holds, then ‖ϕ′(p)‖∞,m≤C naturally follows, and so does ‖p‖∞,m≤C. If (2.4) holds, we employ truncation techniques to ensure that the fluid domain in the porous medium region Ωm satisfies:
Ωm={Ω1m,|p|>1,Ω2m,|p|≤1. | (3.39) |
Applying L'H\^{o}pital's rule, we have
limp→∞∫p0rϕ′(r)drp2=limp→∞pϕ′(p)2p=limp→∞ϕ′(p)2≥C,|p|>1. |
Thus, by (2.4), it follows that
‖p‖2L2(Ω1m)=∫Ω1mp2dx≤∫Ω1mB(p)dx≤C‖B(p)‖L1(Ω1m). |
Using the Gagliardo–Nirenberg inequality, we obtain
‖p‖2∞,m≤C‖∇2p‖2,m‖p‖2,m+‖p‖22,m≤C(1+‖p‖2H2(Ω1m))≤Cϵ0. |
Thus, the proof of (3.7) is complete, and Proposition 3.1 is established by synthesizing (3.9), (3.24), and (3.33).
Lemma 3.5. Under the conditions of Proposition 3.1, it holds that
sup0≤t≤T(‖uxx‖22,f+‖√|ϕ′|pxx‖22,m)+∫T0(‖D(uxx)‖22,f+‖∇pxx‖22,m)dt≤Cϵ0, | (3.40) |
where C depends only on μ1,μ2,λ,G,Ωf,Ωm.
Proof. Differentiate (1.3)1 twice with respect to x, multiply both sides of the resulting equation by uxx, and integrate over Ωf with respect to x. Similarly, differentiate (1.2) twice with respect to x, multiply both sides of the resulting equation by pxx, and integrate over Ωm with respect to x. Then, summing the two resulting equations yields:
12ddt(‖uxx‖22,f+‖√|ϕ′|pxx‖22,m)+2μ1‖D(uxx)‖22,f+μ1G‖uxx⋅τ‖22,i+λμ2‖∇pxx‖22,m≤∫Ωf∂2x(∇|u|22−u⋅∇u)⋅uxxdx−12∫Ωm(ϕ‴ptp2x+2ϕ″ptxpx+ϕ″ptpxx)⋅pxxdx. | (3.41) |
Applying Hölder's inequality, Young's inequality, the Gagliardo–Nirenberg inequality, and Korn's inequality [48], we obtain:
∫Ωf∂2x(∇|u|22−u⋅∇u)⋅uxxdx≤C(‖uxx‖24,f‖∇u‖2,f+‖ux‖4,f‖∇ux‖2,f‖uxx‖4,f+‖u‖4,f‖∇uxx‖2,f‖uxx‖4,f)≤C(‖D(uxx)‖2,f‖u‖2H2(Ωf)+‖D(uxx)‖122,f‖u‖52H2(Ωf)+‖D(uxx)‖322,f‖u‖32H2(Ωf))≤μ1‖D(uxx)‖22,f+C(‖u‖6H2(Ωf)+‖u‖4H2(Ωf)+‖u‖103H2(Ωf)). | (3.42) |
Using Hölder's inequality, Young's inequality, and the Gagliardo–Nirenberg inequality, we can derive that
−12∫Ωm(ϕ‴ptp2x+2ϕ″ptxpx+ϕ″ptpxx)⋅pxxdx≤C(‖√|ϕ′|pt‖2,m‖px‖2H1(Ωm)(‖∇pxx‖232,m‖px‖13H1(Ωm)+‖pxx‖2,m)+C‖√|ϕ′|ptx‖2,m‖px‖H1(Ωm)(‖∇pxx‖122,m‖px‖12H1(Ωm)+‖pxx‖2,m)+C‖√|ϕ′|pt‖2,m(‖∇pxx‖2,m‖px‖H1(Ωm)+‖pxx‖22,m)≤λ2μ2‖∇pxx‖22,m+C(‖√|ϕ′|ptx‖22,m+‖√|ϕ′|pt‖42,m+‖√|ϕ′|pt‖22,m+‖px‖14H1(Ωm)+‖px‖6H1(Ωm)+‖px‖4H1(Ωm)). | (3.43) |
Then, integrating (3.41) over (0,t), and using (3.1), (3.9), (3.24), (3.33), (3.42), and (3.43), we obtain:
12(‖uxx‖22,f+‖√|ϕ′|pxx‖22,m)+∫t0(μ1‖D(uxx)‖22,f+λ2μ2‖∇pxx‖22,m)≤C∫t0(‖u‖6H2(Ωf)+‖u‖4H2(Ωf)+‖u‖103H2(Ωf)+‖√|ϕ′|ptx‖22,m+‖√|ϕ′|pt‖42,m+‖√|ϕ′|pt‖22,m+‖px‖14H1(Ωm)+‖px‖6H1(Ωm)+‖px‖4H1(Ωm))ds≤Cϵ0. |
Thus, we can successfully complete the proof of Lemma 3.5.
It is time to obtain the high-order estimates. We have a refined version of (3.29) given by:
‖√|ϕ′|pt‖4,m≤C(1+‖∇pt‖2,m). | (3.44) |
Let σ(t)=min{1,t}. Then, we obtain the following higher-order estimates.
Lemma 3.6. It holds that
sup0≤t≤Tσ(t)(‖∇3u‖2L2(Ω2)+‖∇3p‖22,m+‖D(ut)‖22,f+‖∇pt‖22,m)+∫T0σ(t)(‖utxx‖22,f+‖√|ϕ′|ptxx‖22,m+‖utt‖22,f+‖√|ϕ′|ptt‖22,m)dt≤C(1+‖u0‖6H3(Ωf)+‖p0‖6H3(Ωm))≜N1, | (3.45) |
where C depends only on μ1,μ2,λ,G,Ωf,Ωm,‖u0‖H2(Ωf),‖p0‖H2(Ωm).
Proof. By considering ‖∇3u‖2,m={uxxx,uxxy,uxyy,uyyy}, we have
‖∇3u‖L2(Ω2)≤C(‖D(uxx)‖2,f+‖∇uyy‖2,f). | (3.46) |
To estimate ‖∇uyy‖2,f, we differentiate (1.3)1 to obtain
‖∇uyy‖22,f≤C(‖D(ut)‖22,f+‖∇(u⋅∇u)‖22,f+‖∇2pf‖22,f+‖D(uxx)‖22,f). |
It follows from Hölder's inequality, Young's inequality, and the Gagliardo–Nirenberg inequality that
‖∇(u⋅∇u)‖2,f≤C(‖u‖122,f‖∇2u‖322,f+‖∇u‖22,f)≤C‖u‖2H2(Ωf). | (3.47) |
For ‖∇2pf‖2,f, using results from the Dirichlet-Neumann problem (Lemma 2.5, [39]), along with Young's inequality, the Trace theorem, and Poincaré's inequality [47], we have
‖∇2pf‖2,f≤‖pf‖H2(Ωf)≤‖∇⋅(u⋅∇u)‖2,f+‖p−2μ1∂xu1‖H32(Γi)≤C(‖u‖2H2(Ωf)+‖p‖H2(Ωm)+‖∇∂xu1‖H1(Ωf))≤C(‖u‖2H2(Ωf)+‖p‖H2(Ωm)+‖uxyy‖2,f+‖D(uxx)‖2,f). | (3.48) |
Combining (3.46)–(3.48), we derive that
‖∇3u‖22,f≤C(‖D(ut)‖22,f+‖u‖4H2(Ωf)+‖uxyy‖22,f+‖D(uxx)‖22,f+‖p‖2H2(Ωm)). | (3.49) |
Next, we will estimate ‖D(ut)‖2,f, ‖uxyy‖2,f, and ‖D(uxx)‖2,f step by step.
Step 1. (L2 estimate of D(ut))
Differentiating (1.3)1 with respect to t, multiplying the resulting equation by utt, and integrating the equation with respect to x over Ωf, simultaneously, differentiating (1.2) with respect to t, multiplying the resulting equation by ptt, and integrating the equation with respect to x over Ωm, and summing up the two resulting equations, we get
‖utt‖22,f+‖√|ϕ′|ptt‖22,m+ddt(μ1‖D(ut)‖22,f+μ12G‖ut⋅τ‖22,i+λ2μ2‖∇pt‖22,m)≤∫Ωf∂t(∇|u|22−u⋅∇u)⋅uttdx−∫Γipt(utt⋅n)dS+∫Γiptt(ut⋅n)dS−∫Ωmϕ″p2tpttdx. | (3.50) |
We have the following inequalities: Hölder's inequality, Young's inequality, Gagliardo–Nirenberg inequality, and Korn's inequality [48]:
∫Ωf∂t(∇|u|22−u⋅∇u)⋅uttdx≤C(‖ut‖4,f‖∇u‖4,f+‖u‖∞,f‖∇ut‖2,f)‖utt‖2,f≤C‖ut‖122,f‖D(ut)‖122,f‖u‖H2(Ωf)+‖u‖H2(Ωf)‖ut‖H1(Ωf))‖utt‖2,f≤12‖utt‖22,f+C(‖ut‖42,f+‖u‖8H2(Ωf)+‖D(ut)‖42,f+1), | (3.51) |
and we obtain, using Hölder's inequality, Young's inequality, Gagliardo–Nirenberg inequality, (2.2), and (3.44):
−∫Ωmϕ″p2tpttdx≤‖ϕ″√|ϕ′|‖∞,m‖√|ϕ′|pt‖24,m‖√|ϕ′|ptt‖2,m≤12‖√|ϕ′|ptt‖22,m+C(‖√|ϕ′|pt‖42,m‖∇p‖4H1(Ωm)+‖∇pt‖42,m+‖√|ϕ′|pt‖42,m). | (3.52) |
Multiplying (3.50) by σ(s) and integrating the result over (0,t), we obtain, using Young's inequality, Hölder's inequality, Gagliardo–Nirenberg inequality, Grönwall's inequality, and (3.51)–(3.54), that
σ(t)(‖D(ut)‖22,f+‖∇pt‖22,m)+∫t0σ(s)(‖utt‖22,f+‖√|ϕ′|ptt‖22,m) ds≤Csup0≤t≤Tσ(t)(‖ut‖4H2(Ωf)+‖u‖8H2(Ωf)+1)+‖D(ut)|t=0‖22,fexp{∫t0‖D(ut)‖22,fds}+Csup0≤t≤T‖√|ϕ′|pt‖42,m(‖∇p‖4H1(Ωm)+1)+‖∇pt|t=0‖22,mexp{∫t0‖∇pt‖22,m}ds+C(sup0≤t≤T‖ut‖22,f+∫t0σ(s)(‖∇pt‖22,m+‖ut‖2H1(Ωf))ds)≤C(1+‖u0‖2H3(Ωf)+‖p0‖2H3(Ωm)), | (3.53) |
where we have used the extension theorem [7] to estimate the interface term as follows:
∫t0σ(s)(−∫Γipt(utt⋅n)dS+∫Γiptt(ut⋅n)dS)ds≤σ(t)(ε‖∇pt‖22,m+C‖ut‖22,f)+C∫t0σ(s)(‖∇pt‖22,m+‖ut‖22,f+‖∇pt‖22,m+η‖utt‖22,f)ds, | (3.54) |
where ε and η are sufficiently small.
Step 2. (L2 estimate of uxyy)
We differentiate (1.3)1 with respect to x to obtain:
‖uxyy‖22,f≤C(‖utx‖22,f+‖∇pfx‖22,f+‖∂x(u⋅∇u)‖22,f+‖D(uxx)‖22,f). | (3.55) |
We know from Young's inequality, the Gagliardo–Nirenberg inequality, and Korn's inequality [48] that
‖∂x(u⋅∇u)‖22,f≤C(‖∇u‖44,f+‖∇u‖2∞,f‖∇uxx‖22,f)≤C‖u‖4H2(Ωf)+‖u‖2H2(Ωf)‖D(uxx)‖22,f. | (3.56) |
For ‖∇pfx‖2,f, by taking the partial derivative of (1.3)1 with respect to x and then applying the divergence operator, and using the incompressibility condition (1.2), we obtain
−Δpfx=div(∂x(u⋅∇u)),inΩf×(0,T). | (3.57) |
Due to (1.4)1, we have the following boundary conditions:
{pfx=0, on ΓU×(0,T),pfx=px−12∂x|u|2−∂2xu1, on Γi×(0,T). | (3.58) |
Similarly to the method used to obtain (3.13), and based on the typical results (Lemma 2.5, [39]) for the Dirichlet–Neumann problem, as well as the Trace theorem and Poincaré's inequality [47], we have
‖∇pfx‖2,f≤C(‖div(∂x(u⋅∇u))‖H−1(Ωf)+‖∇px‖H12(Γi)+‖12∂x|u|2‖H12(Γi)+‖∂2xu1‖H12(Γi))≤C(‖∂x(u⋅∇u)‖2,f+‖∇px‖H1(Ωm)+‖∂2xu1‖H1(Ωm))≤C(‖∇u‖24,f+‖u‖∞,f‖∇ux‖2,f+‖∇px‖H1(Ωm)+‖∂2xu1‖H1(Ωm))≤C(‖u‖2H2(Ωf)+‖∇px‖H1(Ωm)+‖D(uxx)‖2,f). | (3.59) |
By applying the gradient to (1.2) and using (2.2), we obtain
‖∇px‖H1(Ωm)≤C(‖∇pxx‖2,m+‖∇pyy‖2,m+‖∇px‖2,m)≤C(‖∇(ϕ′pt)‖2,m+‖∇px‖2,m)≤C(‖ϕ′‖∞,m‖∇pt‖2,m+‖ϕ″√|ϕ′|‖∞,m‖∇p‖4,m‖√|ϕ′|pt‖4,m+‖∇px‖2,m)≤C(‖∇pt‖2,m+‖∇p‖2H1(Ωm)‖√|ϕ′|pt‖2,m+‖∇p‖H1(Ωm)‖∇pt‖2,m+‖∇p‖H1(Ωm)‖√|ϕ′|pt‖2,m+‖∇p‖H1(Ωm)), | (3.60) |
where we have utilized Hölder's inequality, Young's inequality, Gagliardo–Nirenberg inequality, and (3.44). Thus, combining (3.55), (3.47), (3.59), and (3.60), we conclude that
‖uxyy‖22,f≤C(‖u‖4H2(Ωf)+‖u‖4H2(Ωf)‖D(uxx)‖22,f+‖D(uxx)‖22,f+‖utx‖22,f+‖∇pt‖22,m+‖∇p‖2H1(Ωm)+‖∇p‖4H1(Ωm)‖√|ϕ′|pt‖22,m+‖∇p‖2H1(Ωm)‖√|ϕ′|pt‖22,m+‖∇p‖2H1(Ωm)‖∇pt‖22,m)≤C(‖u‖4H2(Ωf)+‖D(uxx)‖42,f+‖utx‖22,f+‖∇p‖2H1(Ωm)‖∇pt‖22,m+‖∇p‖8H1(Ωm)+‖√|ϕ′|pt‖42,m+1). | (3.61) |
In this case, the term ‖D(uxx)‖42,f is not sufficient to close the estimate. We should proceed to the next step for now.
Step 3. (L2 estimate of D(uxx))
Differentiating (1.3)1 twice with respect to x, multiplying the resulting equation by utxx, and integrating over Ωf with respect to x, and simultaneously differentiating (1.2) twice with respect to x, multiplying the resulting equation by ptxx, and integrating over Ωm with respect to x, we obtain:
‖utxx‖22,f+‖√|ϕ′|ptxx‖22,m+ddt(μ1‖D(uxx)‖22,f+μ12G‖uxx⋅τ‖22,i+λ2μ2‖∇pxx‖22,m)≤∫Ωf∂2x(∇|u|22−u⋅∇u)⋅utxxdx−∫Ωm(ϕ‴p2xpt+ϕ″pxxpt+2ϕ″pxptx)ptxxdx−∫Γipxx(utxx⋅n)dS+∫Γiptxx(uxx⋅n)dS. | (3.62) |
We employ Hölder's inequality, Young's inequality, Gagliardo–Nirenberg inequality, Korn's inequality [48], and (3.61) to obtain:
∫Ωf∂2x(∇|u|22−u⋅∇u)⋅utxxdx≤14‖utxx‖22,f+C‖u‖3H2(Ωf)(‖D(uxx)‖2,f+‖uxyy‖2,f)+C‖u‖2H2(Ωf)‖D(uxx)‖22,f≤14‖utxx‖22,f+C(‖u‖6H2(Ωf)+‖D(uxx)‖42,f+‖utx‖22,f+‖∇p‖2H1(Ωm)‖∇pt‖22,m+‖∇p‖8H1(Ωm)+‖√|ϕ′|pt‖42,m+1). | (3.63) |
Similarly, with the help of Hölder's inequality, Young's inequality, Gagliardo–Nirenberg inequality [46], (2.2), and (3.65)–(3.68), we obtain
−∫Ωm(ϕ‴p2xpt+ϕ″pxxpt+2ϕ″pxptx)ptxxdx≤12‖√|ϕ′|ptxx‖22,m+C(‖px‖46,m‖√|ϕ′|pt‖26,m+‖pxx‖24,m‖√|ϕ′|pt‖24,m+‖px‖2∞,m‖√|ϕ′|ptx‖22,m)≤12‖√|ϕ′|ptxx‖22,m+C(‖∇p‖6H1(Ωm)+‖√|ϕ′|pt‖22,m+‖∇pt‖22,m+1)‖√|ϕ′|ptx‖22,m+C‖∇pxx‖22,m(‖∇p‖2H1(Ωm)‖√|ϕ′|pt‖22,m+‖∇pt‖22,m+‖√|ϕ′|pt‖22,m)+C(‖∇p‖8H1(Ωm)+‖√|ϕ′|pt‖122,m+‖∇pt‖42,m+1). | (3.64) |
The proofs for (3.65)–(3.68) are as follows. By applying the Gagliardo–Nirenberg inequality, we have
‖px‖46,m≤C‖px‖432,m‖px‖83H1(Ωm)≤C‖∇p‖4H1(Ωm). | (3.65) |
Using the Gagliardo–Nirenberg inequality, Young's inequality, Hölder's inequality, along with (2.2) and (3.44), we obtain
‖√|ϕ′|pt‖26,m≤C‖√|ϕ′|pt‖232,m‖∇(√|ϕ′|pt)‖432,m+‖√|ϕ′|pt‖22,m≤C(‖√|ϕ′|pt‖232,m‖∇p‖434,m‖√|ϕ′|pt‖434,m+‖√|ϕ′|pt‖232,m‖∇pt‖432,m+‖√|ϕ′|pt‖22,m)≤C(‖√|ϕ′|pt‖62,m+‖∇p‖8H1(Ωm)+‖∇pt‖22,m+1). | (3.66) |
Using Young's inequality and Hölder's inequality, we obtain
‖pxx‖24,m≤C‖pxx‖2,m‖pxx‖H1(Ωm)≤C(‖∇p‖2H1(Ωm)+‖∇pxx‖22,m). | (3.67) |
By applying the Gagliardo–Nirenberg inequality, Young's inequality, and Hölder's inequality to (1.2), we obtain
‖px‖2∞,m≤C‖∇2px‖2,m‖px‖2,m+‖px‖22,m≤C‖∇(|ϕ′|pt)‖2,m‖px‖2,m+‖px‖22,m≤C(‖∇p‖6H1(Ωm)+‖√|ϕ′|pt‖22,m+‖∇pt‖22,m+1). | (3.68) |
Next, by multiplying (3.62) by σ(s) and integrating with respect to t over (0,t), and applying (3.51), (3.64), and Grönwall's inequality, we obtain
σ(t)(‖∇pxx‖22,m+‖D(uxx)‖22,f)+∫t0σ(s)(‖√|ϕ′|ptxx‖22,m+‖utxx‖22,f)ds≤C∫t0σ(s)(‖u‖6H2(Ωf)+‖D(uxx)‖42,f+‖utx‖22,f+‖∇p‖2H1(Ωm)‖∇pt‖22,m+‖∇p‖8H1(Ωm)+‖√|ϕ′|pt‖42,m+1)ds+C∫t0σ(s)(‖∇p‖6H1(Ωm)+‖√|ϕ′|pt‖22,m+‖∇pt‖22,m+1)‖√|ϕ′|ptx‖22,mds+C∫t0σ(s)‖∇pxx‖22,m(‖∇p‖2H1(Ωm)‖√|ϕ′|pt‖22,m+‖∇pt‖22,m+‖√|ϕ′|pt‖22,m)ds+C∫t0σ(s)(‖∇p‖8H1(Ωm)+‖√|ϕ′|pt‖122,m+‖∇pt‖42,m+1)ds+Cϵ0+∫t0(‖∇pxx‖22,m+‖D(uxx)‖22,f)ds≤C(1+‖u0‖3H3(Ωf)+‖p0‖3H3(Ωm)), | (3.69) |
where we have also utilized the result derived using the extension theorem [7], given as follows:
∫t0σ(s)(−∫Γipxx(utxx⋅n)dS+∫Γiptxx(uxx⋅n)dS)ds≤σ(t)(ε‖∇pxx‖22,m+C‖uxx‖22,f)+C∫t0σ(s)(‖∇pxx‖22,m+‖uxx‖22,f+‖∇pxx‖22,m+η‖utxx‖22,f)ds≤σ(t)ε‖∇pxx‖22,m+η∫t0σ(s)‖utxx‖22,fds+Cϵ0, |
where ε and η are small enough.
Therefore, with (3.8), (3.53), (3.61), (3.69), and Korn's inequality [48], we obtain
σ(t)‖uxyy‖22,f≤Cσ(t)(‖u‖4H2(Ωf)+‖D(uxx)‖42,f+‖D(ut)‖22,f+‖∇p‖2H1(Ωm)‖∇pt‖22,m+‖∇p‖8H1(Ωm)+‖√|ϕ′|pt‖42,m+1)≤C(1+‖u0‖6H3(Ωf)+‖p0‖6H3(Ωm)), | (3.70) |
so combine it with (3.49), we have:
σ(t)‖∇3u‖22,f≤Cσ(t)(‖D(ut)‖22,f+‖u‖4H2(Ωf)+‖uxyy‖22,f+‖D(uxx)‖22,f+‖p‖2H2(Ωm))≤C(1+|u0‖6H3(Ωf)+‖p0‖6H3(Ωm)). | (3.71) |
Knowing from (1.2), (3.8), (3.24), (3.30), (3.53), and (3.44), we can derive, with the aid of the Gagliardo–Nirenberg inequality, Young's inequality, and Hölder's inequality, that
σ(t)‖∇3p‖22,m≤Cσ(t)(‖∇(ϕ′pt)‖22,m≤Cσ(t)(‖ϕ″√|ϕ′|‖2∞,m‖∇p‖24,m‖|√|ϕ′|pt‖24,m+‖ϕ′‖2∞,m‖∇pt‖22,m)≤Cσ(t)(‖∇p‖4H1(Ωm)‖√|ϕ′|pt‖22,m+‖∇p‖2H1(Ωm)‖∇pt‖22,m+‖∇p‖2H1(Ωm)‖√|ϕ′|pt‖22,m+‖∇pt‖22,m)≤C(1+‖u0‖3H3(Ωf)+‖p0‖3H3(Ωm)). |
Therefore, the proof of Lemma 3.6 is complete.
We can establish Lemma 3.7 before deriving the fourth-order estimates.
Lemma 3.7. It holds that
sup0≤t≤Tσ(t)(‖utx‖2L2(Ω2)+‖√|ϕ′|ptx‖22,m+‖√|ϕ′|pt‖24,m+‖uxxx‖22,f+‖√|ϕ′|pxxx‖22,m)+∫T0σ(t)(‖D(utx)‖2L2(Ω2)+‖∇ptx‖22,m+‖√|ϕ′|ptx‖24,m+‖D(uxxx)‖22,f+‖∇pxxx‖22,m)dt≤C(1+N51). | (3.72) |
Proof. Step 1. (L2 estimate of utx)
Apply ∂t∂x to (1.3)1, then multiply the resulting equation by σ(t)utx and integrate over Ωf. Similarly, apply ∂t∂x to (1.2), then multiply the resulting equation by σ(t)ptx and integrate over Ωm. Summing the two resulting equations, we obtain the following result using Hölder's inequality, Gagliardo–Nirenberg inequality, and Young's inequality:
12ddtσ(t)(‖utx‖22,f+‖√|ϕ′|ptx‖22,m)+σ(t)(2μ1‖D(utx)‖22,f+μ1G‖utx⋅τ‖22,i+λμ2‖∇ptx‖22,m)≤σ(t)∫Ωf∂t∂x(∇|u|22−u⋅∇u)⋅utxdx−σ(t)∫Ωm(ϕ‴pxp2tptx+32ϕ″p2txpt+ϕ″pttpxptx)dx+12(‖utx‖22,f+‖√|ϕ′|ptx‖22,m)≤Cσ(t)(‖utx‖22,f‖∇u‖∞,f+‖ut‖4,f‖∇ux‖4,f‖utx‖2,f+‖ux‖4,f‖∇ut‖2,f‖utx‖4,f+‖u‖∞,f‖∇utx‖2,f‖utx‖2,f)+12(‖utx‖22,f+‖√|ϕ′|ptx‖22,m)+Cσ(t)(‖px‖∞,m‖√|ϕ′|pt‖24,m‖√|ϕ′|ptx‖2,m+‖√|ϕ′|ptx‖24,m‖√|ϕ′|pt‖2,m+‖px‖∞,m‖√|ϕ′|ptt‖2,m‖√|ϕ′|ptx‖2,m)≤Cσ(t)(μ1‖D(utx)‖22,f+λ2μ2‖∇ptx‖22,m)+Cσ(t)(‖utx‖22,f‖u‖2H3(Ωf)+‖ut‖6H1(Ωf)‖u‖2H3(Ωf)+‖utx‖22,f+‖∇p‖8H2(Ωm)+‖√|ϕ′|pt‖84,m+‖√|ϕ′|ptx‖42,m+‖∇p‖2H2(Ωm)‖√|ϕ′|ptt‖22,m+‖√|ϕ′|ptx‖22,m+1)+12(‖utx‖22,f+‖√|ϕ′|ptx‖22,m), | (3.73) |
where we have utilized (3.74) and (3.75), which were derived using Hölder's inequality, Gagliardo–Nirenberg inequality, and Young's inequality.
‖√|ϕ′|ptx‖24,m≤C(‖√|ϕ′|ptx‖2,m‖∇(√|ϕ′|ptx)‖2,m+‖√|ϕ′|ptx‖22,m)≤C(‖√|ϕ′|ptx‖2,m‖∇p‖4,m‖√|ϕ′|ptx‖4,m+‖√|ϕ′|ptx‖2,m‖∇ptx‖2,m+‖√|ϕ′|ptx‖22,m)≤ε‖√|ϕ′|ptx‖24,m+η‖∇ptx‖22,m+C(‖√|ϕ′|ptx‖22,m‖∇p‖2H1(Ωm)+‖√|ϕ′|ptx‖22,m), | (3.74) |
where ε and η are small enough. Based on (3.8) and (3.53), we can obtain that
σ(t)‖√|ϕ′|pt‖24,m≤Cσ(t)(‖√|ϕ′|pt‖22,m‖∇p‖2H1(Ωm)+‖∇pt‖22,m+‖√|ϕ′|pt‖22,m)≤N1. | (3.75) |
Thus, integrating (3.73) over (0,t), and using (3.8), (3.74), (3.75), and Grönwall's inequality, we obtain
σ(t)(‖utx‖22,f+‖√|ϕ′|ptx‖22,m)+∫t0σ(s)(‖D(utx)‖22,f+‖∇ptx‖22,m)ds≤sup0≤t≤T‖u‖2H3(Ωf)∫t0σ(s)‖utx‖22,fds+sup0≤t≤T‖u‖2H3(Ωf)sup0≤t≤T‖ut‖6H1(Ωf)+sup0≤t≤T(‖utx‖22,f+‖∇p‖8H2(Ωm)+‖√|ϕ′|pt‖84,m)+‖√|ϕ′|ptx|t=0‖22,mexp{∫t0σ(s)‖√|ϕ′|ptx‖22,mds}+sup0≤t≤T‖∇p‖2H2(Ωm)∫t0σ(s)‖√|ϕ′|ptt‖22,mds+N1≤C(1+N51). | (3.76) |
Additionally, based on equations (3.74) and (3.76), the following results can be derived:
∫t0σ(s)‖√|ϕ′|ptx‖24,mds≤C∫t0σ(s)(‖∇ptx‖22,m+‖√|ϕ′|ptx‖22,m‖∇p‖2H1(Ωm)+‖√|ϕ′|ptx‖22,m)ds≤C(1+N51). | (3.77) |
Step 2. (L2 estimate of uxxx)
By differentiating (1.3)1 three times with respect to x, multiplying both sides of the resulting equation by σ(t)uxxx, and integrating over Ωf with respect to x, and similarly, by differentiating (1.2) three times with respect to x, multiplying both sides of the resulting equation by σ(t)pxxx, and integrating over Ωm with respect to x, we obtain:
12ddtσ(t)(‖uxxx‖22,f+‖√|ϕ′|pxxx‖22,m)+σ(t)(2μ1‖D(uxxx)‖22,f+λμ2‖∇pxxx‖22,m)≤−σ(t)∫Ωm(ϕ(4)p3xpt+3ϕ(3)pxpxxpt+3ϕ(3)p2xptx+3ϕ″pxxptx+3ϕ″pxptxx+ϕ″pxxxpt)pxxxdx+σ(t)∫Ωf∂3x(∇|u|22−u⋅∇u)⋅uxxxdx+12(‖uxxx‖22,f+‖√|ϕ′|pxxx‖22,m). | (3.78) |
Applying Hölder's inequality, Young's inequality, the Gagliardo–Nirenberg inequality, and Korn's inequality [48], we obtain:
∫Ωf∂3x(∇|u|22−u⋅∇u)⋅uxxxdx≤C(‖uxxx‖24,f‖∇u‖2,f+‖uxx‖4,f‖∇ux‖2,f‖uxxx‖4,f+‖ux‖4,f‖∇uxx‖2,f‖uxxx‖4,f+‖u‖4,f‖∇uxxx‖2,f‖uxxx‖4,f)≤C(‖D(uxxx)‖2,f‖u‖2H3(Ωf)+‖D(uxxx)‖122,f‖u‖52H3(Ωf)+‖D(uxxx)‖322,f‖u‖32H3(Ωf))≤μ1‖D(uxxx)‖22,f+C(‖u‖6H3(Ωf)+1). | (3.79) |
And we obtain with (2.2):
−∫Ωm(ϕ(4)p3xpt+3ϕ(3)pxpxxpt+3ϕ(3)p2xptx+3ϕ″pxxptx+3ϕ″pxptxx+ϕ″pxxxpt)pxxxdx≤C(‖px‖3∞,m‖√|ϕ′|pt‖2,m‖pxxx‖2,m+‖px‖∞,m‖pxx‖4,m‖√|ϕ′|pt‖4,m‖pxxx‖2,m+‖px‖2∞,m‖√|ϕ′|ptx‖2,m‖pxxx‖2,m+‖pxx‖4,m‖√|ϕ′|ptx‖4,m‖pxxx‖2,m+‖px‖∞,m‖√|ϕ′|ptxx‖2,m‖pxxx‖2,m+‖pxxx‖24,m‖√|ϕ′|pt‖2,m). | (3.80) |
It can be derived with Gagliardo–Nirenberg inequality that
‖px‖∞,m≤C(‖px‖122,m‖∇2px‖122,m+‖px‖2,m)≤C‖p‖H3(Ωm), | (3.81) |
similarly, we obtain
‖pxx‖4,m≤C(‖pxx‖122,m‖∇pxx‖122,m+‖pxx‖2,m)≤C‖p‖H3(Ωm). | (3.82) |
By applying the Gagliardo–Nirenberg inequality and Young's inequality, we obtain
‖pxxx‖24,m≤C(‖pxxx‖2,m‖∇pxxx‖2,m+‖pxxx‖22,m). | (3.83) |
Then, integrating (3.78) over the interval (0,t) and utilizing (3.8), (3.75)–(3.77), (3.79), (3.80), as well as Grönwall's inequality, we obtain:
σ(t)(‖uxxx‖22,f+‖√|ϕ′|pxxx‖22,m)+∫t0σ(s)(‖D(uxxx)‖22,f+‖∇pxxx‖22,m)ds≤∫t0(‖uxxx‖22,f+‖pxxx‖22,m)ds+sup0≤t≤Tσ(t)(‖u‖6H3(Ωf)+1)+sup0≤t≤Tσ(t)‖p‖6H3(Ωm)sup0≤t≤T‖√|ϕ′|pt‖22,m+sup0≤t≤Tσ(t)‖p‖4H3(Ωm)sup0≤t≤T‖√|ϕ′|pt‖24,m+sup0≤t≤Tσ(t)‖p‖4H3(Ωm)∫t0‖√|ϕ′|ptx‖22,mds+sup0≤t≤Tσ(t)‖p‖2H3(Ωm)∫t0‖√|ϕ′|ptx‖24,mds+sup0≤t≤Tσ(t)‖p‖2H3(Ωm)∫t0‖√|ϕ′|ptxx‖22,mds+sup0≤t≤Tσ(t)‖p‖4H3(Ωm)+sup0≤t≤T‖√|ϕ′|pt‖22,msup0≤t≤Tσ(t)(1+‖p‖4H3(Ωm))≤C(1+N41). |
Thus, the proof of Lemma 3.7 is complete.
Next, it is time to get the fourth-order estimates.
Lemma 3.8. It holds that
sup0≤t≤Tσ(t)2(‖∇4u‖22,f+‖∇2ut‖22,f+‖utt‖22,f+‖∇4p‖22,m+‖∇2pt‖22,m+‖√|ϕ′|ptt‖22,m)+∫T0σ(t)2(‖D(utxxx)‖22,f+‖∇ptxxx‖22,m+‖√|ϕ′|pttx‖22,m+‖D(utt)‖22,f+‖∇ptt‖22,m)dt≤C(1+N61+‖p0‖2H4(Ωm)exp{N51}). | (3.84) |
Proof. It follows from the fact that
(3.85) |
Clearly, by applying Hölder's inequality, Young's inequality, the Gagliardo–Nirenberg inequality [46], Poincaré's inequality [47], Sobolev inequality, and Korn's inequality [48] to (3.11), we obtain:
(3.86) |
For , by leveraging the typical results (Lemma 2.5, [39]) of the Dirichlet–Neumann problem, (3.47), Young's inequality, the Trace theorem, and Poincaré's inequality [47], we have:
(3.87) |
Combining (3.85), (3.86) and (3.87), we have
(3.88) |
Next, we will estimate , , , and step by step.
Step 1. ( estimate of )
By differentiating three times with respect to , multiplying the resulting equation by , and integrating with respect to over , we obtain:
(3.89) |
We employ Hölder's inequality, Young's inequality, the Gagliardo–Nirenberg inequality, and Korn's inequality [48] to obtain:
(3.90) |
Similarly, by differentiating (1.2) three times with respect to , multiplying both sides of the resulting equation by , and integrating with respect to over , we obtain, using Hölder's inequality:
(3.91) |
Then, for , using (3.81)–(3.83), along with Hölder's inequality, Young's inequality, and the Gagliardo–Nirenberg inequality, we obtain:
(3.92) |
where we have used the fact
(3.93) |
In this case, by summing (3.89) and (3.91), multiplying the resulting equation by , and integrating with respect to time over , and utilizing (3.90), (3.92), (3.8), (3.45), and (3.72), we find that:
(3.94) |
In fact, we have applied the extension theorem [7] once more, as done in the previous steps, as follows:
where and are small enough.
Step 2. ( estimate of )
For , according to we have Taking the derivative of with respect to yields:
and
Thus, we have
(3.95) |
For , we apply to and multiply by , then integrate the resulting equation by parts with respect to . Similarly, applying to (1.2) and multiplying by , we integrate this equation by parts with respect to . Summing the results of these two integrations, and then multiplying by and integrating the resulting equation with respect to time over , using (3.75) and (3.77), we obtain:
(3.96) |
Thus, applying Hölder's inequality, Young's inequality, and the Gagliardo–Nirenberg inequality [46], we have:
(3.97) |
and using (2.2), we derive that
(3.98) |
Thus, we can derive
(3.99) |
where we have used the following estimates
where and are small enough.
To complete this step, we need to estimate . We apply to and multiply by , then integrate the resulting equation with respect to over . Similarly, we apply to (1.2) and multiply by , then integrate the resulting equation with respect to over . Finally, summing these results yields:
(3.100) |
We obtain from (3.95), Gagliardo–Nirenberg inequality, Hölder's inequality, Young's inequality, and Korn's inequality [48] that
(3.101) |
Applying the Gagliardo–Nirenberg inequality, Hölder's inequality, Young's inequality, and (2.2), we obtain:
(3.102) |
For , using the Gagliardo–Nirenberg inequality, (3.8), (3.45), and (3.72), we can derive:
For , we have with (3.8), (3.45), (3.72), and (3.75) that
Therefore, by integrating (3.100) with respect to and applying Grönwall's inequality, with the assistance of (3.45) and (3.72), we obtain:
(3.103) |
Thus, we can substitute (3.8), (3.45), (3.99), and (3.103) into (3.95) to obtain:
(3.104) |
Step 3. ( estimate of )
For , by differentiating with respect to and applying the gradient operator to the resulting equation, and utilizing Korn's inequality [48], we obtain:
From (3.87), using the typical results (Lemma 2.5, [39]) of the Dirichlet-Neumann problem, along with (3.47), Young's inequality, the Trace theorem, and Poincaré's inequality [47], we have:
(3.105) |
so we have with (3.45), (3.72), (3.94), and (3.99) that
(3.106) |
For , by applying to and using (3.45) and (3.72), we obtain:
(3.107) |
where we have used the following result:
(3.108) |
Thus, we apply (3.107) into (3.94) to obtain
(3.109) |
Using (3.108) and (3.109), it can be derived that:
(3.110) |
Applying to , with (3.45), (3.72), (3.99), and (3.109) gives
(3.111) |
And with (3.106), (3.107), and (3.111), we have
(3.112) |
From (1.2), we know Using the Gagliardo–Nirenberg inequality, Young's inequality, (3.45), and (3.99), we can derive
(3.113) |
Thus, with (3.45) (3.88), (3.99), (3.103), (3.104), (3.109), and (3.111)–(3.113), the proof of Lemma 3.8 is complete.
According to Lemmas 3.1–3.5, the local existence and uniqueness of the solution to the systems (1.2)–(1.6) on some interval is established. The focus of this study is now to verify the continuity of the strong solution in order to achieve global well-posedness.
In fact, since and , there exists such that (3.6) holds for . Next, we set:
(4.1) |
For , applying Proposition 3.1, Lemmas 3.6–3.8 implies that for any bounded time , i.e., , we have:
Therefore, it can be derived
(4.2) |
since
Now, we set
(4.3) |
By Proposition 3.1 and (3.6), with the Bootstrap principle [7] we know holds. Additionally, (4.2) implies that:
(4.4) |
(4.5) |
which means there exists such that (3.6) holds for . This contradicts the definition of in (4.1), thus .
In this study, we address the global well-posedness of a coupled Navier–Stokes–Darcy model in two-dimensional Euclidean space, incorporating the Beavers–Joseph–Saffman–Jones interface boundary condition. We establish the existence of global strong solutions for both linear and nonlinear cases where porosity depends on pressure. For the linear case, with a porosity function , the Darcy equation simplifies to a standard parabolic equation, aligning with existing research on weak solutions. In the nonlinear case, characterized by , we introduce a novel analysis of the time-dependent Darcy flow, demonstrating that our results are applicable even under complex interface conditions. The constraints considered ensure that our estimations are robust, highlighting the importance of these coupled systems for understanding underground fluid flow and providing a rigorous foundation for future research in porous media modeling and simulation. This research contributes to a deeper understanding of fluid flow phenomena in complex geological formations, offering valuable insights for both theoretical developments and practical applications in hydrogeology and related disciplines.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
We wish to acknowledge the support of the National Natural Science Foundation of China (Grant No. 11931013), and the Natural Science Foundation of Shaanxi Province (Grant No. 2023-JC-QN-0073).
The authors declare there are no conflicts of interest.
[1] | J. R. Fanchi, Principles of Applied Reservoir Simulation, Elsevier, 2005. |
[2] | J. Bear, Dynamics of Fluids in Porous Media, Courier Corporation, 1972. |
[3] |
H. Knüpfer, N. Masmoudi, Well-posedness and uniform bounds for a nonlocal third order evolution operator on an infinite wedge, Commun. Math. Phys., 320 (2013), 395–424. https://doi.org/10.1007/s00220-013-1708-z doi: 10.1007/s00220-013-1708-z
![]() |
[4] | D. A. Nield, A. Bejan, Convection in Porous Media, New York: Springer, 1992. |
[5] | H. K. Versteeg, W. Malalasekera, The finite volume method, in An Introduction to Computational Fluid Dynamics, Pearson Education, 2007. |
[6] |
S. Whitaker, Flow in porous media I: A theoretical derivation of Darcy's law, Transp. Porous Media, 1 (1986), 3–25. https://doi.org/10.1007/BF01036523 doi: 10.1007/BF01036523
![]() |
[7] | L. C. Evans, Measure Theory and Fine Properties of Functions, Routledge, 2018. https://doi.org/10.1201/9780203747940 |
[8] |
A. Çeşmelioğlu, B. Rivière, Analysis of time-dependent Navier-Stokes flow coupled with Darcy flow, J. Numer. Math., 16 (2008), 249–280. https://doi.org/10.1515/JNUM.2008.012 doi: 10.1515/JNUM.2008.012
![]() |
[9] |
A. Çeşmelioğlu, B. Rivière, Primal discontinuous Galerkin methods for time-dependent coupled surface and subsurface flow, J. Sci. Comput., 40 (2009), 115–140. https://doi.org/10.1007/s10915-009-9274-4 doi: 10.1007/s10915-009-9274-4
![]() |
[10] |
P. G. Saffman, On the boundary condition at the surface of a porous medium, Stud. Appl. Math., 50 (1971), 93–101. https://doi.org/10.1002/sapm197150293 doi: 10.1002/sapm197150293
![]() |
[11] |
D. Han, X. He, Q. Wang, Y. Wu, Existence and weak-strong uniqueness of solutions to the Cahn-Hilliard-Navier-Stokes-Darcy system in superposed free flow and porous media, Nonlinear Anal., 211 (2021), 112411. https://doi.org/10.1016/j.na.2021.112411 doi: 10.1016/j.na.2021.112411
![]() |
[12] |
V. Girault, B. Rivière, DG approximation of coupled Navier-Stokes and Darcy equations by Beaver-Joseph-Saffman interface condition, SIAM J. Numer. Anal., 47 (2009), 2052–2089. https://doi.org/10.1137/070686081 doi: 10.1137/070686081
![]() |
[13] |
M. Cai, M. Mu, J. Xu, Numerical solution to a mixed Navier-Stokes/Darcy model by the two-grid approach, SIAM J. Numer. Anal., 47 (2009), 3325–3338. https://doi.org/10.1137/080721868 doi: 10.1137/080721868
![]() |
[14] |
G. Du, L. Zuo, Local and parallel finite element method for the mixed Navier-Stokes/Darcy model with Beavers-Joseph interface conditions, Acta Math. Sci., 37 (2017), 1331–1347. https://doi.org/10.1016/S0252-9602(17)30076-0 doi: 10.1016/S0252-9602(17)30076-0
![]() |
[15] |
C. Qiu, X. He, J. Li, Y. Lin, A domain decomposition method for the time-dependent Navier-Stokes-Darcy model with Beavers-Joseph interface condition and defective boundary condition, J. Comput. Phys., 411 (2020), 109400. https://doi.org/10.1016/j.jcp.2020.109400 doi: 10.1016/j.jcp.2020.109400
![]() |
[16] |
D. Han, D. Sun, X. Wang, Two-phase flows in karstic geometry, Math. Methods Appl. Sci., 37 (2014), 3048–3063. https://doi.org/10.1002/mma.3043 doi: 10.1002/mma.3043
![]() |
[17] |
X. He, J. Li, Y. Lin, J. Ming, A domain decomposition method for the steady-state Navier-Stokes-Darcy model with Beavers-Joseph interface condition, SIAM J. Sci. Comput., 37 (2015), S264–S290. https://doi.org/10.1137/140965776 doi: 10.1137/140965776
![]() |
[18] |
W. Layton, F. Schieweck, I. Yotov, Coupling fluid flow with porous media flow, SIAM J. Numer. Anal., 40 (2003), 2195–2218. https://doi.org/10.1137/S0036142901392766 doi: 10.1137/S0036142901392766
![]() |
[19] |
M. Discacciati, E. Miglio, A. Quarteroni, Mathematical and numerical models for coupling surface and groundwater flows, Appl. Numer. Math., 43 (2002), 57–74. https://doi.org/10.1016/S0168-9274(02)00125-3 doi: 10.1016/S0168-9274(02)00125-3
![]() |
[20] |
B. Rivière, I. Yotov, Locally conservative coupling of Stokes and Darcy flows, SIAM J. Numer. Anal., 42 (2005), 1959–1977. https://doi.org/10.1137/S0036142903427640 doi: 10.1137/S0036142903427640
![]() |
[21] |
M. Discacciati, A. Quarteroni, A. Valli, Robin-Robin domain decomposition methods for the Stokes-Darcy coupling, SIAM J. Numer. Anal., 45 (2007), 1246–1268. https://doi.org/10.1137/06065091X doi: 10.1137/06065091X
![]() |
[22] |
D. Han, Q. Wang, X. Wang, Dynamic transitions and bifurcations for thermal convection in the superposed free flow and porous media, Physica D, 414 (2020), 132687. https://doi.org/10.1016/j.physd.2020.132687 doi: 10.1016/j.physd.2020.132687
![]() |
[23] |
X. Wang, H. Wu, Global weak solutions to the Navier-Stokes-Darcy-Boussinesq system for thermal convection in coupled free and porous media flows, Adv. Differ. Equations, 26 (2021), 1–44. http://doi.org/10.57262/ade/1610420433 doi: 10.57262/ade/1610420433
![]() |
[24] |
Y. Gao, D. Han, X. He, U. Rüde, Unconditionally stable numerical methods for Cahn-Hilliard-Navier-Stokes-Darcy system with different densities and viscosities, J. Comput. Phys., 454 (2022), 110968. https://doi.org/10.1016/j.jcp.2022.110968 doi: 10.1016/j.jcp.2022.110968
![]() |
[25] |
W. Chen, D. Han, X. Wang, Y. Zhang, Uniquely solvable and energy stable decoupled numerical schemes for the Cahn-Hilliard-Navier-Stokes-Darcy-Boussinesq system, J. Sci. Comput., 85 (2020), 45. https://doi.org/10.1007/s10915-020-01341-7 doi: 10.1007/s10915-020-01341-7
![]() |
[26] |
Y. Gao, X. He, L. Mei, X. Yang, Decoupled, linear, and energy stable finite element method for the Cahn-Hilliard-Navier-Stokes-Darcy phase field model, SIAM J. Sci. Comput., 40 (2018), B110–B137. https://doi.org/10.1137/16M1100885 doi: 10.1137/16M1100885
![]() |
[27] |
D. Han, X. Wang, H. Wu, Existence and uniqueness of global weak solutions to a Cahn-Hilliard-Stokes-Darcy system for two phase incompressible flows in karstic geometry, J. Differ. Equations, 257 (2014), 3887–3933. https://doi.org/10.1016/j.jde.2014.07.013 doi: 10.1016/j.jde.2014.07.013
![]() |
[28] |
C. Foias, O. Manley, R. Temam, Attractors for the Bénard problem: existence and physical bounds on their fractal dimension, Nonlinear Anal. Theory Methods Appl., 11 (1987), 939–967. https://doi.org/10.1016/0362-546X(87)90061-7 doi: 10.1016/0362-546X(87)90061-7
![]() |
[29] |
P. Fabrie, Solutions fortes et comportment asymtotique pour un modèle de convection naturelle en milieu poreux, Acta Appl. Math., 7 (1986), 49–77. https://doi.org/10.1007/BF00046977 doi: 10.1007/BF00046977
![]() |
[30] |
H. V. Ly, E. S. Titi, Global Gevrey regularity for the Bénard convection in a porous medium with zero Darcy-Prandtl number, J. Nonlinear Sci., 9 (1999), 333–362. https://doi.org/10.1007/s003329900073 doi: 10.1007/s003329900073
![]() |
[31] |
M. McCurdy, N. Moore, X. Wang, Convection in a coupled free flow-porous media system, SIAM J. Appl. Math., 79 (2019), 2313–2339. https://doi.org/10.1137/19M1238095 doi: 10.1137/19M1238095
![]() |
[32] |
G. S. Beavers, D. D. Joseph, Boundary conditions at a naturally permeable wall, J. Fluid Mech., 30 (1967), 197–207. https://doi.org/10.1017/S0022112067001375 doi: 10.1017/S0022112067001375
![]() |
[33] |
I. P. Jones, Low Reynolds number flow past a porous spherical shell, Math. Proc. Cambridge Philos. Soc., 73 (1973), 231–238. https://doi.org/10.1017/S0305004100047642 doi: 10.1017/S0305004100047642
![]() |
[34] | H. W. Alt, S. Luckhaus, Quasilinear elliptic-parabolic differential equations, Math. Z., 3 (1983), 311–342. |
[35] |
P. Fabrie, M. Langlais, Mathematical analysis of miscible displacement in porous medium, SIAM J. Math. Anal., 23 (1992), 1375–1392. https://doi.org/10.1137/0523079 doi: 10.1137/0523079
![]() |
[36] |
P. Fabrie, T. Gallouët, Modelling wells in porous media flows, Math. Models Methods Appl. Sci., 10 (2000), 673–709. https://doi.org/10.1142/S0218202500000367 doi: 10.1142/S0218202500000367
![]() |
[37] |
F. Marpeau, M. Saad, Mathematical analysis of radionuclides displacement in porous media with nonlinear adsorption, J. Differ. Equations, 228 (2006), 412–439. https://doi.org/10.1016/j.jde.2006.03.023 doi: 10.1016/j.jde.2006.03.023
![]() |
[38] |
P. Liu, W. Liu, Global well-posedness of an initial-boundary value problem of the 2-D incompressible Navier-Stokes-Darcy system, Acta Appl. Math., 160 (2019), 101–128. https://doi.org/10.1007/s10440-018-0197-7 doi: 10.1007/s10440-018-0197-7
![]() |
[39] |
M. Cui, W. Dong, Z. Guo, Global well-posedness of coupled Navier-Stokes and Darcy equations, J. Differ. Equations, 388 (2024), 82–111. https://doi.org/10.1016/j.jde.2023.12.044 doi: 10.1016/j.jde.2023.12.044
![]() |
[40] |
L. Tan, M. Cui, B. Cheng, An approach to the global well-posedness of a coupled 3-dimensional Navier-Stokes-Darcy model with Beavers-Joseph-Saffman-Jones interface boundary condition, AIMS Math., 9 (2024), 6993–7016. http://doi.org/10.3934/math.2024341 doi: 10.3934/math.2024341
![]() |
[41] |
A. Çeşmelioğlu, B. Rivière, Existence of a weak solution for the fully coupled Navier-Stokes/Darcy-transport problem, J. Differ. Equations, 252 (2012), 4138–4175. https://doi.org/10.1016/j.jde.2011.12.001 doi: 10.1016/j.jde.2011.12.001
![]() |
[42] | M. Discacciati, A. Quarteroni, Analysis of a domain decomposition method for the coupling of Stokes and Darcy equations, in Numerical Mathematics and Advanced Applications, Springer Milan, (2003), 3–20. https://doi.org/10.1007/978-88-470-2089-4_1 |
[43] |
A. Çeşmelioğlu, V. Girault, B. Rivière, Time-dependent coupling of Navier-Stokes and Darcy flows, ESAIM. Math. Model. Numer. Anal., 47 (2013), 539–554. https://doi.org/10.1051/m2an/2012034 doi: 10.1051/m2an/2012034
![]() |
[44] |
Y. Hou, D. Xue, Y. Jiang, On the weak solutions to steady-state mixed Navier-Stokes/Darcy model, Acta Math. Sin., 39 (2023), 939–951. https://doi.org/10.1007/s10114-022-9134-9 doi: 10.1007/s10114-022-9134-9
![]() |
[45] | Z. Chen, G. Huan, Y. Ma, Computational Methods for Multiphase Flows in Porous Media, Society for Industrial and Applied Mathematics, 2006. https://doi.org/10.1137/1.9780898718942 |
[46] | L. Nirenberg, On elliptic partial differential equations, Ann. Sc. Norm. Super. Pisa Cl. Sci., 13 (1959), 115–162. Available from: http://eudml.org/doc/83226. |
[47] |
H. Poincaré, Sur les equations aux dérivées partielles de la physique mathématique, Am. J. Math., 12 (1890), 211–294. https://doi.org/10.2307/2369620 doi: 10.2307/2369620
![]() |
[48] |
S. Brenner, Korn's inequalities for piecewise vector fields, Math. Comput., 73 (2004), 1067–1087. https://doi.org/10.1090/S0025-5718-03-01579-5 doi: 10.1090/S0025-5718-03-01579-5
![]() |