
This study focused on investigating the global well-posedness of a coupled Navier-Stokes-Darcy model with the Beavers-Joseph-Saffman-Jones interface boundary condition in the three-dimensional Euclidean space. By utilizing this approach, we successfully obtained the global strong solution of the system in the three-dimensional space. Furthermore, we demonstrated the exponential stability of this strong solution. The significance of such coupled systems lies in their pivotal role in the analysis of subsurface flow problems, particularly in the context of karst aquifers.
Citation: Linlin Tan, Meiying Cui, Bianru 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[J]. AIMS Mathematics, 2024, 9(3): 6993-7016. doi: 10.3934/math.2024341
[1] | Xiaoxia Wang, Jinping Jiang . The pullback attractor for the 2D g-Navier-Stokes equation with nonlinear damping and time delay. AIMS Mathematics, 2023, 8(11): 26650-26664. doi: 10.3934/math.20231363 |
[2] | Ning Duan, Xiaopeng Zhao . On global well-posedness to 3D Navier-Stokes-Landau-Lifshitz equations. AIMS Mathematics, 2020, 5(6): 6457-6463. doi: 10.3934/math.2020416 |
[3] | Yasir Nadeem Anjam . The qualitative analysis of solution of the Stokes and Navier-Stokes system in non-smooth domains with weighted Sobolev spaces. AIMS Mathematics, 2021, 6(6): 5647-5674. doi: 10.3934/math.2021334 |
[4] | Kaile Chen, Yunyun Liang, Nengqiu Zhang . Global existence of strong solutions to compressible Navier-Stokes-Korteweg equations with external potential force. AIMS Mathematics, 2023, 8(11): 27712-27724. doi: 10.3934/math.20231418 |
[5] | Mingyu Zhang . On the Cauchy problem of 3D nonhomogeneous micropolar fluids with density-dependent viscosity. AIMS Mathematics, 2024, 9(9): 23313-23330. doi: 10.3934/math.20241133 |
[6] | Kunquan Li . Analytical solutions and asymptotic behaviors to the vacuum free boundary problem for 2D Navier-Stokes equations with degenerate viscosity. AIMS Mathematics, 2024, 9(5): 12412-12432. doi: 10.3934/math.2024607 |
[7] | T. Caraballo, A. M. Márquez-Durán . Long time dynamics for functional three-dimensional Navier-Stokes-Voigt equations. AIMS Mathematics, 2020, 5(6): 5470-5494. doi: 10.3934/math.2020351 |
[8] | Jingyuan Zhang, Ruikun Zhang, Xue Lin . A stabilized multiple time step method for coupled Stokes-Darcy flows and transport model. AIMS Mathematics, 2023, 8(9): 21406-21438. doi: 10.3934/math.20231091 |
[9] | Jianxia He, Ming Li . Existence of global solution to 3D density-dependent incompressible Navier-Stokes equations. AIMS Mathematics, 2024, 9(3): 7728-7750. doi: 10.3934/math.2024375 |
[10] | Qingkun Xiao, Jianzhu Sun, Tong Tang . Uniform regularity of the isentropic Navier-Stokes-Maxwell system. AIMS Mathematics, 2022, 7(4): 6694-6701. doi: 10.3934/math.2022373 |
This study focused on investigating the global well-posedness of a coupled Navier-Stokes-Darcy model with the Beavers-Joseph-Saffman-Jones interface boundary condition in the three-dimensional Euclidean space. By utilizing this approach, we successfully obtained the global strong solution of the system in the three-dimensional space. Furthermore, we demonstrated the exponential stability of this strong solution. The significance of such coupled systems lies in their pivotal role in the analysis of subsurface flow problems, particularly in the context of karst aquifers.
The study of the coupling between free flow and porous media flow has garnered widespread attention in recent years, owing to its diverse applications in geosciences (e.g., karst aquifers, hyporheic flow, contaminant transport), health sciences (e.g., blood flow), and industrial processes; see [1,2,3,4,5] and the references therein. Insights derived from a comprehensive understanding of the Navier-Stokes-Darcy equations can be readily employed to tackle various engineering challenges. The typical mathematical analysis on the well-posedness of the associated initial boundary value problem has been done by Layton et al. [6] and Discacciati et al. [7]. The mathematical analysis of the miscible displacement problem in the subsurface was done in a seminal paper by Alt-Luckhaus [8] and by others such as Fabrie-Langlais [9], Fabrie-Gallouët [10], and Marpeau-Saad [11].
We select a microelement in the fluid-structure coupling system to consider a plane as the research object, which means that the fluid flow on the cross-section of the interface is isotropic. The schematic diagram is shown in Figure 1:
For the fluid flow in the porous medium, we employ the mass conversation law of the porosity medium and Darcy's law [12,13] to describe the system as follows:
{v=−Πμ2∇P2, in Ωm,∇⋅v=0, in Ωm, | (1.1) |
where v=v(x,t)∈R3 denotes the velocity of the flow in the porosity medium. Obviously, we can get that
−∇⋅(Πμ2∇P2)=0, in Ωm, | (1.2) |
where P2=P2(x,t)∈R3, μ2>0 represents the pressure and the viscosity of the flow in Ωm, respectively, and Π denotes the permeability tensor. We use the incompressible Navier-Stokes equations with constant viscosity μ1>0 to describe the flow in Ωf as the following equations:
{ut−∇⋅(2μ1D(u)−P1I)+u⋅∇u=0, in Ωf,∇⋅u=0, in Ωf, | (1.3) |
where Ωf⊂R3 covers the domain of the free flow, u=u(x,t)=(u1, u2, u3)∈R3 is the velocity of the free flow, P1=P1(x,t) represents the pressure of the flow in Ωf, and μ1 is the viscosity of the free flow.
The interface-boundary and initial conditions are given by
{−n1⋅(2μ1D(u)−P1I)n1=P2, on Γi,−τi⋅(2D(u))n1=α√trΠτi⋅u, on Γi,u=0, on ΓU,P2=0, on ΓL,u⋅n1=v⋅n1, on Γi,u(x,0)=u0,Pi(x,0)=P0i,i=1,2, | (1.4) |
where α is an empirically determined coefficient, τi,i={1, 2} represents two orthogonal tangent vectors in the horizontal direction, and n1 denotes the exterior unit vector normal of ∂Ωf satisfying
u=(u⋅ n1)n1+(u⋅ τ1)τ1+(u⋅ τ2)τ2, |
(1.4)1 is derived by the balance of force in the normal direction and Beavers-Joseph-Saffman-Jones interface boundary condition (1.4)2 states the shear force to the tangential stress of the fluid velocity along Γi.
We acknowledge the pioneering work of researchers who have contributed to the fields of fluid dynamics in porous media, computational methods for solving coupled equations, and the development of interface-boundary conditions [14,15,16,17,18]. For the conditions at the sharp interface, a comprehensive review of these interface selections is provided in [19]. It is known that there are three options for the shear stress conditions in the tangential velocity: The BJ (Beavers-Joseph) condition [14], the BJJ (Beavers-Joseph-Jones) condition [17], and the BJSJ (Beavers-Joseph-Saffman-Jones) condition [18], which is equivalent to what is known as BJS interface condition in some other literature such as [20]. Additionally, two choices are available for the balance of force in the normal direction at the interface: The Lions interface condition and the Rankine-Hudoniot condition.
Research in this domain primarily centers on the classical Navier-Stokes-Darcy equations with sharp interface conditions (refer to [3,6,7,21,22,23,24,25,26,27,28,29,30,31,32,33,34] and the related references), the Navier-Stokes-Darcy-Boussinesq equations involving temperature variations (refer to [35,36] and related references), and the more complex Cahn-Hilliard-Navier-Stokes-Darcy equations with interface mixing (refer to [20,37,38,39,40] and related references).
However, there have been few achievements in the mathematical analysis, especially in the case of well-posedness of strong solutions [41,42]. This is mainly due to the strong coupling of the interface, which makes it difficult to obtain high-order estimates of the system. The convection phenomenon under consideration is notably more intricate than that in a single fluid (see [43] for the free-flow and [44,45] for fluids in a porous medium).
In recent years, researchers have obtained results on non-stationary weak solutions [30,46,47,48]. Cui, Dong, and Guo [41] have studied the strong solutions and exponential decay in the two-dimensional case, and we have extended these results to the problem in the 3-D Euclidean space in this paper. Our primary objective is to conduct an initial analysis on the global well-posedness of a coupled Navier-Stokes-Darcy model in the context of the Beavers-Joseph-Saffman-Jones interface boundary condition and establishing their uniqueness property.
Definition 1. For any T∈(0,+∞], we first define a function space X(0,T) as
X(0,T)={(u,P2)|u∈L∞(0,T;H2(Ω1))∩L∞(τ,T;H4(Ω1)),ut∈L∞(0,T;L2(Ω1))∩L2(0,T;H1(Ω1))∩L∞(τ,T;H2(Ω1)),ut,t∈L∞(τ,T;L2(Ω1))∩L2(τ,T;H1(Ω1)),P2∈L∞(0,T;H3(Ω2))∩L∞(τ,T;H5(Ω2)),∂tP2∈L2(0,T;H1(Ω2))∩L∞(τ,T;H1(Ω2)),∂2tP2∈L2(τ,T;H1(Ω2)),∀τ∈(0,T)}. |
(u,P2)∈X(0,T) is called the strong solution of (1.2)–(1.4), if it satisfies systems (1.2) and (1.3) a.e. in Ω×(0,T), and fulfills the conditions (1.2)–(1.4).
Next, we present the main result of this paper.
Proposition 1.1. (Local well-posedness) Let μ1,μ2 both be positive constants and assume that
λ∣ξ∣2≤Πξ⋅ξ≤Λ∣ξ∣2,∀ξ∈R3, | (1.5) |
the initial data u0∈H2(Ωf) is divergence free, and the compatibility condition holds as follows
u0t=∇⋅(2μ1D(u0)−P01I)−u0⋅∇u0, | (1.6) |
where u0t=ut∣t=0 and P01=P1∣t=0. There exists a time T>0 such that the 3-D coupled Navier-Stokes-Darcy systems (1.2)–(1.4) have a unique strong solution (u,P2)∈X(0,T).
Theorem 1.2. (Global well-posedness) The permeability tensor Π satisfies (1.5) and the initial divergence free velocity field u0∈H2(Ωf) satisfies the compatibility condition (1.6), then there exists a positive constant C depending only on μ1,Ωf, and λ, such that if
‖u0‖L2(Ωf)≤ϵ0=min{μ21C2M,116C2M,18C}, M=max{1,√C‖∇u0‖L2(Ωf)}, |
the three-dimensional coupled Navier-Stokes-Darcy systems (1.2)–(1.4) have a unique global strong solution (u,P2)∈X(0,+∞) as described in Definition 1.
Moreover, the solution (u,P2) has a decay rate
‖u‖Hk(Ωf)+‖P2‖Hk(Ωm)≤Ce−ct, ∀k≥0, | (1.7) |
where the positive constants are C=C(μ1,λ,Ωf,‖u0‖H2(Ωf)) and c=c(μ1,Ωf).
Remark 1. In divergence from the findings of [41,42], this paper makes two primary contributions. First, in terms of analytical techniques, the involvement of more intricate directional derivatives under the three-dimensional model renders estimations challenging and convoluted. Second, regarding outcomes, our research is centered on a strip domain, breaking away from the conventional assumption of periodicity. This marks a pioneering achievement as the first three-dimensional outcome in the rigorous examination of robust solutions for the Navier-Stokes-Darcy system. Notably, our results extend beyond, demonstrating applicability, even in the context of periodic domains.
Furthermore, ϵ0 in Theorem 1.2 will depend on ‖∇u0‖L2(Ωf).
Remark 2. The decay rate obtained in (1.7) indicates that after time t>0, the solution (u,P2) is smooth, and all its derivatives decay for any order.
The existence and uniqueness of local solutions to this problem can be obtained similarly to the approach in [41]. Therefore, our subsequent focus will be on the a priori estimates of the global solution.
As is well known, the global strong solution to the nonlinear partial differential equations can be obtained by combining local solutions with global a priori estimates. The local solution is proved similarly to that in [41] and is omitted here. Instead, we present the crucial a priori estimates pivotal to establishing global well-posedness below. Note that \(K = \frac{\Pi}{\mu_{2}}\) and \(\mathcal{W} = \frac{\mu_{1}\alpha}{\sqrt{\text{tr}\, \Pi}}\) for convenience in the subsequent discussion.
Proposition 2.1. If (u,P2) is a smooth solution of the Navier-Stokes-Darcy systems (1.3) and (1.4) satisfying
sup0≤t≤T‖u‖L2(Ωf)≤2‖u0‖L2(Ωf), sup0≤t≤T‖∇u‖L2(Ωf)≤2M, | (2.1) |
then the following estimates hold:
sup0≤t≤T‖u‖L2(Ωf)≤‖u0‖L2(Ωf),sup0≤t≤T‖∇u‖L2(Ωf)≤M, | (2.2) |
and
sup0<t<T(‖u‖2H2(Ωf)+‖ut‖2H2(Ωf)+‖P2‖2H3(Ωm))+∫T0(‖ut‖2H1(Ωf)+‖∇u‖2H1(Ωf)+‖∇∂tP2‖2L2(Ωm)) dt≤C(‖u0‖2H2(Ωf)+1)exp{‖u0‖2H1(Ωf)+‖u0‖4H1(Ωf)}, | (2.3) |
provided
‖u0‖L2(Ωf)≤ϵ0=min{μ21C2M,116C2M,18C}, M=max{1,√C‖∇u0‖L2(Ωf)}, |
where C depends only on μ1,Ωf, and λ.
The proof of Proposition 2.1 can be successfully summarized by the following Lemmas 2.1–2.4.
Lemma 2.1. Under the conditions of Proposition 2.1, it holds that
sup0≤t≤T‖u‖2L2(Ωf)+2μ1∫T0‖D(u)‖2L2(Ωf) dt≤‖u0‖2L2(Ωf). | (2.4) |
Proof. Multiplying (1.3)1 by u and integrating the result equation on Ωf, then multiplying (1.2) by P2 and integrating it on Ωm, we add up the two resulting equations to have
2μ1‖D(u)‖2L2(Ωf)+W2∑i=1‖u⋅τi‖2L2(Γi)+12ddt‖u‖2L2(Ωf)+λμ2‖∇P2‖2L2(Ωm)≤−∫Ωf(u⋅∇u)⋅u dx≤‖u‖L3(Ωf)‖∇u‖L2(Ωf)‖u‖L6(Ωf)≤C‖u‖12L2(Ωf)‖∇u‖12L2(Ωf)‖∇u‖2L2(Ωf)≤C‖u0‖12L2(Ωf)M12‖D(u)‖2L2(Ωf)≤μ1‖D(u)‖2L2(Ωf), | (2.5) |
where we have used Hölder's inequality, Young's inequality, Gagliardo-Nirenberg inequality, Korn's inequality, and (2.1). Here C depends on the domain Ωf. By integrating (2.5) over (0,t), one gets (2.4) with ϵ0≤μ21C2M and (2.1).
Lemma 2.2. Under the conditions of Proposition 2.1, it holds that
sup0≤t≤T‖D(u)‖2L2(Ωf)+∫T0(‖ut‖2L2(Ωf)+‖∇2u‖2L2(Ωf)) dt≤C‖∇u0‖2L2(Ωf). | (2.6) |
Proof. Multiply (1.3)1 by ut and integrate it over Ωf, while differentiating (1.2) with respect to t, then multiply by P2 and integrate it over Ωm. Now, summing up the two resulting equations yields with Hölder's inequality, Gagliardo-Nirenberg inequality, Young's inequality, Korn's inequality, and (2.1):
‖ut‖2L2(Ωf)+ddt(μ1‖D(u)‖2L2(Ωf)+W22∑i=1‖u⋅τi‖2L2(Γi)+12∫ΩmK∇P2⋅∇P2 dx)≤−∫Ωf(u⋅∇u)⋅ut dx≤‖u‖L3(Ωf)‖ut‖L2(Ωf)‖∇u‖L6(Ωf)≤C‖u‖12L2(Ωf)‖∇u‖12L2(Ωf)(‖∇2u‖L2(Ωf)+‖∇u‖L2(Ωf))‖ut‖L2(Ωf)≤C‖u0‖12L2(Ωf)M12(‖∇2u‖L2(Ωf)+‖∇u‖L2(Ωf))‖ut‖L2(Ωf)≤14‖ut‖2L2(Ωf)+C‖u0‖L2(Ωf)M‖∇2u‖2L2(Ωf)+C‖u0‖L2(Ωf)M‖∇u‖2L2(Ωf)≤14‖ut‖2L2(Ωf)+C‖D(u)‖2L2(Ωf)+C‖∇2u‖2L2(Ωf), | (2.7) |
where C depends on μ1 and Ωf.
Next, we come to estimate ‖∇2u‖L2(Ωf). The fact that
‖∇2u‖L2(Ωf)≤C(‖D(ux)‖L2(Ωf)+‖D(uy)‖L2(Ωf)+‖uz,z‖L2(Ωf)) | (2.8) |
tells us to deduce the estimations on ‖D(ux)‖L2(Ωf) minutely, we omit the details of the estimations on ‖D(uy)‖L2(Ωf) due to the symmetry in the horizontal direction. We know from (2.8) and (2.23)–(2.28) that
‖∇2u‖L2(Ωf)≤C(‖D(ux)‖L2(Ωf)+‖D(uy)‖L2(Ωf)+‖ut‖L2(Ωf)+‖D(u)‖L2(Ωf)+‖u⋅∇u‖L2(Ωf))≤C(‖D(ux)‖L2(Ωf)+‖D(uy)‖L2(Ωf)+‖ut‖L2(Ωf)+‖D(u)‖L2(Ωf)+‖u‖L3(Ωf)‖∇u‖L6(Ωf))≤C(‖D(ux)‖L2(Ωf)+‖D(uy)‖L2(Ωf)+‖ut‖L2(Ωf)+‖D(u)‖L2(Ωf)+‖u‖12L2(Ωf)‖∇u‖12L2(Ωf)‖∇u‖H1(Ωf))≤C‖u0‖12L2(Ωf)M12‖∇u‖H1(Ωf)+C(‖D(ux)‖L2(Ωf)+‖D(uy)‖L2(Ωf)+‖ut‖L2(Ωf)+‖D(u)‖L2(Ωf)), |
thus, we can get from C‖u0‖12L2(Ωf)M12≤12 that
‖∇2u‖L2(Ωf)≤C(‖D(ux)‖L2(Ωf)+‖D(uy)‖L2(Ωf)+‖ut‖L2(Ωf)+‖D(u)‖L2(Ωf)). | (2.9) |
Next, taking the partial derivative of (1.3)1 with respect to x, we get
ut,x−∇⋅(2μ1D(ux)−∂xP1I)+∂x(u⋅∇u)=0 in Ωf×(0,T). | (2.10) |
Multiplying (2.10) by ux and integrating the result inequality with respect to x over Ωm, then using Hölder's inequality, Gagliardo-Nirenberg inequality, Young's inequality, Korn's inequality, and (2.1), one can get
12ddt‖ux‖2L2(Ωf)+2μ1‖D(ux)‖2L2(Ωf)+W2∑i=1‖ux⋅τi‖2L2(Γi)+λμ2‖∇∂xP2‖2L2(Ωm)≤−∫Ωf∂x(u⋅∇u)⋅ux dx≤−∫Ωf((ux⋅∇u)⋅ux+(u⋅∇ux)⋅ux) dx≤‖u‖L3(Ωf)‖ux‖L6(Ωf)‖∇ux‖L2(Ωf)+∫Γi−K∇P2⋅n1(u⋅ux) dS≤C‖u‖12L2(Ωf)‖∇u‖12L2(Ωf)‖∇ux‖2L2(Ωf)+C‖∇∂xP2‖L4(Γi)‖u‖L2(Γi)‖ux‖L4(Γi)≤C‖u0‖12L2(Ωf)M12‖D(ux)‖2L2(Ωf)+C‖∇u‖H1(Ωf)(‖u‖L2(Ωf)+‖u‖12L2(Ωf)‖∇u‖12L2(Ωf))‖ux‖H1(Ωf)≤C‖u0‖12L2(Ωf)M12‖D(ux)‖2L2(Ωf)+C‖∇u‖H1(Ωf)(‖u0‖L2(Ωf)+‖u0‖12L2(Ωf)M12)≤C‖∇u‖2H1(Ωf)(‖u0‖L2(Ωf)+‖u0‖12L2(Ωf)M12), | (2.11) |
and, similarly, we have
12ddt‖uy‖2L2(Ωf)+2μ1‖D(uy)‖2L2(Ωf)+W2∑i=1‖uy⋅τi‖2L2(Γi)+λμ2‖∇∂yP2‖2L2(Ωm)≤C‖∇u‖2H1(Ωf)(‖u0‖L2(Ωf)+‖u0‖12L2(Ωf)M12). | (2.12) |
Therefore, plugging (2.9) into (2.7) and combining (2.1), (2.11), and (2.12), one gets
ddt(μ1‖D(u)‖2L2(Ωf)+W22∑i=1‖u⋅τi‖2L2(Γi)+12∫ΩmK∇P2⋅∇P2 dx+12‖ux‖2L2(Ωf)+12‖uy‖2L2(Ωf))+34‖ut‖2L2(Ωf)+2μ1‖D(ux)‖2L2(Ωf)+2μ1‖D(uy)‖2L2(Ωf)≤C(‖u0‖L2(Ωf)+‖u0‖12L2(Ωf)M12+‖u0‖L2(Ωf)M)‖∇2u‖2L2(Ωf)+C‖D(u)‖2L2(Ωf)≤12(34‖ut‖2L2(Ωf)+2μ1‖D(ux)‖2L2(Ωf)+2μ1‖D(uy)‖2L2(Ωf))+C‖D(u)‖2L2(Ωf), | (2.13) |
where we have used C(‖u0‖L2(Ωf)+‖u0‖12L2(Ωf)M12+‖u0‖L2(Ωf)M)≤12. Integrating (2.13) over (0,t), we can obtain (2.6) with Lemma 2.1, Young's inequality, and
ϵ0=min{μ21C2M,116C2M,18C}, M=max{1,√C‖∇u0‖L2(Ωf)}. |
Thus, the proof of Lemma 2.2 is completed.
While combining Lemmas 2.1 and 2. 2, we have
sup0≤t≤T‖∇u‖L2(Ωf)≤√C‖∇u0‖L2(Ωf)≤M. |
Therefore, the proof of (2.2) is complete.
Lemma 2.3. Under the conditions of Proposition 2.1, it holds that
sup0≤t≤T‖ut‖2L2(Ωf)+∫T0(‖D(ut)‖2L2(Ωf)+‖∇∂tP2‖2L2(Ωm)) dt≤C‖u0‖2H2(Ωf)exp{‖u0‖4H1(Ωf)}. | (2.14) |
Proof. Differentiate (1.3)1 with respect to t, multiply by ut, then integrate over Ωf. Differentiate (1.2) with respect to t, multiply by ∂tP2, then integrate over Ωm. Adding up the two result equations, we obtain that
12ddt‖ut‖2L2(Ωf)+2μ1‖D(ut)‖2L2(Ωf)+W2∑i=1‖ut⋅τi‖2L2(Γi)+λμ2‖∇∂tP2‖2L2(Ωm)≤−∫Ωf∂t(u⋅∇u)⋅ut dx≤C‖∇u‖L2(Ωf)‖ut‖2L4(Ωf)+C‖u‖L6(Ωf)‖ut‖L3(Ωf)‖∇ut‖L2(Ωf)≤C‖∇u‖L2(Ωf)‖ut‖12L2(Ωf)‖∇ut‖32L2(Ωf)≤μ1‖D(ut)‖2L2(Ωf)+C‖D(u)‖4L2(Ωf)‖ut‖2L2(Ωf), | (2.15) |
where we have used Hölder's inequality, Young's inequality, Gagliardo-Nirenberg inequality, and Korn's inequality, then we can obtain by Grönwall's inequality that
sup0≤t≤T‖ut‖2L2(Ωf)+∫T0(‖D(ut)‖2L2(Ωf)+‖∇∂tP2‖2L2(Ωm)) ds≤‖ut,0‖2L2(Ωf)exp{C∫T0‖D(u)‖4L2(Ωf) ds}≤C‖u0‖2H2(Ωf)exp{sup0≤t≤T‖D(u)‖2L2(Ωf)∫T0‖D(u)‖2L2(Ωf) ds}≤C‖u0‖2H2(Ωf)exp{C‖∇u0‖2L2(Ωf)‖u0‖2L2(Ωf)}≤C‖u0‖2H2(Ωf)exp{‖u0‖4H1(Ωf)}. |
The proof is completed with the supports of Lemmas 2.1 and 2.2.
Lemma 2.4. Under the conditions of Proposition 2.1, it holds that
sup0≤t≤T(‖∇2u‖2L2(Ωf)+‖P2‖2H3(Ωm))+∫T0(‖ut,x‖2L2(Ωf)+‖∇3u‖2L2(Ωm)) dt≤C(‖u0‖2H2(Ωf)+1)exp{‖u0‖2H1(Ωf)}. | (2.16) |
Proof. Put ∂x on (1.3)1 and multiply ut,x by both sides of it, then integrate it on Ωf; meanwhile, apply ∂x∂t to (1.2), then multiply by ∂xP2 and integrate it on Ωm. Adding up the two resulting equations gives that
ddt(μ1‖D(ux)‖2L2(Ωf)+W22∑i=1‖ux⋅τi‖2L2(Γi)+12∫ΩmK∇P2,x⋅∇P2,x dx)+‖ut,x‖2L2(Ωf)≤−∫Ωf∂x(u⋅∇u)⋅ut,x dx≤12‖ut,x‖2L2(Ωf)+12‖ux‖2L6(Ωf)‖∇u‖2L3(Ωf)+C‖u‖2L∞(Ωf)‖∇ux‖2L2(Ωf)≤12‖ut,x‖2L2(Ωf)+C‖∇ux‖2L2(Ωf)‖∇u‖2L3(Ωf)+C‖u‖2L∞(Ωf)‖∇ux‖2L2(Ωf)≤12‖ut,x‖2L2(Ωf)+C‖∇ux‖2L2(Ωf)(‖∇u‖L2(Ωf)‖∇2u‖L2(Ωf)+‖∇u‖2L2(Ωf))+C(‖u‖12L2(Ωf)‖∇2u‖32L2(Ωf)+‖u‖2L2(Ωf))‖∇ux‖2L2(Ωf)≤12‖ut,x‖2L2(Ωf)+C‖∇ux‖2L2(Ωf)(‖∇u‖2L2(Ωf)+‖∇2u‖2L2(Ωf)+‖u‖2L2(Ωf))≤12‖ut,x‖2L2(Ωf)+C‖D(ux)‖2L2(Ωf)(‖D(u)‖2L2(Ωf)+‖∇2u‖2L2(Ωf)), | (2.17) |
where we have used Poincaré's inequality, Hölder's inequality, Young's inequality, Gagliardo-Nirenberg inequality, and Korn's inequality. Similarly, we can derive that
ddt(μ1‖D(uy)‖2L2(Ωf)+W22∑i=1‖ux⋅τi‖2L2(Γi)+12∫ΩmK∇P2,y⋅∇P2,y dx)+‖ut,y‖2L2(Ωf)≤12‖ut,y‖2L2(Ωf)+C‖D(uy)‖2L2(Ωf)(‖D(u)‖2L2(Ωf)+‖∇2u‖2L2(Ωf)), | (2.18) |
then we add up (2.17) and (2.18), and use Grönwall's inequality, Lemma 2.1, and Lemma 2.2 to get
sup0≤t≤T(‖D(ux)‖2L2(Ωf)+‖D(uy)‖2L2(Ωf))+∫T0(‖ut,x‖2L2(Ωf)+‖ut,y‖2L2(Ωf)) ds≤C‖u0‖2H2(Ωf)exp{∫T0(‖D(u)‖2L2(Ωf)+‖∇2u‖2L2(Ωf)) ds}≤C‖u0‖2H2(Ωf)exp{‖u0‖2L2(Ωf)+‖∇u0‖2L2(Ωf)}. | (2.19) |
Therefore, know from (2.9) and (2.19) that
sup0≤t≤T‖uz,z‖2L2(Ωf)≤C∫T0(‖D(ux)‖2L2(Ωf)+‖D(uy)‖2L2(Ωf)+‖ut‖2L2(Ωf)+‖D(u)‖2L2(Ωf)) ds≤C‖u0‖2H2(Ωf)exp{‖u0‖2L2(Ωf)+‖∇u0‖2L2(Ωf)}+‖u0‖2L2(Ωf)+‖∇u0‖2L2(Ωf)≤C(‖u0‖2H2(Ωf)+1)exp{‖u0‖2H1(Ωf)}, | (2.20) |
and it is easy to get that (see [42])
‖∇P2‖2Hk(Ωm)≤C‖u‖2Hk−1(Ωf),∀k≥2, | (2.21) |
thus, the proof of Lemma 2.4 is complete with (2.4), (2.6), (2.9), (2.14), (2.19), and (2.20).
The proof of Proposition 2.1 has been finished. Next, we will proceed with the higher-order estimation involving the time weighting.
Let σ(t)=min{1,t} and, from now on, the generic positive constant is defined by the right term of (2.3) as N≜C(‖u0‖2H2(Ωf)+1)exp{‖u0‖2H1(Ωf)+‖u0‖4H1(Ωf)}.
We have the following third-order estimates.
Lemma 2.5. It holds that
sup0≤t≤Tσ(t)(‖∇3u‖2L2(Ωf)+‖D(ut)‖2L2(Ωf)+‖∇∂tP2‖2L2(Ωm)+‖∇4P2‖2L2(Ωm))+∫T0σ(t)(‖D(ut,x)‖2L2(Ωf)+‖D(ut,y)‖2L2(Ωf)+‖ut,t‖2L2(Ωf)+‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,x,y)‖2L2(Ωf)+‖D(ux,y,y)‖2L2(Ωf)+‖D(uy,y,y)‖2L2(Ωf)+‖∇∂t∂xP2‖2L2(Ωm)+‖∇∂t∂yP2‖2L2(Ωm)) dt≤C(‖u0‖10H2(Ωf)+1)exp{‖u0‖2H1(Ωf)+‖u0‖4H1(Ωf)}. | (2.22) |
Proof. It follows from the fact that
‖∇3u‖L2(Ωf)≤C(‖D(ux,x)‖L2(Ωf)+‖D(ux,y)‖L2(Ωf)+‖D(uy,y)‖L2(Ωf)+‖∇uz,z‖L2(Ωf)). |
First, we focus on uz,z. By (1.3)1, we have
‖uz,z‖2L2(Ωf)≤‖1μ1(ut+u⋅∇u+∇P1−ux,x−uy,y)‖2L2(Ωf)≤C(‖ut‖2L2(Ωf)+‖u⋅∇u‖2L2(Ωf)+‖∇P1‖2L2(Ωf)+‖ux,x‖2L2(Ωf)+‖uy,y‖2L2(Ωf)), | (2.23) |
and taking divergence to (1.3)1, the elliptic problem can be obtained as follows:
−ΔP1=∇⋅(u⋅∇u) in Ωf×(0,T). | (2.24) |
With the boundary condition obtained by (1.3)2 and (1.4)1, we have
{P1⋅n1=0 on ΓU,P1=P2+2μ1∂zu3=P2−2μ1∂xu1−2μ1∂yu2 on Γi. | (2.25) |
It is clear to show by Lemma 2.5 [41] and the Trace theorem that
‖∇P1‖2L2(Ωf)≤C‖P1‖2H1(Ωf)≤C(‖∇⋅(u⋅∇u)‖2H−1(Ωf)+‖P2−2μ1∂xu1−2μ1∂yu2‖2H12(Γi))≤C(‖u⋅∇u‖2L2(Ωf)+‖∇P2‖2L2(Ωm)+‖∂xu1‖2H1(Ωf)+‖∂yu2‖2H1(Ωf))≤C(‖u⋅∇u‖2L2(Ωf)+‖∇P2‖2L2(Ωm)+‖D(∂xu1)‖2L2(Ωf)+‖D(∂yu2)‖2L2(Ωf))≤C(‖u⋅∇u‖2L2(Ωf)+‖∇P2‖2L2(Ωm)+‖D(ux)‖2L2(Ωf)+‖D(uy)‖2L2(Ωf)). | (2.26) |
The estimation of ‖u⋅∇u‖2L2(Ωf) can be estimated by Hölder's inequality, Gagliardo-Nirenberg inequality, Young's inequality, and Korn's inequality as follows:
‖u⋅∇u‖2L2(Ωf)≤‖u‖2L∞(Ωf)‖∇u‖2L2(Ωf)≤C‖∇u‖L2(Ωf)‖∇2u‖L2(Ωf)‖∇u‖2L2(Ωf)≤C(ϵ)‖D(u)‖6L2(Ωf)+ϵ‖∇2u‖2L2(Ωf). | (2.27) |
Thus, we can derive with (2.23), (2.27), and (2.21), by Young's inequality, that
‖∇u‖2H1(Ωf)≤C(‖ut‖2L2(Ωf)+‖D(ux)‖2L2(Ωf)+‖D(uy)‖2L2(Ωf)+‖D(u)‖6L2(Ωf)+1). | (2.28) |
Obviously, from (2.23) with Hölder's inequality, Young's inequality, Gagliardo-Nirenberg inequality, Poincáre's inequality, Sobolev inequality, and Korn's inequality we have
‖∇uz,z‖L2(Ωf)≤C(‖∇ut‖L2(Ωf)+‖∇(u⋅∇u)‖L2(Ωf)+‖∇2P1‖L2(Ωf)+‖∇ux,x‖L2(Ωf)+‖∇uy,y‖L2(Ωf))≤C(‖D(ut)‖L2(Ωf)+‖D(ux,x)‖L2(Ωf)+‖D(uy,y)‖L2(Ωf)+‖D(ux,y)‖L2(Ωf)+‖ux,z,z‖L2(Ωf)+‖uy,z,z‖L2(Ωf)+‖u‖2H2(Ωf)), |
where we have used the estimation obtained based on (2.27) as follows:
‖u⋅∇u‖H1(Ωf)≤‖u‖L∞(Ωf)‖∇2u‖L2(Ωf)+‖∇u‖2L4(Ωf)+‖u⋅∇u‖L2(Ωf)≤‖∇u‖12L2(Ωf)‖∇2u‖32L2(Ωf)+‖∇u‖2H1(Ωf)≤C‖u‖2H2(Ωf), | (2.29) |
and, again, using Lemma 2.5 [41] together with the Trace theorem to get that
‖∇2P1‖L2(Ωf)≤‖P1‖H2(Ωf)≤‖∇⋅(u⋅∇u)‖L2(Ωf)+‖P2−2μ1∂xu1−2μ1∂yu2‖H32(Γi)≤C(‖u‖2H2(Ωf)+‖P2‖2H2(Ωm)+‖∇∂xu1‖H1(Ωf)+‖∇∂yu2‖H1(Ωf))≤C(‖D(ux,x)‖L2(Ωf)+‖D(uy,y)‖L2(Ωf)+‖D(ux,y)‖L2(Ωf)+‖ux,z,z‖L2(Ωf)+‖uy,z,z‖L2(Ωf)+‖u‖2H2(Ωf)), | (2.30) |
such that we have
‖∇3u‖2L2(Ωf)≤C(‖D(ut)‖2L2(Ωf)+‖D(ux,x)‖2L2(Ωf)+‖D(uy,y)‖2L2(Ωf)+‖D(ux,y)‖2L2(Ωf)+‖ux,z,z‖2L2(Ωf)+‖uy,z,z‖2L2(Ωf)+‖u‖4H2(Ωf)). | (2.31) |
Next, we will get the bound of each term at the righthand of (2.31) step by step.
Step 1. Next, we apply ∂x∂y to (1.3)1 and multiply ut,x,y before integrating the result equations over Ωf; meanwhile, we apply ∂x∂y to (1.2) and integrate it over Ωm after multiplying P2,x,y on it. Finally, summing up the two resulting equations would come to (2.32), with the help of Hölder's inequality, Gagliardo-Nirenberg inequality, Young's inequality, and Korn's inequality,
ddt(μ1‖D(ux,y)‖2L2(Ωf)+W22∑i=1‖ux,y⋅τi‖2L2(Γi)+12∫ΩmK∇P2,x,y⋅∇P2,x,y dx)+‖ut,x,y‖2L2(Ωf)≤−∫Ωf∂x∂y(u⋅∇u)⋅ut,x,y dx≤12‖ut,x,y‖2L2(Ωf)+‖∇u‖2L∞(Ωf)‖ux,y‖2L2(Ωf)+‖ux‖2L∞(Ωf)‖∇uy‖2L2(Ωf)+‖uy‖2L∞(Ωf)‖∇ux‖2L2(Ωf)+‖u‖2L∞(Ωf)‖∇ux,y‖2L2(Ωf)≤12‖ut,x,y‖2L2(Ωf)+‖ux,y‖2L2(Ωf)‖∇u‖H1(Ωf)‖∇2u‖H1(Ωf)+‖ux‖H1(Ωf)‖∇ux‖H1(Ωf)‖∇uy‖2L2(Ωf)+‖uy‖H1(Ωf)‖∇uy‖H1(Ωf)‖∇ux‖2L2(Ωf)+‖∇u‖L2(Ωf)‖∇2u‖L2(Ωf)‖∇ux,y‖2L2(Ωf)≤12‖ut,x,y‖2L2(Ωf)+C(‖u‖3H2(Ωf)‖∇2u‖H1(Ωf)+‖u‖2H2(Ωf)‖D(ux,y)‖2L2(Ωf)). | (2.32) |
Multiplying (2.32) by σ(t) and integrating the result over (0,t), with integration by parts and Young's inequality, obtains that
σ(t)‖D(ux,y)‖2L2(Ωf)+12∫t0σ(s)‖ut,x,y‖2L2(Ωf) ds≤C∫t0‖u‖3H2(Ωf)‖∇2u‖H1(Ωf) ds+C∫t0(1+‖u‖2H2(Ωf))‖D(ux,y)‖2L2(Ωf) ds≤C(1+N2). | (2.33) |
Similarly, we can get
σ(t)‖D(ux,x)‖2L2(Ωf)+12∫t0σ(s)‖ut,x,x‖2L2(Ωf) ds≤C(1+N2), | (2.34) |
and
σ(t)‖D(uy,y)‖2L2(Ωf)+12∫t0σ(s)‖ut,y,y‖2L2(Ωf) ds≤C(1+N2). | (2.35) |
Step 2. For ‖D(ut)‖2L2(Ωf), first differentiate (1.3)1 with respect to t and multiply by ∂2tu, then integrate it on Ωf. Meanwhile, apply ∂2t to (1.2) and multiply by ∂2tP2, then integrate it on Ωm. At last, summing up the two resulting equations with Hölder's inequality, Gagliardo-Nirenberg inequality, and Young's inequality gains that
‖ut,t‖2L2(Ωf)+ddt(μ1‖D(ut)‖2L2(Ωf)+W22∑i=1‖ut⋅τi‖2L2(Γi)+12∫ΩmK∇∂tP2⋅∇∂tP2 dx)≤−∫Ωf∂t(u⋅∇u)⋅ut,t dx≤(‖ut⋅∇u‖L2(Ωf)+‖u⋅∇ut‖L2(Ωf))‖ut,t‖L2(Ωf)≤‖ut‖2L4(Ωf)‖∇u‖2L4(Ωf)+‖u‖2L∞(Ωf)‖∇ut‖2L2(Ωf)+12‖ut,t‖2L2(Ωf)≤C‖∇ut‖2L2(Ωf)‖∇u‖2H1(Ωf)+12‖ut,t‖2L2(Ωf). | (2.36) |
Multiplying (2.36) by σ(t) and integrating the result over (0,t), with Lemma 2.3 and Young's inequality, obtains that
σ(t)(‖D(ut)‖2L2(Ωf)+‖∇∂tP2‖2L2(Ωf))+∫t0σ(s)‖ut,t‖2L2(Ωf) ds≤C∫t0(1+‖u‖2H2(Ωf))‖D(ut)‖2L2(Ωf) ds+C∫t0‖∇∂tP2‖2L2(Ωf) ds≤C(1+N2). | (2.37) |
Step 3. Undoubtedly, from (2.25), (2.29), and (2.30), we get that
‖ux,z,z‖2L2(Ωf)+‖uy,z,z‖2L2(Ωf)≤‖1μ1∂x(ut+u⋅∇u+∇P1−ux,x−uy,y)‖2L2(Ωf)+‖1μ1∂y(ut+u⋅∇u+∇P1−ux,x−uy,y)‖2L2(Ωf)≤C(‖ut,x‖2L2(Ωf)+‖ut,y‖2L2(Ωf)+‖∂x(u⋅∇u)‖2L2(Ωf)+‖∂y(u⋅∇u)‖2L2(Ωf)+‖∇∂xP1‖2L2(Ωf)+‖∇∂yP1‖2L2(Ωf)+‖D(ux,x)‖2L2(Ωf)+‖D(uy,y)‖2L2(Ωf)≤C(‖ut,x‖2L2(Ωf)+‖ut,y‖2L2(Ωf)+‖u‖4H2(Ωf)+‖D(ux,x)‖2L2(Ωf)+‖D(uy,y)‖2L2(Ωf)+‖D(ux,y)‖2L2(Ωf)+1), | (2.38) |
where we have used the estimation:
‖∂x(u⋅∇u)‖2L2(Ωf)+‖∂y(u⋅∇u)‖2L2(Ωf)≤‖ux⋅∇u‖2L2(Ωf)+‖u⋅∇ux‖2L2(Ωf)+‖uy⋅∇u‖2L2(Ωf)+‖u⋅∇uy‖2L2(Ωf)≤C(‖ux‖2L4(Ωf)‖∇u‖2L4(Ωf)+‖u‖2L∞(Ωf)‖∇ux‖2L2(Ωf)+‖uy‖2L4(Ωf)‖∇u‖2L4(Ωf)+‖u‖2L∞(Ωf)‖∇uy‖2L2(Ωf))≤C(‖ux‖12L2(Ωf)‖ux‖32H1(Ωf)‖∇u‖2H1(Ωf)+‖∇u‖L2(Ωf)‖∇2u‖L2(Ωf)‖∇ux‖2L2(Ωf)+‖uy‖12L2(Ωf)‖uy‖32H1(Ωf)‖∇u‖2H1(Ωf)+‖∇u‖L2(Ωf)‖∇2u‖L2(Ωf)‖∇uy‖2L2(Ωf))≤C‖u‖4H2(Ωf) |
and
‖∇∂xP1‖2L2(Ωf)+‖∇∂yP1‖2L2(Ωf)≤C(‖∂x(u⋅∇u)‖2L2(Ωf)+‖∂xP2‖2H1(Ωm)+‖∂2xu1‖2H1(Ωf)+‖∂x∂yu1‖2H1(Ωf))+C(‖∂y(u⋅∇u)‖2L2(Ωf)+‖∂yP2‖2H1(Ωm)+‖∂2yu2‖2H1(Ωf)+‖∂x∂yu2‖2H1(Ωf))≤C(‖u‖4H2(Ωf)+‖D(ux,x)‖2L2(Ωf)+‖D(uy,y)‖2L2(Ωf)+‖D(ux,y)‖2L2(Ωf)), | (2.39) |
which are derived by Hölder's inequality, Gagliardo-Nirenberg inequality, and Young's inequality. Thus, from (2.38), we could derive that
σ(t)(‖ux,z,z‖2L2(Ωf)+‖uy,z,z‖2L2(Ωf))≤C(1+N2). | (2.40) |
Step 4. Apply ∂t∂x to (1.3)1, then multiply σ(t)ut,x and integrate the result equation over Ωf. Meanwhile, apply ∂t∂x to (1.2), then multiply σ(t)∂t∂xP2 and integrate the result equation over Ωm. Summing up the two resulting equations, we obtain with Hölder's inequality, Gagliardo-Nirenberg inequality, and Young's inequality that
ddt(σ(t)‖ut,x‖2L2(Ωf))+σ(t)W2∑i=1‖ut,x⋅τi‖2L2(Γi)+σ(t)‖D(ut,x)‖2L2(Ωf)+σ(t)‖∇∂t∂xP2‖2L2(Ωm)≤C‖ut,x‖2L2(Ωf)+Cσ(t)∫Ωf∂t∂x(u⋅∇u)⋅ut,x dx≤C‖ut,x‖2L2(Ωf)+Cσ(t)‖ut,x‖2L4(Ωf)‖∇u‖L2(Ωf)+Cσ(t)‖ux‖L4(Ωf)‖∇ut‖L2(Ωf)‖ut,x‖L4(Ωf)+Cσ(t)‖ut‖L4(Ωf)‖∇ux‖L2(Ωf)‖ut,x‖L4(Ωf)+Cσ(t)‖u‖L4(Ωf)‖∇ut,x‖L2(Ωf)‖ut,x‖L4(Ωf)≤C‖ut,x‖2L2(Ωf)+Cσ(t)‖ut,x‖12L2(Ωf)‖ut,x‖32H1(Ωf)‖∇u‖L2(Ωf)+Cσ(t)‖ux‖14L2(Ωf)‖ux‖34H1(Ωf)‖∇ut‖L2(Ωf)‖ut,x‖14L2(Ωf)‖ut,x‖34H1(Ωf)+Cσ(t)‖ut‖14L2(Ωf)‖ut‖34H1(Ωf)‖∇ux‖L2(Ωf)‖ut,x‖14L2(Ωf)‖ut,x‖34H1(Ωf)+Cσ(t)‖u‖14L2(Ωf)‖∇u‖34L2(Ωf)‖∇ut,x‖L2(Ωf)‖ut,x‖14L2(Ωf)‖ut,x‖34H1(Ωf)≤C‖ut,x‖2L2(Ωf)+12σ(t)‖D(ut,x)‖2L2(Ωf)+Cσ(t)(1+‖u‖8H2(Ωf))‖ut‖2H1(Ωf), |
which means that
∫t0σ(s)(‖D(ut,x)‖2L2(Ωf)+‖∇∂t∂xP2‖2L2(Ωm)) ds≤C(1+N5), | (2.41) |
and, similarly, we have
∫t0σ(s)(‖D(ut,y)‖2L2(Ωf)+‖∇∂t∂yP2‖2L2(Ωm)) ds≤C(1+N5). | (2.42) |
Finally, conclusions can be derived from Proposition 2.1, (2.21), (2.31), (2.33), (2.34), (2.35), (2.37), (2.40), (2.41), and (2.42) that
σ(t)(‖∇3u‖2L2(Ωf)+‖D(ut)‖2L2(Ωf)+‖∇∂tP2‖2L2(Ωm)+‖∇4P2‖2L2(Ωf))+∫t0σ(s)(‖D(ut,x)‖2L2(Ωf)+‖D(ut,y)‖2L2(Ωf)+‖ut,t‖2L2(Ωf)+‖∇∂t∂xP2‖2L2(Ωm)+‖∇∂t∂yP2‖2L2(Ωm)) ds≤C(1+N5). |
So far, it's time to complete the 3-order estimates in the next step.
Step 5. We detailedly display the estimate of ∫t0σ(s)‖D(ux,x,y)‖2L2(Ωm) ds in this step, then the estimates of ∫t0σ(s)‖D(ux,x,x)‖2L2(Ωm) ds, ∫t0σ(s)‖D(ux,y,y)‖2L2(Ωm) ds, and ∫t0σ(s)‖D(uy,y,y)‖2L2(Ωm) ds can be obtained in a similar derivation.
Apply ∂2x∂y to (1.3)1 and multiply by σ(t)ux,x,y before integrating the resulting equations over Ωf. Meanwhile, apply ∂2x∂y to (1.2) and multiply σ(t)∂2x∂yP2 on it after that, integrate over Ωm. Finally, summing up the above two resulting equations could come to
ddt(σ(t)‖ux,x,y‖2L2(Ωf))+σ(t)‖D(ux,x,y)‖2L2(Ωf)≤C‖ux,x,y‖2L2(Ωf)+Cσ(t)∫Ωf∂2x∂y(u⋅∇u)⋅ux,x,y dx≤C‖ux,x,y‖2L2(Ωf)+Cσ(t)‖ux,x,y‖2L4(Ωf)‖∇u‖L2(Ωf)+Cσ(t)‖ux,x‖L4(Ωf)‖∇uy‖L2(Ωf)‖ux,x,y‖L4(Ωf)+Cσ(t)‖ux,y‖L4(Ωf)‖∇ux‖L2(Ωf)‖ux,x,y‖L4(Ωf)+Cσ(t)‖ux‖L4(Ωf)‖∇ux,y‖L2(Ωf)‖ux,x,y‖L4(Ωf)+Cσ(t)‖uy‖L4(Ωf)‖∇ux,x‖L2(Ωf)‖ux,x,y‖L4(Ωf)+Cσ(t)‖u‖L4(Ωf)‖∇ux,x,y‖L2(Ωf)‖ux,x,y‖L4(Ωf)≤C‖ux,x,y‖2L2(Ωf)+Cσ(t)‖ux,x,y‖32H1(Ωf)‖ux,x,y‖12L2(Ωf)‖∇u‖L2(Ωf)+Cσ(t)‖ux,x‖34H1(Ωf)‖ux,x‖14L2(Ωf)‖∇uy‖L2(Ωf)‖ux,x,y‖34H1(Ωf)‖ux,x,y‖14L2(Ωf)+Cσ(t)‖ux,y‖34H1(Ωf)‖ux,y‖14L2(Ωf)‖∇ux‖L2(Ωf)‖ux,x,y‖34H1(Ωf)‖ux,x,y‖14L2(Ωf)+Cσ(t)‖ux‖34H1(Ωf)‖ux‖14L2(Ωf)‖∇ux,y‖L2(Ωf)‖ux,x,y‖34H1(Ωf)‖ux,x,y‖14L2(Ωf)+Cσ(t)‖uy‖34H1(Ωf)‖uy‖14L2(Ωf)‖∇ux,x‖L2(Ωf)‖ux,x,y‖34H1(Ωf)‖ux,x,y‖14L2(Ωf)+Cσ(t)‖∇u‖34L2(Ωf)‖u‖14L2(Ωf)‖∇ux,x,y‖L2(Ωf)‖ux,x,y‖34H1(Ωf)‖ux,x,y‖14L2(Ωf)≤C‖ux,x,y‖2L2(Ωf)+12σ(t)‖D(ux,x,y)‖2L2(Ωf)+Cσ(t)(1+‖D(ux,x)‖2L2(Ωf)+‖D(ux,y)‖2L2(Ωf))(1+‖u‖8H2(Ωf)), |
the proof of which is reckoned from the Gagliardo-Nirenberg inequality and Young's inequality. It is easy to find that
∫t0σ(s)‖D(ux,x,y)‖2L2(Ωf) ds≤C(1+N5), |
then, similarly, we have
∫t0σ(s)‖D(ux,y,y)‖2L2(Ωf) ds≤C(1+N5), |
and it is more concisely to be derived that
∫t0σ(s)(‖D(ux,x,x)‖2L2(Ωf)+‖D(uy,y,y)‖2L2(Ωf)) ds≤C(1+N5), |
so that the proof of Lemma 2.5 is completed with Young's inequality.
Using a similar argument as that in the proof of Lemma 2.5, we can easily obtain the following fourth-order estimates.
Lemma 2.6. It holds that
sup0≤t≤Tσ(t)2(‖∇4u‖2L2(Ωf)+‖∇2ut‖2L2(Ωf)+‖ut,t‖2L2(Ωf)+‖∇5P2‖2L2(Ωm)+‖∇∂t∂xP2‖2L2(Ωm)+‖∇∂t∂yP2‖2L2(Ωm))+∫t0σ(t)2(‖ut,x,x,x‖2L2(Ωf)+‖ut,x,y,y‖2L2(Ωf)+‖ut,y,y,y‖2L2(Ωf)+‖D(ut,t)‖2L2(Ωf)+‖∇∂2tP2‖2L2(Ωm)) ds≤C(‖u0‖18H2(Ωf)+1)exp{‖u0‖2H1(Ωf)+‖u0‖4H1(Ωf)}. | (2.43) |
Proof. It follows from the fact that
‖∇4u‖2L2(Ωf)≤C(‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,x,y)‖2L2(Ωf)+‖D(ux,y,y)‖2L2(Ωf)+‖D(uy,y,y)‖2L2(Ωf)+‖∇2uz,z‖2L2(Ωf)). |
Obviously, from (2.23) with Hölder's inequality, Young's inequality, Gagliardo-Nirenberg inequality, Poincáre's inequality, Sobolev inequality, and Korn's inequality we have that
‖∇2uz,z‖2L2(Ωf)≤C(‖∇2ut‖2L2(Ωf)+‖∇2(u⋅∇u)‖2L2(Ωf)+‖∇3P1‖2L2(Ωf)+‖∇2ux,x‖2L2(Ωf)+‖∇2uy,y‖2L2(Ωf)), |
where
‖u⋅∇u‖2H2(Ωf)≤‖∇2(u⋅∇u)‖2L2(Ωf)+‖u⋅∇u‖2H1(Ωf)≤C‖∇2u⋅∇u‖2L2(Ωf)+‖u⋅∇3u‖2L2(Ωf)+C‖u‖4H2(Ωf),≤‖∇2u‖2L4(Ωf)‖∇u‖2L4(Ωf)+‖u‖2L∞(Ωf)‖∇3u‖2L2(Ωf)+C‖u‖4H2(Ωf)≤C(‖∇2u‖2H1(Ωf)‖∇u‖2H1(Ωf)+‖u‖2H2(Ωf)‖∇3u‖2L2(Ωf)+‖u‖4H2(Ωf))≤C(‖u‖2H2(Ωf)‖∇3u‖2L2(Ωf)+‖u‖4H2(Ωf)). | (2.44) |
Additionally, we leverage Lemma 2.5 [41], coupled with the Trace theorem. The methodology employed here mirrors that of (2.26). We obtain
‖∇3P1‖2L2(Ωf)≤‖P1‖2H3(Ωf)≤‖∇⋅(u⋅∇u)‖2H1(Ωf)+‖P2−2μ1∂xu1−2μ1∂yu2‖2H52(Γi)≤C(‖u‖2H2(Ωf)‖∇3u‖2L2(Ωf)+‖u‖4H2(Ωf)+‖u‖2H1(Ωf)+‖∂xu1‖2H3(Ωf)+‖∂yu2‖2H3(Ωf))≤C(‖u‖2H2(Ωf)‖∇3u‖2L2(Ωf)+‖u‖4H2(Ωf)+‖∇ux,z,z‖2L2(Ωf)+‖∇uy,z,z‖2L2(Ωf)+‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,x,y)‖2L2(Ωf)+‖D(ux,y,y)‖2L2(Ωf)+‖D(uy,y,y)‖2L2(Ωf)), |
such that one gets
‖∇4u‖2L2(Ωf)≤C(‖∇2ut‖2L2(Ωf)+‖u‖2H2(Ωf)‖∇3u‖2L2(Ωf)+‖u‖4H2(Ωf)+‖∇ux,z,z‖2L2(Ωf)+‖∇uy,z,z‖2L2(Ωf)+‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,x,y)‖2L2(Ωf)+‖D(ux,y,y)‖2L2(Ωf)+‖D(uy,y,y)‖2L2(Ωf)). | (2.45) |
Step 1. We detail work on σ(t)2‖D(ux,x,y)‖2L2(Ωf) in this step to get σ(t)2‖D(ux,x,x)‖2L2(Ωf), σ(t)2‖D(ux,y,y)‖2L2(Ωf), and σ(t)2‖D(uy,y,y)‖2L2(Ωf).
Apply ∂2x∂y to the Eq (1.3)1 and multiply by ∂2x∂yut. Meanwhile, apply ∂2x∂y to (1.2) and multiply by ∂2x∂yP2, integrate the two equations with respect to x by parts, then add up the resulting formulas to get
‖ut,x,x,y‖2L2(Ωf)+ddt(μ1‖D(ux,x,y)‖2L2(Ωf)+W22∑i=1‖ux,x,y⋅τi‖2L2(Γi)+12∫ΩmK∇P2,x,x,y⋅∇P2,x,x,y dx)≤−∫Ωf∂2x∂y(u⋅∇u)⋅∂2x∂yut dx≤12‖ut,x,x,y‖2L2(Ωf)+‖ux,x,y‖2L4(Ωf)‖∇u‖2L4(Ωf)+‖ux,x‖2L4(Ωf)‖∇uy‖2L4(Ωf)+‖ux,y‖2L4(Ωf)‖∇ux‖2L4(Ωf)+‖uy‖2L∞(Ωf)‖∇ux,x‖2L2(Ωf)+‖ux‖2L∞(Ωf)‖∇ux,y‖2L2(Ωf)+‖u‖2L∞(Ωf)‖∇ux,x,y‖2L2(Ωf)≤12‖ut,x,x,y‖2L2(Ωf)+‖ux,x,y‖12L2(Ωf)‖ux,x,y‖32H1(Ωf)‖u‖2H2(Ωf)+‖∇2u‖2H1(Ωf)‖∇2u‖2L2(Ωf)+‖u‖2H2(Ωf)(‖∇ux,x‖2L2(Ωf)+‖∇ux,y‖2L2(Ωf))+‖u‖2H2(Ωf)‖∇ux,x,y‖2L2(Ωf), | (2.46) |
then by multiplying by σ(s)2 and integrating over (0,t), one has
σ(t)2‖D(ux,x,y)‖2L2(Ωf)+12∫t0σ(s)2‖ut,x,x,y‖2L2(Ωf) ds≤C(1+N6). | (2.47) |
We can get the similar results on σ(t)2‖D(ux,x,x)‖2L2(Ωf),σ(t)2‖D(ux,y,y)‖2L2(Ωf), and σ(t)2‖D(uy,y,y)‖2L2(Ωf); thus, we have
σ(t)2(‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,y,y)‖2L2(Ωf)+‖D(uy,y,y)‖2L2(Ωf))+12∫t0σ(s)2(‖ut,x,x,x‖2L2(Ωf)+‖ut,x,y,y‖2L2(Ωf)+‖ut,y,y,y‖2L2(Ωf)) ds≤C(1+N6). | (2.48) |
Step 2. For ‖∇2ut‖L2(Ωf), we have
‖∇2ut‖L2(Ωf)≤C(‖∇ut,x‖L2(Ωf)+‖∇ut,y‖L2(Ωf)+‖ut,z,z‖L2(Ωf)). | (2.49) |
First, we apply ∂x∂t to the Eq (1.3)1 and multiply by ut,t,x. Meanwhile, apply ∂x∂t to (1.2) and multiply by ∂x∂tP2, integrate the two equations with respect to x by parts, then we add up the resulting formulas to get
‖ut,t,x‖2L2(Ωf)+ddt(μ1‖D(ut,x)‖2L2(Ωf)+W22∑i=1‖ut,x⋅τi‖2L2(Γi))+12∫ΩmK∇∂t∂xP2⋅∇∂t∂xP2dx≤−∫Ωf∂t∂x(u⋅∇u)⋅ut,t,xdx≤−∫Ωf(ut,x⋅∇u+ut⋅∇ux+ux⋅∇ut+u⋅∇ut,x)⋅ut,t,xdx≤‖ut,x‖2L4(Ωf)‖∇u‖2L4(Ωf)+‖ut‖2L4(Ωf)‖∇ux‖2L4(Ωf)+‖ux‖2L∞(Ωf)‖∇ut‖2L2(Ωf)+‖u‖2L∞(Ωf)‖∇ut,x‖2L2(Ωf)+12‖ut,t,x‖2L2(Ωf)≤C(‖ut,x‖12L2(Ωf)‖ut,x‖32H1(Ωf)‖u‖2H2(Ωf)+‖ut‖2H1(Ωf)‖u‖2H3(Ωf)+‖∇ut,x‖2L2(Ωf)‖u‖2H2(Ωf))+12‖ut,t,x‖2L2(Ωf). | (2.50) |
Multiplying by σ(s)2 on (2.50) and integrating over (0,t) with (2.3) and (2.22) yields:
σ(t)2(‖D(ut,x)‖2L2(Ωf)+‖∇∂t∂xP2‖2L2(Ωm))+12∫t0σ(s)2‖ut,t,x‖2L2(Ωf) ds≤C(1+N6), | (2.51) |
then, similarly, we have
σ(t)2(‖D(ut,y)‖2L2(Ωf)+‖∇∂t∂yP2‖2L2(Ωm))+12∫t0σ(s)2‖ut,t,y‖2L2(Ωf) ds≤C(1+N6). | (2.52) |
Second, we try to get a bound of ‖ut,z,z‖L2(Ωf), and differentiating (2.23) with respect to t shows that
‖ut,z,z‖2L2(Ωf)≤‖1μ1(ut,t+∂t(u⋅∇u)+∇∂tP1−ut,x,x−ut,y,y)‖2L2(Ωf)≤C(‖ut,t‖2L2(Ωf)+‖ut‖2H1(Ωf)‖u‖2H2(Ωf)+‖∇∂tP1‖2L2(Ωf)+‖ut,x,x‖2L2(Ωf)+‖ut,y,y‖2L2(Ωf)). | (2.53) |
For ‖ut,t‖2L2(Ωf), first apply ∂2t to (1.3)1 and multiply by σ(t)2∂2tu. Meanwhile, apply ∂2t to (1.2) and multiply by σ(t)2∂2tP2, then integrate it on Ωm. At last, integrate the two resulting equations with respect to x and t and sum them up to arrive at
σ(t)2‖ut,t‖2L2(Ωf)+∫t0σ(s)2∫Ωf(‖D(ut,t)‖2L2(Ωf)+‖∇∂2tP2‖2L2(Ωm))dxds≤C∫t0σ(s)2‖ut,t‖2L2(Ωf)ds+C∫t0σ(s)2(u⋅∇u)tt⋅ut,tds≤C(1+N5)+C∫t0σ(s)2‖ut,t‖L2(Ωf)‖∇u‖L4(Ωf)‖ut,t‖L4(Ωf) ds+C∫t0σ(s)2(‖ut‖L4(Ωf)‖∇ut‖L2(Ωf)+‖u‖L4(Ωf)‖∇ut,t‖L2(Ωf))‖ut,t‖L4(Ωf) ds≤C(1+N5)+12∫t0σ(s)2‖D(ut,t)‖2L2(Ωf) ds+C∫t0σ(s)2(‖ut,t‖2L2(Ωf)(1+‖u‖8H2(Ωf))+‖ut‖4H1(Ωf)) ds≤C(1+N9)+12∫t0σ(s)2‖D(ut,t)‖2L2(Ωf) ds, | (2.54) |
where we have used Lemma 2.5. Next, it's time to deal with ‖∇∂tP1‖L2(Ωf). We differentiate (2.24) and (2.25) with respect to t, and then using the standard L2-estimate for the elliptic system with (2.3), (2.51), and (2.52), we have
σ(t)2‖∇∂tP1‖2L2(Ωf)≤Cσ(t)2(‖ut‖2H1(Ωf)‖u‖2H2(Ωf)+‖∂tP2‖2H1(Ωm)+‖ut,x‖2H1(Ωf)+‖ut,y‖2H1(Ωf))≤C(1+N6). | (2.55) |
Next, plugging (2.54) and (2.55) into (2.53), we can get that
σ(t)2‖ut,z,z‖2L2(Ωf)≤C(1+N9), | (2.56) |
Similarly, we can plug (2.56), (2.51), and (2.52) to (2.49) to obtain
σ(t)2‖∇2ut‖2L2(Ωf)≤C(1+N9). | (2.57) |
Step 3. We work on ‖∇ux,z,z‖2L2(Ωf)+‖∇uy,z,z‖2L2(Ωf) in this step. Applying the gradient operator ∇ to (2.38), then multiplying by σ(s)2 and integrating over (0,t) yields
σ(t)2(‖∇ux,z,z‖2L2(Ωf)+‖∇uy,z,z‖2L2(Ωf))≤1μ1σ(t)2‖∇∂x(ut+u⋅∇u+∇P1−ux,x−uy,y)‖2L2(Ωf)+1μ1σ(t)2‖∇∂y(ut+u⋅∇u+∇P1−ux,x−uy,y)‖2L2(Ωf)≤C(1+N9)+σ(t)2(‖∇2∂xP1‖2L2(Ωf)+‖∇2∂yP1‖2L2(Ωf), | (2.58) |
where we have used Proposition 2.1, (2.44), (2.47), (2.48), (2.51), and (2.52).
We now estimate ‖∇2∂xP1‖2L2(Ωf)+‖∇2∂yP1‖2L2(Ωf). We apply Lemma 2.5 in [41] together with the Trace theorem on (2.24) and (2.25), similarly as (2.39), to get that
σ(t)2(‖∇2∂xP1‖2L2(Ωf)+‖∇2∂yP1‖2L2(Ωf))≤Cσ(t)2(‖div(∂x(u⋅∇u))‖2L2(Ωf)+‖div(∂y(u⋅∇u))‖2L2(Ωf)+‖∂xP2‖2H32(Γi)+‖∂yP2‖2H32(Γi)+‖∂2xu1‖2H32(Ωf)+‖∂x∂yu1‖2H32(Ωf)+‖∂2yu2‖2H32(Ωf)+‖∂x∂yu2‖2H32(Ωf))≤Cσ(t)2(‖u⋅∇u‖2H2(Ωf)+‖P2‖2H3(Ωm)+‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,x,y)‖2L2(Ωf)+‖D(ux,y,y)‖2L2(Ωf)+‖ux,x,z,z‖2L2(Ωf)+‖ux,y,z,z‖2L2(Ωf)+‖uy,y,z,z‖2L2(Ωf))≤C(1+N6)+Cσ(t)2(‖ux,x,z,z‖2L2(Ωf)+‖ux,y,z,z‖2L2(Ωf)+‖uy,y,z,z‖2L2(Ωf)). | (2.59) |
Now, we respectively apply ∂2x, ∂x∂y, ∂2y to (2.23) and multiply by σ(s)2 to get
σ(t)2(‖ux,x,z,z‖2L2(Ωf)+‖ux,y,z,z‖2L2(Ωf)+‖uy,y,z,z‖2L2(Ωf))≤Cσ(t)2(‖D(ut,x)‖2L2(Ωf)+‖D(ut,y)‖2L2(Ωf)+‖u⋅∇u‖2H2(Ωf)+‖∇∂2xP1‖2L2(Ωf)+‖∇∂x∂yP1‖2L2(Ωf)+‖∇∂2yP1‖2L2(Ωf)+‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,x,y)‖2L2(Ωf))≤C(1+N9)+Cσ(t)2(‖∇∂2xP1‖2L2(Ωf)+‖∇∂x∂yP1‖2L2(Ωf)+‖∇∂2yP1‖2L2(Ωf)), |
where we have used Lemma 2.5, (2.44), (2.51), and (2.52), and it remains to estimate the last three terms on the righthand side. Again, we apply Lemma 2.5 [41] together with the Trace theorem on (2.24) and (2.25), similarly as (2.39), to get that
σ(t)2(‖∇∂2xP1‖2L2(Ωf)+‖∇∂x∂yP1‖2L2(Ωf)+‖∇∂2yP1‖2L2(Ωf))≤Cσ(t)2(‖∂2x(u⋅∇u)‖2L2(Ωf)+‖∂x∂y(u⋅∇u)‖2L2(Ωf)+‖∂2y(u⋅∇u)‖2L2(Ωf)+‖∂2xP2‖2H1(Ωm)+‖∂2yP2‖2H1(Ωm)+‖∂x∂yP2‖2H1(Ωm)+‖∂3xu1‖2H1(Ωf)+‖∂2x∂yu1‖2H1(Ωf)+‖∂x∂2yu1‖2H1(Ωf)+‖∂2x∂yu2‖2H1(Ωf)+‖∂x∂2yu2‖2H1(Ωf)+‖∂3yu2‖2H1(Ωf))≤Cσ(t)2(‖u⋅∇u‖2H2(Ωf)+‖P2‖2H3(Ωm)+‖D(ux,x,x)‖2L2(Ωf)+‖D(ux,x,y)‖2L2(Ωf)+‖D(ux,y,y)‖2L2(Ωf)+‖D(uy,y,y)‖2L2(Ωf))≤C(1+N6), | (2.60) |
thus, combining the above steps, we can complete the proof of Lemma 2.6.
First of all, we know that the systems (1.3) and (1.4) have a unique local strong solution (u,P2) on Ω×(0,T∗] for some T∗>0, so it's time to verify the continuity of the strong solution to extend it globally in time with counter-evidence.
It follows from the fact that u0∈H2(Ωf) and Proposition 1.1 that there exists a T1∈(0,T∗] such that (2.1) holds for T=T1. Next, we set
T∗=sup{T∣(u,P2) is a strong solution on Ω×(0,T] and (2.1) holds}, | (3.1) |
then T∗≥T1≥0. Hence, for any 0<τ<T≤T∗ with T finite, we can get
u∈L∞(τ,T;H4(Ωf)), ut∈L∞(τ,T;H2(Ωf)), |
from Proposition 2.1, and Lemmas 2.5 and 2.6. Thus one can deduce that
u∈C([τ,T];C2(Ωf))∩C([τ,T];H3(Ωf)) | (3.2) |
from
L∞(τ,T;H4(Ωf))∩W1,∞(τ,T;H2(Ωf))↪C([τ,T];C2(Ωf))∩C([τ,T];H3(Ωf)). |
Now, we suppose that
T∗<∞, | (3.3) |
then T=T∗ holds by Proposition 2.1 and (2.2). It follows from (3.2) that
u(x,T∗):=limt→T∗u(x,t)∈H2(Ωf), | (3.4) |
thus, the initial data condition in Proposition 1.1 is satisfied, which gives that there exists a T∗∗>T∗ such that (2.1) holds for T=T∗∗. This contradicts the definition of T∗ in (3.1), so T∗=∞.
Finally, to finish the proof of Theorem 1.2, we have from (2.5) that
12ddt‖u‖2L2(Ωf)+μ1CΩf‖∇u‖2L2(Ωf)≤0. | (3.5) |
According to the Poincar inequality for three-dimensional cases, we have
‖u‖2L2(Ωf)≤2√3∣Ωf∣13‖∇u‖2L2(Ωf), | (3.6) |
then substituting (3.6) into (3.5), we get
ddt‖u‖2L2(Ωf)+3μ12∣Ωf∣23CΩf‖u‖2L2(Ωf)≤0. | (3.7) |
Therefore, we have
‖u‖2L2(Ωf)≤Cϵ20e−ct. | (3.8) |
Here, the positive constants are C=C(μ1,λ,Ωf,‖u0‖H2(Ωf)) and c=c(μ1,Ωf).
In conclusion, this study has made some progress in understanding the global well-posedness of a coupled Navier-Stokes-Darcy model with the Beavers-Joseph-Saffman-Jones interface boundary condition in three-dimensional Euclidean space. Through our investigation, we have achieved the establishment of a global strong solution for the system, marking a crucial advancement in the field. Moreover, we have demonstrated the exponential stability of this strong solution, further reinforcing its reliability. The implications of our findings extend to the analysis of subsurface flow problems, notably in the realm of karst aquifers, where such coupled systems play a pivotal role. By shedding light on the dynamics and behaviors of these systems, our 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, 12061081), the Natural Science Foundation of Shaanxi Province (Grant No. 2023-JC-QN-0073), the Natural Science Foundation of Yulin (Grant No. CXY-2022-76), and the Natural Science Foundation of Yulin University (Grant No. 21GK07).
The authors have no conflicts to disclose.
[1] | J. R. Fanchi, Principles of applied reservoir simulation, Elsevier, 2005. http://dx.doi.org/10.1016/B978-075067933-6/50014-3 |
[2] | J. Bear, Dynamics of fluids in porous media, Courier Corporation, 1972. http://dx.doi.org/10.1097/00010694-197508000-00022 |
[3] |
M. Discacciati, A. Quarteroni, Navier-Stokes/Darcy coupling: Modeling, analysis, and numerical approximation, Rev. Mat. Comput., 2 (2009), 315–426. http://dx.doi.org/10.5209/rev_REMA.2009.v22.n2.16263 doi: 10.5209/rev_REMA.2009.v22.n2.16263
![]() |
[4] |
D. Han, D. Sun, X. Wang, Two-phase flows in karstic geometry, Math. Method. Appl. Sci., 37 (2014), 3048–3063. http://dx.doi.org/10.1002/mma.3043 doi: 10.1002/mma.3043
![]() |
[5] |
M. Alam, H. M. Byrne, G. P. R. Sekhar, Existence and uniqueness results on biphasic mixture model for an in-vivo tumor, Appl. Anal., 101 (2022), 5442–5468. http://dx.doi.org/10.1080/00036811.2021.1895122 doi: 10.1080/00036811.2021.1895122
![]() |
[6] |
W. Layton, F. Schieweck, I. Yotov, Coupling fluid flow with porous media flow, SIAM J. Numer. Anal., 40 (2003), 2195–2218. http://dx.doi.org/10.1137/S0036142901392766 doi: 10.1137/S0036142901392766
![]() |
[7] |
M. Discacciati, E. Miglio, A. Quarteroni, Mathematical and numerical models for coupling surface and groundwater flows, Appl. Numer. Math., 43 (2001), 57–74. http://dx.doi.org/10.1016/S0168-9274(02)00125-3 doi: 10.1016/S0168-9274(02)00125-3
![]() |
[8] |
H. W. Alt, S. Luckhaus, Quasilinear elliptic-parabolic differential equation, Math. Z., 3 (1983), 311–341. http://dx.doi.org/10.1070/SM1968v006n03ABEH001065 doi: 10.1070/SM1968v006n03ABEH001065
![]() |
[9] |
P. Fabrie, M. Langlais, Modelling wells in porous media flows, SIAM J. Math. Anal., 23 (1992), 1375–1392. http://dx.doi.org/10.1142/S0218202500000367 doi: 10.1142/S0218202500000367
![]() |
[10] |
P. Fabrie, T. Gallouët, Modelling wells in porous media flows, Math. Mod. Method. Appl. Sci., 10 (2000), 673–709. http://dx.doi.org/10.1142/S0218202500000367 doi: 10.1142/S0218202500000367
![]() |
[11] |
F. Marpeau, M. Saad, Mathematical analysis of radionuclides displacement in porous media with nonlinear adsorption, J. Differ. Equations, 228 (2006), 412–439. http://dx.doi.org/10.1016/j.jde.2006.03.023 doi: 10.1016/j.jde.2006.03.023
![]() |
[12] |
K. Hans, 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. http://dx.doi.org/10.1007/s00220-013-1708-z doi: 10.1007/s00220-013-1708-z
![]() |
[13] | D. A. Nield, A. Bejan, Convection in porous media, 1 Eds., New York: Springer, 1992. http://dx.doi.org/10.1007/978-3-319-49562-0 |
[14] |
G. S. Beavers, D. D. Joseph, Boundary conditions at a naturally permeable wall, J. Fluid Mech., 30 (1967), 197–207. http://dx.doi.org/10.1017/s0022112067001375 doi: 10.1017/s0022112067001375
![]() |
[15] | H. K. Versteeg, W. Malalasekera, An introduction to computational fluid dynamics: The finite volume method, Pearson Education, 2007. |
[16] |
S. Whitaker, Flow in porous media I: A theoretical derivation of Darcy's law, Transport Porous Med., 1 (1986), 3–25. http://dx.doi.org/10.1007/BF01036523 doi: 10.1007/BF01036523
![]() |
[17] |
I. P. Jones, Low Reynolds number flow past a porous spherical shell, Math. Proc. Cambridge, 73 (1973), 231–238. http://dx.doi.org/10.1017/S0305004100047642 doi: 10.1017/S0305004100047642
![]() |
[18] |
P. G. Saffman, On the boundary condition at the surface of a porous medium, Stud. Appl. Math., 50 (1971), 93–101. http://dx.doi.org/10.1002/sapm197150293 doi: 10.1002/sapm197150293
![]() |
[19] |
M. McCurdy, N. Moore, X. Wang, Convection in a coupled free flow-porous media system, SIAM J. Appl. Math., 6 (2019), 2313–2339. http://dx.doi.org/10.1137/19M1238095 doi: 10.1137/19M1238095
![]() |
[20] |
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. http://dx.doi.org/10.1016/j.na.2021.112411 doi: 10.1016/j.na.2021.112411
![]() |
[21] |
A. Çeşmelioğlu, B. Rivière, Primal discontinuous Galerkin methods for time-dependent coupled surface and subsurface flow, J. Sci. Comput., 40 (2009), 115–140. http://dx.doi.org/10.1007/s10915-009-9274-4 doi: 10.1007/s10915-009-9274-4
![]() |
[22] |
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. http://dx.doi.org/10.1137/070686081 doi: 10.1137/070686081
![]() |
[23] |
Y. Cao, Y. Chu, X. He, M. Wei, Decoupling the stationary Navier-Stokes-Darcy system with the Beavers-Joseph-Saffman interface condition, Abstr. Appl. Anal., 2013. http://dx.doi.org/10.1155/2013/136483 doi: 10.1155/2013/136483
![]() |
[24] |
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. http://dx.doi.org/10.1137/080721868 doi: 10.1137/080721868
![]() |
[25] |
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. http://dx.doi.org/10.1016/S0252-9602(17)30076-0 doi: 10.1016/S0252-9602(17)30076-0
![]() |
[26] |
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). http://dx.doi.org/10.1016/j.jcp.2020.109400 doi: 10.1016/j.jcp.2020.109400
![]() |
[27] | Y, Qin, Y. Hou, Optimal error estimates of a decoupled scheme based on two-grid finite element for mixed Navier-Stokes/Darcy model, Acta Math. Sci., 38 (2018), 1361–1369. |
[28] |
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. http://dx.doi.org/10.1137/140965776 doi: 10.1137/140965776
![]() |
[29] |
L. Zuo, G. Du, A parallel two-grid linearized method for the coupled Navier-Stokes-Darcy problem, Numer. Algorithms, 77 (2018), 151–165, http://dx.doi.org/10.1007/s11075-017-0308-y doi: 10.1007/s11075-017-0308-y
![]() |
[30] | M. Discacciati, A. Quarteroni, Analysis of a domain decomposition method for the coupling of Stokes and Darcy equations, In: F. Brezzi, A. Buffa, S. Corsaro, A. Murli, Numerical mathematics and advanced applications, Milan: Springer, 2003. http://dx.doi.org/10.1007/978-88-470-2089-4-1 |
[31] |
B. Rivière, I. Yotov, Locally conservative coupling of Stokes and Darcy flows, SIAM J. Numer. Anal., 42 (2004), 1959–1977. http://dx.doi.org/10.1137/S0036142903427640 doi: 10.1137/S0036142903427640
![]() |
[32] |
B. Rivière, Analysis of a discontinuous finite element method for the coupled Stokes and Darcy problems, J. Sci. Comput., 22 (2005), 479–500. http://dx.doi.org/10.1007/s10915-004-4147-3 doi: 10.1007/s10915-004-4147-3
![]() |
[33] | B. Rivière, Analysis of a multi-numerics/multi-physics problem, In: M. Feistauer, V. Doleší, P. Knobloch, K. Najzar, Numerical Mathematics and Advanced Applications, Berlin: Springer, 2004,726–735. http://dx.doi.org/10.1007/978-3-642-18775-9_71 |
[34] |
M. Discacciati, A. Quarteroni, A. Valli, Robin-Robin domain decomposition methods for the Stokes-Darcy coupling, SIAM J. Numer. Anal., 45 (2007), 1246–1268. http://dx.doi.org/10.1137/06065091X doi: 10.1137/06065091X
![]() |
[35] |
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). http://dx.doi.org/10.1016/j.physd.2020.132687 doi: 10.1016/j.physd.2020.132687
![]() |
[36] |
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, 2021. http://dx.doi.org/10.57262/ade/1610420433 doi: 10.57262/ade/1610420433
![]() |
[37] |
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). http://dx.doi.org/10.1016/j.jcp.2022.110968 doi: 10.1016/j.jcp.2022.110968
![]() |
[38] |
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). http://dx.doi.org/10.1007/s00211-017-0870-1 doi: 10.1007/s00211-017-0870-1
![]() |
[39] |
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. http://dx.doi.org/10.1137/16M1100885 doi: 10.1137/16M1100885
![]() |
[40] |
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. http://dx.doi.org/10.1016/j.jde.2014.07.013 doi: 10.1016/j.jde.2014.07.013
![]() |
[41] |
M. Cui, W. Dong, Z. Guo, Global well-posedness of coupled Navier-Stokes and Darcy equations, J. Differ. Equations, 388 (2024), 82–111. http://dx.doi.org/10.1016/j.jde.2023.12.044 doi: 10.1016/j.jde.2023.12.044
![]() |
[42] | P. Liu, W. Liu, Global well-posedness of an initial-boundary value problem of the 2-D incompressible Navier-Stokes-Darcy system, Netherland: Springer, 1 (2019). http://dx.doi.org/10.1007/s10440-018-0197-7 |
[43] |
C. Foias, O. Manley, R. Temam, Attractors for the Bénard problem: Existence and physical bounds on their fractal dimension, Nonlinear Anal., 11 (1987), 939–967. http://dx.doi.org/10.1016/0362-546X(87)90061-7 doi: 10.1016/0362-546X(87)90061-7
![]() |
[44] |
P. Fabrie, Solutions fortes et comportment asymtotique pour un modèle de convection naturelle en milieu poreux, Acta Appl. Math., 7 (1986), 49–77. http://dx.doi.org/10.1007/BF00046977 doi: 10.1007/BF00046977
![]() |
[45] |
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. http://dx.doi.org/10.1007/s003329900073 doi: 10.1007/s003329900073
![]() |
[46] |
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. http://dx.doi.org/10.1016/j.jde.2011.12.001 doi: 10.1016/j.jde.2011.12.001
![]() |
[47] |
A. Çeşmelioğlu, V. Girault, B. Rivière, Time-dependent coupling of Navier-Stokes and Darcy flows, ESAIM Math. Model. Num., 47 (2013), 540–555. http://dx.doi.org/10.1051/m2an/2012034 doi: 10.1051/m2an/2012034
![]() |
[48] |
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. http://dx.doi.org/10.1007/s10114-022-9134-9 doi: 10.1007/s10114-022-9134-9
![]() |
1. | Linlin Tan, Bianru Cheng, Global well-posedness of 2D incompressible Navier–Stokes–Darcy flow in a type of generalized time-dependent porosity media, 2024, 32, 2688-1594, 5649, 10.3934/era.2024262 |