Loading [MathJax]/jax/output/SVG/jax.js
Special Issues

A conforming-nonconforming mixed immersed finite element method for unsteady Stokes equations with moving interfaces

  • In this article, we develop a new mixed immersed finite element discretization for two-dimensional unsteady Stokes interface problems with unfitted meshes. The proposed IFE spaces use conforming linear elements for one velocity component and non-conforming linear elements for the other velocity component. The pressure is approximated by piecewise constant. Unisolvency, among other fundamental properties of the new vector-valued IFE functions, is analyzed. Based on the new IFE spaces, semi-discrete and full-discrete schemes are developed for solving the unsteady Stokes equations with a stationary or a moving interface. Re-meshing is not required in our numerical scheme for solving the moving-interface problem. Numerical experiments are carried out to demonstrate the performance of this new IFE method.

    Citation: Derrick Jones, Xu Zhang. A conforming-nonconforming mixed immersed finite element method for unsteady Stokes equations with moving interfaces[J]. Electronic Research Archive, 2021, 29(5): 3171-3191. doi: 10.3934/era.2021032

    Related Papers:

    [1] Derrick Jones, Xu Zhang . A conforming-nonconforming mixed immersed finite element method for unsteady Stokes equations with moving interfaces. Electronic Research Archive, 2021, 29(5): 3171-3191. doi: 10.3934/era.2021032
    [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] Kun Wang, Lin Mu . Numerical investigation of a new cut finite element method for Stokes interface equations. Electronic Research Archive, 2025, 33(4): 2503-2524. doi: 10.3934/era.2025111
    [4] Youngmok Jeon, Dongwook Shin . Immersed hybrid difference methods for elliptic boundary value problems by artificial interface conditions. Electronic Research Archive, 2021, 29(5): 3361-3382. doi: 10.3934/era.2021043
    [5] Yue Feng, Yujie Liu, Ruishu Wang, Shangyou Zhang . A conforming discontinuous Galerkin finite element method on rectangular partitions. Electronic Research Archive, 2021, 29(3): 2375-2389. doi: 10.3934/era.2020120
    [6] Bin Wang, Lin Mu . Viscosity robust weak Galerkin finite element methods for Stokes problems. Electronic Research Archive, 2021, 29(1): 1881-1895. doi: 10.3934/era.2020096
    [7] Hongsong Feng, Shan Zhao . A multigrid based finite difference method for solving parabolic interface problem. Electronic Research Archive, 2021, 29(5): 3141-3170. doi: 10.3934/era.2021031
    [8] Linlin Tan, Bianru Cheng . Global well-posedness of 2D incompressible Navier–Stokes–Darcy flow in a type of generalized time-dependent porosity media. Electronic Research Archive, 2024, 32(10): 5649-5681. doi: 10.3934/era.2024262
    [9] Yujia Chang, Yi Jiang, Rongliang Chen . A parallel domain decomposition algorithm for fluid-structure interaction simulations of the left ventricle with patient-specific shape. Electronic Research Archive, 2022, 30(9): 3377-3396. doi: 10.3934/era.2022172
    [10] 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
  • In this article, we develop a new mixed immersed finite element discretization for two-dimensional unsteady Stokes interface problems with unfitted meshes. The proposed IFE spaces use conforming linear elements for one velocity component and non-conforming linear elements for the other velocity component. The pressure is approximated by piecewise constant. Unisolvency, among other fundamental properties of the new vector-valued IFE functions, is analyzed. Based on the new IFE spaces, semi-discrete and full-discrete schemes are developed for solving the unsteady Stokes equations with a stationary or a moving interface. Re-meshing is not required in our numerical scheme for solving the moving-interface problem. Numerical experiments are carried out to demonstrate the performance of this new IFE method.



    Let ΩR2 be an open bounded domain separated by a time-dependent smooth interface Γ(t). The evolving interface Γ(t) divides the domain Ω into two open subdomains Ω+(t) and Ω(t) such that Ω=Ω+(t)Ω(t)Γ(t), see Figure 1. Consider the following initial-boundary-value problems of the Stokes equation

    ut(μupI)=f in Ω×[0,T], (1)
    u=0 in Ω×[0,T], (2)
    u=0 on Ω×[0,T], (3)
    u(x,0)=u0,p(x,0)=p0 on Ω, (4)
    Figure 1.  The geometrical setup of a moving interface problem.

    where u and p denote the flow velocity and the pressure, respectively. Functions f, u0, and p0 are the given body force, the initial velocity, and the initial pressure, respectively. I denotes the identity tensor. The movement of the interface is assumed to be guided by a given velocity field v(x,t) as follows

    dxdt=v(x,t),onΓ(t)×[0,T]. (5)

    The viscosity function μ(x) is assumed to have a finite jump across the interface Γ(t). For simplicity, we assume that μ(x) is a piecewise constant function

    μ(x)={μ in Ω(t),μ+ in Ω+(t), (6)

    where μ±>0 and x=(x,y). At any time t, the velocity and the stress tensors satisfy the following homogeneous interface jump conditions

    [[u]]Γ=0, (7)
    [[(μupI)n]]Γ=0, (8)

    where the jump [[v(x)]]Γ:=v+(x)|Γv(x)|Γ, and n denotes the unit normal vector to the interface Γ pointing from Ω(t) to Ω+(t).

    Numerical approximations of the Stokes equation have been extensively studied for many years due to its wide applicability to model natural phenomena such as airflow, water flow, and ocean currents. The family of Taylor-Hood finite elements [35] uses conforming Pk-Pk1 pairs to approximate the velocity and pressure requiring the polynomial degree k2. To use low-order polynomials and preserve the mass conservation property, nonconforming finite element methods [4] and discontinuous Galerkin methods [33] have been widely used. Crouzeix and Raviart introduced the lowest-order nonconforming P1-P0 finite element [6], which is well-known as the CR finite element. On quadrilateral meshes, Rannacher and Turek developed a nonconforming rotated-Q1 element in [32]. These nonconforming finite elements make use of low-order polynomials and they are elementwise divergence-free [2,23]. A mixed conforming-nonconforming finite element space was introduced in [21]. In this discretization, one velocity component is approximated by the conforming P1 element and the other one is approximated by the nonconforming P1 element, while the pressure is piecewise constant. This mixed FEM construction has advantages over the CR element in handling the Neumann boundary condition [21], and it is computationally less expensive than the CR element. This mixed conforming-nonconforming FEM has also been extended to the 3D Stokes equation [38]. For more details regarding numerical methods for Stokes equations, we refer readers to [8,19].

    Traditional numerical methods use interface-fitted meshes for solving interface problems. For fluid flow interface problems, the arbitrary Lagrangian-Eulerian (ALE)-based finite element is a popular numerical method [7,22,37]. Recently, there has been a growing interest in developing unfitted-mesh numerical methods for solving a variety of interface problems, see Figure 2. Comparing with conventional fitted-mesh methods, such as classical FE and DG methods, the unfitted-mesh methods do not require the alignment of the mesh with a prescribed nontrivial interface; hence it is more desirable for time-dependent problems with moving interfaces. In the past decades, several unfitted-mesh methods have been developed for solving Stokes interface problems, such as CutFEM [15], Nitsche's FEM [36], XFEM[9], fictitious domain FEM [31,34], to name only a few. The immersed finite element method (IFEM) [24,26,18,11,14,30] is a class of unfitted-mesh finite element methods for solving interface problems. The main idea of IFEM is to incorporate the interface jump conditions in the construction of IFE basis functions. Unlike other aforementioned unfitted-mesh methods, the IFE space is isomorphic to the standard FE space with no interface. Consequently, not only is the mesh independent of the interface in an IFEM, but also the number and the location of the degrees of freedom are interface-independent. For time-dependent interface problems with a moving interface, the linear system has the same size at each time level and the nonzero entries remain at the same locations [10,12,13,16,17]. Moreover, the method-of-lines technique can be utilized together with IFEM for solving moving interface problems [25].

    Figure 2.  From left: an interface-fitted mesh and an unfitted mesh.

    There have been some IFE methods developed for steady-state Stokes interface problems. In [1,3] the Q1-Q0 immersed DG method was introduced. The velocity is approximated by the broken Q1 functions while the pressure is approximated by the piecewise constant functions. The computational framework is based on the interior penalty DG method [33]. Based on the nonconforming finite element framework [6,32], a class of nonconforming IFE approximations was developed [20]. Recently, a P2-P1 Taylor-Hood IFE space was introduced in [5]. The partially penalized IFE scheme is used with ghost penalty for enhancing the stability of numerical scheme especially for the pressure approximation.

    The goal of this paper is two-fold. First, we develop a lowest-order conforming-nonconforming mixed IFE space for the Stokes equation based on [21]. Comparing with the IDG method [1] and the Taylor-Hood IFE method [5], our new IFE method has no additional consistency and stability terms, so the numerical formulation is much simpler to implement. Comparing with the CR-P0 IFE space [20], there are significantly less degrees of freedom due to the conformity of one velocity component. In fact, on the same triangular mesh, only two-thirds of degrees of freedom are required for velocity in this new mixed IFEM. Besides, the mixed conforming-nonconforming finite element is robust for handling both Dirichlet and Neumann boundary conditions, while the CR finite element space is only stable for Dirichlet boundary conditions [21].

    The second goal is to apply this mixed IFE method for solving unsteady Stokes equations with a moving interface. We will use the new vector-valued IFE spaces for semi-discretization, and use the prototypical backward-Euler and Crank-Nicolson scheme for full-discretization. Our method does not require re-meshing at any time level. Since the degrees of freedom are also independent of the interface, there is no need to overhaul the global matrices at each time level. Instead, only local modification is carried out on elements where the interface configuration changed during two consecutive time steps.

    The rest of the paper is organized as follows. In Section 2, we construct the new mixed IFE spaces for Stokes equations. In Section 3, we report some fundamental properties of the new IFE spaces. In section 4, we present the semi-discrete and the full-discrete IFE method for solving unsteady Stokes interface problems with a moving interface. Some numerical examples are reported in Section 5. A brief conclusion is given in Section 6.

    In this section, we introduce the mixed conforming-nonconforming IFE spaces for Stokes equations. Let Th={T} be an interface-unfitted triangulation of a polygonal domain Ω. Let Nh and Eh denote the collections of nodes and edges of the mesh Th, respectively. Elements in Th are divided into two categories: an interface element if T is cut through by the interface Γ, and a non-interface element otherwise. The collections of interface elements and non-interface elements are denoted by Tih and Tnh, respectively. Similarly, for each edge eEh, if e intersects the interface, it is called an interface edge; otherwise it is a non-interface edge. The collections of interface edges and non-interface edges are denoted by Eih and Enh, respectively. Additionally, we let ˚Eh and Ebh be the collections of internal edges and boundary edges, respectively. Let ˚Nh and Nbh be the collections of internal nodes and boundary nodes, respectively. We also assume that the triangulation Th satisfies the following hypotheses [28]:

    (H1) The interface Γ cannot intersect an edge of any element at more than two points unless the edge is part of Γ.

    (H2) If Γ intersects the boundary of an element at two points, these intersection points must be on different edges of this element.

    (H3) The interface Γ is a piecewise C2-continuous function, and the mesh Th is formed such that the subset of Γ in every interface element is C2-continuous.

    Let TTnh be a non-interface element with vertices A1, A2, A3 oriented counterclockwise. We label the edges of T by e1=¯A1A2, e2=¯A2A3, and e3=¯A3A1. Let λj,TP1 be the Lagrange linear nodal basis functions such that

    λj,T(Ai)=δij,i,j=1,2,3, (9)

    where δij is the Kronecker function. Define ψj,T=12λkj,T with k1=3, k2=1, and k3=2. It can be verified that ψj,T satisfies the mean-value conditions, namely,

    1|ei|eiψj,T(x,y)ds=δij,i,j=1,2,3. (10)

    Thus ψj,T, j=1,2,3 are nonconforming-P1 (CR) basis functions on T. The pressure is approximated by the piecewise constant function space denoted by P0. On each non-interface triangle TTnh, the vector-valued CR-P1-P0 finite element space can be written as Snh(T)=P1×P1×P0, or quivalently, Snh(T)=span{ψi,T:1i7} where the vector-valued basis functions are given below

    ψj,T=(ψj,T00),j=1,2,3,ψj,T=(0λj3,T0),j=4,5,6,ψ7,T=(001). (11)

    Similarly, we can also form the P1-CR-P0 finite element space using conforming-P1 bases for the first component, and the nonconforming-P1 in the second component, then the basis functions are

    ˜ψj,T=(λj,T00),j=1,2,3,˜ψj,T=(0ψj3,T0),j=4,5,6,˜ψ7,T=(001). (12)

    The P1-CR-P0 finite element space is ˜Snh(T)=span{˜ψi,T:1i7}. Note that these two spaces Snh(T) and ˜Snh(T) are identical, both equal P1×P1×P0. However, the degrees of freedom of Snh(T) and ˜Snh(T) are different, as indicated in (9) and (10). For more details of the conforming-nonconforming finite elements, we refer readers to [21].

    In this subsection, we extend these conforming-nonconforming finite elements to the IFE spaces on each interface triangle TTih. Let Ai=(xi,yi), i=1,2,3 be the vertices of T. Without loss of generality, we consider the reference triangle whose vertices are given by

    ˆA1=(0,0),ˆA2=(1,0),ˆA3=(0,1).

    Note that an arbitrary triangle with vertices Ai=(xi,yi), i=1,2,3 can be mapped to this reference triangle by the following mapping

    (ˆxˆy)=(x2x1x3x1y2y1y3y1)1(xx1yy1). (13)

    To simplify the notation, we still use x, y, rather than ˆx, ˆy on the reference triangle. According to the hypotheses (H1)-(H3), there are two distinct intersection points on each interface triangle, denoted by D=(xd,yd) and E=(xe,ye), on two different edges. There are generally three types of interface triangles as depicted in Figure 3. The line segment ¯DE is used to approximate the actual interface curve ΓT, and it divides the element T into two subelements, denoted by T+ and T. For example, on a Type Ⅰ interface element, D=A1+d(A2A1) and E=A1+e(A3A1) where 0<d,e<1.

    Figure 3.  Types of interface elements. From left: Type Ⅰ, Type Ⅱ, Type Ⅲ.

    We construct the vector-valued IFE shape functions in terms of the FE functions ψi,T in (11). To be more precise, we have

    ϕj,T(x,y)={7i=1c+ijψi,T(x,y),if(x,y)T+,7i=1cijψi,T(x,y),if(x,y)T,j=1,2,,7. (14)

    It can be observed that each vector-valued IFE shape function ϕj,T has 14 unknown coefficients csij, with 1i7 and s=+,. These coefficients are determined by seven local degrees of freedom (prescribed nodal values, edge values, and the mean pressure value), six interface jump conditions, and a divergence free condition stated below.

    ● Three edge-value conditions:

    1|ek|ekϕj,Tds=(δjk00),k=1,2,3. (15)

    ● Three nodal-value conditions:

    ϕj,T(Ak3)=(0δjk0),k=4,5,6. (16)

    ● One mean-pressure-value condition:

    1|T|Tϕj,Tdxdy=(00δjk),k=7. (17)

    ● Four continuity conditions of the velocity to incorporate (7):

    [[ϕ1,j(D)]]=[[ϕ2,j(D)]]=[[ϕ1,j(E)]]=[[ϕ2,j(E)]]=0. (18)

    ● Two stress continuity conditions to incorporate (8):

    [[μ(xϕ1,jn1+yϕ1,jn2)ϕp,jn1]]¯DE=0, (19)
    [[μ(xϕ2,jn1+yϕ2,jn2)ϕp,jn2]]¯DE=0. (20)

    ● One continuity of the divergence condition to incorporate (2):

    [[xϕ1,j+yϕ2,j]]¯DE=0. (21)

    Here, in (18)-(21), the scalar function ϕi,j denotes the i-th component of ϕj,T. More precisely, we have ϕj,T=(ϕ1,j,ϕ2,j,ϕp,j) such that ϕj,T|Ts=ϕsj,T=(ϕs1,j,ϕs2,j,ϕsp,j) P1×P1×P0, with s=+,. Combining the conditions (15)-(21) yields a linear system of fourteen unknowns. On Type Ⅰ interface element, we have

    MIcj=ej (22)

    where the coefficient matrix MI is written as the first seven columns and the next seven seven columns due to width limit of the page:

    MI(:,1:7)=(dd2ddd200000000000ee2e2ee0000000100000000000000000000000de112d2d10000000d1d002e112e10000000e10e02d2d+4e4edd0e2e2e02dee2dd0221010)

    and

    MI(:,8:14)=(1ddd2d2d00000100000e2eee21e00000000000000010000000100000001de12d112d00000001dd0012e2e110000001e0e02dρ2(d+2e)ρ4eρdρdρ0e2eρ2eρ0(2d+e)ρeρ2dρd0221010)

    with ρ=μ+/μ being the jump ratio. The unknown vector cj and the right-hand-side vector ej take the form

    cj=(c+1j,c+2j,c+3j,c+4j,c+5j,c+6j,c+7j,c1j,c2j,c3j,c4j,c5j,c6j,c7j)t,ej=(δj1,δj2,δj3,δj4,δj5,δj6,δj7,0,0,0,0,0,0,0)t.

    We can obtain the vector-valued IFE shape functions ϕj,T by solving for cj with each vector ej, j=1,2,,7. Note that the matrices for Type Ⅱ and Type Ⅲ interface elements, denoted by MII and MIII, can be derived in a similar fashion; hence, we omit the details in this paper.

    As an illustration, we plot the three components of the CR-P1-P0 IFE shape function ϕ4,T in Figure 4. As a comparison, we plot the standard CR-P1-P0 FE shape function ψ4,T. We note that both FE and IFE shape functions are such that their second velocity components have the value one at the node A1. However, due to the coupled stress jump condition (8), the first velocity component and the pressure component of the IFE shape function ϕ4,T are not completely zero, as the FE shape function. This is a similar phenomenon that also occurs in other vector-valued IFE functions [1,20,27,29].

    Figure 4.  A comparison of the vector-valued IFE shape function ϕ4,T with μ=1, μ+=5 (top), and the corresponding FE shape function ψ4,T (bottom) on the reference triangle.

    The local CR-P1-P0 IFE space is formed by Sih(T)=span{ϕj,T:1j7}, and the global CR-P1-P0 IFE space is defined to be

    Sh(Th)={v=(v1,v2,vp)t[L2(Ω)]3:vsatisfies conditionsC1-C3}. (23)

    C1: v|TSnh(T),TTnh, and v|TSih(T)TTih.

    C2: e[[v1]]ds=0,e˚Eh.

    C3: v2 is continuous at every internal point (x,y)˚Nh.

    We can construct the P1-CR-P0 IFE space in a similar manner. In this case the edge-value conditions (15) are imposed on the second velocity component, and the nodal-value conditions (16) will apply to the first velocity component. The remaining conditions (17)-(21) are the same. Let ˜Sih(T) be the local P1-CR-P0 IFE space, then the corresponding global IFE space ˜Sh(Th) is defined as follows

    ˜Sh(Th)={v=[v1,v2,vp]t[L2(Ω)]3:satisfies conditionsC4-C6}. (24)

    C4: v|T˜Snh(T),TTnh, and v|T˜Sih(T)TTih.

    C5: v1 is continuous at every internal point (x,y)˚Nh.

    C6: e[[v2]]ds=0,e˚Eh.

    Remark 1. In many cases, the momentum equation (1) of the Stokes system is written as

    ut(2μϵ(u)pI)=f (25)

    where the stress is expressed using the strain tensor ϵ(u)=(u+(u)t)/2. In this setup, the stress jump condition (8) should also be changed to

    [[(2μϵ(u)pI)n]]Γ=0. (26)

    Since the viscosity coefficient μ(x) is a (piecewise) constant, the incompressibility condition (2) yields

    2μϵ(u)=μΔu.

    Hence, these two equations are equivalent in this case. In construction of IFE shape functions, only (19)-(20) need to be replaced by the following two conditions

    [[μ(2xϕ1,jn1+(yϕ1,j+xϕ2,j)n2)ϕp,jn1]]¯DE=0, (27)
    [[μ((xϕ2,j+yϕ1,j)n1+2yϕ2,jn2)ϕp,jn2]]¯DE=0. (28)

    The local CR-P1-P0 IFE functions, denoted by ϕϵj,T, and local P1-CR-P0 IFE functions, denoted by ˜ϕϵj,T can be constructed accordingly. The corresponding global IFE spaces are denoted by Sϵh(Th) and ˜Sϵh(Th).

    In this section, we present some basic properties of the mixed conforming-nonconforming IFE spaces.

    Theorem 3.1 (Unisolvency). The CR-P1-P0 IFE shape functions ϕj,T, 1j7 can be uniquely determined by the prescribed edge values, the nodal values, and the mean pressure value, regardless of the interface locations and the jumps of viscosity coefficients μ±>0.

    Proof. We show the unisolvency by considering the invertibility of the coefficient matrices MI, MII, and MIII. For the Type Ⅰ interface triangle, by direct calculation we have

    det(MI)=4(d4(1de)+d2e2(2de)+e4(1d)+ρde(d4+de2+d2e2+e3))<0.

    For the Type Ⅱ interface element, we have

    det(MII)=D1+ρD2

    where

    D1=4(1d)e((1+d)4+4(1+d)3e+7(1+d)2e2+(5+6d)e3+2e4)=4(1d)e((1d)34(1d)3e+7(1d)2e26(1d)e3+e3+2e4)4(1d)e((1d)34(1d)3e+7(1d)2e26(1d)e3+3e4)=4(1d)e((1d)2(1d2e)2+3e2(1de)2)<0,

    and with s=1d, we have

    D2=4(4e2(es)2+s2(2es)2)4es(2e4+e3(16s)+7e2s24es3+s4)4(4e2(es)2+s2(2es)2)4es(3e46e3s+7e2s24es3+s4)=4(4e2(es)2+s2(2es)2)4es(3e2(es)2+s2(2es)2)<0.

    For the Type Ⅲ interface element, we have

    det(MIII)=D3+ρD4

    where

    D3=4(1+d)2(12(1+d)2d+d(4+d(1+2d))e(2+d)(1+2d)e2+(2+d)e3)=4s2(s(1t)2t+s2(1t)+t(2s3+t22s2t))4s2(s(1t)2t+s2(1t)+ts3+t(s2t)2)<0

    and with s=1d and t=1e,

    D4=4(3s49s3t+s4t2s5t+8s2t2+2s3t2+2s4t24st3s2t3s3t3+t4)<0.

    The determinants of coefficient matrices are uniformly nonzero for all 0d1, 0e1, ρ>0 and for all three types of interface elements. This ensures the unisolvency of the IFE functions.

    The following theorems provide basic properties of the new IFE functions. The proofs of these results can be verified by direct calculation, hence we omit the proof in this paper. For more details, we refer the readers to some earlier references [1,20].

    Theorem 3.2 (Consistency). Let TTih be an interface triangle.

    If μ+=μ, the IFE shape functions ϕj,T become the FE shape functions ψj,T, 1j7.

    If the interface moves out of a triangle T, i.e.,

    min{|T|,|T+|}|T|0, (29)

    the IFE shape functions ϕj,T become the FE shape functions ψj,T, 1j7.

    Remark 2. The consistency (29) enables us to use IFE functions for solving Stokes moving interface problem efficiently. In fact, as the interface moves out of an element, the IFE functions smoothly convert to the FE functions. No extra condition is needed to enforce this transition.

    Theorem 3.3 (Continuity of Velocity). Let TTih be an interface element and ϕj,T be the vector-valued shape functions. Then the velocity components ϕi,jC(T), for i=1,2, and j=1,2,,7.

    Theorem 3.4 (Partition of Unity). Let TTih be an interface element. The vector-valued IFE shape functions ϕj,T, j=1,2,,7, satisfy the partition of unity property, namely: for any (x,y)T.

    3j=1ϕj,T(x,y)=(100),6j=4ϕj,T(x,y)=(010),ϕ7,T(x,y)=(001). (30)

    In this section, we first derive the weak form of the unsteady Stokes interface problem (1)-(8), and then develop the semi-discrete and full-discrete IFE schemes. We use (,)ω to denote the L2 inner product on a subset ωΩ. We will omit the subscript ω if ω=Ω.

    Taking the inner product with v[H10(Ω)]2 on the equation (1) and integrating by parts over Ω yields,

    (ut,v)Ω+(μupI,v)Ω((μupI)nΩ,v)Ω=(f,v)Ω.

    Here the second term is the inner product of two tensors A=[Aij] and B=[Bij], which is defined by (A,B):=i,j(Aij,Bij). Note that nΓ is pointing from Ω to Ω+ and v vanishes on the outer boundary Ω. We have

    (ut,v)Ω+(μupI,v)Ω((μupI)nΓ,v)Γ=(f,v)Ω.

    Similar argument applying to the subdomain Ω+ yields

    (ut,v)Ω++(μupI,v)Ω++((μupI)nΓ,v)Γ=(f,v)Ω+.

    Adding the above two equations together, and applying the interface jump condition (8), we have

    (ut,v)+(μu,v)(p,v)=(f,v).

    Multiplying qL2(Ω) to (2), and integrating by parts we have

    (q,u)=0. (31)

    Define the bilinear form and the linear form

    a(w,v)=(μw,v),w,v[H10(Ω)]2, (32)
    b(v,q)=(q,v),v[H10(Ω)]2,qL20(Ω). (33)

    Here, L20(Ω)={qL2(Ω):Ωqdx=0}. The weak form of the unsteady Stokes interface problem (1)-(8) is given as follows.

    Weak Form: Find uH1(0,T;[H10(Ω)]2) and pL2(0,T;L20(Ω)) such that for each t[0,T]

    (ut,v)+a(u,v)+b(v,p)=(f,v),v[H10(Ω)]2, (34)
    b(u,q)=0,qL20(Ω), (35)

    and subject to the initial conditions u(x,0)=u0(x), p(x,0)=p0(x).

    For semi-discretization in space, we use the CR-P1-P0 IFE space Sh(Th) to approximate to approximate [H10(Ω)]2×L2(Ω). We write the vector-valued IFE space Sh(Th)=U1h×U2h×Wh. Then we propose the semi-discrete scheme as follows.

    Semi-discrete IFE Scheme: Find (uh,ph):=(u1h,u2h,ph)H1(0,T;U1h)×H1(0,T;U2h)×L2(0,T;Wh) such that

    (tuh,vh)+a(uh,vh)+b(vh,ph)=(fh,vh),vhU1h×U2h, (36)
    b(uh,qh)=0,qhWh, (37)

    and subject to the initial conditions

    uh(x,0)=u0,h(x),p(x,0)=p0,h(x), (38)

    where u0,h and p0,h are some approximations (e.g. the interpolation) of u0 and p0 in U1h×U2h and Wh. We rewrite the semi-discrete scheme in the following matrix form.

    Matrix Form: Find the vector function U(t) such that

    M(t)U(t)+A(t)U(t)=F(t), (39)
    U(0)=U0, (40)

    where M(t) and A(t) denote the IFE mass and stiffness matrices, and F(t) is the vector corresponding to the right-hand side of (36)-(37). The initial vector U0 takes the values of the coefficients of the interpolation Ih(u0,p0). More details will be given in Section 5.

    Remark 3. Since the interface Γ(t) is a function of time t, the IFE spaces Sh(Th)=U1h×U2h×Wh depend on the interface location; hence they are time-dependent. Although the background mesh Th is time-independent, the collections of interface elements Ti(t)h and non-interface elements Tn(t)h vary by time. That is why the mass matrix M(t) and stiffness matrix A(t) are both time-dependent.

    Let 0=t0<t1<<tN1<tN=T be a partition of the time interval [0,T] with the uniform step size τ, i.e., τ=T/N, and tn=nτ. Evaluating (39) at t=tn+θ:=tn+θΔt, we have

    M(tn+θ)U(tn+θ)+A(tn+θ)U(tn+θ)=F(tn+θ). (41)

    Using the following finite-difference approximations in (41)

    M(tn+θ)U(tn+θ)M(tn+θ)U(tn+1)U(tn)τ1τ(M(tn+1)U(tn+1)M(tn)U(tn)), (42)
    A(tn+θ)U(tn+θ)(1θ)A(tn)U(tn)+θA(tn+1)U(tn+1), (43)
    F(tn+θ)(1θ)F(tn)+θF(tn+1), (44)

    we can obtain the following full-discrete IFE scheme.

    Full-discrete IFE Scheme: Given initial vector U0, find Un+1 for each n=0,1,,N1 in

    (1τMn+1+θAn+1)Un+1=(1τMn(1θ)An)Un+(1θ)Fn+θFn+1. (45)

    Note that when θ=1, the method becomes the Backward-Euler method:

    (1τMn+1+An+1)Un+1=1τMnUn+Fn+1. (46)

    When θ=12, the method is the Crank-Nicolson method:

    (1τMn+1+12An+1)Un+1=(1τMn12An)Un+12(Fn+Fn+1). (47)

    Remark 4. For the time-dependent Stokes interface problem with a stationary interface, i.e. Γ is time-independent, the matrices M and A in the full-discrete scheme (45) will remain unchanged as time evolves. As a result, at each time level, only the vector Fn needs to be updated.

    Remark 5. For the time-dependent Stokes interface problem with a moving interface, although the matrices Mn and An depend on the location of interface, which further depends on time, these matrices can be efficiently generated by locally modifying the matrices from the previous time step. A unique feature of IFEM is that not only is the computational mesh interface independent, but the number as well as the location of the unknowns also remain unchanged. In two consecutive time steps, only a small portion of elements change their interface configurations, as shown in Figure 5 marked in dark yellow color. Consequently, we only need to modify local stiffness and mass matrices on those elements. The majority of the global matrices remain unchanged. This feature is also important in the error analysis of IFE methods for moving interface problems, see [10].

    Figure 5.  An illustration of a moving interface in two consecutive steps. Elements in dark yellow indicate interface configuration changes, and elements in dark blue remain unchanged.

    In this section, we report some numerical experiments for the mixed conforming-nonconforming IFE methods for the Stokes interface problems. We test both the interpolation and the IFE solution with various configurations of the interface and coefficient jumps. All of our numerical experiments are performed on a family of Cartesian triangular meshes which are obtained by first partitioning the domain into Ns×Ns congruent rectangles, and then further dividing each rectangle into two triangles by its diagonal with the positive slope.

    We investigate the approximation property of IFE space by the interpolation. Define the CR-P1-P0 IFE interpolation operator is defined to be Ih:H1(Ω)×C(Ω)×L2(Ω)Sh(Th) such that

    Ih(u,p)|T=Ih,T(u,p)={7j=1cjϕj,T, if TTih,7j=1cjψj,T, if TTnh, (48)

    where ϕj,T and ψj,T are the local IFE/FE shape functions given in (14) and (11), respectively. For a fixed t, the coefficients cj take the values

    cj=1|ej|eju1(x,y)ds,1j3,cj=u2(Aj3),4j6,

    and

    c7=1|T|Tp(x,y)dxdy,

    where Aj and ej, j=1,2,3 are the vertices and edges of the the triangle T, respectively. The P1-CR-P0 interpolation can be defined similarly. The errors of the IFE interpolations are measured in L2 and semi-H1 norms as follows

    e0(u1,I)=u1u1,IL2(Ω),e0(u2,I)=u2u2,IL2(Ω),e0(pI)=ppIL2(Ω),
    e1(u1,I)=|u1u1,I|H1(Ω),e1(u2,I)=|u2u2,I|H1(Ω),

    where u1,I, u2,I, pI are components of the vector-valued function Ih(u,p). In the tables below, we report the convergence rate based on two consecutive meshes Th and Th/2, as well as the overall convergence rate among all meshes using the linear regression.

    Example 5.1 (Interpolation Accuracy). In this example, we test the approximation capability of the new vector valued IFE space using interpolation. Since the interpolation is a time-independent procedure, we use a steady-state solution given in [1,20] for this experiment. Let the domain be Ω=[1,1]2 and the interface be Γ={(x,y):x2+y2=0.3}. The circular interface separates the domain Ω into two subdomains Ω={(x,y):x2+y2<0.3} and Ω+={(x,y):x2+y2>0.3}. The exact solutions u1, u2 and p are defined as follows:

    u(x,y)={u1={y(x2+y20.3)μ+, if (x,y)Ω+,y(x2+y20.3)μ, if (x,y)Ω,u2={x(x2+y20.3)μ+, if (x,y)Ω+,x(x2+y20.3)μ, if (x,y)Ω,andp(x,y)=110(x3y3). (49)

    We first test a moderate coefficient contrast with μ=1 and μ+=10. Tables 1-2 report the interpolation errors using CR-P1-P0 and P1-CR-P0 IFE functions, respectively. We can see from these tables that the accuracy of these two IFE spaces are similar. Both of these interpolation errors obey

    e0(ui,I)O(h2),e1(ui,I)O(h),e0(pI)O(h), (50)
    Table 1.  CR-P1-P0 IFE Interpolation errors for Example 5.1 with μ=1 and μ+=10.
    N e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 5.36e-3 n/a 1.15e-2 n/a 7.02e-2 n/a 1.21e-1 n/a 1.54e-1 n/a
    16 1.39e-3 1.95 3.03e-3 1.92 3.14e-2 1.16 5.80e-2 1.06 7.32e-2 1.06
    32 3.59e-4 1.95 7.84e-4 1.95 1.46e-2 1.10 2.85e-2 1.02 3.73e-2 0.96
    64 9.20e-5 1.96 2.03e-4 1.95 5.28e-3 1.47 1.45e-2 0.98 1.91e-2 0.97
    128 2.33e-5 1.98 5.14e-5 1.98 2.10e-3 1.33 7.34e-3 0.98 9.66e-3 0.98
    256 5.85e-6 1.99 1.29e-5 1.99 8.47e-4 1.31 3.68e-3 1.00 4.85e-3 0.99
    rate 1.98 1.96 1.29 1.00 0.99

     | Show Table
    DownLoad: CSV
    Table 2.  P1-CR-P0 IFE Interpolation errors for Example 5.1 with μ=1 and μ+=10.
    N e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 1.16e-2 n/a 5.44e-3 n/a 1.44e-1 n/a 1.49e-1 n/a 1.30e-1 n/a
    16 3.08e-3 1.92 1.42e-3 1.94 5.93e-2 1.29 7.47e-2 1.00 5.80e-2 1.16
    32 5.15e-4 1.95 2.36e-4 1.96 2.14e-2 1.18 3.08e-2 0.96 2.37e-2 0.98
    64 7.94e-4 1.96 3.65e-4 1.97 2.70e-2 1.14 3.76e-2 0.99 2.88e-2 1.00
    128 5.15e-5 1.99 2.34e-5 1.99 3.56e-3 1.43 9.69e-3 0.98 7.35e-3 0.99
    256 1.29e-5 1.99 5.86e-6 2.00 1.32e-3 1.43 4.86e-3 0.99 3.68e-3 1.00
    rate 1.89 1.90 1.31 0.95 0.98

     | Show Table
    DownLoad: CSV

    where i=1,2.

    Next, we test a larger coefficient jump (μ=1 and μ+=200) and a flipped coefficients case (μ=10 and μ+=1). We only report the P1-CR-P0 IFE interpolation, since the CR-P1-P0 IFE results are close. Errors for large jump and flipped jump cases are listed in Tables 3-4, respectively. The convergence rates are again consistent with (50).

    Table 3.  P1-CR-P0 IFE Interpolation errors for Example 5.1 with μ=1 and μ+=200.
    N e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 1.01e-2 n/a 4.86e-2 n/a 2.81e-0 n/a 1.35e-1 n/a 1.26e-1 n/a
    16 2.73e-3 1.88 1.28e-3 1.92 1.21e-0 1.21 6.77e-2 1.00 5.31e-2 1.24
    32 7.19e-4 1.93 3.33e-4 1.95 5.75e-1 1.08 3.43e-2 0.98 2.66e-2 1.00
    64 1.86e-4 1.95 8.59e-5 1.97 1.98e-2 1.54 1.75e-2 0.97 1.34e-2 0.99
    128 4.73e-5 1.98 2.15e-5 1.98 7.26e-2 1.45 8.91e-3 0.98 6.79e-3 0.98
    256 1.19e-5 1.99 5.40e-6 1.99 2.59e-2 1.49 4.49e-3 0.99 3.41e-3 0.99
    rate 1.95 1.90 1.45 0.98 1.03

     | Show Table
    DownLoad: CSV
    Table 4.  P1-CR-P0 IFE Interpolation errors for Example 5.1 with μ=10 and μ+=1.
    Ns e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 5.11e-2 n/a 2.32e-2 n/a 3.38e-1 n/a 6.04e-1 n/a 4.61e-1 n/a
    16 1.29e-2 1.99 5.82e-3 1.99 9.59e-2 1.82 3.02e-1 1.00 2.29e-1 1.01
    32 3.23e-3 1.99 1.46e-3 2.00 2.36e-2 2.03 1.51e-1 1.00 1.15e-1 1.00
    64 8.09e-4 2.00 3.66e-4 2.00 1.07e-2 1.14 7.58e-2 1.00 5.73e-2 1.00
    128 2.02e-4 2.00 9.14e-5 2.00 3.41e-3 1.65 3.79e-2 1.00 2.87e-2 1.00
    256 5.06e-5 2.00 2.29e-5 2.00 1.37e-3 1.32 1.90e-2 1.00 1.43e-2 1.00
    rate 2.00 2.00 1.58 1.00 1.00

     | Show Table
    DownLoad: CSV

    Example 5.2 (Unsteady Stokes Equation with Fixed Interface). In this example, we consider a time-dependent Stokes equation with a fixed interface. The domain Ω and interface Γ is the same as in Example 5.1. The time domain is set to be [0,1], and it is partitioned uniformly to Nt subintervals. We use both backward-Euler and Crank-Nicolson schemes with the time step size τ=2h. The errors are measured at the final time t=1. The initial data u0, p0, the boundary condition, and the source term f are chosen so that the exact solutions of this problem are as follows

    u(x,y,t)={u1={y(x2+y20.3)μ+e3t, if (x,y)Ω+,y(x2+y20.3)μe3t, if (x,y)Ω,u2={x(x2+y20.3)μ+e3t, if (x,y)Ω+,x(x2+y20.3)μe3t, if (x,y)Ω,p(x,y)=110(x3y3). (51)

    Table 5 and Table 6 report the backward-Euler and the Crank-Nicolson IFE solutions at the final time t=1, respectively. The numerical results indicate that the errors of Crank-Nicolson are a little smaller than those of backward-Euler. They obey the expected convergence rates

    e0(uih)O(h2+τk),e1(uih)O(h+τk),e0(ph)O(h+τk), (52)
    Table 5.  P1-CR-P0 backward-Euler IFE solutions for Example 5.2 at t=1 with μ=1 and μ+=10.
    N e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 2.49e-1 n/a 1.72e-1 n/a 9.46e-0 n/a 2.95e-0 n/a 2.83e-0 n/a
    16 6.86e-2 1.86 4.70e-2 1.87 4.70e-0 1.01 1.51e-0 0.97 1.38e-0 1.03
    32 1.69e-2 2.02 1.18e-2 1.99 2.44e-0 0.95 7.65e-1 0.98 7.14e-1 0.96
    64 3.87e-3 2.13 3.54e-3 1.74 1.15e-0 1.08 3.94e-1 0.96 3.69e-1 0.95
    128 1.57e-3 1.31 1.65e-3 1.10 6.23e-1 0.88 2.04e-1 0.95 1.91e-1 0.95
    256 8.69e-4 0.85 9.07e-4 0.86 3.35e-1 0.90 1.07e-1 0.93 1.02e-1 0.91
    rate 1.69 1.54 0.97 0.96 0.96

     | Show Table
    DownLoad: CSV
    Table 6.  P1-CR-P0 Crank-Nicolson IFE solutions for Example 5.2 at t=1 with μ=1 and μ+=10.
    N e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 2.51e-1 n/a 1.72e-1 n/a 9.02e-0 n/a 2.94e-0 n/a 2.79e-0 n/a
    16 7.25e-2 1.79 5.02e-2 1.77 4.51e-0 1.00 1.50e-0 0.97 1.36e-0 1.04
    32 1.92e-2 1.92 1.39e-2 1.85 2.34e-0 0.94 7.62e-1 0.98 6.98e-1 0.96
    64 4.33e-3 2.15 3.27e-3 2.09 1.11e-0 1.08 3.92e-1 0.96 3.61e-1 0.95
    128 9.96e-4 2.12 7.94e-4 2.04 5.97e-1 0.89 2.03e-1 0.95 1.87e-1 0.95
    256 2.39e-4 2.06 2.33e-4 1.76 3.20e-1 0.90 1.06e-1 0.93 1.02e-1 0.91
    rate 2.03 1.93 0.97 0.96 0.96

     | Show Table
    DownLoad: CSV

    where i=1,2, and k=1 for backward-Euler, and k=2 for Crank-Nicolson. As before, we report only the P1-CR-P0 IFE solutions, and the results for the CR-P1-P0 IFE solution are similar.

    Example 5.3 (Unsteady Stokes Equation: Circular Moving Interface). In this example we test our mixed IFE method on a Stokes moving interface problem. The interface curve is a circle centered at origin with a varying radius. The function for the interface curve is given as

    Γ(x,y,t)=x2+y20.3(12sin(2πt)+1).

    It can be seen that at time t=0, the interface is the same as Example 5.2 with a radius of r=0.54772. As the time t increases, the radius will first increase, then decrease, and finally return to the original one. The maximum and minimum radius rmax=0.67082 and rmin=0.3873 occur at t=0.25 and t=0.75, as shown in Figure 6. The exact solution u is written in terms of the level-set interface function Γ=0:

    u(x,y,t)={u1={1μ+yΓ(x,y,t), if (x,y)Ω+(t),1μyΓ(x,y,t), if (x,y)Ω(t),u2={1μ+xΓ(x,y,t), if (x,y)Ω+(t),1μxΓ(x,y,t), if (x,y)Ω(t),p(x,y)=110(x3y3). (53)
    Figure 6.  CR-P1-P0 IFE Solution of Example 5.3 with μ=1 and μ+=10 on the 64×64 mesh at times t=0.25, 0.75, and 1. Top plots: Interfaces, middle: IFE solutions u1h, bottom: IFE solutions u2h.

    In this experiment, we set the time step size τ=h. We first test the moderate jump case for this moving interface problem. Table 7 reports the errors at the final time level of the backward-Euler IFE solutions. The error decay is observed to converge in an optimal order, as stated in (52). Figure 6 shows the IFE solution u1 and u2 at time t=0.25, t=0.75, and t=1, respectively, on the 64×64 mesh. For a larger jump case, the errors are reported in Table 8.

    Table 7.  CR-P1-P0 Backward-Euler IFE solution for Example 5.3 at t=1 with μ=1 and μ+=10.
    N e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 7.85e-3 n/a 1.14e-2 n/a 4.83e-1 n/a 1.36e-1 n/a 1.51e-1 n/a
    16 2.05e-3 1.94 2.95e-3 1.95 2.41e-1 1.00 7.02e-2 0.95 7.45e-2 1.02
    32 5.13e-4 2.00 6.54e-4 2.17 1.24e-1 0.96 3.57e-2 0.98 3.82e-2 0.96
    64 1.68e-4 1.61 1.32e-4 2.30 5.78e-2 1.10 1.84e-2 0.96 1.96e-2 0.96
    128 8.54e-5 0.98 6.68e-5 0.99 3.12e-2 0.89 9.52e-3 0.95 1.01e-2 0.95
    rate 1.67 1.93 1.00 0.96 0.97

     | Show Table
    DownLoad: CSV
    Table 8.  CR-P1-P0 Backward-Euler IFE solution for Example 5.3 at t=1 with μ=1 and μ+=200.
    N e0(u1,I) rate e0(u2,I) rate e0(pI) rate e1(u1,I) rate e1(u2,I) rate
    8 1.17e-2 n/a 1.29e-2 n/a 1.25e-0 n/a 1.44e-1 n/a 1.41e-1 n/a
    16 3.86e-3 1.60 4.56e-3 1.50 8.16e-1 0.61 7.99e-2 0.85 7.01e-1 1.01
    32 1.20e-3 1.69 1.42e-3 1.69 5.10e-1 0.68 3.80e-2 1.07 3.56e-2 0.98
    64 2.02e-4 2.57 2.50e-4 2.50 2.00e-1 1.35 1.74e-2 1.12 1.78e-2 1.00
    128 3.48e-5 2.54 4.21e-5 2.57 8.70e-2 1.20 8.43e-3 1.05 9.01e-3 0.98
    rate 2.10 2.07 0.97 1.04 1.04

     | Show Table
    DownLoad: CSV

    The condition numbers of the IFE systems are reported in Tables 9 and 10. We monitor the condition numbers at t=0.25, t=0.75, and t=1 which correspond to the interface circle listed in Figure 6. We test different contrast ratios by fixing the coefficient μ=1 and varying the other coefficient μ+=0.01, 0.1, 1, 10, and 100. Note that when μ+=1, there is no jump in coefficient, hence the IFE scheme becomes the standard FE scheme. Even in this no-jump case, we observe that the condition number is of order O(h4). We also observe that the condition number increases as the jump ratio enlarges. No significant differences have been noticed for backward-Euler and Crank-Nicolson in terms of the conditions numbers.

    Table 9.  Condition Number for Backward-Euler CR-P1-P0 Example 5.3 with μ=1.
    Ns μ+=0.01 μ+=0.1 μ+=1 μ+=10 μ+=100
    t=0.25 8 3.03e+05 5.94e+04 2.80e+05 1.38e+07 1.31e+09
    16 1.04e+06 7.82e+05 4.36e+06 1.11e+08 1.40e+10
    32 2.69e+08 6.06e+06 6.87e+07 9.06e+08 4.64e+11
    64 6.51e+10 7.07e+07 1.09e+09 8.46e+09 8.48e+12
    128 1.30e+12 7.27e+08 1.74e+10 8.15e+10 6.26e+14
    t=0.75 8 2.07e+04 4.24e+04 2.80e+05 1.22e+07 1.78e+09
    16 1.04e+06 7.82e+05 4.36e+06 1.64e+08 2.22e+10
    32 1.15e+08 9.36e+06 6.87e+07 1.67e+09 1.79e+11
    64 2.44e+09 1.11e+08 1.09e+09 1.62e+10 7.07e+13
    128 1.22e+10 9.29e+08 1.74e+10 1.16e+11 2.66e+15
    t=1 8 2.34e+06 3.68e+04 2.80e+05 1.26e+07 1.10e+09
    16 7.76e+06 5.65e+05 4.36e+06 1.08e+08 1.94e+10
    32 4.30e+07 8.53e+06 6.87e+07 1.41e+09 1.59e+13
    64 2.99e+08 9.10e+07 1.09e+09 1.05e+10 3.93e+13
    128 7.30e+11 8.24e+08 1.74e+10 9.94e+10 2.72e+15

     | Show Table
    DownLoad: CSV
    Table 10.  Condition Number for Crank-Nicolson CR-P1-P0 Example 5.3 with μ=1.
    Ns μ+=0.01 μ+=0.1 μ+=1 μ+=10 μ+=100
    t=0.25 8 4.92e+05 7.56e+04 2.88e+05 1.39e+07 1.32e+09
    16 1.43e+06 9.20e+05 4.42e+06 1.12e+08 1.40e+10
    32 3.29e+08 6.58e+06 6.92e+07 9.08e+08 4.65e+11
    64 7.54e+10 7.37e+07 1.10e+09 8.48e+09 8.48e+12
    128 1.43e+12 7.42e+08 1.74e+10 8.16e+10 6.26e+14
    t=0.75 8 3.52e+04 5.61e+04 2.88e+05 1.22e+07 1.78e+09
    16 7.29e+05 8.50e+05 4.42e+06 1.64e+08 2.22e+10
    32 1.51e+08 1.04e+07 6.92e+07 1.67e+09 1.79e+11
    64 3.00e+09 1.17e+08 1.10e+09 1.62e+10 7.08e+13
    128 1.39e+10 9.50e+08 1.74e+10 1.16e+11 2.66e+15
    t=1 8 3.29e+06 4.49e+04 2.88e+05 1.26e+07 1.10e+09
    16 1.02e+07 6.54e+05 4.42e+06 1.08e+08 1.95e+10
    32 5.58e+07 9.37e+06 6.92e+07 1.41e+09 1.59e+13
    64 3.71e+08 9.56e+07 1.10e+09 1.06e+10 3.93e+13
    128 8.04e+11 8.42e+08 1.74e+10 9.95e+10 2.73e+15

     | Show Table
    DownLoad: CSV

    In this paper, we developed a mixed conforming-nonconforming immersed finite element method for unsteady Stokes interface problems. The proposed vector-valued IFE spaces use conforming P1 approximation for one velocity component and nonconforming P1 approximation for the other. The pressure is approximated by piecewise constant. Unisolvency of the vector-valued IFE functions is proved. Interpolation errors are observed to be optimal which indicates these new IFE spaces have sufficient approximation capabilities. This new IFE function space can be used to solve steady-state, and unsteady Stokes interface problems. In addition, we have extended the application to a moving interface problem, and our numerical results show the optimal convergence in some benchmark tests. The proposed IFE method can be extended to some more general fluid flow interface problems such as Navier-Stokes equations, which will be an interesting future topic to explore.



    [1] An immersed discontinuous finite element method for Stokes interface problems. Comput. Methods Appl. Mech. Engrg. (2015) 293: 170-190.
    [2] On nonconforming linear-constant elements for some variants of the Stokes equations. Istit. Lombardo Accad. Sci. Lett. Rend. A (1993) 127: 83-93.
    [3] N. Chaabane, Immersed and Discontinuous Finite Element Methods, Thesis (Ph.D.)-Virginia Polytechnic Institute and State University. 2015.
    [4] Z. Chen, Finite Element Methods and their Applications, Scientific Computation. Springer-Verlag, Berlin, 2005.
    [5] A P2-P1 partially penalized immersed finite element method for Stokes interface problems. Int. J. Numer. Anal. Model. (2021) 18: 120-141.
    [6] Conforming and nonconforming finite element methods for solving the stationary Stokes equations. I. Rev. Franç caise Automat. Informat. Recherche Opérationnelle Sér. Rouge (1973) 7: 33-75.
    [7] Arbitrary Lagrangian-Eulerian method for Navier-Stokes equations with moving boundaries. Comput. Methods Appl. Mech. Engrg. (2004) 193: 4819-4836.
    [8] V. Girault and P.-A. Raviart, Finite Element Methods for Navier-Stokes Equations, volume 5 of Springer Series in Computational Mathematics, Springer-Verlag, Berlin, 1986. Theory and algorithms. doi: 10.1007/978-3-642-61623-5
    [9] An extended pressure finite element space for two-phase incompressible flows with surface tension. J. Comput. Phys. (2007) 224: 40-58.
    [10] Solving parabolic moving interface problems with dynamical immersed spaces on unfitted meshes: Fully discrete analysis. SIAM J. Numer. Anal. (2021) 59: 797-828.
    [11] A group of immersed finite element spaces for elliptic interface problems. IMA J. Numer. Anal. (2019) 39: 482-511.
    [12] A fixed mesh method with immersed finite elements for solving interface inverse problems. J. Sci. Comput. (2019) 79: 148-175.
    [13] R. Guo, T. Lin and Y. Lin, Recovering elastic inclusions by shape optimization methods with immersed finite elements, J. Comput. Phys., 404 (2020), 109123, 24 pp. doi: 10.1016/j.jcp.2019.109123
    [14] Improved error estimation for the partially penalized immersed finite element methods for elliptic interface problems. Int. J. Numer. Anal. Model. (2019) 16: 575-589.
    [15] A cut finite element method for a Stokes interface problem. Appl. Numer. Math. (2014) 85: 90-114.
    [16] Immersed finite element methods for elliptic interface problems with non-homogeneous jump conditions. Int. J. Numer. Anal. Model. (2011) 8: 284-301.
    [17] Immersed finite element methods for parabolic equations with moving interface. Numer. Methods Partial Differential Equations (2013) 29: 619-646.
    [18] Residual-based a posteriori error estimation for immersed finite element methods. J. Sci. Comput. (2019) 81: 2051-2079.
    [19] V. John, Finite Element Methods for Incompressible Flow Problems, volume 51 of Springer Series in Computational Mathematics, Springer, Cham, 2016. doi: 10.1007/978-3-319-45750-5
    [20] D. Jones and X. Zhang, A class of nonconforming immersed finite element methods for Stokes interface problems, J. Comput. Appl. Math., 392 (2021), 113493. doi: 10.1016/j.cam.2021.113493
    [21] A linear nonconforming finite element method for nearly incompressible elasticity and Stokes flow. Comput. Methods Appl. Mech. Engrg. (1995) 124: 195-212.
    [22] R. Lan and P. Sun, A monolithic arbitrary Lagrangian-Eulerian finite element analysis for a Stokes/parabolic moving interface problem, J. Sci. Comput., 82 (2020), Paper No. 59, 36 pp. doi: 10.1007/s10915-020-01161-9
    [23] A new local stabilized nonconforming finite element method for the Stokes equations. Computing (2008) 82: 157-170.
    [24] New Cartesian grid methods for interface problems using the finite element formulation. Numer. Math. (2003) 96: 61-98.
    [25] A method of lines based on immersed finite elements for parabolic moving interface problems. Adv. Appl. Math. Mech. (2013) 5: 548-568.
    [26] Partially penalized immersed finite element methods for elliptic interface problems. SIAM J. Numer. Anal. (2015) 53: 1121-1144.
    [27] A locking-free immersed finite element method for planar elasticity interface problems. J. Comput. Phys. (2013) 247: 228-247.
    [28] A nonconforming immersed finite element method for elliptic interface problems. J. Sci. Comput. (2019) 79: 442-463.
    [29] Linear and bilinear immersed finite elements for planar elasticity interface problems. J. Comput. Appl. Math. (2012) 236: 4681-4699.
    [30] T. Lin and Q. Zhuang, Optimal error bounds for partially penalized immersed finite element methods for parabolic interface problems, J. Comput. Appl. Math., 366 (2020), 112401, 11 pp. doi: 10.1016/j.cam.2019.112401
    [31] Distributed Lagrange multiplier-fictitious domain finite element method for Stokes interface problems. Int. J. Numer. Anal. Model. (2019) 16: 939-963.
    [32] Simple nonconforming quadrilateral Stokes element. Numer. Methods Partial Differential Equations (1992) 8: 97-111.
    [33] B. Rivière, Discontinuous Galerkin Methods for Solving Elliptic and Parabolic Equations, volume 35 of Frontiers in Applied Mathematics, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2008. Theory and implementation. doi: 10.1137/1.9780898717440
    [34] Fictitious domain finite element method for Stokes/elliptic interface problems with jump coefficients. J. Comput. Appl. Math. (2019) 356: 81-97.
    [35] A numerical solution of the Navier-Stokes equations using the finite element technique. Internat. J. Comput. & Fluids (1973) 1: 73-100.
    [36] A nonconforming Nitsche's extended finite element method for Stokes interface problems. J. Sci. Comput. (2019) 81: 342-374.
    [37] N. K. Yamaleev, D. C. Del Rey Fernández, J. Lou and M. H. Carpenter, Entropy stable spectral collocation schemes for the 3-D Navier-Stokes equations on dynamic unstructured grids, J. Comput. Phys., 399 (2019), 108897, 27 pp. doi: 10.1016/j.jcp.2019.108897
    [38] A 3D conforming-nonconforming mixed finite element for solving symmetric stress Stokes equations. Int. J. Numer. Anal. Model. (2017) 14: 730-743.
  • This article has been cited by:

    1. Mehdi Jamei, Mehdi Mosharaf-Dehkordi, Hamid Reza Ghafouri, A Sequentially- Hybridized Locally Conservative Non-conforming Finite Element Scheme for Two-phase Flow Simulation through Heterogeneous Porous Media, 2022, 162, 03091708, 104155, 10.1016/j.advwatres.2022.104155
    2. Yujia Chang, Yi Jiang, Rongliang Chen, A parallel domain decomposition algorithm for fluid-structure interaction simulations of the left ventricle with patient-specific shape, 2022, 30, 2688-1594, 3377, 10.3934/era.2022172
    3. Yuan Chen, Xu Zhang, Solving Navier–Stokes Equations with Stationary and Moving Interfaces on Unfitted Meshes, 2024, 98, 0885-7474, 10.1007/s10915-023-02414-z
    4. Yuan Chen, Songming Hou, Xu Zhang, Semi and fully discrete error analysis for elastodynamic interface problems using immersed finite element methods, 2023, 147, 08981221, 92, 10.1016/j.camwa.2023.07.014
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2688) PDF downloads(245) Cited by(4)

Figures and Tables

Figures(6)  /  Tables(10)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog