• 検索結果がありません。

side,butalsoallowsustoweaklyimposeaconditiononthejumpofthenormalderivativeinlieuoftheprescribedforcingfunctionintheearlierwork.Ournumericalexperiments intotwonon-overlappingsubdomains.Thistechniquenotonlyseparatesthestressoneach ],Nitsche’sformulationwase

N/A
N/A
Protected

Academic year: 2022

シェア "side,butalsoallowsustoweaklyimposeaconditiononthejumpofthenormalderivativeinlieuoftheprescribedforcingfunctionintheearlierwork.Ournumericalexperiments intotwonon-overlappingsubdomains.Thistechniquenotonlyseparatesthestressoneach ],Nitsche’sformulationwase"

Copied!
27
0
0

読み込み中.... (全文を見る)

全文

(1)

AN UNCONDITIONALLY STABLE SEMI-IMPLICIT CUTFEM FOR AN INTERACTION PROBLEM BETWEEN AN ELASTIC MEMBRANE AND AN

INCOMPRESSIBLE FLUID

KYLE DUNN†‡, ROGER LUI, ANDMARCUS SARKIS

Abstract.In this paper we introduce a finite element method for the Stokes equations with a massless immersed membrane. This membrane applies normal and tangential forces affecting the velocity and pressure of the fluid.

Additionally, the points representing this membrane move with the local fluid velocity. We design and implement a high-accuracy cut finite element method (CutFEM) which enables the use of a structured mesh that is not aligned with the immersed membrane, and we formulate a time discretization that yields an unconditionally energy stable scheme. We prove that the stability is not restricted by the parameter choices that constrained previous finite element immersed boundary methods and illustrate the theoretical results with numerical simulations.

Key words.immersed boundary method, finite element method, numerical stability, CutFEM, unfitted methods AMS subject classifications.65N12, 65N30, 74F10

1. Introduction. Fluid dynamics problems with immersed boundaries have arisen in many real world scenarios such as cardiac blood flow [40,42] and cell mechanics [2]. Two prevalent ideas are theimmersed boundary methodintroduced by Peskin [41] and theimmersed interface method by LeVeque and Li [31,32]. These are both finite difference methods developed for very involved problems. We note that the immersed interface method was also extended to the finite element method by imposing flux conservation and continuity of the solution strongly at certain points ofΓ; see [1,11,17,18,21,22,26,27,33,34]. In the immersed boundary method, the interface applies a local force when computing the fluid velocity and pressure globally at each time step. The right-hand side function is defined only on the interface and contains a Dirac delta function whose main purpose is to pass information between Eulerian and Lagrangian coordinates. Peskin’s use of a finite difference method requires a smoothing of the effects of the force applied by the membrane. Boffi and Gastaldi extended these ideas to the finite element method in [5]. In their work, a variational formulation in weak form is introduced, and the action of the forcing function, due to bending and stretching, is now written as an integral over the immersed membrane. One can also show that when the problem is written in the strong form, the force applied by the membrane to the fluid is equal to the jump in the normal stress [29] of the fluid across the membrane. The conditional energy stability of the method proposed in [5] was proved later in [6].

The framework of our finite element method begins with Nitsche’s formulation [39] in order to weakly impose Dirichlet boundary conditions on fitted meshes. In [23], Nitsche’s formulation was extended to the case where the domain boundary does not align with the underlying finite element mesh. In our work, we employ one particular fictitious domain finite element method known as CutFEM [8,9,12,13], which allows us to divide the global domain into two non-overlapping subdomains. This technique not only separates the stress on each side, but also allows us to weakly impose a condition on the jump of the normal derivative in lieu of the prescribed forcing function in the earlier work. Our numerical experiments

Received April 18, 2019. Accepted December 8, 2020. Published online on April 13, 2021. Recommended by Roland Herzog. This research was supported by the National Science Foundation Division of Mathematical Sciences Grant No. 1522663 Higher-Order Methods for Interface Problems with Non-Aligned Meshes.

Now at Cold Regions Research and Engineering Laboratory - U.S. Army ERDC, Hanover, NH 03755 (Kyle.G.Dunn@usace.army.mil).

Mathematical Sciences Department, Worcester Polytechnic Institute, Worcester, MA 01609 ({rlui, msarkis}@wpi.edu).

296

(2)

show that optimal spatial convergence can be obtained using CutFEM when the interface is described by a static, smooth parameterization.

CutFEM was implemented for the Stokes equation with an immersed boundary and a P1-iso-P2 element in [24], where Hansbo et al. use a known a priori level set method to track the interface. In that article it is noted that the optimality of their approach is independent of the interface representation, which moves with a prescribed velocity. Some additional work has been done on problems with a known interface velocity [25]. Our approach focuses on the movement of the interface not known a priori. In other words, we let the interface move with the velocity of the fluid, which is not known prior to solving the system at each time step. Recently, this approach was considered in [16], wherein the interface was considered as the zero level set of the function that is the solution of a pure advective partial differential equation with the vector field equal to the fluid velocity. Here in this paper, we implement a Q2-P1 time-depedent Stokes element [28,30,35,36,44] and track the immersed boundary by updating the position of a fixed number of points sampled from the initial curve. One of the contributions of our work here is the semi-implicit discretization of the force term and the corresponding energy estimates for this fluid-membrane problem. As can be seen on the right-hand side of (3.2), the implicitness of a fully-implicit method would appear in two places: a) the position of the interface at timetn+1, which is needed for integration, and b) the interface deformation, which depends on the velocity attn+1. The idea of the semi-implicit method is to make the method explicit in a) and implicit via extrapolation in b). This idea can be found in general free-boundary problems such as in [19]; see also [3,38,43].

Using techniques similar to those presented in [3,19,38,43], we show that our method is energy stable on the finite difference immersed boundary method. We believe that the method presented in this paper can be extended to time-depedent two-phase flows using results from static interface problems with unfitted meshes for two-phase flows, which has been used by many groups; see, e.g., [4,9,14,15].

This paper is organized as follows. In Section2the model is derived, and the strong formulation of the spatially continuous problem is introduced. Then in Section3we look into the time discretization of the problem and prove stability results, building to the fully discretized problem. In Section4we introduce the necessary notation and spaces of functions en route to defining our finite element methods. We proceed to prove energy stability of the proposed finite element problem, which is unconditional for our semi-implicit method and yields a CFL-like condition for the explicit method. The results of some numerical tests are shown and discussed in Section5, and we draw conclusions in Section6.

2. Model formulation. Consider a domainΩinRd,d= 2,3, which can be any Lips- chitz domain. For simplicity, we will defineΩ := (0,1)2. The following equations model an elastic material insideΩusing the time-dependent incompressible Stokes equations. The stress tensor is defined byT:=−µε(u) +Ip, whereε(u) =12(∇u+ (∇u)T)andpis the pressure. To reduce notational clutter, defineµto be twice the traditional dynamic viscosity.

We assume thatµ >0in order to impose proper boundary and interface conditions based on integration by parts of the viscous term. InsideΩthere will be a closed curveΓrepresenting a massless, elastic interface between two non-overlapping subdomainsΩ1andΩ2. Throughout this paper, we letΩ1denote the region exterior to the curveΓsuch that∂Ω1=∂Ω∪Γand Ω2denote the interior region encapsulated byΓ. Throughout this work,µis assumed to be constant. We note that our results, with some minor modifications, hold also for the case where µjumps acrossΓand varies mildly inside eachΩi; see [11,14,24] and Remark4.4.

The description of the interfaceΓand the model for the jump of the stress acrossΓare based on the immersed boundary method; see Peskin [41] and Boffi et al. [6]. As we will see in Remark3.1, it is advantageous to describeΓ, and therefore the jump of the stress, at time

(3)

tin parametric formΓ(s, t)fors∈[0, L]and fixedLindependent of time. In general,sis notan arc-length parameterizaton ofΓat any timet. We useX(s, t)to denote the Cartesian coordinates ofΓ(s, t)corresponding to a pointsfor any given timet. Since this is a closed curve,X(0, t) =X(L, t)for all time. To constructΓ(s, t), we first defineΓ(s,0)given by a parametrizations∈[0, L]. The fact thatΓis not necessarily parameterized by arc-length allows us to define an initial elastic membrane not only with bending, but also with stretching, that is,|∂X/∂s|is not necessarily equal to one. For timet >0, we letX(s, t)be the material point on the elastic membrane that moves from an initial positionX(s,0), and also we assume that the movement of a pointX(s, t)on the interface is given by the fluid velocity at that point.

Hence, we impose continuity of the velocity [[u]] = 0

on the interface. For a quantityφdefined overΩ, we denoteφ1=φ|1andφ2=φ|2. Then [[φ]] = (φ1−φ2)|Γdenotes the jump ofφacrossΓat a given point. We also impose a no-slip condition on the interface, that is,

∂X(s, t)

∂t =u(X(s, t), t).

The unit tangent vector, chosen to be in the direction of the parameterization, is defined in terms ofsby

τ = 1

|∂X/∂s|

∂X

∂s.

The boundary tensionT(s, t)of the elastic membrane is modeled using a generalized Hooke’s law, where

T(s, t) =σ(|∂X/∂s|;s, t),

and the functionσis defined below. By computing the elastic force on an arbitrary segment between two pointsaandb, we find that

(Tτ)(b, t)−(Tτ)(a, t) = Z b

a

∂s(Tτ)(s, t)ds.

Since this equality holds for any choice ofaandb, we know the force onΓis defined in terms ofsby

(2.1) F= ∂

∂s(Tτ).

A slight modification of the proof of [29, Theorem 1] shows that for a forceFdefined in terms ofs∈[0, L],

[[pn]] =−(F·n)n

|∂X/∂s| µ[[ε(u)n]] = F−(F·n)n

|∂X/∂s| .

It follows from (2.1) that if we chooseσ(|∂X/∂s|;s, t)to be proportional to|∂X/∂s|, i.e., σ(|∂X/∂s|;s, t) =κ|∂X/∂s|, then the jump condition is defined by

(2.2) [[(µε(u)−p)n]] = F(s, t)

∂X∂s(s, t)

= κ

∂X∂s(s, t)

2X

∂s2(s, t).

(4)

Physically, (2.2) means that the elastic interface will apply a force as it is stretched or bent at a given point. Here, the jump condition is defined in terms of the respective quantities restricted toΓ(t). For example,

[[(µε(u)−p)n]] = (µε(u1(X(s, t), t))−p1(X(s, t), t))

−(µε(u2(X(s, t), t))−p2(X(s, t), t))

n(X(s, t), t) for anys ∈[0, L]. To ease the notation, we denoten=n1, i.e., the unit normal pointing outward from the exteriorΩ1.

We further impose a homogeneous Dirichlet condition on∂Ωand homogeneous initial conditions. Combining the Stokes equations with the continuity of velocity and (2.2), the strong form of the equations to be solved is given in Problem1.

Problem 1: Strong formulation

FindX(s, t),ui(x, t), andpi(x, t), fori= 1,2, such that for allt∈(0, T),

∂ui

∂t −µ∇ ·ε(ui) +∇pi= 0 inΩi(t), i= 1,2, (2.3a)

∇ ·ui= 0 inΩi(t), i= 1,2, (2.3b)

[[(µε(u)−p)n]] = κ ∂X∂s

2X(s, t)

∂s2 fors∈[0, L], (2.3c)

[[u]] = 0 fors∈[0, L], (2.3d)

u1= 0 on∂Ω,

(2.3e)

∂X(s, t)

∂t =u(X(s, t), t) fors∈[0, L].

(2.3f)

u= 0 inΩi(0), i= 1,2, (2.3g)

REMARK 2.1. Note the time dependence of each subdomain and the location of the interface. When deriving the weak formulation, our spaces of test functions depend on time as well.

3. Discrete-time approximation. Given∆t, we consider equally-spaced time steps tn = n∆t, for 0 ≤ n ≤ Nt, whereNtis the number of time steps. We also let un = u(x, tn)andpn=p(x, tn)be the discrete time approximations to the velocity and pressure, respectively, to simplify notation. Then for eachn, letΓn = Γ(tn)be the interface separating the two subdomainsΩni = Ωi(tn). By our choice of notation, the location of the interfaceΓn in discrete time is updated usinguni, which has been solved within the subdomain from the previous time step,Ωn−1i . Below in the temporally discrete variations of (2.3a)–(2.3f), we use a backward-difference approximation for∂tui. In other words, the derivative with respect to time attn+1is approximated by∂tun+1i = (un+1i −uni)/∆t. We note thatuni is solved in Ωn−1i and thenΩni is obtained by movingΓ(tn−1)with speeduni. Sinceuni is obtained in Ωn−1i anduni must be integrated overΩni to solve forun+1i , we define

˜ un :=

un1(x), forx∈Ωn−11 , un2(x), forx∈Ωn−12 .

(5)

Thus,u˜nwill be used in the integration of the backward difference in all temporally discrete weak formulations below; see Problem2.

REMARK 3.1. In Nitsche’s formulation of the interface problem, we must substitute (2.3c) into an integral overΓ(t). We note that in its original form, we are integrating with respect to a time-dependent arc length parameterization of the interface. Since (2.3c) is defined in terms ofs, we transform this integral overΓ(t)to an integral over[0, L]. We have the following equalities:

[[(µε(u)−p)n]],{v}

Γ(t)= Z

[[(µε(u)−p)n]]· {v(x)}δ(x−X(s, t))dx

= Z L

0

[[(µε(u)−p)n]]· {v(X(s, t))}

∂X

∂s

ds

=− Z L

0

κ∂X

∂s(s, t)· ∂

∂s{v(X(s, t))}ds, where we denote the average of a functionφby{φ}= 12(φ|1+φ|2).

The weak formulation can be obtained by the usual integration by parts on (2.3a)–(2.3b) after multiplication by a test function. To symmetrize the problem for increased accuracy and computational efficiency, we add consistent terms to the weak formulation as seen in [10,23,24]. A nonsymmetric interior penalty method may also be used; see, e.g., [10,24].

Recall that each integral overΓnwill be expressed in terms ofs. We also writeXn(s) = X(s, tn)to simplify the notation. In the spatially continuous case, we simplify the inner products involving the jump condition on the interface as follows:

• Explicit method:

(3.1) [[(µε(u)−p)n]],{v}

Γn=−κ Z L

0

∂Xn(s)

∂s

∂s{v(Xn(s))}ds.

• Semi-implicit method:

[[(µε(u)−p)n]],{v}

Γn=−κ Z L

0

∂Xn+1

∂s

∂s{v(Xn(s))}ds

=−κ Z L

0

∂Xn

∂s + ∆t∂

∂s

un+1(Xn(s)) ∂

∂s{v(Xn(s))} ds.

(3.2)

Observe that the difference between (3.1) and (3.2) is the extrapolation used in the semi- implicit method, where we solve forXn+1in (2.3f). Note thatun+1(Xn) =

un+1(Xn) in the spatially continuous problem, and the average is included for comparison to the discrete case. The expression for the forcing function (3.2) incorporates the unknown velocity of the interface at the current time step.

We formulate the continuous-space, discrete-time Problem2, lettingY=Xn+1for the semi-implicit method andY=Xnfor the explicit method. For completeness, we note that the implicit method setsY=Xn+1and integrates overΩn+1andΓn+1in Problem2instead ofΩnandΓn, respectively. Additional challenges arise with the implicit method because Γn+1andΩn+1are not known prior to integration, so we omit this discretization. We have included the terms involving[[un+1]]in Problem2to compare to the one further discretized in space, Problem3, although[[un+1]] = 0in the current continuous setting. For the same reason, we include

∂sun+1(Xn(s)) =∂s un+1(Xn(s)).

(6)

Problem 2: Discrete-time weak formulation Given(u01,u02)∈[H1(Ω01)]2×[H1(Ω02)]2, where

u01|∂Ω= 0, (p01, p02)∈(L2(Ω01)×L2(Ω02))/R, X0: [0, L]→Ω.

Find for all 1 ≤ n ≤ Nt −1 solutions (un+11 ,un+12 ) ∈ [H1(Ωn1)]2 × [H1(Ωn2)]2, (pn+11 , pn+12 )∈ L2(Ωn1)×L2(Ωn2)

/R, andXn+1: [0, L]→Ωsuch that un+11 = 0on∂Ω, [[un+1]] = 0onΓn, and

2

X

i=1

1

∆t(un+1i −u˜n,vi)n

i +µ(ε(un+1i ),ε(vi))n

i −(pn+1i ,∇ ·vi)n

i

− {(µε(un+1)−pn+1)n},[[v]]

Γn− [[un+1]],{µε(v)n}

Γn

− (µε(un+11 )−pn+11 )n,v1

∂Ω− un+11 , µε(v1)n

∂Ω

= Z L

0

κ∂2Y

∂s2 (s){v(Xn(s))}ds (3.3a)

2

X

i=1

−(∇ ·un+1i , qi)ni + [[un+1]],{qn}

Γn+ (un+11 , q1n)∂Ω= 0 (3.3b)

for all(v1,v2)∈[H1(Ωn1)]2×[H1(Ωn2)]2and(q1, q2)∈L2(Ωn1)×L2(Ωn2), and Xn+1(s)−Xn(s)

∆t ={un+1(Xn(s))} fors∈[0, L].

(3.3c)

3.1. Energy estimates. The proposed semi-implicit method combines the analytical simplicity and stability of the implicit method in [6] with the computational convenience of the explicit method. For a quantityφ(s)defined onΓ, we define the norm over the reference configurationR= [0, L]to be

(3.4) kφk2L2(R):=

Z L 0

φ(s)2 ds.

If we define total energy to be the sum of the kinetic and elastic energies

(3.5) En :=1

2kunk2L2(Ω)+1 2κ

∂Xn

∂s

2

L2(R)

,

then the following lemma shows that the energy of the system computed usingY=Xn+1is monotonically decreasing.

LEMMA3.2.Letun+1,pn+1, andXn+1be solutions to(3.3a)–(3.3c)at timetn+1with Y=Xn+1. Then the following equality holds:

En+1=En−1 2

un+1−un

2

L2(Ω)−∆t

2

X

i=1

µ

ε(un+1i )

2 L2(Ω)

−1

2κ(∆t)2

Γun+1

2 L2n). (3.6)

(7)

Proof. Begin by lettingv = un+1 andq = pn+1in (3.3a)–(3.3b) and subtract (3.3b) from (3.3a), whereun+1stands forun+1=un+1i onΩni, fori= 1,2. We note that for each time step we have[[un+1]] = 0onΓn andun+11 = 0on∂Ω. Thus, these boundary terms disappear in (3.3a)–(3.3b). Using the symmetry of the bilinear form we are able to simplify the difference of (3.3a) and (3.3b) to

1

∆t(un+1−un,un+1)+µ(ε(un+1),ε(un+1))

+κ Z L

0

∂Xn+1

∂s (s)∂

∂s

un+1(Xn(s)) ds= 0.

(3.7)

First, we can rewrite

2

X

i=1

µ(ε(un+1i ),ε(un+1i ))ni =

2

X

i=1

µ

ε(un+1i )

2

L2(Ωni)

ε(un+1)

2 L2(Ω).

Now simplifying the forcing term in (3.7), we have

κ Z L

0

∂Xn+1

∂s (s)∂

∂s

un+1(Xn(s)) ds

= κ

∆t Z L

0

∂Xn+1

∂s

∂Xn+1

∂s −∂Xn

∂s

ds

= κ

2∆t Z L

0

∂Xn+1

∂s +∂Xn+1

∂s −∂Xn

∂s +∂Xn

∂s

∂Xn+1

∂s −∂Xn

∂s

ds

= κ

2∆t Z L

0

"

∂Xn+1

∂s 2

+

∂Xn+1

∂s −∂Xn

∂s 2

− ∂Xn

∂s 2#

ds

= κ

2∆t

∂Xn+1

∂s

2

L2(R)

+

∂Xn+1

∂s −∂Xn

∂s

2

L2(R)

∂Xn

∂s

2

L2(R)

!

= κ

2∆t

∂Xn+1

∂s

2

L2(R)

+

∆t ∂

∂sun+1(Xn)

2

L2(R)

∂Xn

∂s

2

L2(R)

!

= κ

2∆t

∂Xn+1

∂s

2

L2(R)

+ ∆t2

Γun+1

2 L2n)

∂Xn

∂s

2

L2(R)

! .

Using a similar manipulation for the first term on the left-hand side of (3.7), we obtain by a simple calculation that

(un+1−un,un+1)=1 2

un+1

2

L2(Ω)+

un+1−un

2

L2(Ω)− kunk2L2(Ω)

.

Applying the above simplifications to each term in (3.7) and multiplying by∆twe have (3.6).

We now turn to the explicit method, whose solution must satisfy equations (3.3a)–(3.3c) withY =Xn. The velocityun+1and pressurepn+1are computed by explicitly using the interface locationΓnand subdomainsΩni determined in the previous time step. The energy estimate for the explicit method is similar to that of the semi-implicit method but lacks the

(8)

stabilizing contribution of the extrapolation used to compute the force of the membrane in (3.2).

We have the following lemma.

LEMMA3.3.Letun+1,pn+1, andXn+1be solutions to(3.3a)–(3.3c)at timetn+1with Y=Xn. Then the following equality holds:

En+1=En−1 2

un+1−un

2

L2(Ω)−∆t

2

X

i=1

µ

ε(un+1)

2 L2(Ω)

+κ 2(∆t)2

Γun+1

2 L2n). (3.8)

Proof.We begin with the simplification made in the previous proof:

1

∆t(un+1−un,un+1)+µ(ε(un+1),ε(un+1))

+κ Z L

0

∂Xn

∂s (s)∂

∂s

un+1(Xn(s)) = 0.

(3.9)

The proof for the explicit case is identical to the proof in the semi-implicit case with one important difference in the treatment of the final term in (3.9). We have

κ Z L

0

∂Xn

∂s (s) ∂

∂s

un+1(Xn(s))

= κ

∆t Z L

0

∂Xn

∂s

∂Xn+1

∂s −∂Xn

∂s

ds

= κ

2∆t Z L

0

∂Xn+1

∂s −∂Xn+1

∂s +∂Xn

∂s +∂Xn

∂s

∂Xn+1

∂s −∂Xn

∂s

ds

= κ

2∆t Z L

0

"

∂Xn+1

∂s 2

∂Xn+1

∂s −∂Xn

∂s 2

− ∂Xn

∂s 2#

ds

= κ

2∆t

∂Xn+1

∂s

2

L2(R)

∂Xn+1

∂s −∂Xn

∂s

2

L2(R)

∂Xn

∂s

2

L2(R)

! .

The simplification of the last term in (3.9) shown above is almost identical to Lemma3.2, but the important difference is that the middle term in the final line above is negative. Now the energy may not be decreasing.

To make more sense of the norm involving bothXn+1andXn, we can write it in terms of the surface gradient of the velocity onΓas follows:

κ 2∆t

∂Xn+1

∂s −∂Xn

∂s

2

L2(R)

= κ

2∆t

∆t ∂

∂sun+1(Xn)

2

L2(R)

=κ∆t 2

Γun+1

2 L2n).

We substitute the final expression into (3.9) along with the simplification of the time-derivative term in the proof of Lemma3.2to get (3.8).

We note that for the discrete case, a trace theorem and an inverse inequality can be used for ∇Γun+1

L2n)to establish conditional stability; see [6]. From now on, we focus only on the unconditionally stable semi-implicit method since the explicit case can be treated similarly.

(9)

4. Discrete-space finite element approximation. The spatial discretization of the prob- lem requires two steps. First, the interfaceΓis discretized. Recall that we create a mapping from the interval[0, L]toΓ(s, t)withX(0, t) =X(L, t)that is not necessarily an arc-length parameterization. We choose equally-spaced points0 =s0< s1<· · ·< sm=Lby letting

˜h=L/mandsj =j˜h. We note that the set of points{sj}mj=0need not be evenly spaced but is chosen so for computational convenience. Then the initial immersed boundaryΓ0is approximated by a polygonΓ0˜

hwithmvertices, where thejth vertex is obtained by evaluat- ingX0j =X(sj,0). While{sj}mj=0may be equally-spaced for computational convenience, dist(Xn(sj),Xn(sj+1))may not be uniform. The interface is then approximated by linear segments between the points{Xnj}.

Second, we discretize the bulk fluid. The polygonal approximationΓ0˜

hdividesΩinto the two approximate subdomains. As the discrete interface moves, these subdomains will change and are denoted byΩn

i,˜hat timetn. LetThpartitionΩinto squares with side lengthh. Then the subset ofThthat overlaps eachΩn

i,h˜is denoted by

Ti,hn :={K∈ Th:meas2(K∩Ωni,˜h)>0},

where measd denotes the Lebesgue measure inddimensions. These sets of elements are further decomposed into two disjoint sets,Ti,hn,IandThn,Γ. We define the set of elements of Ti,hn strictly interior toΩi,˜hby

Ti,hn,I :={K∈ Ti,hn :K⊂Ωni,h˜}.

Similarly, the set of elements ofTi,hn whose interior is intersected by the interfaceΓn˜

his defined by

Thn,Γ:={K∈ Th:meas1(K∩Γn˜h)>0}.

Thus, for eachiandnthe relationshipTi,hn =Ti,hn,I∪ Thn,Γholds. Consider the union of all elements inTi,hn; we define the interior of each union to be the extended subdomainΩn,ei,h. These subdomainsΩn,ei,h depend on bothΓn˜

handhand can be formally defined by Ωn,ei,h :=Int

 [

K∈Ti,hn

K

.

Many approximate quantities depend onhand˜h, although only one is used as a subscript. For example,uni,h,pni,h,Thn,Γ, and others depend on bothhand˜h. The set of points{Xnj}mj=0 depends only onh, and the polygon will be refined as˜ ˜hdecreases for fixedL.

It is worth noting that adjacent points on the discrete interface will not form clusters in the dynamic system and need not be manually redistributed. As a set of adjacent points begin to cluster, other points on the interface must be stretched away from one another due to the incompressibility of the fluid inside the interface. The tangential elastic forces generated along the interface by the stress jump condition (2.2) pull the clustering points away from one another to alleviate localized higher tension; see Example4below.

4.1. Finite element problem. For each set of elementsTi,hn we define the finite element spaces

Vi,hn :=

v∈[C0(Ωn,ei,h)]2:v|K ∈[Q2(K)]2,∀K∈ Ti,hn

(10)

FIG. 4.1.Plot of example meshes withFi,hn,Γhighlighted with a red dotted line for each subdomain.

and

Mi,hn :={q∈L2(Ωn,ei,h) :q|K ∈ P1(K), ∀K∈ Ti,hn}.

RecallP1(K)is the space of linear functions defined on an elementK. A generalq∈Mi,hn is discontinuous across each edge of the elements since a linear function in two variables is defined by its value at three points. The spaceQ2(K)consists of biquadratic functions defined on the elementKwith nine local degrees of freedom. A generalv ∈Vi,hn has components which are continuous function across the elements.

Additional “ghost" penalty terms are included to mitigate the jumps of the flux and pressure across the faces of elements, particularly to minimize spiking at the ghost nodes and spurious oscillations. To add these to the minimizing functional, we first need to define the sets of edges over which these jumps will be minimized, denotedFi,hn,Γ. Informally, we describe eachFi,hn,Γas the union of all edges shared by two elements inTi,hn, where at least one of the elements is inThn,Γ. Formally, these sets are defined by

Fi,hn,Γ={K∩K0 :K6=K0, andK∈ Thn,Γ, K0 ∈ Ti,hn}.

Figure4.1showsFi,hn,Γfor each subdomain.

LetKandK0be adjacent square elements withF =K∩K0and defineφonKandφ0 onK0. Below,[φ] =φ|F −φ0|F denotes the jump of a function over the faceF. Then the stabilizing ghost penalty terms are defined by

ji,h(ui,h,vi,h) =

1

X

`=0

X

F∈Fi,hn,Γ

Z

F

h2`+1h

(`)nF(ε(ui,h)nF)i

·h

n(`)F(ε(vi,h)nF)i ,

Ji,h(pi,h, qi,h) =

1

X

`=0

X

F∈Fi,hn,Γ

Z

F

h2`+1h

(`)n

Fpi,hi h

n(`)

Fqi,hi ,

where ∂(0)nFφ = φand∂n(1)Fφstands for the derivative of each component ofφin thenF

direction.

Since the interface cuts through elements, we must weakly impose interface conditions acrossΓn˜

h. The jump of the stress is incorporated naturally by substitution into the integral

(11)

resulting from integration by parts. To impose the weak interface continuity condition and also the weak flux continuity condition, we must add mathematically consistent penalty terms

(4.1) γ1

µ h Z

Γn˜

h

[[uh]]·[[vh]] and γ2

h

∆t Z

Γn˜

h

[[uh·n]][[vh·n]]

for someγ1 >0 andγ2 > 0. We note that the second penalty term above is required to establish the inf-sup condition whenh/∆tdominatesµ/h. Ifuis the exact solution, then the jump ofuis equal to zero and (4.1) will vanish for the velocity satisfying the system of equations (2.3a)–(2.3f). Thus, addition of (4.1) will keep the variational formulation consistent with the original problem. We enforce the Dirichlet boundary condition and zero-flux on∂Ω by adding the penalty terms

(4.2) γ1

µ h Z

(u1,h−0)·v1,h and γ2

h

∆t Z

∂Ω

(u1,h−0)·nv1,h·n.

The parameterization coordinate of thejth vertex ofΓn˜

his denotedsj, where0≤j≤m, and the corresponding Cartesian coordinate pairs areXn˜

h,j =X˜h(sj, n∆t). Additionally, s0 = 0andsm=Lso thatXn˜

h,0 =Xn˜

h,m for alln. To ease the notation, we will letXnj denote the coordinate pairXn˜h,j on the discrete interface.

For both explicit and semi-implicit temporal discretizations, we will find the following simplification of (3.3a) withY=Xnuseful. After integration by parts onΓn˜

h, we simplify the right-hand side of (3.3a) using the fact that ∂X∂sn is constant on each edge of the polygon Γn˜

h. If we define

∂Xnj

∂s = Xnj+1−Xnj sj+1−sj

,

then the resulting simplification is written

−κ Z L

0

∂Xn

∂s

∂s v(Xn˜

h(s)) ds=−κ

m−1

X

j=0

∂Xnj

∂s Z sj+1

sj

∂s v(Xn˜

h(s)) ds

=−κ

m−1

X

j=0

∂Xnj

∂s

v(Xnj+1) − v(Xnj)

m−1

X

j=0

∂Xnj+1

∂s −∂Xnj

∂s

v(Xnj+1) . (4.3)

To simplify notation in Problem3we drop the “h" or “˜h" subscript from the discrete approximation of the subdomainsΩnh,i, and the quantitiesunh,vhn,pnh,qhn, andXn˜

h. The definition ofγ1, γ2, γu, andγpare given later in the paper; see (4.10).

REMARK 4.1. To distinguish the difference between the explicit and semi-implicit methods in Problem3, we define a parameterνwhich can be set to either 0 or 1. Settingν = 1 yields the semi-implicit method, whileν = 0leaves us with the explicit method; see the last term of the left-hand side of (4.4b). We note that for the numerical tests we consider both cases.

(12)

Problem 3: Discrete-time finite element formulation 1. Solve forun+1i andpn+1i in

Fn+1,v

m−1

X

j=0

∂Xnj+1

∂s −∂Xnj

∂s

vn+1(Xnj+1) (4.4a)

2

X

i=1

1

∆t(un+1i −u˜ni,vi)ni + (µε(un+1i ),ε(vi))ni −(pn+1i ,∇ ·vi)ni

− {(µε(un+1)−pn+1)n},[[v]]

Γn˜

h

− [[un+1]],{µε(v)n}

Γn˜

h

− (µε(un+11 )−pn+11 )n,v1

∂Ω− un+11 , µε(v1)n

∂Ω

uji,hn (un+1i ,vi) +γ1

µ

h([[un+1]],[[v]])Γn˜

h1

µ

h(un+11 ,v1)∂Ω

2

h

∆t([[un+1·n]],[[v·n]])Γn˜

h2

h

∆t(un+11 ·n,v1·n)∂Ω

+νκ∆t Z n

Γ˜h

∂un+1(Xn)

∂s

∂s{v(Xn(s))} ds=

Fn+1,v (4.4b)

2

X

i=1

−(∇ ·un+1i , qi)n

i + [[un+1]],{qn}

Γn˜

h

+ (un+11 , q1n)∂Ω

−γpJi,hn (pn+1i , qi) = 0 (4.4c)

for allvn+1i ∈Vi,hn andqin+1∈Mi,hn ,i= 1,2.

2. Solve forXn+1j using

Xn+1j =Xnj + ∆t{un+1(Xnj)} forj = 0, . . . , m.

(4.4d)

4.2. Energy stability of the FEM. We are able to prove the unconditional stability of the semi-implicit method in Problem3under the assumption below, similar to that seen in [37].

ASSUMPTION1. GivenK∈ Thn,Γ, there existsK0 ∈ Ti,hn,I, an integerN >0, andN elements{Kk}Nk=1such thatK1=K,KN =K0, andKk∩Kk+1⊂ Fi,hn,Γ.

The next lemma is necessary to bound the strain on the extended subdomainΩn,ei,h by the strain on the original subdomainΩn

i,˜h. The result shows why it is necessary to include ji,h(u,v)for stability. Here and in the rest of the paper we do not use the fact that the mesh is structured.

LEMMA4.2.Supposev∈Vi,hn is continuous onΩn,ei,h and Assumption1holds forTi,hn. Then we have the following estimate:

kε(v)k2L2(Ωn,ei,h)≤Cε

kε(v)k2L2(Ωn

i,˜h)+ji,h(v,v)

,

whereCεdepends on neithervnorh.

Proof.LetK1andK2be neighboring square elements with a shared edgeF =K1∩K2. Note thatkε(v)k2L2(Ωn,ei,h) andji,h(v,v)do not depend on the directions of the Cartesian coordinate axes in which we representv. Therefore, without loss of generality, we assume

(13)

nF = (1,0)andv = (u, v), and their derivatives are using the canonical representationx- andy-axis. We make use of the following result from [37, Lemma 5.1]:

(4.5) kvk2L2(K1)≤Cm

kvk2L2(K2)+ X

0≤`≤p

h2`+1 Z

F

|[∂n`Fv]|2

,

wherev|K1,v|K2are polynomial functions of degree less than or equal top. WithVi,hn defined as above, the summation in (4.5) simplifies top = 1. Denotingv = (u, v), we apply the inequality (4.5) directly to each term ofε(v) :ε(v) = (∂xu)2+12(∂xv+∂yu)2+ (∂yv)2 and add the inequalities. Note that on any vertical edge, the jump of ally-derivatives will be zero becausevis continuous acrossF andv|F is simply a polynomial in each component.

Also on a vertical edge, the unit normal vector isnF = (1,0), and it follows that [∂n`F(ε(v)nF)]·[∂`nF(ε(v)nF)] = ([∂x`xu])2+1

4([∂x`(∂xv+∂yu)])2. The resulting inequality is

kε(v)k2L2(K1)= Z

K1

(∂xu)2+1

2(∂xv+∂yu)2+ (∂yv)2

≤Cm Z

K2

(∂xu)2+1

2(∂xv+∂yu)2+ (∂yv)2 +

1

X

`=0

h2`+1 Z

F

[∂n`

Fxu]2+1 2[∂n`

F(∂xv+∂yu)]2

≤2Cm kε(v)k2L2(K2)+

1

X

`=0

h2`+1 Z

F

[∂n`F(ε(v)nF)]2

! .

Similarly for any horizontal edge, we letnF = (0,1)and use the fact that[∂xu] = 0to get the same inequality.

Using Assumption1, we are able to find a sequence of at mostN adjacent elements leading from an elementK1∈ Thn,Γto an elementKN ∈ Ti,hn,I. Applying the above inequality across each of the edgesFk=Kk∩Kk+1, for1≤k≤N−1, we have

(4.6) kε(v)k2L2(K1)≤(2Cm)N kε(v)k2L2(KN)+

N

X

k=1 1

X

`=0

Z

Fk

[∂n`

Fkε(v)nFk]2

! .

Repeating (4.6) for allK∈ Thn,Γand denotingCε= (2Cm)N completes the proof.

Recall the definition of the norm over the reference configurationR = [0, L]. For a quantityφ(s)that is constant on each linear segment of the polygonalΓn˜h, we can simplify (3.4) to

kφk2L2(R)=

m−1

X

j=0

(φ(sj))2 (sj+1−sj).

Above,φ(sj)denotes the value ofφ(s)on the interval(sj, sj+1). Using the definition of the energy (3.5), we are able to prove the following theorem.

(14)

THEOREM4.3. Letun+1i,h ,pn+1i,h , andXn+1˜

h be solutions to(4.4a)–(4.4d)at timetn+1 withν = 1and assume Assumption1holds. Then the following inequality holds:

En+1≤En−1 2

un+1h −unh

2

L2(Ω)+1 2κ

∂Xn+1˜

h

∂s −∂Xn˜

h

∂s

2

L2(R)

+ ∆t

γ1

µ h

Z

Γn˜

h

[[un+1h ]]2ds+γ1

µ h Z

∂Ω

(un+11,h )2ds +γ2

h

∆t Z

Γn˜

h

[[un+1h ·n]]2ds+γ2

h

∆t Z

∂Ω

(un+11,h ·n)2ds

+

2

X

i=1

γuji,h(un+1i,h ,un+1i,h ) +γpJi,h(pn+1i,h , pn+1i,h ) . (4.7)

Proof.We first letvh=un+1h andqh=pn+1h in (4.4a)–(4.4c) and subtract (4.4c) from (4.4b). After cancellation due to the symmetry of theL2-inner product and some simplification we have

2

X

i=1

1

∆t

un+1i,h −uni,h,un+1i,h

n

i,˜h

+

2

X

i=1

µ

ε(un+1i,h ) L2(Ωn

i,˜h)

−2 [[un+1h ]],{µε(un+1h )n}

Γn˜

h

−2

un+11,h , µε(un+11,h )n1

∂Ω

ujh(un+1i,h ,un+1i,h ) +γpJh(pn+1i,h , pn+1i,h ) +γ1

µ h Z

Γn˜

h

[[un+1h ]]21

µ h Z

∂Ω

(un+11,h )22 h

∆t Z

Γn˜

h

[[un+1h ]]22 h

∆t Z

∂Ω

(un+11,h )2

m−1

X

j=0

∂Xn+1j+1

∂s −∂Xn+1j

∂s

!

un+1h (Xnj+1) . (4.8)

The term on the right-hand side of (4.8) is obtained using (4.3) and (4.4d) as follows:

κ

m−1

X

j=0

∂Xn+1j+1

∂s −∂Xn+1j

∂s

!

un+1h (Xnj+1)

=−κ Z L

0

∂Xn+1

∂s

∂s

un+1h (Xnh˜(s)) ds

=−κ Z L

0

∂Xn

∂s

∂s

un+1h (Xn˜h(s)) ds

−κ∆t Z L

0

∂un+1h (Xn)

∂s

·

∂un+1h (Xn)

∂s

ds

m−1

X

j=0

∂Xnj+1

∂s −∂Xnj

∂s

un+1h (Xnj+1)

−κ∆t 2

Z L 0

∂un+1h (Xn)

∂s

·

∂un+1h (Xn)

∂s

ds.

参照

関連したドキュメント

The first paper, devoted to second order partial differential equations with nonlocal integral conditions goes back to Cannon [4].This type of boundary value problems with

Based on these results, we first prove superconvergence at the collocation points for an in- tegral equation based on a single layer formulation that solves the exterior Neumann

the existence of a weak solution for the problem for a viscoelastic material with regularized contact stress and constant friction coefficient has been established, using the

In Section 2, we discuss existence and uniqueness of a solution to problem (1.1). Section 3 deals with its discretization by the standard finite element method where, under a

“Breuil-M´ezard conjecture and modularity lifting for potentially semistable deformations after

Let F be a simple smooth closed curve and denote its exterior by Aco.. From here our plan is to approximate the solution of the problem P using the finite element method. The

Let F be a simple smooth closed curve and denote its exterior by Aco.. From here our plan is to approximate the solution of the problem P using the finite element method. The

We introduce a new iterative method for finding a common element of the set of solutions of a generalized equilibrium problem with a relaxed monotone mapping and the set of common