This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A new framework of high-order unfitted finite element methods using ALE maps for moving-domain problems

Wenhao Lu [email protected] Chuwen Ma111This author was supported in part by China Postdoctoral Science Foundation (No. 2023M732248) and Postdoctoral Innovative Talents Support Program (No. 20230219). [email protected] Weiying Zheng222This author was supported in part by National Key R & D Program of China 2019YFA0709600 and 2019YFA0709602 and by the National Science Fund for Distinguished Young Scholars 11725106. [email protected] LSEC, NCMIS, Institute of Computational Mathematics and Scientific/Engineering Computing, Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing, 100190, China. School of Mathematical Science, University of Chinese Academy of Sciences, Beijing, 100049, China. Institute of Natural Sciences, Shanghai Jiao Tong University, Shanghai 200240, China.
Abstract

As a sequel to our previous work [C. Ma, Q. Zhang and W. Zheng, SIAM J. Numer. Anal., 60 (2022)], [C. Ma and W. Zheng, J. Comput. Phys. 469 (2022)], this paper presents a generic framework of arbitrary Lagrangian-Eulerian unfitted finite element (ALE-UFE) methods for partial differential equations (PDEs) on time-varying domains. The ALE-UFE method has a great potential in developing high-order unfitted finite element methods. The usefulness of the method is demonstrated by a variety of moving-domain problems, including a linear problem with explicit velocity of the boundary (or interface), a PDE-domain coupled problem, and a problem whose domain has a topological change. Numerical experiments show that optimal convergence is achieved by both third- and fourth-order methods on domains with smooth boundaries, but is deteriorated to the second order when the domain has topological changes.

keywords:
Arbitrary Lagrangian-Eulerian unfitted finite element (ALE-UFE) method, moving domain problems, high-order schemes. MSC:   65M60,76R99,76T99

1 Introduction

Multiphase flows with time-varying domains or free-surface fluids have an extremely wide range of applications in science and engineering. Transient deformations of material regions pose a great challenge to the design of high-order numerical methods for solving such problems. In terms of the relative position of a moving domain to its partition mesh, current numerical methods can be roughly classified into two regimes. In body-fitted methods, the mesh is arranged to follow the moving phase, and the implementation of boundary conditions becomes easy. In the popular arbitrary Lagrangian-Eulerian (ALE) approach [13, 5, 7, 6, 9, 14], the mesh velocity can be chosen independent of local fluid velocity. One needs to transform the problem from a moving domain to a fixed reference domain through an arbitrary mapping and further mesh the reference domain. The method can also be used in combination with space-time Galerkin formulations[23, 25]. However, these conveniences incur the cost of mesh regeneration and data migration across the entire computational domain at each time step. Another major concern of body-fitted methods is how to maintain high accuracy in the presence of abrupt motions or large deformations of the phase [24].

In the other regime of unfitted methods, the mesh for the bulk phase is fixed while the moving interface is allowed to cross the fixed mesh, resulting in cut cells near the interface where the degrees of freedom are doubled with additional penalty terms or basis functions are modified to enforce the interface conditions weakly. Unfitted methods may encounter significant challenges for dynamic interface problems. Since the computational domain is time-varying, traditional methods for time integration may not be applicable [8, 29]. To overcome this difficulty, the space-time method in [16] combines a discontinuous Galerkin technique in time with an extended finite element method. The immersed finite element method exploits the invariant degrees of freedom and uses backward Euler for the time discretization [10, 1]. Methods in [15, 19, 27, 4] extend the discrete solution at each time-step by using a ghost penalty, which enables the use of a backward differentiation formula. Most recently, von Wahl and Richter developed an error estimate for this Eulerian time-stepping scheme for a PDE on a moving domain, where the domain motion is driven by an ordinary differential equation (ODE) coupled to the PDE [26]. In addition, the characteristic approach is applied in [21, 20, 22] to develop high-order unfitted finite element methods and in [12] for convection-diffusion problems on time-dependent surfaces.

This paper is inspired by the ALE method, where the mesh velocity is typically derived from boundary motion, which allows the fluid to move with respect to the mesh. We exploit the ALE map to construct a backward flow map and then approximate the time derivative, which makes it easier to obtain high-order convergence rates even for large deformations. The basic idea behind is to replace the partial time derivative with a material derivative,

tu(x,tn)=ddtu(𝒜n(x,t),t)|t=tn𝒘n(𝒙,tn)u(x,tn),\partial_{t}u(x,t_{n})=\frac{\mathrm{d}}{\mathrm{d}t}u(\mathcal{A}^{n}(x,t),t\big{)}\big{|}_{t=t_{n}}-{\bm{w}}^{n}({\bm{x}},t_{n})\cdot\nabla u(x,t_{n}),

where 𝒜n(x,t)\mathcal{A}^{n}(x,t) is an arbitrary mapping that maps 𝒙Ωtn{\bm{x}}\in\Omega_{t_{n}} to y=𝒜n(x,t)Ωty=\mathcal{A}^{n}(x,t)\in\Omega_{t}, for any ttnt\leq t_{n}, and 𝒘n=t𝒜n(x,t){\bm{w}}^{n}=\partial_{t}\mathcal{A}^{n}(x,t) is the velocity of the moving domain. The arbitrary feature of the algorithm arises from the fact that this application dose not follow trajectories of bulk fluid particles, but only of boundary fluid particles. In particular, if 𝒘n{\bm{w}}^{n} coincides with the fluid velocity, the method turns out to be the unfitted characteristic finite element (UCFE) method proposed in [21, 22, 20]. Compared with the UCFE method, the proposed method – ALE-UFE has two advantages:

  1. (i)

    In the framework of unfitted finite element method, it is difficult to calculate the integration of the numerical solutions from early time steps at the present time step, since they do not belong to the present finite element space, especially in nonlinear problems where the moving interface depends on the solution of the equation. The ALE-UFE method overcomes this difficulty by choosing a simple ALE mapping to construct a backward flow map (see section 4.3).

  2. (ii)

    Since the construction of the ALE map requires only the position of the moving interface and not the internal variation of the region as in the UCFE approach, the ALE-UFE method can be applied to more models, including those where the moving region does not maintain the same volume (see section 6.1) and the topology changes (see section 6.4).

The main contributions of this work are summarized as follows.

  1. (a)

    Based on the heat equation in a time-evolving domain, we propose a new framework of designing high-order unfitted finite element methods with ALE maps. We prove the stability of the numerical solution in the energy norm and establish optimal error estimates, arising from both interface-tracking algorithms and spatial-temporal discretization of the equations.

  2. (b)

    The competitive performance of the ALE-UFE method is demonstrated by numerical experiments on largely deforming domains, including a PDE-domain coupled model, a two-fluid model with moving interface, and a problem with topologically changing domain. Optimal convergence of the method is obtained for domains with smooth boundaries or interface, but is deteriorated for domains with topological changes.

The rest of the paper is organized as follows. In section 2, we introduce the forward and backward flow maps. In section 3, we propose the ALE-UFE method for the heat equation on a moving domain, and establish the stability and error estimates of the numerical solution. A general framework of the ALE-UFE method is presented. In section 4, we apply the ALE-UFE method to a PDE-domain coupled problem. In section 5, the ALE-UFE method is applied to a two-phase flow problem. In section 6, we present numerical results to demonstrate optimal convergence of both the third- and fourth-order methods, and apply the method to more challenging problems.

Throughout this paper, vector-valued quantities and matrix-valued quantities are denoted by boldface symbols and blackboard bold symbols, respectively, such as 𝑳2(Ω)=L2(Ω)2{\bm{L}}^{2}({\Omega})=L^{2}({\Omega})^{2} and 𝕃(Ω)=L(Ω)2×2\mathbb{L}^{\infty}(\Omega)=L^{\infty}({\Omega})^{2\times 2}. The notation fgf\lesssim g means that fCgf\leq Cg holds with a constant C>0C>0 independent of sensitive quantities, such as the segment size η\eta for interface-tracking, the spatial mesh size hh, the time-step size τ\tau, and the number of time steps nn. Moreover, we use the notations (,)Ω(\cdot,\cdot)_{\Omega} and ,Γ\left<\cdot,\cdot\right>_{\Gamma} to denote the inner product on L2(Ω)L^{2}(\Omega) and L2(Γ)L^{2}(\Gamma), respectively.

2 Flow maps

In this section, we introduce the forward boundary map for interface tracking and the backward flow map for time integration based on a given velocity field which has compact support in space. Throughout the theoretical analysis, we assume the velocity is smooth such that 𝒗𝑪r(2×[0,T]){\bm{v}}\in{\bm{C}}^{r}(\mathbb{R}^{2}\times[0,T]), rk+1r\geq k+1 with 2k42\leq k\leq 4 being the order of time integration.

2.1 Forward boundary map

Suppose the initial domain Ω02\Omega_{0}\subset\mathbb{R}^{2} is bounded and has a CrC^{r}-smooth boundary Γ0=Ω0\Gamma_{0}=\partial\Omega_{0}. The variation of the domain boundary Γt\Gamma_{t} has the form

Γt={𝑿F(t;0,𝒙0):𝒙0Γ0},\Gamma_{t}=\{{\bm{X}}_{F}(t;0,{\bm{x}}_{0}):\forall{\bm{x}}_{0}\in\Gamma_{0}\}, (1)

where 𝑿F{\bm{X}}_{F} is a forward boundary map, defined by

ddt𝑿F(t;s,𝒙s)=𝒗(𝑿F(t;s,𝒙s),t),𝑿F(s;s,𝒙s)=𝒙s.\frac{\mathrm{d}}{\mathrm{d}t}{\bm{X}}_{F}(t;s,{\bm{x}}_{s})={\bm{v}}({\bm{X}}_{F}(t;s,{\bm{x}}_{s}),t),\quad{\bm{X}}_{F}(s;s,{\bm{x}}_{s})={\bm{x}}_{s}. (2)

The moving domain Ωt\Omega_{t} is surrounded by the boundary Γt\Gamma_{t}, i.e., Γt=Ωt\Gamma_{t}=\partial\Omega_{t}. For theoretical analysis, we assume that 𝒗{\bm{v}} is 𝑪r{\bm{C}}^{r}-smooth, thus 𝑿F{\bm{X}}_{F} is a diffeomorphism and Ωt\Omega_{t} has same topological properties as Ω0\Omega_{0}. Meanwhile, we also assume that Γt\Gamma_{t} is CrC^{r} smooth for all t[0,T]t\in[0,T].

Consider a uniform partition of the interval [0,T][0,T], given by tn=nτt_{n}=n\tau, n=0,1,,Nn=0,1,\cdots,N, where τ=T/N\tau=T/N. Denote the forward boundary maps, the transient boundaries, and the transient domains at time tnt_{n}, respectively, by

𝑿Fn1,n:=𝑿F(tn;tn1,),Γn=Γtn,Ωn=Ωtn,n>0.{\bm{X}}_{F}^{n-1,n}:={\bm{X}}_{F}(t_{n};t_{n-1},\cdot),\quad\Gamma^{n}=\Gamma_{t_{n}},\quad\Omega^{n}=\Omega_{t_{n}},\quad n>0.

The well-posedness of (2) implies that 𝑿Fn1,n{\bm{X}}_{F}^{n-1,n}: Γn1Γn\Gamma^{n-1}\to\Gamma^{n} is one-to-one.

2.2 Backward flow map based on the ALE map

For fixed tnt_{n}, we choose Ωn\Omega^{n} as a reference domain and are going to define a backward flow map 𝒜n(,t)\mathcal{A}^{n}(\cdot,t) and the corresponding fluid velocity 𝒘n{\bm{w}}^{n} by means of ALE map: for ttnt\leq t_{n},

𝒜n(𝒙,t):Ω¯nΩ¯t,𝒜n(𝒙,tn):=𝒙,𝒘n(𝒙,t)=t𝒜n(𝒙,t).\mathcal{A}^{n}({\bm{x}},t):\bar{\Omega}^{n}\to\bar{\Omega}_{t},\quad\mathcal{A}^{n}({\bm{x}},t_{n}):={\bm{x}},\qquad{\bm{w}}^{n}({\bm{x}},t)=\partial_{t}\mathcal{A}^{n}({\bm{x}},t). (3)

In practice, we only need the ALE map at discrete time steps. First we construct a discrete boundary map either by 𝒈n,n1(𝒙)=𝑿F(tn1;tn,𝒙){\bm{g}}^{n,n-1}({\bm{x}})={\bm{X}}_{F}(t_{n-1};t_{n},{\bm{x}}) or by the closet point mapping 𝒈n,n1(𝒙)=argmin𝒚Γn1|𝒚𝒙|{\bm{g}}^{n,n-1}({\bm{x}})=\operatorname*{argmin}\limits_{{\bm{y}}\in\Gamma^{n-1}}|{\bm{y}}-{\bm{x}}| (see [12]), for all 𝒙Γn{\bm{x}}\in\Gamma^{n}. The backward flow map is defined by the solution to the harmonic equation

Δ𝑿n,n1=0inΩn,𝑿n,n1=𝒈n,n1onΓn.-\Delta{\bm{X}}^{n,n-1}=\textbf{0}\quad\text{in}\;\;\Omega^{n},\qquad{\bm{X}}^{n,n-1}={\bm{g}}^{n,n-1}\quad\text{on}\;\;\Gamma^{n}. (4)

The maximum principle implies that 𝑿n,n1{\bm{X}}^{n,n-1} maps Ωn¯\overline{\Omega^{n}} to Ωn1¯\overline{\Omega^{n-1}}. We then define a multi-step map from Ωn¯\overline{\Omega^{n}} to Ωni¯\overline{\Omega^{n-i}}, 2ik2\leq i\leq k, by

𝑿n,ni=𝑿ni+1,ni𝑿n,n1.{\bm{X}}^{n,n-i}={\bm{X}}^{n-i+1,n-i}\circ\cdots\circ{\bm{X}}^{n,n-1}. (5)

Let lniPk([tnk,tn])l_{n}^{i}\in P_{k}([t_{n-k},t_{n}]) be the basis functions of Lagrange interpolation satisfying lni(tnj)=δijl_{n}^{i}(t_{n-j})=\delta_{ij}, with δij\delta_{ij} the Kronecker delta. Define the semi-discrete ALE map by

𝒜kn(𝒙,t)=i=0klni(t)𝑿n,ni(𝒙)(𝒙,t)Ωn¯×[tnk,tn].\mathcal{A}^{n}_{k}({\bm{x}},t)=\sum\limits_{i=0}^{k}l_{n}^{i}(t){\bm{X}}^{n,n-i}({\bm{x}})\quad\forall\,({\bm{x}},t)\in\overline{\Omega^{n}}\times[t_{n-k},t_{n}]. (6)

Clearly 𝒜n(,tni)=𝒜kn(tni)=𝑿n,ni\mathcal{A}^{n}(\bm{\cdot},t_{n-i})=\mathcal{A}^{n}_{k}(\cdot\,t_{n-i})={\bm{X}}^{n,n-i}. The artificial fluid velocity is defined by

𝒘kn(𝒙,t):=t𝒜kn(t,𝒙)=i=0k(lni)(t)𝑿n,ni(𝒙).{\bm{w}}^{n}_{k}({\bm{x}},t):=\partial_{t}\mathcal{A}^{n}_{k}(t,{\bm{x}})=\sum_{i=0}^{k}(l_{n}^{i})^{\prime}(t){\bm{X}}^{n,n-i}({\bm{x}}). (7)
Remark 2.3.

The construction of the ALE map is not unique. For example, one can represent the motion of a domain by considering the domain as elastic or viscoelastic solid, and solve the problem by resorting to the equations of elastic dynamics.

3 The ALE-UFE method

The purpose of this section is to propose a high-order finite element method for solving PDEs on time-moving domains based on ALE map. For clearness, we first focus on the heat equation and will extend the result to more complex problems in sections 46.

3.1 The heat equation on a moving domain

Based on the forward boundary map and backward flow map presented in the previous section, we now design the numerical scheme for solving the heat equation on a time-evolving domain

utΔu=finΩt,u=0onΓt,u(,0)=u0inΩ0,\displaystyle\frac{\partial u}{\partial t}-\Delta u=f\quad\text{in}\;\;\Omega_{t},\qquad u=0\quad\text{on}\;\;\Gamma_{t},\qquad u(\cdot,0)=u_{0}\quad\text{in}\;\;\Omega_{0}, (8)

where Ωt2\Omega_{t}\subset\mathbb{R}^{2} is the time-varying domain, Γt=Ωt\Gamma_{t}=\partial\Omega_{t} the moving boundary defined in (1), u(𝒙,t)u({\bm{x}},t) the tracer transported by the fluid, u0u_{0} the initial value, and f(𝒙,t)f({\bm{x}},t) the source term distributed in 2\mathbb{R}^{2} and having a compact support.

By the chain rule and (3), the first equation of (8) can be written as

ddtu(𝒜n(𝒙,t),t)|t=tn𝒘n(𝒙,tn)u(𝒙,tn)Δu(𝒙,tn)=f(tn).\frac{\mathrm{d}}{\mathrm{d}t}u(\mathcal{A}^{n}({\bm{x}},t),t)\Big{|}_{t=t_{n}}-{\bm{w}}^{n}({\bm{x}},t_{n})\cdot\nabla u({\bm{x}},t_{n})-\Delta u({\bm{x}},t_{n})=f(t_{n}). (9)

From (9), 𝒜n(𝒙,t)\mathcal{A}^{n}({\bm{x}},t) is not necessarily differentiable in 𝒙{\bm{x}}. This inspires us to construct a discrete ALE map in practice. The semi-discrete scheme is given by

1τΛ𝒘k𝑼¯nΔun=fn,\frac{1}{\tau}\Lambda^{k}_{{\bm{w}}}\underline{{\bm{U}}}^{n}-\Delta u^{n}=f^{n}, (10)

where unu^{n}, fnf^{n} denote u(𝒙,tn)u({\bm{x}},t_{n}) and f(𝒙,tn)f({\bm{x}},t_{n}), respectively, and τ1Λ𝒘k\tau^{-1}\Lambda_{{\bm{w}}}^{k} stands for the kthk^{\text{th}}-order time finite difference operator, given by

1τΛ𝒘k𝑼¯n=1τΛk𝑼¯n𝒘kn(𝒙,tn)UnwithΛk𝑼¯n=i=0kλikUni,n.\frac{1}{\tau}\Lambda_{{\bm{w}}}^{k}\underline{{\bm{U}}}^{n}=\frac{1}{\tau}\Lambda^{k}\underline{{\bm{U}}}^{n}-{\bm{w}}_{k}^{n}({\bm{x}},t_{n})\cdot\nabla U^{n}\quad\hbox{with}\;\;\Lambda^{k}\underline{{\bm{U}}}^{n}=\sum_{i=0}^{k}\lambda_{i}^{k}U^{n-i,n}. (11)

Here τ1Λk\tau^{-1}\Lambda^{k} denotes the BDF-kk finite difference operator defined by (cf. [18]) and

𝑼¯=[Unk,,Un],Uni=u(𝒜kn(𝒙,tni),tni),0ik.\underline{{\bm{U}}}=[U^{n-k},\cdots,U^{n}],\quad U^{n-i}=u(\mathcal{A}^{n}_{k}({\bm{x}},t_{n-i}),t_{n-i}),\quad 0\leq i\leq k.

The coefficients λik\lambda_{i}^{k} for k=2,3,4k=2,3,4 are listed in Table 1.

Table 1: Coefficients of λik\lambda_{i}^{k} in BDF{\rm BDF} schemes.
           kk ii λik\lambda_{i}^{k}           0           11           22           33           44
          22           3/23/2           2-2           1/21/2           0           0
          33           11/611/6           3-3           3/23/2           1/3-1/3           0
          44           25/1225/12           4-4           33           4/3-4/3           1/41/4

3.2 Interface-tracking approximation

Even the velocity of Γt\Gamma_{t} is known explicitly, we still need to approximate it with an approximate boundary due to computational complexity in practice. The approximate boundary Γηn\Gamma^{n}_{\eta} can be constructed by either explicit algorithms such as front-tracking methods [17] and cubic MARS (Mapping and Adjusting Regular Semi-analytic sets) methods [28], or implicit algorithms such as level set methods [2]. The domain enclosed by Γηn\Gamma^{n}_{\eta} is denoted by Ωηn\Omega_{\eta}^{n} with η\eta being the parameter of approximation.

Let 𝝌n(l)\bm{\chi}_{n}(l) and 𝝌^n(l)\hat{\bm{\chi}}_{n}(l), l[0,L]l\in[0,L], denote the parametric representations of Γηn\Gamma_{\eta}^{n} and Γn\Gamma^{n}, respectively. In order to get optimal error estimates, we make an assumption that the approximation of the boundary is of high order in the sense that

|𝝌n𝝌^n|Cμ([0,L])τk+1μ,μ=0,1.\left|{\bm{\chi}_{n}-\hat{\bm{\chi}}_{n}}\right|_{C^{\mu}([0,L])}\lesssim\tau^{k+1-\mu},\quad\mu=0,1. (12)

For a rigorous proof of the error estimate, we refer to [21, section 3.2] where the cubic MARS method is used for interface tracking [28]. Using (12), we have the error estimate between the exact domain and the approximate domain

area(Ωn\Ωηn)+area(Ωηn\Ωn)=O(τk+1).\text{area}(\Omega^{n}\backslash\Omega_{\eta}^{n})+\text{area}(\Omega_{\eta}^{n}\backslash\Omega^{n})=O(\tau^{k+1}).
Remark 3.3.

Assumption (12) is used only for analysis. In practice, we require the Hausdorff distance between Γηn\Gamma_{\eta}^{n} and Γn\Gamma^{n} to be small, i.e., dist(Γηn,Γn)τk+1\text{dist}(\Gamma_{\eta}^{n},\Gamma^{n})\lesssim\tau^{k+1}.

3.4 Finite element spaces

First we define a δ\delta-neighborhood of Ωηn\Omega_{\eta}^{n} by

Ωδn=:{𝒙2:min𝒚Ωηn|𝒙𝒚|0.5τ}.\Omega_{\delta}^{n}=:\big{\{}{\bm{x}}\in\mathbb{R}^{2}:\min_{{\bm{y}}\in\Omega_{\eta}^{n}}|{\bm{x}}-{\bm{y}}|\leq 0.5\tau\big{\}}. (13)

It is easy to see ΩηnΩδn\Omega_{\eta}^{n}\subset\Omega_{\delta}^{n} (see Fig. 1). Let DD be an open square satisfying ΩtΩδnD\Omega_{t}\cup\Omega^{n}_{\delta}\subset D for all 0tT0\leq t\leq T and 0nN0\leq n\leq N. Let 𝒯h\mathcal{T}_{h} be the uniform partition of D¯\bar{D} into closed squares of side-length hh. It generates two covers of Ωδn\Omega^{n}_{\delta} and ΓηnΩδn\Gamma^{n}_{\eta}\cup\partial\Omega_{\delta}^{n}, respectively,

𝒯hn\displaystyle\mathcal{T}^{n}_{h} :={K𝒯h:K¯Ωδn},\displaystyle:=\left\{K\in\mathcal{T}_{h}:\;\overline{K}\cap\Omega^{n}_{\delta}\neq\emptyset\right\}, (14)
𝒯h,Bn\displaystyle\mathcal{T}^{n}_{h,B} :={K𝒯hn:K¯ΓηnorK¯Ωδn}.\displaystyle:=\left\{K\in\mathcal{T}^{n}_{h}:\;\overline{K}\cap\Gamma^{n}_{\eta}\neq\emptyset\;\;\text{or}\;\;\overline{K}\cap\partial\Omega_{\delta}^{n}\neq\emptyset\right\}. (15)

The cover 𝒯hn\mathcal{T}^{n}_{h} generates a fictitious domain Ω~n:=interior(K𝒯hnK)\tilde{\Omega}^{n}:=\mathrm{interior}\big{(}\cup_{K\in\mathcal{T}^{n}_{h}}K\big{)}. Let h\mathcal{E}_{h} be the set of all edges in 𝒯h\mathcal{T}_{h}. The set of interior edges of boundary elements are denoted by

h,Bn={Eh:K𝒯h,Bns.t.EK\Ω~n}.\mathcal{E}_{h,B}^{n}=\big{\{}E\in\mathcal{E}_{h}:\;\exists K\in\mathcal{T}^{n}_{h,B}\;\;\hbox{s.t.}\;\;E\subset\partial K\backslash\partial\tilde{\Omega}^{n}\big{\}}. (16)
DDΩηn\Omega_{\eta}^{n}Ωδn\Omega_{\delta}^{n}
Ω~n\tilde{\Omega}^{n}𝒯h,Bn\mathcal{T}_{h,B}^{n}
Figure 1: Left: the domain DD and its partition 𝒯h\mathcal{T}_{h}, the approximate domain Ωηn\Omega_{\eta}^{n} and its δ\delta-neighborhood Ωδn\Omega_{\delta}^{n}. Right: the set of red and yellow squares 𝒯hn\mathcal{T}^{n}_{h}, the set of red squares 𝒯h,Bn\mathcal{T}^{n}_{h,B}, the set of blue edges h,Bn\mathcal{E}_{h,B}^{n}, the union of red and yellow squares Ω~n\tilde{\Omega}^{n}.

Next we define the finite element spaces on DD and Ω~n\tilde{\Omega}^{n}, respectively, as follows

V(k,𝒯h):=\displaystyle V(k,\mathcal{T}_{h}):=\, {vH1(D):v|KQk(K),K𝒯h},\displaystyle\big{\{}v\in H^{1}({D}):v|_{K}\in Q_{k}(K),\;\forall\,K\in\mathcal{T}_{h}\big{\}},
V(k,𝒯hn):=\displaystyle V(k,\mathcal{T}^{n}_{h}):=\, {v|Ω~n:vV(k,𝒯h)},\displaystyle\big{\{}v|_{\tilde{\Omega}^{n}}:v\in V(k,\mathcal{T}_{h})\big{\}},

where QkQ_{k} is the space of polynomials whose degrees are no more than kk for each variable. The space of piecewise regular functions over 𝒯hn\mathcal{T}^{n}_{h} is defined by

Hm(𝒯hn):={vL2(Ω~n):v|KHm(K),K𝒯hn},m1.H^{m}(\mathcal{T}^{n}_{h}):=\big{\{}v\in L^{2}({\tilde{\Omega}^{n}}):\;v|_{K}\in H^{m}(K),\;\forall\,K\in\mathcal{T}^{n}_{h}\big{\}},\qquad m\geq 1. (17)

3.5 Construction of an approximate ALE map

Let 𝑿τn,n1{\bm{X}}_{\tau}^{n,n-1} denote the approximation of the inverse of 𝑿F(tn,tn1,𝒙){\bm{X}}_{F}(t_{n},t_{n-1},{\bm{x}}). This is computed by solving (2) with the RK-(k+1)(k+1) scheme (the (k+1)th(k+1)^{\text{th}}-order Runge-Kutta scheme) from tnt_{n} to tn1t_{n-1}. For any 𝒙nΓηn{\bm{x}}^{n}\in\Gamma_{\eta}^{n}, the point 𝒙n1=𝑿τn,n1(𝒙n){\bm{x}}^{n-1}={\bm{X}}_{\tau}^{n,n-1}({\bm{x}}^{n}) is calculated as follows:

{𝒙(1)=𝒙n,𝒙(i)=𝒙nτj=1i1aijk+1𝒗(𝒙(j),tncjk+1τ),2ink+1,𝒙n1=𝒙nτi=1nk+1dik+1𝒗(𝒙(i),t(i)).\begin{cases}{\bm{x}}^{(1)}={\bm{x}}^{n},\\ {\bm{x}}^{(i)}={\bm{x}}^{n}-\tau\sum\limits_{j=1}^{i-1}a^{k+1}_{ij}{\bm{v}}({\bm{x}}^{(j)},t_{n}-c_{j}^{k+1}\tau),\quad 2\leq i\leq n_{k+1},\vspace{0.5mm}\\ {\bm{x}}^{n-1}={\bm{x}}^{n}-\tau\sum\limits_{i=1}^{n_{k+1}}d_{i}^{k+1}{\bm{v}}({\bm{x}}^{(i)},t^{(i)}).\end{cases} (18)

Here aijk+1,dik+1,cik+1a_{ij}^{k+1},\;d_{i}^{k+1},\;c_{i}^{k+1} are the coefficients of the RK-(k+1k+1) scheme, satisfying aijk+1=0a_{ij}^{k+1}=0 if jij\geq i, and nk+1n_{k+1} is the number of stages. Recall from section 2.2 that 𝒈n,n1{\bm{g}}^{n,n-1} maps Γn\Gamma^{n} to Γn1\Gamma^{n-1}. We define its approximation by 𝒈ηn,n1=𝑿τn,n1|Γηn{\bm{g}}_{\eta}^{n,n-1}={\bm{X}}_{\tau}^{n,n-1}|_{\Gamma_{\eta}^{n}}.

A discrete approximation of 𝑿n,n1{\bm{X}}^{n,n-1} is defined by solving the discrete problem: find 𝑿hn,n1𝑽(k,𝒯hn):=(V(k,𝒯hn))2{\bm{X}}_{h}^{n,n-1}\in{\bm{V}}(k,\mathcal{T}_{h}^{n}):=(V(k,\mathcal{T}_{h}^{n}))^{2} such that

𝒜hn(𝑿hn,n1,𝒗h)=Γηn(𝒈ηn,n1,𝒗h)𝒗h𝑽(k,𝒯hn).\mathscr{A}_{h}^{n}({\bm{X}}_{h}^{n,n-1},{\bm{v}}_{h})=\mathcal{F}_{\Gamma_{\eta}^{n}}({\bm{g}}_{\eta}^{n,n-1},{\bm{v}}_{h})\quad\;\forall\,{\bm{v}}_{h}\in{\bm{V}}(k,\mathcal{T}_{h}^{n}). (19)

where Γηn(𝒘,𝒗h)=𝒘,γ0h1𝒗h𝒏𝒗hΓηn\mathcal{F}_{\Gamma_{\eta}^{n}}({\bm{w}},{\bm{v}}_{h})=\left<{\bm{w}},\gamma_{0}h^{-1}{\bm{v}}_{h}-\partial_{{\bm{n}}}{\bm{v}}_{h}\right>_{\Gamma_{\eta}^{n}} and the bilinear forms are defined by

𝓐hn(𝒖,𝒗):=\displaystyle\bm{\mathscr{A}}_{h}^{n}({\bm{u}},{\bm{v}}):=\, 𝒜hn(u1,v1)+𝒜hn(u2,v2),\displaystyle\mathscr{A}_{h}^{n}(u_{1},v_{1})+\mathscr{A}_{h}^{n}(u_{2},v_{2}), (20)
𝒜hn(u,v):=\displaystyle\mathscr{A}^{n}_{h}(u,v):=\, (u,v)Ωηn+𝒮hn(u,v)+𝒥0n(u,v)+𝒥1n(u,v),\displaystyle(\nabla u,\nabla v)_{\Omega^{n}_{\eta}}+\mathscr{S}^{n}_{h}(u,v)+\mathscr{J}_{0}^{n}(u,v)+\mathscr{J}_{1}^{n}(u,v), (21)
𝒮hn(u,v):=\displaystyle\mathscr{S}^{n}_{h}(u,v):=\, v,𝒏uΓηnu,𝒏vΓηn,\displaystyle-\left<v,\partial_{\bm{n}}u\right>_{\Gamma_{\eta}^{n}}-\left<u,\partial_{\bm{n}}v\right>_{\Gamma_{\eta}^{n}}, (22)
𝒥0n(u,v):=\displaystyle\mathscr{J}^{n}_{0}(u,v):=\, γ0hu,vΓηn,𝒥1n(u,v):=Eh,Bnl=1kh2l1𝒏lu,𝒏lvE.\displaystyle\frac{\gamma_{0}}{h}\left<u,v\right>_{\Gamma_{\eta}^{n}},\quad\mathscr{J}^{n}_{1}(u,v):=\sum_{E\in\mathcal{E}_{h,B}^{n}}\sum_{l=1}^{k}h^{2l-1}\left<\left\llbracket{\partial_{{\bm{n}}}^{l}u}\right\rrbracket,\left\llbracket{\partial_{{\bm{n}}}^{l}v}\right\rrbracket\right>_{E}. (23)

In (22), 𝒏u=u𝒏\partial_{{\bm{n}}}u=\frac{\partial u}{\partial{\bm{n}}} denotes the normal derivative of uu on Γηn\Gamma_{\eta}^{n}. In (23), 𝒥0n\mathscr{J}^{n}_{0} is used to impose the Dirichlet boundary condition weakly with γ0>0\gamma_{0}>0, 𝒏lv\partial_{{\bm{n}}}^{l}v denotes the ll-th order normal derivative of vv, and v|E=v|K1v|K2\left\llbracket{v}\right\rrbracket|_{E}=v|_{K_{1}}-v|_{K_{2}} denotes the jump of vv across edge EE where K1,K2𝒯hK_{1},K_{2}\in\mathcal{T}_{h} are the two elements sharing EE.

Similar to (6), we define the fully discrete ALE map and artificial velocity by

𝒜k,hn(𝒙,t)=i=0klni(t)𝑿hn,ni,𝒘k,hn(𝒙,t)=i=0k(lni)(t)𝑿hn,ni,\mathcal{A}^{n}_{k,h}({\bm{x}},t)=\sum_{i=0}^{k}l_{n}^{i}(t){\bm{X}}_{h}^{n,n-i},\quad{\bm{w}}_{k,h}^{n}({\bm{x}},t)=\sum_{i=0}^{k}(l_{n}^{i})^{\prime}(t){\bm{X}}_{h}^{n,n-i}, (24)

where 𝑿hn,n=𝑰{\bm{X}}_{h}^{n,n}={\bm{I}} and the multi-step backward flow map is defined by

𝑿hn,ni:=𝑿hn1,ni𝑿hn,n1=𝑿hni+1,ni𝑿hn,n1.\displaystyle{\bm{X}}_{h}^{n,n-i}:={\bm{X}}_{h}^{n-1,n-i}\circ{\bm{X}}_{h}^{n,n-1}={\bm{X}}_{h}^{n-i+1,n-i}\circ\cdots\circ{\bm{X}}_{h}^{n,n-1}. (25)

The recursive definition only needs to compute the one-step map 𝑿hn,n1{\bm{X}}_{h}^{n,n-1} at each tnt_{n}.

3.6 The fully discrete scheme

Given the finite element function 𝒘k,hn{\bm{w}}_{k,h}^{n}, we define the kthk^{\text{th}}-order time finite difference operator as

1τΛ𝒘hk𝑼hn¯=1τΛk𝑼hn¯𝒘k,hnUhn,n,\frac{1}{\tau}\Lambda^{k}_{{\bm{w}}_{h}}\underline{{\bm{U}}^{n}_{h}}=\frac{1}{\tau}\Lambda^{k}\underline{{\bm{U}}^{n}_{h}}-{\bm{w}}_{k,h}^{n}\cdot\nabla U^{n,n}_{h}, (26)

where 𝑼hn¯=[Uhnk,n,,Uhn,n]\underline{{\bm{U}}^{n}_{h}}=\big{[}U^{n-k,n}_{h},\cdots,U^{n,n}_{h}\big{]}^{\top} and Uhni,n=uhni𝑿hn,niU_{h}^{n-i,n}=u_{h}^{n-i}\circ{\bm{X}}_{h}^{n,n-i}. Note that Uhn,n=uhnU_{h}^{n,n}=u_{h}^{n}. The discrete approximation to problem (8) is to seek uhnV(k,𝒯hn)u^{n}_{h}\in V(k,\mathcal{T}_{h}^{n}) such that

1τ(Λ𝒘hk𝑼hn¯,vh)Ωηn+𝒜hn(uhn,vh)=(fn,vh)ΩηnvhV(k,𝒯hn).\displaystyle\frac{1}{\tau}\big{(}\Lambda^{k}_{{\bm{w}}_{h}}\underline{{\bm{U}}^{n}_{h}},v_{h}\big{)}_{\Omega^{n}_{\eta}}+\mathscr{A}^{n}_{h}(u_{h}^{n},v_{h})=(f^{n},v_{h})_{\Omega^{n}_{\eta}}\quad\forall\,v_{h}\in V(k,\mathcal{T}_{h}^{n}). (27)

In view of (27), the stiffness matrix corresponding to 𝒜hn\mathscr{A}_{h}^{n} has already been obtained when computing the discrete ALE map in (19). The δ\delta-neighborhood Ωδn\Omega_{\delta}^{n} is chosen to ensure that uhni𝑿hn,niu_{h}^{n-i}\circ{\bm{X}}_{h}^{n,n-i} is well-defined for 0ik0\leq i\leq k.

3.7 Well-posedness and error estimates

In this section, we show the well-posedness, stability, and error estimates of the discrete problems in the appendix. Firstly, we define the mesh-dependent norms

|v|Ωηn=(|v|H1(Ωηn)2+h1vL2(Γηn)2+h𝒏vL2(Γηn)2)1/2,\displaystyle\left\|\left|{v}\right|\right\|_{\Omega^{n}_{\eta}}=\big{(}{\left|{v}\right|}_{H^{1}({\Omega^{n}_{\eta}})}^{2}+h^{-1}\left\|{v}\right\|_{L^{2}({\Gamma^{n}_{\eta}})}^{2}+h\left\|{\partial_{{\bm{n}}}v}\right\|_{L^{2}({\Gamma^{n}_{\eta}})}^{2}\big{)}^{1/2},
|v|𝒯hn=(|v|H1(Ωηn)2+𝒥0n(v,v)+𝒥1n(v,v))1/2.\displaystyle\left\|\left|{v}\right|\right\|_{\mathcal{T}^{n}_{h}}=\big{(}{\left|{v}\right|}_{H^{1}({\Omega^{n}_{\eta}})}^{2}+\mathscr{J}^{n}_{0}(v,v)+\mathscr{J}^{n}_{1}(v,v)\big{)}^{1/2}.

Clearly ||𝒯hn\left\|\left|{\cdot}\right|\right\|_{\mathcal{T}^{n}_{h}} is a norm on H1(Ω~n)Hk+1(𝒯hn)H^{1}(\tilde{\Omega}^{n})\cap H^{k+1}(\mathcal{T}^{n}_{h}). From [11], we have the following norm inequalities: for any vhV(k,𝒯hn)v_{h}\in V(k,\mathcal{T}_{h}^{n}),

vhL2(Ω~n)2vhL2(Ωηn)2+h2𝒥1n(vh,vh),|vh|Ωηn|vh|𝒯hn.\displaystyle\left\|{v_{h}}\right\|_{L^{2}({\tilde{\Omega}^{n}})}^{2}\lesssim\left\|{v_{h}}\right\|_{L^{2}({\Omega^{n}_{\eta}})}^{2}+h^{2}\mathscr{J}_{1}^{n}(v_{h},v_{h}),\qquad\left\|\left|{v_{h}}\right|\right\|_{\Omega^{n}_{\eta}}\lesssim\left\|\left|{v_{h}}\right|\right\|_{\mathcal{T}^{n}_{h}}. (28)

Suppose γ0\gamma_{0} is large enough. It is standard to prove the coercivity and continuity of the bilinear form (see [21]): for any vhV(k,𝒯hn)v_{h}\in V(k,\mathcal{T}^{n}_{h}) and uHk+1(𝒯hn)H1(Ωηn)u\in H^{k+1}(\mathcal{T}_{h}^{n})\cap H^{1}(\Omega_{\eta}^{n}),

𝒜hn(vh,vh)Ca|vh|𝒯hn2,|𝒜hn(u,vh)||u|Ωηn|vh|𝒯hn.\displaystyle\mathscr{A}_{h}^{n}(v_{h},v_{h})\geq C_{a}\left\|\left|{v_{h}}\right|\right\|_{\mathcal{T}^{n}_{h}}^{2},\qquad\left|{\mathscr{A}_{h}^{n}(u,v_{h})}\right|\lesssim\left\|\left|{u}\right|\right\|_{\Omega_{\eta}^{n}}\left\|\left|{v_{h}}\right|\right\|_{\mathcal{T}^{n}_{h}}.

By Corollary B.2, the artificial velocity 𝒘k,hn{\bm{w}}_{k,h}^{n} is bounded. This gives the theorem.

Theorem 3.8.

There is a positive constant τ0Ca𝐰k,hn𝐋(Ωηn)2\tau_{0}\lesssim C_{a}\|{\bm{w}}_{k,h}^{n}\|_{{\bm{L}}^{\infty}(\Omega_{\eta}^{n})}^{-2} small enough such that the problem (27) has a unique solution for any τ(0,τ0]\tau\in(0,\tau_{0}].

Theorem 3.9.

Assume the approximate boundary satisfies (12) and the penalty parameter γ0\gamma_{0} in 𝒜hn\mathscr{A}^{n}_{h} is large enough. Let uhnu_{h}^{n} be the solution to the discrete problem (27) based on the pre-calculated initial values {uh0,,uhk1}\{u_{h}^{0},\cdots,u^{k-1}_{h}\}. There is an h0>0h_{0}>0 small enough such that, for any h=O(τ)(0,h0]h=O(\tau)\in(0,h_{0}] and mkm\geq k,

uhmL2(Ωηn)2+n=kmτ|uhn|𝒯hn2n=kmτfnL2(Ωηn)2+i=0k1[uhiL2(Ωηi)2+τuhiH1(Ω~i)2].\displaystyle\|{u^{m}_{h}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+\sum_{n=k}^{m}\tau\left\|\left|{u_{h}^{n}}\right|\right\|^{2}_{\mathcal{T}_{h}^{n}}\lesssim\,\sum_{n=k}^{m}\tau\|{f^{n}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+\sum_{i=0}^{k-1}\left[\|{u^{i}_{h}}\|^{2}_{L^{2}(\Omega^{i}_{\eta})}+\tau\|u^{i}_{h}\|_{H^{1}(\tilde{\Omega}^{i})}^{2}\right].

The proof of Theorem 3.9 is presented in C.

For vWμ,(0,T;Hs(Ωt))v\in W^{\mu,\infty}(0,T;H^{s}(\Omega_{t})), one has v(t)Hs(Ωt)Wμ,([0,T])\|v(t)\|_{H^{s}(\Omega_{t})}\in W^{\mu,\infty}([0,T]) for μ=0,1\mu=0,1 and s=1,,k+1s=1,\cdots,k+1. Suppose fL(0,T;H1(D))f\in L^{\infty}(0,T;H^{1}(D)). The exact solution satisfies

uHk+1(QT)L(0,T;Hk+1(Ωt))W1,(0,T;H1(Ωt)),u\in H^{k+1}(Q_{T})\cap L^{\infty}(0,T;H^{k+1}(\Omega_{t}))\cap W^{1,\infty}(0,T;H^{1}(\Omega_{t})), (29)

where QT={(𝒙,t):𝒙Ωt,t[0,T]}Q_{T}=\{({\bm{x}},t):{\bm{x}}\in\Omega_{t},t\in[0,T]\}. Assume 𝒜n\mathcal{A}^{n} is smooth in time such that

tk+1𝒜n(𝒙,t)L2(Ωt)M0t[tnk,tn],\|\partial_{t}^{k+1}\mathcal{A}^{n}({\bm{x}},t)\|_{L^{2}(\Omega_{t})}\leq M_{0}\quad\forall\,t\in[t_{n-k},t_{n}], (30)

where kk is the order of the BDF scheme and M0M_{0} is independent of τ\tau, η\eta, hh and nn. Since 𝒜kn\mathcal{A}^{n}_{k} is the kthk^{\text{th}}-order Lagrange interpolation of 𝒜n\mathcal{A}^{n} in [tnk,tn][t_{n-k},t_{n}], we have

i=0kti𝒜k,hnL2(Ωηn)i=0kti𝒜knL2(Ωt)tk+1𝒜nL2(Ωt)t[tnk,tn].\sum_{i=0}^{k}\|\partial_{t}^{i}\mathcal{A}^{n}_{k,h}\|_{L^{2}(\Omega_{\eta}^{n})}\lesssim\sum_{i=0}^{k}\|\partial_{t}^{i}\mathcal{A}^{n}_{k}\|_{L^{2}(\Omega_{t})}\lesssim\|\partial_{t}^{k+1}\mathcal{A}^{n}\|_{L^{2}(\Omega_{t})}\quad\forall\,t\in[t_{n-k},t_{n}]. (31)
Theorem 3.10.

Suppose the assumptions in Theorem 3.9 hold and the initial solutions satisfy uiuhiL2(Ωηi)2+τ|uiuhi|Ωηi2τ2k\|{u^{i}-u_{h}^{i}}\|^{2}_{L^{2}(\Omega_{\eta}^{i})}+\tau\||{u^{i}-u_{h}^{i}}|\|_{\Omega_{\eta}^{i}}^{2}\lesssim\tau^{2k} for ik1\leq i\leq k-1. Then

umuhmL2(Ωηm)2+n=kmτ|unuhn|𝒯hn2τ2k,kmN.\|{u^{m}-u_{h}^{m}}\|^{2}_{L^{2}(\Omega^{m}_{\eta})}+\sum_{n=k}^{m}\tau\||{u^{n}-u_{h}^{n}}|\|_{\mathcal{T}_{h}^{n}}^{2}\lesssim\tau^{2k},\qquad k\leq m\leq N. (32)

By [21], there is an extension u~\tilde{u} of uu such that, for 1mk+11\leq m\leq k+1,

{u~(t)Hm(D)u(t)Hm(Ωt),u~(t)Hk+1(D×[0,T])u(t)Hk+1(QT),tu~(t)H1(D)u(t)H2(Ωt)+tu(t)H1(Ωt).\begin{cases}\|\tilde{u}(t)\|_{H^{m}(D)}\lesssim\|u(t)\|_{H^{m}(\Omega_{t})},\vspace{0.5mm}\\ \|\tilde{u}(t)\|_{H^{k+1}(D\times[0,T])}\lesssim\|u(t)\|_{H^{k+1}(Q_{T})},\vspace{0.5mm}\\ \|\partial_{t}\tilde{u}(t)\|_{H^{1}(D)}\lesssim\|u(t)\|_{H^{2}(\Omega_{t})}+\|\partial_{t}u(t)\|_{H^{1}(\Omega_{t})}.\end{cases} (33)

Then (31) and (33) show dk+1dtk+1u~(𝒜hn(x,t),tn)L2(Ωηn)M0uHk+1(Ωt×[tnk,tn])\big{\|}\frac{\mathrm{d}^{k+1}}{\mathrm{d}t^{k+1}}\tilde{u}(\mathcal{A}_{h}^{n}(x,t),t_{n})\big{\|}_{L^{2}(\Omega_{\eta}^{n})}\lesssim M_{0}\|u\|_{H^{k+1}(\Omega_{t}\times[t_{n-k},t_{n}])}. It is easy to see that the extension u~\tilde{u} satisfies

1τ(Λwhk𝑼¯~,vh)Ωηn+𝒜hn(u~,vh)=u~n,(γ0/hn)vhΓηn+(f~n+Rn,vh)Ωηn,\displaystyle\frac{1}{\tau}(\Lambda^{k}_{w_{h}}\tilde{\underline{{\bm{U}}}},v_{h})_{\Omega_{\eta}^{n}}+\mathscr{A}_{h}^{n}(\tilde{u},v_{h})=\left<\tilde{u}^{n},\,(\gamma_{0}/h-\partial_{n})v_{h}\right>_{\Gamma_{\eta}^{n}}+(\tilde{f}^{n}+R^{n},v_{h})_{\Omega_{\eta}^{n}}, (34)

where 𝑼¯~=[u~nk𝒜k,hn(,tnk),,u~n]\tilde{\underline{{\bm{U}}}}=[\tilde{u}^{n-k}\circ\mathcal{A}_{k,h}^{n}(\cdot,t_{n-k}),\cdots,\tilde{u}^{n}], f~n=tu~(tn)Δu~n\tilde{f}^{n}=\partial_{t}\tilde{u}(t_{n})-\Delta\tilde{u}^{n}, and

Rn=τ1Λk𝑼¯~ddtu~𝒜k,hn(,t)|t=tn.\displaystyle R^{n}=\tau^{-1}\Lambda^{k}\tilde{\underline{{\bm{U}}}}-\frac{\mathrm{d}}{\mathrm{d}t}\tilde{u}\circ\mathcal{A}_{k,h}^{n}(\cdot,t)|_{t=t_{n}}.

The proof of Theorem 3.10 is parallel to [21, Section 6] by subtracting (27) from (34). We omit the details here.

3.11 The ALE-UFE framework

Now we conclude the ALE-UFE framework for solving PDEs on varying domains. It consists of four steps.

1. Track the varying interface by the forward boundary map (2). 2. Construct a one-step backward flow map with (19), a discrete ALE map and an artificial velocity with (24), and a multi-step map 𝑿hn,ni{\bm{X}}^{n,n-i}_{h} with (25). 3. Define the finite difference operator in (26) by combining the BDF-kk scheme and the backward flow map. 4. Construct the fully discrete scheme as in (27).

The framework can be conveniently applied to various moving-domain problems. The operations in each step are adapted to a specific problem.

4 A domain-PDE-coupled problem

In this section we apply the ALE-UFE framework to a nonlinear problem where the moving domain depends on the solution. Unless otherwise specified, the finite element spaces and bilinear forms are defined in the same way as in the previous section. We consider the following model:

𝒗tΔ𝒗=𝒇inΩt,𝒏𝒗=𝒈NonΓt,𝒗(𝒙,t0)=𝒗0,inΩ0,\frac{\partial{\bm{v}}}{\partial t}-\Delta{\bm{v}}={\bm{f}}\quad\text{in}\;\;\Omega_{t},\quad\partial_{{\bm{n}}}{\bm{v}}={\bm{g}}_{N}\quad\text{on}\;\Gamma_{t},\qquad{\bm{v}}({\bm{x}},t_{0})={\bm{v}}_{0},\;\;\text{in}\;\Omega_{0}, (35)

where Ωt\Omega_{t} is the domain surrounded by a varying boundary Γt\Gamma_{t}, i.e. Γt=Ωt\Gamma_{t}=\partial\Omega_{t}. The Neumann boundary condition 𝒈N{\bm{g}}_{N} can be viewed as an applied force on Γt\Gamma_{t}, such as a surface tension of fluids [20]. For simplicity, we treat it as a given function.

4.1 Flow maps

The variation of Γt\Gamma_{t} has the form of (1), namely Γt=𝑿F(t;0,Γ0)\Gamma_{t}={\bm{X}}_{F}(t;0,\Gamma_{0}), and the forward boundary map is defined by

ddt𝑿F(t;s,𝒙s)=𝒗(𝑿F(t;s,𝒙s),t),𝑿F(s;s,𝒙s)=𝒙s.\frac{\mathrm{d}}{\mathrm{d}t}{\bm{X}}_{F}(t;s,{\bm{x}}_{s})={\bm{v}}({\bm{X}}_{F}(t;s,{\bm{x}}_{s}),t),\quad{\bm{X}}_{F}(s;s,{\bm{x}}_{s})={\bm{x}}_{s}. (36)

Note that 𝒗{\bm{v}} is the solution of the problem. Based on an ALE map 𝒜n(𝒙,t):ΩnΩt\mathcal{A}^{n}({\bm{x}},t):\Omega^{n}\to\Omega_{t}, (ttn)(t\leq t_{n}), the equation in (35) can be written as follows

d𝒗(𝒜n(𝒙,t),t)dt|t=tn=𝒘n(𝒙,tn)𝒗(𝒙,tn)+Δ𝒗(𝒙,tn)+𝒇(𝒙,tn),\frac{\mathrm{d}\,{\bm{v}}(\mathcal{A}^{n}({\bm{x}},t),t)}{\mathrm{d}t}\Big{|}_{t=t_{n}}={\bm{w}}^{n}({\bm{x}},t_{n})\nabla{\bm{v}}({\bm{x}},t_{n})+\Delta{\bm{v}}({\bm{x}},t_{n})+{\bm{f}}({\bm{x}},t_{n}), (37)

where 𝒘n(𝒙,tn)=t𝒜n(𝒙,t)|t=tn{\bm{w}}^{n}({\bm{x}},t_{n})=\partial_{t}\mathcal{A}^{n}({\bm{x}},t)|_{t=t_{n}}. Here, the unknown variables are the flow velocity 𝒗{\bm{v}} and the moving boundary Γt\Gamma_{t}. Equation (36) governs the evolution of the boundary through the forward boundary map and Equation (37) governs the dynamics of the fluid through an ALE map. They form a nonlinear system. We adopt semi-implicit BDF schemes for solving them: 𝑿F{\bm{X}}_{F} is computed explicitly at each time step, while the update of 𝒗{\bm{v}} is done implicitly. The semi-discrete scheme reads as follows,

1τi=0kaik𝑿Fn1,ni(𝒙)=i=1kbik𝒗ni𝑿Fn1,ni(𝒙)𝒙Γn1,\displaystyle\frac{1}{\tau}\sum_{i=0}^{k}a_{i}^{k}{\bm{X}}_{F}^{n-1,n-i}({\bm{x}})=\sum_{i=1}^{k}b_{i}^{k}{\bm{v}}^{n-i}\circ{\bm{X}}_{F}^{n-1,n-i}({\bm{x}})\quad\forall{\bm{x}}\in\Gamma^{n-1}, (38)
1τi=0kaik𝒗ni𝒜kn(,tni)=𝒘kn(𝒙,tn)𝒗n+Δ𝒗n+𝒇n,\displaystyle\frac{1}{\tau}\sum_{i=0}^{k}a_{i}^{k}{\bm{v}}^{n-i}\circ\mathcal{A}_{k}^{n}(\cdot,t_{n-i})={\bm{w}}_{k}^{n}({\bm{x}},t_{n})\nabla{\bm{v}}^{n}+\Delta{\bm{v}}^{n}+{\bm{f}}^{n}, (39)

where the coefficients aika_{i}^{k}, bikb_{i}^{k} are listed in Table 2 (cf. [3]), 𝒗ni=𝒗(,tni){\bm{v}}^{n-i}={\bm{v}}(\cdot,t_{n-i}), and 𝒇n=𝒇(,tn){\bm{f}}^{n}={\bm{f}}(\cdot,t_{n}). Here 𝑿Fn1,ni=(𝑿Fni,n1)1{\bm{X}}_{F}^{n-1,n-i}=({\bm{X}}_{F}^{n-i,n-1})^{-1} is the inverse of 𝑿Fni,n1{\bm{X}}_{F}^{n-i,n-1}. From (38), an explicit forward map on the boundary is defined by

𝑿Fn1,n(𝒙)=1a0ki=1k[τbik𝒗ni𝑿Fn1,ni(𝒙)aik𝑿Fn1,ni(𝒙)].{\bm{X}}_{F}^{n-1,n}({\bm{x}})=\frac{1}{a_{0}^{k}}\sum_{i=1}^{k}\big{[}\tau b_{i}^{k}{\bm{v}}^{n-i}\circ{\bm{X}}_{F}^{n-1,n-i}({\bm{x}})-a_{i}^{k}{\bm{X}}_{F}^{n-1,n-i}({\bm{x}})\big{]}.
Table 2: Coefficients for aik,bika_{i}^{k},b_{i}^{k} schemes.
        kk ii (aik,bik)(a_{i}^{k},b_{i}^{k})        0        11        22        33        44
       22        (3/2,)(3/2,\bm{\cdot})        (2,2)(-2,2)        (1/2,1)(1/2,-1)        (0,0)(0,0)        (0,0)(0,0)
       33        (11/6,)({11}/{6},\bm{\cdot})        (3,3)(-3,3)        (3/2,3)(3/2,-3)        (1/3,1)(-1/3,1)        (0,0)(0,0)
       44        (25/12,)({25}/{12},\bm{\cdot})        (4,4)(-4,4)        (3,6)(3,-6)        (4/3,4)(-4/3,4)        (1/4,1)(1/4,-1)

4.2 Surface tracking algorithm via forward boundary map

In practice, we construct a discrete approximation 𝑿hn1,n{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n} to 𝑿Fn1,n{\bm{X}}_{F}^{n-1,n} for surface tracking. Suppose that the approximate boundary Γηni\Gamma_{\eta}^{n-i}, the discrete solutions 𝒗hni𝑽(k,𝒯hni){\bm{v}}^{n-i}_{h}\in{\bm{V}}(k,\mathcal{T}_{h}^{n-i}), and the approximate backward boundary maps 𝑿hn1,ni:Γηn1Γηni{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n-i}:\Gamma_{\eta}^{n-1}\to\Gamma_{\eta}^{n-i} have been obtained for 1i<k1\leq i<k. We use the information to construct the forward boundary map 𝑿hn1,n:Γηn12{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n}:\Gamma_{\eta}^{n-1}\to\mathbb{R}^{2}, explicitly, as follows

𝑿hn1,n(𝒙)=1a0ki=1k[τbik𝒗hni𝑿hn1,ni(𝒙)aik𝑿hn1,ni(𝒙)]𝒙Γηn1.{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n}({\bm{x}})=\frac{1}{a_{0}^{k}}\sum_{i=1}^{k}\big{[}\tau b_{i}^{k}{\bm{v}}_{h}^{n-i}\circ{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n-i}({\bm{x}})-a_{i}^{k}{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n-i}({\bm{x}})\big{]}\quad\forall{\bm{x}}\in\Gamma_{\eta}^{n-1}. (40)

Let 𝒫0={𝒑j0:0jJ0}\mathcal{P}^{0}=\left\{{\bm{p}}^{0}_{j}:0\leq j\leq J^{0}\right\} be the set of control points on the initial boundary Γη0:=Γ0\Gamma^{0}_{\eta}:=\Gamma_{0}. Suppose that the arc length of Γη0\Gamma^{0}_{\eta} between 𝒑00{\bm{p}}^{0}_{0} and 𝒑j0{\bm{p}}^{0}_{j} equals to Lj0=jηL^{0}_{j}=j\eta for 1jJ01\leq j\leq J^{0}, where η:=L0/J0\eta:=L^{0}/J^{0} and L0L^{0} is the arc length of Γη0\Gamma^{0}_{\eta}. For all 0m<n0\leq m<n, suppose we are given with the set of control points 𝒫m={𝒑jm:0jJm}Γηm\mathcal{P}^{m}=\{{\bm{p}}^{m}_{j}:0\leq j\leq J^{m}\}\subset\Gamma^{m}_{\eta} and the parametric representation 𝝌m\bm{\chi}_{m} of Γηm\Gamma^{m}_{\eta}, which satisfies

𝝌m(Ljm)=𝒑jm,Ljm=i=0j|𝒑i+1m𝒑im|,0jJm.\bm{\chi}_{m}(L^{m}_{j})={\bm{p}}^{m}_{j},\quad L^{m}_{j}=\sum_{i=0}^{j}\left|{{\bm{p}}^{m}_{i+1}-{\bm{p}}^{m}_{i}}\right|,\quad 0\leq j\leq J^{m}.

The set 𝒫n={𝒑jn:j=0,,Jn}\mathcal{P}^{n}=\{{\bm{p}}_{j}^{n}:j=0,\cdots,J^{n}\} are obtained by (40) and satisfy |𝒑j+1n𝒑jn|η|{\bm{p}}_{j+1}^{n}-{\bm{p}}_{j}^{n}|\sim\eta. We adopt the surface-tracking method in [22, Algorithm 3.1 ] to construct a C2C^{2} smooth boundary Γηn\Gamma_{\eta}^{n} with cubic spline interpolation.

4.3 Construction of an approximate ALE map

Since ALE maps depend on boundary motions, the first task is to construct an approximation of the inverse map (𝑿hn1,n)1({\bm{X}}_{\mathcal{F}_{h}}^{n-1,n})^{-1}, denoted by 𝑿hn,n1{\bm{X}}_{\mathcal{F}_{h}}^{n,n-1} without causing confusions. The backward flow map 𝑿hn,n1{\bm{X}}_{\mathcal{F}_{h}}^{n,n-1} in (40) will also be used to construct Γηn+1\Gamma_{\eta}^{n+1} at the next time step. The construction of 𝑿hn,n1{\bm{X}}_{\mathcal{F}_{h}}^{n,n-1} consists of two steps.

Step 1. Construct an approximation of Xhn1,n{\bm{X}}^{n-1,n}_{\mathcal{F}_{h}}. Let Γη,Kn1=Γηn1K\Gamma^{n-1}_{\eta,K}=\Gamma^{n-1}_{\eta}\cap K and let MM be the number of control points in the interior of KK, which divide the curve into M+1M+1 segments, namely,

Γη,Kn1=ΓK,0n1ΓK,1n1ΓK,Mn1,Γ̊K,ln1Γ̊K,mn1=,lm.\Gamma^{n-1}_{\eta,K}=\Gamma_{K,0}^{n-1}\cup\Gamma_{K,1}^{n-1}\cup\cdots\cup\Gamma_{K,M}^{n-1},\quad\mathring{\Gamma}_{K,l}^{n-1}\cap\mathring{\Gamma}_{K,m}^{n-1}=\emptyset,\quad l\neq m. (41)

For each ΓK,mn1\Gamma^{n-1}_{K,m}, we take k+1k+1 nodal points 𝑨0,,𝑨k{\bm{A}}_{0},\cdots,{\bm{A}}_{k} quasi-uniformly on ΓK,mn1\Gamma^{n-1}_{K,m} with 𝑨0,𝑨k𝒫n1{\bm{A}}_{0},{\bm{A}}_{k}\in\mathcal{P}^{n-1} (see Fig. 2(a)). Let I^=[0,1]\hat{I}=[0,1] be the reference interval. The two isoparametric transforms are defined as

𝑭K,m(ξ):=i=0k𝑨ibi(ξ),𝑮K,m(ξ):=i=0k𝑨inbi(ξ)ξI^,\displaystyle{\bm{F}}_{{K,m}}(\xi):=\sum_{i=0}^{k}{\bm{A}}_{i}b_{i}(\xi),\quad{\bm{G}}_{{K,m}}(\xi):=\sum_{i=0}^{k}{\bm{A}}_{i}^{n}b_{i}(\xi)\quad\forall\,\xi\in\hat{I}, (42)

where biPk(I^)b_{i}\in P_{k}(\hat{I}) satisfies bi(l/k)=δi,lb_{i}(l/k)=\delta_{i,l} and 𝑨in=𝑿hn1,n(𝑨i){\bm{A}}_{i}^{n}={\bm{X}}_{\mathcal{F}_{h}}^{n-1,n}\big{(}{\bm{A}}_{i}\big{)}. They define a homeomorphism from ΓK,mn1\Gamma_{K,m}^{n-1} to Γ~K,mn:={GK,m(ξ):ξI^}\tilde{\Gamma}_{K,m}^{n}:=\big{\{}G_{{K,m}}(\xi):\xi\in\hat{I}\big{\}}:

𝑿~K,mn1,n:=GK,mFK,m1.\tilde{\bm{X}}^{n-1,n}_{{K,m}}:=G_{{K,m}}\circ F_{{K,m}}^{-1}. (43)

Define Γ~Kn:=m=0MΓ~K,mn\tilde{\Gamma}_{K}^{n}:=\cup_{m=0}^{M}\tilde{\Gamma}_{K,m}^{n}, and Γ~ηn:=K𝒯h,Bn1Γ~Kn\tilde{\Gamma}^{n}_{\eta}:=\cup_{K\in\mathcal{T}^{n-1}_{h,B}}\tilde{\Gamma}_{K}^{n}. We obtain a homeomorphism 𝑿~hn1,n\tilde{\bm{X}}^{n-1,n}_{\mathcal{F}_{h}}: Γηn1Γ~ηn\Gamma^{n-1}_{\eta}\to\tilde{\Gamma}^{n}_{\eta} which is defined piecewisely as follows (see Fig. 2(b))

𝑿~hn1,n=𝑿~Kn1,nonKΓηn1,𝑿~Kn1,n|ΓK,mn1=𝑿~K,mn1,n.\tilde{\bm{X}}^{n-1,n}_{\mathcal{F}_{h}}=\tilde{\bm{X}}^{n-1,n}_{K}\quad\hbox{on}\;\;K\cap\Gamma_{\eta}^{n-1},\qquad\tilde{\bm{X}}^{n-1,n}_{K}\big{|}_{\Gamma_{K,m}^{n-1}}=\tilde{\bm{X}}^{n-1,n}_{{K,m}}. (44)
𝑨0{\bm{A}}_{0}𝑨1{\bm{A}}_{1}𝑨2{\bm{A}}_{2}𝑨3{\bm{A}}_{3}ΓK,0n1\Gamma_{K,0}^{n-1}ΓK,1n1\Gamma_{K,1}^{n-1}ΓK,2n1\Gamma_{K,2}^{n-1}
(a) A cut element with markers
ΓK,1n1\Gamma^{n-1}_{K,1}𝑿hn1,n{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n}𝑨0{\bm{A}}_{0}𝑨1{\bm{A}}_{1}𝑨2{\bm{A}}_{2}𝑨3{\bm{A}}_{3}𝑨0n{\bm{A}}_{0}^{n}𝑨1n{\bm{A}}_{1}^{n}𝑨2n{\bm{A}}_{2}^{n}𝑨3n{\bm{A}}_{3}^{n}Γ~K,1n\tilde{\Gamma}^{n}_{K,1}Γηn{\Gamma}^{n}_{\eta}
(b) An illustration of 𝑿~hn1,n\tilde{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n} for k=3k=3.
Figure 2: Left: A cut element with markers 𝒫n1\mathcal{P}^{n-1}(blue points), 33 segments of Γη,Kn1\Gamma_{\eta,K}^{n-1}, (red lines), quasi-uniformly distributed points 𝑨i{\bm{A}}_{i} on ΓK,1n1\Gamma_{K,1}^{n-1}. Right: The black line Γ~K,mn\tilde{\Gamma}_{K,m}^{n} is obtained by Lagrange interpolation based on the points 𝑨in{\bm{A}}_{i}^{n}, 0i30\leq i\leq 3 (green points), which is approximation of Γηn\Gamma_{\eta}^{n} (blue line).

Step 2. Construct the backward flow map Xhn,n1{\bm{X}}^{n,n-1}_{\mathcal{F}_{h}}. It is easy to see that 𝑿~Kn1,n\tilde{\bm{X}}_{K}^{n-1,n} is the kthk^{\text{th}}-order Lagrange interpolation of 𝑿hn1,n(Γη,Kn1){\bm{X}}_{\mathcal{F}_{h}}^{n-1,n}(\Gamma_{\eta,K}^{n-1}). It is natural to use the inverse of 𝑿~hn1,n\tilde{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n} to approximate (𝑿hn1,n)1({\bm{X}}_{\mathcal{F}_{h}}^{n-1,n})^{-1}. Finally, we define the backward flow map by 𝑿hn,n1:=(𝑿~hn1,n)1𝑷n{\bm{X}}_{\mathcal{F}_{h}}^{n,n-1}:=(\tilde{\bm{X}}_{\mathcal{F}_{h}}^{n-1,n})^{-1}\circ{\bm{P}}_{n}, where 𝑷n:ΓηnΓ~ηn{\bm{P}}_{n}:\Gamma_{\eta}^{n}\to\tilde{\Gamma}_{\eta}^{n} is the projection defined by 𝑷n(𝒙):=argmin𝒚Γ~ηn|𝒚𝒙|{\bm{P}}_{n}({\bm{x}}):=\operatorname*{argmin}\limits_{{\bm{y}}\in\tilde{\Gamma}_{\eta}^{n}}\left|{{\bm{y}}-{\bm{x}}}\right| for any 𝒙Γηn{\bm{x}}\in\Gamma_{\eta}^{n}.

The construction of 𝑿hn,n1{\bm{X}}_{\mathcal{F}_{h}}^{n,n-1} is similar to [22, Section 3.4] but much more computationaly economic, since we only need the boundary map from Γηn\Gamma_{\eta}^{n} to Γηn1\Gamma_{\eta}^{n-1} instead of the volume map from Ωηn\Omega_{\eta}^{n} to Ωηn1\Omega_{\eta}^{n-1}. Finally, the ALE map 𝑿hn,n1𝑽(k,𝒯hn){\bm{X}}_{h}^{n,n-1}\in{\bm{V}}(k,\mathcal{T}_{h}^{n}) is defined as in (19) by letting 𝒈ηn,n1=𝑿hn,n1{\bm{g}}_{\eta}^{n,n-1}={\bm{X}}_{\mathcal{F}_{h}}^{n,n-1}, and 𝒜k,hn\mathcal{A}_{k,h}^{n}, 𝒘k,hn{\bm{w}}_{k,h}^{n}, and 𝑿hn,ni{\bm{X}}_{h}^{n,n-i} are obtained similarly as in (24)–(25) for i=2,,ki=2,\cdots,k.

4.4 The discrete scheme

In view of (39), the discrete scheme for solving (35) is to find 𝒗hn𝑽(k,𝒯hn){\bm{v}}_{h}^{n}\in{\bm{V}}(k,\mathcal{T}_{h}^{n}) such that

1τ(𝚲𝒘hk𝑽hn¯,𝝋h)Ωηn+(𝒗hn,𝝋h)Ωηn+𝒥1n(𝒗hn,𝝋h)=(𝒇n,𝝋h)Ωηn+𝒈N,𝝋hΓηn,\frac{1}{\tau}(\bm{\Lambda}_{{\bm{w}}_{h}}^{k}\underline{{\bm{V}}_{h}^{n}},\bm{\varphi}_{h})_{\Omega_{\eta}^{n}}+(\nabla{\bm{v}}_{h}^{n},\nabla\bm{\varphi}_{h})_{\Omega_{\eta}^{n}}+\mathscr{J}_{1}^{n}({\bm{v}}_{h}^{n},\bm{\varphi}_{h})=({\bm{f}}^{n},\bm{\varphi}_{h})_{\Omega_{\eta}^{n}}+\left<{\bm{g}}_{N},\bm{\varphi}_{h}\right>_{\Gamma_{\eta}^{n}},

for any 𝝋h𝑽(k,𝒯hn)\bm{\varphi}_{h}\in{\bm{V}}(k,\mathcal{T}_{h}^{n}), where

1τ𝚲𝒘hk𝑽hn¯=1τi=0kaik𝒗hni𝑿hn,ni𝒘k,hn𝒗hn.\frac{1}{\tau}\bm{\Lambda}_{{\bm{w}}_{h}}^{k}\underline{{\bm{V}}_{h}^{n}}=\frac{1}{\tau}\sum_{i=0}^{k}a_{i}^{k}{\bm{v}}_{h}^{n-i}\circ{\bm{X}}_{h}^{n,n-i}-{\bm{w}}_{k,h}^{n}\cdot\nabla{\bm{v}}_{h}^{n}.

5 A two-phase problem

In this section, we focus on a two-phase model with a time-varying interface. Let D2D\subset\mathbb{R}^{2} be an open square with boundary Σ=D\Sigma=\partial D. For any t0t\geq 0, let Ω1,t\Omega_{1,t} and Ω2,t\Omega_{2,t} be two time-varying sub-domains of DD occupied by two immiscible fluids, respectively. We assume Γt=Ω1,tD\Gamma_{t}=\partial\Omega_{1,t}\subset D and Ω2,t=ΣΓt\partial\Omega_{2,t}=\Sigma\cup\Gamma_{t}. Consider the linear interface problem:

uitνiΔui=fiinΩi,t,i=1,2,u=gDonΓt,𝒏u=gNonΓt,\frac{\partial u_{i}}{\partial t}-\nu_{i}\Delta u_{i}=f_{i}\;\;\;\text{in}\;\Omega_{i,t},\;\;i=1,2,\quad\left\llbracket{u}\right\rrbracket=g_{D}\;\;\text{on}\;\Gamma_{t},\quad\left\llbracket{\partial_{{\bm{n}}}u}\right\rrbracket=g_{N}\;\;\text{on}\;\Gamma_{t}, (45)

where u=u1u2\left\llbracket{u}\right\rrbracket=u_{1}-u_{2} denotes the jump of uu across Γt\Gamma_{t}, the viscosities νi\nu_{i} are positive constants, and 𝒏{\bm{n}} is the unit normal on Γt\Gamma_{t} pointing to Ω2,t\Omega_{2,t}. The interface Γt\Gamma_{t} is driven by 𝒗{\bm{v}} and has the same form as (1). We treat the model as two free-boundary problems which are coupled with the interface conditions.

5.1 Finite element spaces

Suppose we have obtained the approximate interface Γηn\Gamma_{\eta}^{n} by using some interface-tracking algorithm mentioned previously. Let Ωη,1n\Omega_{\eta,1}^{n} be an approximate domain of Ω1n=Ω1,tn\Omega_{1}^{n}=\Omega_{1,t_{n}} such that Γηn=Ωη,1n\Gamma_{\eta}^{n}=\partial\Omega_{\eta,1}^{n}. Define Ωη,2n=D\Ωη,1n\Omega_{\eta,2}^{n}=D\backslash\Omega_{\eta,1}^{n}. Similar to the single phase case, we define

Ωδ,in=:{𝒙2:min𝒚Ωη,in|𝒙𝒚|0.5τ},i=1,2.\Omega_{\delta,i}^{n}=:\Big{\{}{\bm{x}}\in\mathbb{R}^{2}:\min_{{\bm{y}}\in\Omega_{\eta,i}^{n}}|{\bm{x}}-{\bm{y}}|\leq 0.5\tau\Big{\}},\quad i=1,2. (46)

Let 𝒯h\mathcal{T}_{h} be the uniform partition of DD into closed squares of side-length hh. It generates the covers of Ωδ,in\Omega^{n}_{\delta,i}, i=1,2i=1,2 and the cover of Γηn\Gamma^{n}_{\eta}

𝒯h,in:={K𝒯hn:K¯Ω¯δ,in},𝒯B,in:={K𝒯h:K¯(ΓηnΩδ,in)}.\displaystyle\mathcal{T}^{n}_{h,i}:=\left\{K\in\mathcal{T}_{h}^{n}:\bar{K}\cap\bar{\Omega}_{\delta,i}^{n}\neq\emptyset\right\},\quad\mathcal{T}^{n}_{B,i}:=\left\{K\in\mathcal{T}_{h}:\;\bar{K}\cap(\Gamma^{n}_{\eta}\cup\partial\Omega_{\delta,i}^{n})\neq\emptyset\right\}.

We define Ω~in:=K𝒯h,inK\tilde{\Omega}_{i}^{n}:=\bigcup_{K\in\mathcal{T}^{n}_{h,i}}K. Clearly Ωη,inΩ~in\Omega^{n}_{\eta,i}\subset\tilde{\Omega}^{n}_{i}. Define

i,Bn={Eh:EΩ~inandK𝒯B,ins.t.EK}.\mathcal{E}_{i,B}^{n}=\big{\{}E\in\mathcal{E}_{h}:\;E\not\subset\partial\tilde{\Omega}^{n}_{i}\;\;\hbox{and}\;\;\exists K\in\mathcal{T}^{n}_{B,i}\;\;\hbox{s.t.}\;\;E\subset\partial K\big{\}}.

The mesh is shown in Fig. 3. We define the finite element spaces

𝒲h,0={vH01(D),v|KQk(K),K𝒯h},\displaystyle\mathcal{W}_{h,0}=\{v\in H^{1}_{0}(D),v|_{K}\in Q_{k}(K),\forall K\in\mathcal{T}_{h}\},
𝓦hn:={(vh,1,vh,2):vh,i=vh|Ω~in,vh𝒲h,0,i=1,2}.\displaystyle\mathcal{{\bm{W}}}_{h}^{n}:=\big{\{}(v_{h,1},v_{h,2}):v_{h,i}=v_{h}|_{\tilde{\Omega}^{n}_{i}},\;v_{h}\in\mathcal{W}_{h,0},\;i=1,2\big{\}}.
Ωη,1n\Omega_{\eta,1}^{n}Ωδ,1n\partial\Omega_{\delta,1}^{n}
Ωη,2n\Omega_{\eta,2}^{n}Ωδ,2n\partial\Omega_{\delta,2}^{n}
Figure 3: Left figure: Ω~1n\tilde{\Omega}_{1}^{n} (red and yellow squares), 𝒯B,1n\mathcal{T}_{B,1}^{n} (the set of red squares), 1,Bn\mathcal{E}_{1,B}^{n} (blue edges) and Ωδ,1n\partial\Omega_{\delta,1}^{n} (green edge). Right figure: Ω~2n\tilde{\Omega}_{2}^{n} (red and yellow squares), 𝒯B,2n\mathcal{T}_{B,2}^{n} (the set of red squares), 2,Bn\mathcal{E}_{2,B}^{n} (blue edges) and Ωδ,2n\partial\Omega_{\delta,2}^{n} (green edge).

5.2 Discrete ALE map and fully discrete scheme

The ALE-UFE framework is used to each phase of the interface problem, and leads to the fully discrete scheme by coupling the discrete formulations of both phases with interface conditions.

First we construct ALE maps piecewise. The discrete ALE map 𝒜k,h,1n\mathcal{A}_{k,h,1}^{n}, 𝒘k,h,1n{\bm{w}}_{k,h,1}^{n} are exactly the same as 𝒜k,hn\mathcal{A}_{k,h}^{n} and 𝒘k,hn{\bm{w}}_{k,h}^{n} defined in section 3.5. To construct 𝒜k,h,2n\mathcal{A}_{k,h,2}^{n}, we let 𝑿h,2n,n1𝑽(k,𝒯h,2n){\bm{X}}_{h,2}^{n,n-1}\in{\bm{V}}(k,\mathcal{T}_{h,2}^{n}) be the solution satisfying 𝑿h,2n,n1|D=𝑰{\bm{X}}_{h,2}^{n,n-1}\big{|}_{\partial D}={\bm{I}} and

𝒜hn(𝑿h,2n,n1,𝒗h)=Γηn(𝒈ηn,n1,𝒗h)𝒗h𝑽0(k,𝒯h,2n),\mathscr{A}_{h}^{n}({\bm{X}}_{h,2}^{n,n-1},{\bm{v}}_{h})=\mathcal{F}_{\Gamma_{\eta}^{n}}({\bm{g}}_{\eta}^{n,n-1},{\bm{v}}_{h})\qquad\forall{\bm{v}}_{h}\in{\bm{V}}_{0}(k,\mathcal{T}_{h,2}^{n}), (47)

where 𝑽0(k,𝒯h,2n):={𝒗𝑽(k,𝒯h,2n);𝒗|D=0}{\bm{V}}_{0}(k,\mathcal{T}_{h,2}^{n}):=\{{\bm{v}}\in{\bm{V}}(k,\mathcal{T}_{h,2}^{n});{\bm{v}}|_{\partial D}=0\} and 𝒈ηn,n1{\bm{g}}_{\eta}^{n,n-1} is defined in section 3.5. Then we define 𝒜k,h,2n=i=0klni𝑿h,2n,ni\mathcal{A}_{k,h,2}^{n}=\sum_{i=0}^{k}l_{n}^{i}{\bm{X}}_{h,2}^{n,n-i} and 𝒘k,h,2n=i=0k(lni)𝑿h,2n,ni{\bm{w}}_{k,h,2}^{n}=\sum_{i=0}^{k}(l_{n}^{i})^{\prime}{\bm{X}}_{h,2}^{n,n-i}.

Next we use the backward flow maps to discretize the time derivatives

1τΛ𝒘h,jk𝑼h,jn¯=1τΛk𝑼h,jn¯𝒘k,h,jnuh,jn,Λk𝑼h,jn¯:=i=0kλikuh,jni𝑿h,jn,ni.\frac{1}{\tau}\Lambda_{{\bm{w}}_{h,j}}^{k}\underline{{\bm{U}}^{n}_{h,j}}=\frac{1}{\tau}\Lambda^{k}\underline{{\bm{U}}^{n}_{h,j}}-{\bm{w}}_{k,h,j}^{n}\cdot\nabla u_{h,j}^{n},\quad\Lambda^{k}\underline{{\bm{U}}^{n}_{h,j}}:=\sum_{i=0}^{k}\lambda_{i}^{k}u_{h,j}^{n-i}\circ{\bm{X}}_{h,j}^{n,n-i}.

The discrete scheme is to find uhn𝑾hnu_{h}^{n}\in{\bm{W}}_{h}^{n} such that

j=12[(1τΛ𝒘h,jk𝑼h,jn¯fhn,vj)Ωη,jn+𝒜jn(uh,jn,vj)]+𝒮In(uhn,v)+𝒥In(uhn,v)=0,\displaystyle\sum_{j=1}^{2}\Big{[}\Big{(}\frac{1}{\tau}\Lambda^{k}_{{\bm{w}}_{h,j}}\underline{{\bm{U}}^{n}_{h,j}}-f^{n}_{h},v_{j}\Big{)}_{\Omega_{\eta,j}^{n}}+\mathscr{A}_{j}^{n}(u_{h,j}^{n},v_{j})\Big{]}+\mathscr{S}_{I}^{n}(u_{h}^{n},v)+\mathscr{J}_{I}^{n}(u_{h}^{n},v)=0,

for all v𝑾hnv\in{\bm{W}}_{h}^{n} with vj=v|Ωη,jnv_{j}=v|_{\Omega_{\eta,j}^{n}}, where

𝒜jn(u,v)\displaystyle\mathscr{A}_{j}^{n}(u,v) :=(νju,v)Ωη,jn+𝒥1,j(u,v),\displaystyle:=(\nu_{j}\nabla u,\,\nabla v)_{\Omega_{\eta,j}^{n}}+\mathscr{J}_{1,j}(u,v),
𝒥1,j(u,v)\displaystyle\mathscr{J}_{1,j}(u,v) :=Ej,Bnl=1kh2l1νj𝒏luj,𝒏lvjE.\displaystyle:=\sum_{E\in\mathcal{E}_{j,B}^{n}}\sum_{l=1}^{k}h^{2l-1}\left<\nu_{j}\left\llbracket{\partial_{{\bm{n}}}^{l}u_{j}}\right\rrbracket,\left\llbracket{\partial_{{\bm{n}}}^{l}v_{j}}\right\rrbracket\right>_{E}.

Moreover, 𝒮In\mathscr{S}_{I}^{n} and 𝒥In\mathscr{J}_{I}^{n} are bilinear forms which combine the two phases

𝒮In(u,v):=\displaystyle\mathscr{S}_{I}^{n}(u,v):=\, {{ν𝒏u}},vΓηn{{ν𝒏v}},uΓηn,\displaystyle-\left<\left\{\hskip-3.5pt\left\{{\nu\partial_{{\bm{n}}}u}\right\}\hskip-3.5pt\right\},\left\llbracket{v}\right\rrbracket\right>_{{\Gamma_{\eta}^{n}}}-\left<\left\{\hskip-3.5pt\left\{{\nu\partial_{{\bm{n}}}v}\right\}\hskip-3.5pt\right\},\left\llbracket{u}\right\rrbracket\right>_{{\Gamma_{\eta}^{n}}},
𝒥In(u,v):=\displaystyle\mathscr{J}_{I}^{n}(u,v):=\, γ0h1{{ν}}u,vΓηn.\displaystyle\gamma_{0}h^{-1}\left\{\hskip-3.5pt\left\{{\nu}\right\}\hskip-3.5pt\right\}\left<\left\llbracket{u}\right\rrbracket,\left\llbracket{v}\right\rrbracket\right>_{\Gamma_{\eta}^{n}}.

where γ0\gamma_{0} is a positive penalty coefficient and 𝒏{\bm{n}} is the unit outward normal to Γηn\Gamma_{\eta}^{n} from Ωη,1n\Omega_{\eta,1}^{n} to Ωη,2n\Omega_{\eta,2}^{n}. Here we have used the average operator

{{a}}=κ1a1+κ2a2,κ1=ν2/(ν1+ν2),κ2=ν1/(ν1+ν2).\left\{\hskip-3.5pt\left\{{a}\right\}\hskip-3.5pt\right\}=\kappa_{1}a_{1}+\kappa_{2}a_{2},\quad\kappa_{1}=\nu_{2}/({\nu_{1}+\nu_{2}}),\quad\kappa_{2}={\nu_{1}}/({\nu_{1}+\nu_{2}}).

6 Numerical experiments

In this section, we demonstrate the ALE-UFE method with all the models in the previous sections. Throughout the section, we choose γ0=1000\gamma_{0}=1000. To simplify computations, we set the pre-calculated initial values by the exact solution, namely, uhj=u(,tj)u^{j}_{h}=u(\cdot,t_{j}), for 0jk10\leq j\leq k-1.

6.1 One-phase linear problem

In this section, we consider a one-phase problem where the velocity of the moving boundary is given by an analytic function. The exact solution is set by u=sin(π(x1+t))sin(π(x2+t))u=\sin(\pi(x_{1}+t))\sin(\pi(x_{2}+t)). The right-hand side and the boundary values in (8) are set by uu. The mesh of the evolution domain is shown in Fig. 4. The boundary is varying according to

x1(t)=x1(0)1+0.2sin(2t)+sin(2t)16,x2(t)=x2(0)10.25sin(2t)+sin(2t)16.x_{1}(t)=\frac{x_{1}(0)}{1+0.2\sin(2t)}+\frac{\sin(2t)}{16},\quad x_{2}(t)=\frac{x_{2}(0)}{1-0.25\sin(2t)}+\frac{\sin(2t)}{16}.

The initial boundary is a circle centered at (0.5,0.5)(0.5,0.5)^{\top} with radius 1/81/8. The numerical error is measured by

eN=(u(,T)uhNL2(ΩηN)2+τn=kN|uuhn|H1(Ωηn)2)1/2.e^{N}=\bigg{(}\|u(\cdot,T)-u_{h}^{N}\|_{L^{2}(\Omega_{\eta}^{N})}^{2}+\tau\sum_{n=k}^{N}|u-u_{h}^{n}|_{H^{1}(\Omega_{\eta}^{n})}^{2}\bigg{)}^{1/2}.
Refer to caption
Figure 4: The moving domain Ωηn\Omega_{\eta}^{n} at tn=0t_{n}=0, 0.20.2, 0.40.4, 0.60.6, 0.80.8 and 11 (h=1/16h=1/16).

Numerical results for k=3,4k=3,4 are shown in Tables 3. Optimal convergence rates eN=O(τk)e^{N}=O(\tau^{k}) are obtained for both the third- and fourth-order methods. Although the location of the boundary is given explicitly, the UCFE method in [21] is not applicable in this case due to the lack of flow velocity inside the domain and the absence of moving domains that maintain the same volume. Therefore, the ALE-UFE method has a wider range of applications than the UCFE mehtod.

Table 3: Convergence rates for section 6.1.
           h=τh=\tau            eN(k=3)e^{N}\;(k=3)            rate            eN(k=4)e^{N}\;(k=4)            rate
           1/16            6.16e-03            -            1.91e-3            -
           1/32            7.94e-04            2.95            1.25e-4            3.93
           1/64            1.00e-04            2.98            9.97e-6            3.97
           1/128            1.25e-05            2.99            5.01e-7            3.98

6.2 Nonlinear problem

Now we demonstrate the ALE-UFE method with a nonlinear problem where the boundary motion is specified by the solution. We require ΩT=Ω0\Omega_{T}=\Omega_{0} to check the accuracy of interface-tracking. The exact solution is set by

𝒗=cos(πt/3)[sin(2πx1)sin(2πx2),cos(2πx1)cos(2πx2)].{\bm{v}}=\cos(\pi t/3)\big{[}\sin(2\pi x_{1})\sin(2\pi x_{2}),\cos(2\pi x_{1})\cos(2\pi x_{2})\big{]}^{\top}.

The initial domain is a disk with radius R=0.15R=0.15 and centered at (0.5,0.5)(0.5,0.5). The domain is stretched into an inverted U shape at t=1.5t=1.5 and returns to its initial shape at T=3T=3 (see Fig. 5).

Refer to caption
Figure 5: Approximate domains at tn=0t_{n}=0, 0.60.6, 1.21.2, 1.81.8, 2.42.4 and 33 (h=1/16h=1/16).

The source term and the Neumann condition of (35) are given by

𝒇=t𝒗Δ𝒗inΩt,𝒈N=𝒗𝒏on Γt.{\bm{f}}=\partial_{t}{\bm{v}}-\Delta{\bm{v}}\quad\text{in}\;\Omega_{t},\quad{\bm{g}}_{N}=\nabla{\bm{v}}\cdot{\bm{n}}\quad\text{on }\;\Gamma_{t}.

Define the boundary-tracking errors

e0,Ω\displaystyle e_{0,\Omega} =K𝒯h|area(ΩTK)area(ΩηNK)|,\displaystyle=\sum_{K\in\mathcal{T}_{h}}|\text{area}(\Omega_{T}\cap K)-\text{area}(\Omega_{\eta}^{N}\cap K)|,
e1,Ω\displaystyle e_{1,\Omega} =[n=kNτ(K𝒯h|area(ΩrefnK)area(ΩηnK)|)2]1/2.\displaystyle=\bigg{[}\sum_{n=k}^{N}\tau\Big{(}\sum_{K\in\mathcal{T}_{h}}\big{|}\text{area}(\Omega_{\text{ref}}^{n}\cap K)-\text{area}(\Omega_{\eta}^{n}\cap K)\big{|}\Big{)}^{2}\bigg{]}^{1/2}.

Here Ωrefn\Omega_{\text{ref}}^{n} is the reference domain computed with the finest grid and the smallest time step. Approximation errors are computed on the numerically tracked domain

e0=(n=kNτ𝒗(,tn)𝒗hn𝑳2(Ωrefn)2)1/2,e1=(n=kNτ|𝒗(,tn)𝒗hn|𝑯1(Ωrefn)2)1/2.e_{0}=\bigg{(}\sum_{n=k}^{N}\tau\|{\bm{v}}(\cdot,t_{n})-{\bm{v}}_{h}^{n}\|^{2}_{{\bm{L}}^{2}(\Omega_{\text{ref}}^{n})}\bigg{)}^{1/2},\quad e_{1}=\bigg{(}\sum_{n=k}^{N}\tau|{\bm{v}}(\cdot,t_{n})-{\bm{v}}_{h}^{n}|^{2}_{{\bm{H}}^{1}(\Omega_{\text{ref}}^{n})}\bigg{)}^{1/2}.

Tables 4 and  5 show that optimal convergence rates are obtained for the numerical solutions and the boundary-tracking algorithm as hh (or τ\tau) approaches 0.

Table 4: Convergence rates for section 6.2 (k=3k=3).
h=τh=\tau e0e_{0} rate e1e_{1} rate e0,Ωe_{0,\Omega} rate e1,Ωe_{1,\Omega} rate
1/32 1.11e-04 - 4.06e-04 - 1.21e-03 - 1.40e-03 -
1/64 1.20e-05 3.21 4.42e-05 3.19 1.65e-04 2.88 1.71e-04 3.03
1/128 1.41e-06 3.08 5.50e-06 3.00 2.13e-05 2.95 2.11e-05 3.01
1/256 1.72e-07 3.03 6.88e-07 3.00 2.67e-06 2.99 2.63e-06 3.00
Table 5: Convergence rates for section 6.2 (k=4k=4)
h=τh=\tau e0e_{0} rate e1e_{1} rate e0,Ωe_{0,\Omega} rate e1,Ωe_{1,\Omega} rate
1/32 4.72e-05 - 2.34e-04 - 4.05e-04 - 4.48e-04 -
1/64 4.14e-06 3.51 1.47e-05 3.99 3.93e-05 3.36 3.84e-05 3.54
1/128 2.82e-07 3.87 4.39e-07 5.06 2.73e-06 3.85 2.78e-06 3.78
1/256 1.80e-08 3.96 2.76e-08 3.99 1.75e-07 3.96 1.86e-07 3.90

6.3 Two-phase flow

Refer to caption
Figure 6: Two moving subdomains Ωη,1n\Omega_{\eta,1}^{n}, Ωη,2n\Omega_{\eta,2}^{n} at tn=0t_{n}=0, 0.30.3, 0.60.6, 0.90.9, 1.21.2 and 1.51.5 (h=1/16h=1/16).

We use the cubic MARS algorithm in [28] to track the interface and construct Ωη,1n\Omega_{\eta,1}^{n} and Ωη,2n=D\Ωη,1n\Omega_{\eta,2}^{n}=D\backslash\Omega_{\eta,1}^{n}. The initial domain Ω1,0\Omega_{1,0} is a disk of radius R=0.15R=0.15 at (0.5,0.75)(0.5,0.75). The flow velocity is set by

𝒗=cos(πt/3)[sin2(πx1)sin(2πx2),sin2(πx2)sin(2πx1)]{\bm{v}}=\cos(\pi t/3)\big{[}\sin^{2}(\pi x_{1})\sin(2\pi x_{2}),-\sin^{2}(\pi x_{2})\sin(2\pi x_{1})\big{]}^{\top}

The domain is stretched into a snake-like region at T=1.5T=1.5 (see Fig. 6). The viscosities are set by ν1=1000\nu_{1}=1000 and ν2=1\nu_{2}=1. The exact solution is set by

u1=sin(π(x1+t))sin(π(x2+t)),u2=exsin(π(x2+t)).u_{1}=\sin(\pi(x_{1}+t))\sin(\pi(x_{2}+t)),\quad u_{2}=e^{x}\sin(\pi(x_{2}+t)).

The numerical error is measured by eNe^{N}, where

(eN)2=i=12[u(T)uhNL2(Ωη,iN)2+τn=kNνi|u(tn)uhn|H1(Ωη,in)2].(e^{N})^{2}=\sum_{i=1}^{2}\bigg{[}\|u(T)-u_{h}^{N}\|_{L^{2}(\Omega_{\eta,i}^{N})}^{2}+\tau\sum_{n=k}^{N}\nu_{i}|u(t_{n})-u_{h}^{n}|_{H^{1}(\Omega_{\eta,i}^{n})}^{2}\bigg{]}.

Convergence orders for k=3k=3 and 44 are shown in Table 6. For such a large deformation of the domain, the method still yields optimal convergence rates.

Table 6: Convergence rates for section 6.3.
           h=τh=\tau            eN(k=3)e^{N}\;(k=3)            rate            eN(k=4)e^{N}\;(k=4)            rate
           1/16            6.05e-03            -            7.87e-3            -
           1/32            8.78e-04            2.78            5.17e-5            3.92
           1/64            1.15e-04            2.92            3.27e-6            3.97
           1/128            1.47e-05            2.97            2.07e-7            3.97

6.4 Domain with topological change

Finally we consider the domain having a topological change. Its boundary is given by the level set of the function

ϕ(𝒙,t)=min{|𝒙𝒄1(t)|,|𝒙𝒄2(t)|}0.15,\displaystyle\phi({\bm{x}},t)=\min\{|{\bm{x}}-{\bm{c}}_{1}(t)|,|{\bm{x}}-{\bm{c}}_{2}(t)|\}-0.15,

where 𝒄1(t)=[0.5,0.750.5t]{\bm{c}}_{1}(t)=[0.5,0.75-0.5t]^{\top} and 𝒄2(t)=[0.5,0.25+0.5t]{\bm{c}}_{2}(t)=[0.5,0.25+0.5t]^{\top} are the centers of the two circles. The finial time is set by T=1T=1 which satisfies ϕ(𝒙,0)=ϕ(𝒙,T)\phi({\bm{x}},0)=\phi({\bm{x}},T). The evolution of the domain is shown in Fig. 7.

Refer to caption
Figure 7: The moving domain Ωηn\Omega_{\eta}^{n} at tn=0t_{n}=0, 0.20.2, 0.40.4, 0.60.6, 0.80.8, and 11 (h=1/16h=1/16).
Table 7: Convergence rates for section 6.4 (k=3k=3).
            h=τh=\tau             e0e_{0}             rate             e1e_{1}             rate
            1/16             4.95e-05             -             6.37e-4             -
            1/32             1.21e-05             2.03             1.52e-4             2.02
            1/64             3.06e-06             1.98             4.36e-5             1.85
            1/128             7.80e-07             1.98             1.09e-5             2.00

The exact solution is u=sin(π(x1+t))sin(π(x2+t))u=\sin(\pi(x_{1}+t))\sin(\pi(x_{2}+t)). In the implementation of the ALE map, we choose 𝒈ηn,n1{\bm{g}}_{\eta}^{n,n-1} by the closet point mapping. Again we compute the L2L^{2}- and H1H^{1}-errors of the solution on the approximate domain Ωηn\Omega_{\eta}^{n}. Since the evolution of the domain is discontinuous, the assumption in (30) does not hold anymore. Table  7 shows that the convergence rate deteriorates into second order.

7 Conclusions

An arbitrary Lagrangian-Eulerian unfitted finite element (ALE-UFE) method has been presented for solving PDEs on time-varying domains. High-order convergence is obtained by adopting BDF schemes and ALE maps for time integration and unfitted finite element method for spatial discretization. The method is applied to various models, including a varying interface problem, a PDE-domain coupled problem, and a problem with topologically changing domain. The ALE-UFE method has the potential for solving a variety of moving-domain problems, such as fluid dynamics and FSI problems. These will be our future work.

Appendix A Useful estimates of the ALE mapping 𝑿n,n1{\bm{X}}^{n,n-1}

Lemma A.1.

Suppose 𝐠n,n1(𝐱)=𝐗F(tn1;tn,𝐱){\bm{g}}^{n,n-1}({\bm{x}})={\bm{X}}_{F}(t_{n-1};t_{n},{\bm{x}}) and 𝐗F{\bm{X}}_{F} is given by (2). The Jacobi matrices 𝕁n,ni=𝐱𝐗n,ni\mathbb{J}^{n,n-i}=\partial_{{\bm{x}}}{\bm{X}}^{n,n-i}, 1ik1\leq i\leq k, of ALE maps admit

𝕁n,ni𝕀L(Ωn)τ,𝒘kn(𝒙,tn).\|\mathbb{J}^{n,n-i}-\mathbb{I}\|_{L^{\infty}(\Omega^{n})}\lesssim\tau,\quad\|{\bm{w}}_{k}^{n}({\bm{x}},t_{n})\|\lesssim\infty.
Proof.

For ease of notation, we write 𝒅𝑿=𝑿n,n1𝑿n,n{\bm{d}}_{\bm{X}}={\bm{X}}^{n,n-1}-{\bm{X}}^{n,n}. Then (4) implies

Δ𝒅𝑿=0inΩn,𝒅𝑿=𝒈n,n1𝒈n,nonΓn.\displaystyle-\Delta{\bm{d}}_{\bm{X}}=\textbf{0}\quad\text{in}\;\Omega^{n},\qquad{\bm{d}}_{\bm{X}}={\bm{g}}^{n,n-1}-{\bm{g}}^{n,n}\quad\text{on}\;\Gamma^{n}.

Since 𝒈n,n=𝑰{\bm{g}}^{n,n}={\bm{I}} and both Γn\Gamma^{n} and 𝒗{\bm{v}} are CrC^{r}-smooth, we have 𝒈n,n1𝒈n,n𝑯r(Γn)τ\|{\bm{g}}^{n,n-1}-{\bm{g}}^{n,n}\|_{{\bm{H}}^{r}(\Gamma^{n})}\lesssim\tau. The regularity result of elliptic equations yields

𝒅𝑿𝑯r+1/2(Ωn)C𝒈n,n1𝒈n,n𝑯r(Γn)τ.\left\|{{\bm{d}}_{\bm{X}}}\right\|_{{\bm{H}}^{r+1/2}(\Omega^{n})}\leq C\left\|{{\bm{g}}^{n,n-1}-{\bm{g}}^{n,n}}\right\|_{{\bm{H}}^{r}(\Gamma^{n})}\lesssim\tau. (48)

Using 𝑿n,n=𝑰{\bm{X}}^{n,n}={\bm{I}}, Sobolev’s inequality, and (48), we find that

𝕁n,n1𝕀𝑳(Ωn)=𝒅𝑿𝑾1,(Ωn)𝒅𝑿𝑯r+1/2(Ωn)τ.\|{\mathbb{J}^{n,n-1}-\mathbb{I}}\|_{{\bm{L}}^{\infty}(\Omega^{n})}=\|{{\bm{d}}_{\bm{X}}}\|_{{\bm{W}}^{1,\infty}(\Omega^{n})}\lesssim\|{{\bm{d}}_{\bm{X}}}\|_{{\bm{H}}^{r+1/2}(\Omega^{n})}\lesssim\tau.

By (5) and the chain rule, we obtain 𝕁n,ni𝕀𝑳(Ωn)τ\|{\mathbb{J}^{n,n-i}-\mathbb{I}}\|_{{\bm{L}}^{\infty}(\Omega^{n})}\lesssim\tau for 2ik2\leq i\leq k.

From the definition of 𝒘kn{\bm{w}}^{n}_{k} in (7), a simple calculation gives

𝒘kn(𝒙,tn)|Γn=i=0k(lni)(tn)𝑿n,ni|Γn=1τi=0kλik𝑿n,ni|Γn=1τi=0kλik𝒈n,ni,{\bm{w}}_{k}^{n}({\bm{x}},t_{n})|_{\Gamma^{n}}=\sum_{i=0}^{k}(l_{n}^{i})^{\prime}(t_{n}){\bm{X}}^{n,n-i}|_{\Gamma^{n}}=\frac{1}{\tau}\sum_{i=0}^{k}\lambda_{i}^{k}{\bm{X}}^{n,n-i}|_{\Gamma^{n}}=\frac{1}{\tau}\sum_{i=0}^{k}\lambda_{i}^{k}{\bm{g}}^{n,n-i},

where λik\lambda_{i}^{k} are the coefficients of BDF-kk. Clearly 𝒘kn(𝒙,tn)|Γn{\bm{w}}_{k}^{n}({\bm{x}},t_{n})|_{\Gamma^{n}} is an approximation to t𝑿F(t;tn,)|t=tn\partial_{t}{\bm{X}}_{F}(t;t_{n},\cdot)|_{t=t_{n}}. It is easy to see from (4) that

Δ𝒘kn(𝒙,tn)=0inΩn,𝒘kn(𝒙,tn)=τ1i=0kλik𝒈n,nionΓn.-\Delta{\bm{w}}^{n}_{k}({\bm{x}},t_{n})=0\quad\text{in}\;\;\Omega^{n},\qquad{\bm{w}}^{n}_{k}({\bm{x}},t_{n})=\tau^{-1}\sum_{i=0}^{k}\lambda_{i}^{k}{\bm{g}}^{n,n-i}\quad\text{on}\;\;\Gamma^{n}. (49)

By Sobolev’s inequalities and regularity theories of elliptic equations, we obtain

𝒘kn(tn)𝑳(Ωn)1τi=0kλik𝒈n,ni𝑯r(Γn)t𝑿F(t;tn,)|t=tn𝑯r(Γn).\displaystyle\|{\bm{w}}^{n}_{k}(t_{n})\|_{{\bm{L}}^{\infty}(\Omega^{n})}\lesssim\frac{1}{\tau}\Big{\|}\sum_{i=0}^{k}\lambda_{i}^{k}{\bm{g}}^{n,n-i}\Big{\|}_{{\bm{H}}^{r}(\Gamma^{n})}\lesssim\|\partial_{t}{\bm{X}}_{F}(t;t_{n},\cdot)|_{t=t_{n}}\|_{{\bm{H}}^{r}(\Gamma^{n})}.

Since the computations and numerical analysis are performed on the approximate domain Ωηn\Omega_{\eta}^{n}, which is different from the exact one Ωn\Omega^{n} in general, we have to extend 𝑿n,n1{\bm{X}}^{n,n-1} from Ωn\Omega^{n} to the fictitious domain Ω~n\tilde{\Omega}^{n}.

Lemma A.2.

There exits an extension of 𝐗n,n1𝐇r+1/2(Ωn){\bm{X}}^{n,n-1}\in{\bm{H}}^{r+1/2}(\Omega^{n}), denoted by 𝐗~n,n1𝐇r+1/2(Ω~n)\tilde{\bm{X}}^{n,n-1}\in{\bm{H}}^{r+1/2}(\tilde{\Omega}^{n}), such that

𝑿~n,n1|Ωn=𝑿n,n1,𝑿~n,n1𝑯r+1/2(Ω~n).\tilde{\bm{X}}^{n,n-1}|_{\Omega^{n}}={\bm{X}}^{n,n-1},\qquad\|\tilde{\bm{X}}^{n,n-1}\|_{{\bm{H}}^{r+1/2}(\tilde{\Omega}^{n})}\leq\infty. (50)

Moreover, the Jacobi matrix 𝕁~n,n1=𝐱𝐗~n,n1\tilde{\mathbb{J}}^{n,n-1}=\partial_{{\bm{x}}}\tilde{{\bm{X}}}^{n,n-1} satisfies

𝕁~n,n1|Ωn=𝕁n,n1,𝕁~n,n1𝕀𝕃(Ω~n)τ.\tilde{\mathbb{J}}^{n,n-1}|_{\Omega^{n}}=\mathbb{J}^{n,n-1},\qquad\|\tilde{\mathbb{J}}^{n,n-1}-\mathbb{I}\|_{\mathbb{L}^{\infty}(\tilde{\Omega}^{n})}\lesssim\tau.
Proof.

By the Sobolev extension theorem and (48), there exits an extension of 𝒅𝑿{\bm{d}}_{{\bm{X}}}, denoted by 𝒅𝑿~𝑯r+1/2(Ω~n)\widetilde{{\bm{d}}_{{\bm{X}}}}\in{\bm{H}}^{r+1/2}(\tilde{\Omega}^{n}), such that

𝒅𝑿~|Ωn=𝒅𝑿,𝒅𝑿~𝑯r+1/2(Ω~n)𝒅𝑿𝑯r+1/2(Ωn)τ.\widetilde{{\bm{d}}_{{\bm{X}}}}|_{\Omega^{n}}={\bm{d}}_{\bm{X}},\qquad\|\widetilde{{\bm{d}}_{\bm{X}}}\|_{{\bm{H}}^{r+1/2}(\tilde{\Omega}^{n})}\lesssim\|{\bm{d}}_{\bm{X}}\|_{{\bm{H}}^{r+1/2}(\Omega^{n})}\lesssim\tau. (51)

Then the extension of 𝑿n,n1{\bm{X}}^{n,n-1} is defined by 𝑿~n,n1:=𝑰+τ𝒅𝑿~\tilde{{{\bm{X}}}}^{n,n-1}:={\bm{I}}+\tau\widetilde{{\bm{d}}_{\bm{X}}}. It follows that

𝑿~n,n1|Ωn=𝑿n,n1,𝑿~n,n1𝑯r+12(Ω~n)area(D)+τ.\tilde{{\bm{X}}}^{n,n-1}|_{\Omega^{n}}={{\bm{X}}}^{n,n-1},\qquad\|\tilde{{\bm{X}}}^{n,n-1}\|_{{\bm{H}}^{r+\frac{1}{2}}(\tilde{\Omega}^{n})}\lesssim\text{area}(D)+\tau. (52)

Since rk+13r\geq k+1\geq 3, from Sobolev’s inequality and (51), for μ=0\mu=0, 11, we get

𝑿~n,n1𝑰𝑾μ,(Ω~n)τ𝒅𝑿~𝑯2+μ(Ω~n)τ2.\|\tilde{{\bm{X}}}^{n,n-1}-{\bm{I}}\|_{{\bm{W}}^{\mu,\infty}(\tilde{\Omega}^{n})}\lesssim\tau\|\widetilde{{{\bm{d}}}_{{\bm{X}}}}\|_{{\bm{H}}^{2+\mu}(\tilde{\Omega}^{n})}\lesssim\tau^{2}. (53)

This implies 𝕁~n,n1𝕀𝕃(Ω~n)τ\big{\|}\tilde{\mathbb{J}}^{n,n-1}-\mathbb{I}\big{\|}_{\mathbb{L}^{\infty}(\tilde{\Omega}^{n})}\lesssim\tau by noting 𝕁~n,n1=𝒙𝑿~n,n1\tilde{\mathbb{J}}^{n,n-1}=\partial_{{\bm{x}}}\tilde{\bm{X}}^{n,n-1}. ∎

Appendix B Useful estimates of the discrete ALE map 𝑿hn,ni{\bm{X}}_{h}^{n,n-i}

Lemma B.1.

Let 𝕁hn,n1:=𝐱𝐗hn,n1\mathbb{J}_{h}^{n,n-1}:=\partial_{{\bm{x}}}{\bm{X}}_{h}^{n,n-1} be the Jacobi matrix of 𝐗hn,n1{\bm{X}}_{h}^{n,n-1}. Upon hidden constants independent of τ,h,n\tau,h,n, and η\eta, there hold

𝕁hn,n1𝕀𝕃(Ω~n)τ,𝑿hn,n1𝑿~n,n1𝑳(Ω~n)τk1/2.\|{\mathbb{J}_{h}^{n,n-1}-\mathbb{I}}\|_{\mathbb{L}^{\infty}(\tilde{\Omega}^{n})}\lesssim\tau,\qquad\|{{\bm{X}}_{h}^{n,n-1}-\tilde{\bm{X}}^{n,n-1}}\|_{{\bm{L}}^{\infty}(\tilde{\Omega}^{n})}\lesssim\tau^{k-1/2}. (54)
Proof.

Since ΩηnΩn\Omega_{\eta}^{n}\neq\Omega^{n}, multiplying both sides of (4) by 𝒗h𝑽(k,𝒯hn){\bm{v}}_{h}\in{\bm{V}}(k,\mathcal{T}_{h}^{n}) and using 𝑿~n,n1𝑯r+1/2(Ω~n)Hk+1(Ω~n)\tilde{\bm{X}}^{n,n-1}\in{\bm{H}}^{r+1/2}(\tilde{\Omega}^{n})\subset H^{k+1}(\tilde{\Omega}^{n}), we have

𝒜hn(𝑿~n,n1,𝒗h)=(Δ𝑿~n,n1,𝒗h)Ωηn\Ωn+Γηn(𝑿~n,n1,𝒗h).\mathcal{A}_{h}^{n}(\tilde{\bm{X}}^{n,n-1},{\bm{v}}_{h})=-(\Delta\tilde{{\bm{X}}}^{n,n-1},{\bm{v}}_{h})_{\Omega_{\eta}^{n}\backslash\Omega^{n}}+\mathcal{F}_{\Gamma_{\eta}^{n}}(\tilde{\bm{X}}^{n,n-1},{\bm{v}}_{h}). (55)

For convenience, let hn\mathcal{I}_{h}^{n} be the Lagrange interpolation operator and define

𝒆=𝑿hn,n1𝑿~n,n1,𝝃h=𝑿hn,n1hn𝑿~n,n1,𝜼=hn𝑿~n,n1𝑿~n,n1.{\bm{e}}={\bm{X}}_{h}^{n,n-1}-\tilde{\bm{X}}^{n,n-1},\;\bm{\xi}_{h}={\bm{X}}_{h}^{n,n-1}-\mathcal{I}_{h}^{n}\tilde{\bm{X}}^{n,n-1},\;\bm{\eta}=\mathcal{I}_{h}^{n}\tilde{\bm{X}}^{n,n-1}-\tilde{\bm{X}}^{n,n-1}.

It is easy to see that 𝒆=𝝃h+𝜼{\bm{e}}=\bm{\xi}_{h}+\bm{\eta}. Subtracting (55) from (19), one has

𝒜hn(𝒆,𝒗h)=(Δ𝑿~n,n1,𝒗h)Ωηn\Ωn+Γηn(𝒈ηn,n1𝑿~n,n1,𝒗h).\displaystyle\mathcal{A}_{h}^{n}({\bm{e}},{\bm{v}}_{h})=(\Delta\tilde{{\bm{X}}}^{n,n-1},{\bm{v}}_{h})_{\Omega_{\eta}^{n}\backslash\Omega^{n}}+\mathcal{F}_{\Gamma_{\eta}^{n}}({\bm{g}}_{\eta}^{n,n-1}-\tilde{\bm{X}}^{n,n-1},{\bm{v}}_{h}). (56)

For any 𝒗h𝑽(k,𝒯hn){\bm{v}}_{h}\in{\bm{V}}(k,\mathcal{T}_{h}^{n}), Poincare´\acute{\text{e}} inequality and (28) show

𝒗h𝑳2(Ω~n)|𝒗h|𝑯1(Ω~n)+𝒥0n(𝒗h,𝒗h)|𝒗h|𝒯hn.\left\|{{\bm{v}}_{h}}\right\|_{{\bm{L}}^{2}({\tilde{\Omega}^{n}})}\lesssim{\left|{{\bm{v}}_{h}}\right|}_{{\bm{H}}^{1}({\tilde{\Omega}^{n}})}+\mathscr{J}_{0}^{n}({\bm{v}}_{h},{\bm{v}}_{h})\lesssim\left\|\left|{{\bm{v}}_{h}}\right|\right\|_{\mathcal{T}_{h}^{n}}. (57)

Since area(Ωηn\Ωn)=τk+1\text{area}(\Omega_{\eta}^{n}\backslash\Omega^{n})=\tau^{k+1}, by [21, Lemma A4] and (57), we have

(Δ𝑿~n,n1,𝒗h)Ωηn\Ωn\displaystyle(\Delta\tilde{{\bm{X}}}^{n,n-1},{\bm{v}}_{h})_{\Omega_{\eta}^{n}\backslash\Omega^{n}} τk+1𝑿~n,n1𝑯3(Ω~n)|𝒗h|𝒯hn.\displaystyle\lesssim\tau^{k+1}\|{\tilde{\bm{X}}^{n,n-1}}\|_{{\bm{H}}^{3}(\tilde{\Omega}^{n})}\||{{\bm{v}}_{h}}|\|_{\mathcal{T}_{h}^{n}}. (58)

Since the approximate boundary satisfies dist(Γn,Γηn)τk+1\text{dist}(\Gamma^{n},\Gamma_{\eta}^{n})\lesssim\tau^{k+1}, and 𝒈ηn,n1{\bm{g}}_{\eta}^{n,n-1} is the (k+1)th(k+1)^{\text{th}}-order approximation to 𝒈n,n1{\bm{g}}^{n,n-1}, there holds

𝒈ηn,n1𝑿~n,n1L2(Γηn)τk+1.\|{\bm{g}}_{\eta}^{n,n-1}-\tilde{\bm{X}}^{n,n-1}\|_{L^{2}(\Gamma_{\eta}^{n})}\lesssim\;\tau^{k+1}. (59)

From the trace inequality, inverse estimate, and (59), we have

Γηn(𝑿τn,n1𝑿~n,n1,𝒗h)τk+12|𝒗h|𝒯hn.\displaystyle\mathcal{F}_{\Gamma_{\eta}^{n}}({\bm{X}}_{\tau}^{n,n-1}-\tilde{\bm{X}}^{n,n-1},{\bm{v}}_{h})\lesssim\tau^{k+\frac{1}{2}}\||{{\bm{v}}_{h}}|\|_{\mathcal{T}_{h}^{n}}. (60)

Insert (58) and (60) into (56). By Lemma 3.8 and interpolation error estimates, we obtain

|𝝃h|𝒯hn2𝒜hn(𝝃h,𝝃h)=𝒜hn(𝒆,𝝃h)𝒜hn(𝜼,𝝃h)(τk+12+hk)|𝝃h|𝒯hn.\displaystyle\||{\bm{\xi}_{h}}|\|^{2}_{\mathcal{T}_{h}^{n}}\lesssim\mathscr{A}_{h}^{n}(\bm{\xi}_{h},\bm{\xi}_{h})=\mathcal{A}_{h}^{n}({\bm{e}},\bm{\xi}_{h})-\mathcal{A}_{h}^{n}(\bm{\eta},\bm{\xi}_{h})\lesssim(\tau^{k+\frac{1}{2}}+h^{k})\||{\bm{\xi}_{h}}|\|_{\mathcal{T}_{h}^{n}}. (61)

Thus together with (28) and (61), one has

|𝝃h|𝑯1(Ω~n)|𝝃h|𝒯hnhk+τk+12.|{\bm{\xi}_{h}}|_{{\bm{H}}^{1}(\tilde{\Omega}^{n})}\lesssim\||{\bm{\xi}_{h}}|\|_{\mathcal{T}_{h}^{n}}\lesssim h^{k}+\tau^{k+\frac{1}{2}}. (62)

These yield |𝒆|𝑯1(Ω~n)hk+τk+12|{{\bm{e}}}|_{{\bm{H}}^{1}(\tilde{\Omega}^{n})}\lesssim h^{k}+\tau^{k+\frac{1}{2}}. Since 𝑿~n,n1𝑾k,(Ω~n)𝑿~n,n1𝑯r+12(Ω~n)\|{\tilde{\bm{X}}^{n,n-1}}\|_{{\bm{W}}^{k,\infty}(\tilde{\Omega}^{n})}\lesssim\|{\tilde{\bm{X}}^{n,n-1}}\|_{{\bm{H}}^{r+\frac{1}{2}}(\tilde{\Omega}^{n})}, inverse estimates imply

𝒆𝑾1,(Ω~n)h1𝝃h𝑯1(Ω~n)+hk1𝑿~n,n1𝑯r+12(Ω~n)hk1+τk12.\displaystyle\|{{\bm{e}}}\|_{{\bm{W}}^{1,\infty}(\tilde{\Omega}^{n})}\lesssim h^{-1}\|{\bm{\xi}_{h}}\|_{{\bm{H}}^{1}(\tilde{\Omega}^{n})}+h^{k-1}\|{\tilde{\bm{X}}^{n,n-1}}\|_{{\bm{H}}^{r+\frac{1}{2}}(\tilde{\Omega}^{n})}\lesssim h^{k-1}+\tau^{k-\frac{1}{2}}. (63)

Let 𝕁hn,n1=𝒙𝑿~hn,n1\mathbb{J}_{h}^{n,n-1}=\partial_{{\bm{x}}}\tilde{\bm{X}}_{h}^{n,n-1}. From (63) and (53), we have

𝕁hn,n1𝕀𝑳(Ω~n)\displaystyle\|{\mathbb{J}_{h}^{n,n-1}-\mathbb{I}}\|_{{\bm{L}}^{\infty}(\tilde{\Omega}^{n})} 𝕁hn,n1𝕁~n,n1𝑳(Ω~n)+𝕁~n,n1𝕀𝑳(Ω~n)𝒆𝑾1,(Ω~n)+τ𝑿n,n1𝑯r+12(Ωn)τ.\displaystyle\leq\|{\mathbb{J}_{h}^{n,n-1}-\tilde{\mathbb{J}}^{n,n-1}}\|_{{\bm{L}}^{\infty}(\tilde{\Omega}^{n})}+\|{\tilde{\mathbb{J}}^{n,n-1}-\mathbb{I}}\|_{{\bm{L}}^{\infty}(\tilde{\Omega}^{n})}\leq\|{{\bm{e}}}\|_{{\bm{W}}^{1,\infty}(\tilde{\Omega}^{n})}+\tau\|{{\bm{X}}^{n,n-1}}\|_{{\bm{H}}^{r+\frac{1}{2}}(\Omega^{n})}\lesssim\tau.

From Sobolev’s inequality and inverse estimates, for any vhH1(Ω~n)v_{h}\in H^{1}(\tilde{\Omega}^{n}), we get

vhL(Ω~n)vhW1,4(Ω~n)h12vhH1(Ω~n).\left\|{v_{h}}\right\|_{L^{\infty}({\tilde{\Omega}^{n}})}\lesssim\left\|{v_{h}}\right\|_{W^{1,4}(\tilde{\Omega}^{n})}\lesssim h^{-\frac{1}{2}}{\left\|{v_{h}}\right\|}_{H^{1}({\tilde{\Omega}^{n}})}. (64)

Moreover, it yields from (57), (62), (64) that

𝒆𝑳(Ω~n)\displaystyle\left\|{{\bm{e}}}\right\|_{{\bm{L}}^{\infty}(\tilde{\Omega}^{n})} 𝝃h𝑳(Ω~n)+𝜼𝑳(Ω~n)h12𝝃hH1(Ω~n)+hk𝑿n,n1𝑾k,(Ω~n)\displaystyle\leq\left\|{\bm{\xi}_{h}}\right\|_{{\bm{L}}^{\infty}(\tilde{\Omega}^{n})}+\left\|{\bm{\eta}}\right\|_{{\bm{L}}^{\infty}(\tilde{\Omega}^{n})}\lesssim h^{-\frac{1}{2}}{\left\|{\bm{\xi}_{h}}\right\|}_{H^{1}({\tilde{\Omega}^{n}})}+h^{k}\left\|{{\bm{X}}^{n,n-1}}\right\|_{{\bm{W}}^{k,\infty}(\tilde{\Omega}^{n})}
h12|𝝃h|𝒯hn+hkhk12+h12τk+12.\displaystyle\lesssim h^{-\frac{1}{2}}\left\|\left|{\bm{\xi}_{h}}\right|\right\|_{\mathcal{T}_{h}^{n}}+h^{k}\lesssim h^{k-\frac{1}{2}}+h^{-\frac{1}{2}}\tau^{k+\frac{1}{2}}.

The proof is finished by assuming h=O(τ)h=O(\tau). ∎

Corollary B.2.

Let 𝐰k,hn{\bm{w}}^{n}_{k,h} be defined in (24) and write 𝕁hn,ni:=𝐱𝐗hn,ni\mathbb{J}_{h}^{n,n-i}:=\partial_{{\bm{x}}}{\bm{X}}^{n,n-i}_{h} and 𝕁hni,n:=(𝕁hn,ni)1\mathbb{J}_{h}^{n-i,n}:=(\mathbb{J}_{h}^{n,n-i})^{-1} for 1ik1\leq i\leq k. Then 𝐰k,hn𝐋(Ωηn)𝐰kn𝐇r+1/2(Ωn)\|{\bm{w}}_{k,h}^{n}\|_{{\bm{L}}^{\infty}(\Omega_{\eta}^{n})}\lesssim\left\|{{\bm{w}}^{n}_{k}}\right\|_{{\bm{H}}^{r+1/2}(\Omega^{n})} and

𝕁hn,ni𝕀𝕃(Ω~n)τ,det(𝕁hn,ni)=1+O(τ),\displaystyle\|\mathbb{J}_{h}^{n,n-i}-\mathbb{I}\|_{\mathbb{L}^{\infty}(\tilde{\Omega}^{n})}\lesssim\tau,\hskip 54.06023pt\operatorname{det}(\mathbb{J}_{h}^{n,n-i})=1+O(\tau), (65)
𝕁hni,n𝕀𝕃(𝑿hn,ni(Ωηn))τ,det(𝕁hni,n)=1+O(τ).\displaystyle\|{\mathbb{J}_{h}^{n-i,n}-\mathbb{I}}\|_{\mathbb{L}^{\infty}({\bm{X}}_{h}^{n,n-i}(\Omega_{\eta}^{n}))}\lesssim\tau,\hskip 34.1433pt\operatorname{det}(\mathbb{J}_{h}^{n-i,n})=1+O(\tau). (66)
Proof.

Thanks to (13) and Lemma B.1, we have 𝑿hn,nj(Ωηn)Ω~nj{\bm{X}}_{h}^{n,n-j}(\Omega_{\eta}^{n})\subset\tilde{\Omega}^{n-j} for 1j<k1\leq j<k. So (25) is well defined. The proofs of (65)–(66) are direct consequences of (54). The stability of Sovolev extension shows that 𝒘~kn=1τi=0kλik𝑿~n,ni\tilde{\bm{w}}^{n}_{k}=\frac{1}{\tau}\sum\limits_{i=0}^{k}\lambda_{i}^{k}\tilde{\bm{X}}^{n,n-i} satisfies

𝒘~kn|Ωn=𝒘kn,𝒘~kn𝑯r+1/2(Ω~n)𝒘kn𝑯r+1/2(Ωn).\tilde{\bm{w}}^{n}_{k}|_{\Omega^{n}}={\bm{w}}^{n}_{k},\qquad\|\tilde{\bm{w}}^{n}_{k}\|_{{\bm{H}}^{r+1/2}(\tilde{\Omega}^{n})}\lesssim\|{\bm{w}}^{n}_{k}\|_{{\bm{H}}^{r+1/2}(\Omega^{n})}.

Using Lemma B.1, we have

𝒘k,hn𝒘~kn𝑳(Ωηn)τ1i=0k𝑿hn,ni𝑿~n,ni𝑳(Ωηn)τk3/2.\|{\bm{w}}_{k,h}^{n}-\tilde{\bm{w}}^{n}_{k}\|_{{\bm{L}}^{\infty}(\Omega_{\eta}^{n})}\lesssim\tau^{-1}\sum_{i=0}^{k}\|{\bm{X}}_{h}^{n,n-i}-\tilde{\bm{X}}^{n,n-i}\|_{{\bm{L}}^{\infty}(\Omega_{\eta}^{n})}\lesssim\tau^{k-3/2}. (67)

The proof is finished by the triangle inequality. ∎

Finally, we present estimates of the discrete ALE mapping by using Lemma B.1 and Corollary B.2.

Lemma B.3.

Assuming 𝐗hn,n1{\bm{X}}_{h}^{n,n-1} satisfies (54) and τ=O(h)\tau=O(h), there exits a constant CC independent of n,τn,\tau such that, for any vhV(k,𝒯hni)v_{h}\in V(k,\mathcal{T}_{h}^{n-i}), and μ=0,1\mu=0,1,

μ(vh𝑿hn,ni)L2(Ωηn)2\displaystyle\|{\nabla^{\mu}\big{(}v_{h}\circ{\bm{X}}_{h}^{n,n-i}\big{)}}\|^{2}_{L^{2}(\Omega_{\eta}^{n})} μ(vh𝑿hnl,ni)L2(Ωηnl)2+ChμvhL2(Ω~ni)2,\displaystyle\leq\;\|{\nabla^{\mu}(v_{h}\circ{\bm{X}}_{h}^{n-l,n-i})}\|^{2}_{L^{2}(\Omega_{\eta}^{n-l})}+Ch\|\nabla^{\mu}v_{h}\|_{L^{2}(\tilde{\Omega}^{n-i})}^{2}, (68)
vh𝑿hn,n1L2(Γηn)2\displaystyle\|{v_{h}\circ{\bm{X}}_{h}^{n,n-1}}\|^{2}_{L^{2}(\Gamma_{\eta}^{n})} (1+Cτ)vhL2(Γηn1)2+Cτk1vhH1(Ω~n1)2,\displaystyle\leq(1+C\tau)\left\|{v_{h}}\right\|_{L^{2}({\Gamma_{\eta}^{n-1}})}^{2}+C\tau^{k-1}{\left\|{v_{h}}\right\|}_{H^{1}({\tilde{\Omega}^{n-1}})}^{2}, (69)
|vh𝑿hn,n1|H1(Γηn)2\displaystyle\big{|}{v_{h}\circ{\bm{X}}_{h}^{n,n-1}}\big{|}^{2}_{H^{1}(\Gamma_{\eta}^{n})} h1|vh|H1(Ω~n1)2.\displaystyle\lesssim h^{-1}{\left|{v_{h}}\right|}_{H^{1}({\tilde{\Omega}^{n-1}})}^{2}. (70)

Appendix C The stability of numerical solutions

Since the computational domain is time-varying, we have to deal with the issue that uhnjV(k,𝒯hn)u^{n-j}_{h}\notin V(k,\mathcal{T}^{n}_{h}) for 1jk1\leq j\leq k when proving the stability and convergence of the discrete solution. To overcome this difficulty, we introduce a modified Ritz projection operator 𝒫hn:Y(Ωηn)V(k,𝒯hn)\mathcal{P}^{n}_{h}:Y(\Omega^{n}_{\eta})\to V(k,\mathcal{T}^{n}_{h}) proposed in [21] which satisfies

𝒜hn(𝒫hnw,vh)=ahn(w,vh)vhV(k,𝒯hn),\displaystyle\mathscr{A}^{n}_{h}(\mathcal{P}^{n}_{h}w,v_{h})=a^{n}_{h}(w,v_{h})\qquad\forall\,v_{h}\in V(k,\mathcal{T}^{n}_{h}), (71)

where Y(Ωηn):={vH1(Ωηn):|v|Ωηn<}Y(\Omega_{\eta}^{n}):=\big{\{}v\in H^{1}({\Omega^{n}_{\eta}}):\left\|\left|{v}\right|\right\|_{\Omega^{n}_{\eta}}<\infty\big{\}}. It satisfies

|𝒫hnw|Ωηn|w|Ωηn,w𝒫hnwL2(Ωηn)h|w|Ωηn,wY(Ωηn).\displaystyle\left\|\left|{\mathcal{P}_{h}^{n}w}\right|\right\|_{\Omega_{\eta}^{n}}\lesssim\left\|\left|{w}\right|\right\|_{\Omega_{\eta}^{n}},\qquad\left\|{w-\mathcal{P}^{n}_{h}w}\right\|_{L^{2}({\Omega^{n}_{\eta}})}\lesssim h\left\|\left|{w}\right|\right\|_{\Omega^{n}_{\eta}},\quad\forall w\in Y(\Omega^{n}_{\eta}). (72)

C.1 Proof of Theorem 3.9

Proof.

We only prove the theorem for k=4k=4. The proofs for other cases are similar. The rest of the proof is parallel to the proof of [21, Theorem 5.2], we just sketch the proof and omit details.

Step1: Selection of a particular test function. Write U~hn1,n=𝒫hn(Uhn1,n)\tilde{U}^{n-1,n}_{h}=\mathcal{P}^{n}_{h}(U^{n-1,n}_{h}). Since Uhn1,nV(k,𝒯hn)U^{n-1,n}_{h}\notin V(k,\mathcal{T}^{n}_{h}), we choose vh=2uhnU~hn1,nv_{h}=2u^{n}_{h}-\tilde{U}^{n-1,n}_{h} as a test function, the discrete problem for k=4k=4 has the form

i=15𝚿i4(𝑼hn¯)L2(Ωηn)2i=14𝚽i4(𝑼hn¯)L2(Ωηn)2+τ𝒜hn(uhn,vh)+τhn(uhn,vh)=τR1n+R2n,\sum_{i=1}^{5}\|\bm{\Psi}_{i}^{4}(\underline{{\bm{U}}_{h}^{n}})\|_{L^{2}(\Omega^{n}_{\eta})}^{2}-\sum_{i=1}^{4}\|\bm{\Phi}_{i}^{4}(\underline{{\bm{U}}_{h}^{n}})\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+\tau\mathscr{A}^{n}_{h}(u_{h}^{n},v_{h})+\tau\mathscr{B}_{h}^{n}(u_{h}^{n},v_{h})=\tau R^{n}_{1}+R^{n}_{2}, (73)

where hn(uhn,vh)=(𝒘k,hnuhn,vh)Ωηn\mathscr{B}_{h}^{n}(u_{h}^{n},v_{h})=(-\nabla{\bm{w}}_{k,h}^{n}\cdot\nabla u_{h}^{n},v_{h})_{\Omega_{\eta}^{n}}, we have used [18, Section 2 and Appendix A], and

𝚿i4(𝑼hn¯)=j=1ici,j4Uhn+1j,n,𝚽i4(𝑼hn¯)=j=1ici,j4Uhnj,n,\displaystyle\bm{\Psi}^{4}_{i}\big{(}\underline{{\bm{U}}^{n}_{h}}\big{)}=\sum_{j=1}^{i}c_{i,j}^{4}U_{h}^{n+1-j,n},\quad\bm{\Phi}_{i}^{4}\big{(}\underline{{\bm{U}}^{n}_{h}}\big{)}=\sum_{j=1}^{i}c_{i,j}^{4}U_{h}^{n-j,n},
R1n=(fn,2uhnU~hn1,n)Ωηn,R2n=(Λ4𝑼hn¯,U~hn1,nUhn1,n)Ωηn.\displaystyle R^{n}_{1}=\big{(}f^{n},2u^{n}_{h}-\tilde{U}^{n-1,n}_{h}\big{)}_{\Omega^{n}_{\eta}},\quad R^{n}_{2}=\big{(}\Lambda^{4}\underline{{\bm{U}}_{h}^{n}},\tilde{U}^{n-1,n}_{h}-U^{n-1,n}_{h}\big{)}_{\Omega^{n}_{\eta}}.

The trace inequality and inverse estimates show 𝒏uhnL2(Γηn)h12|uhn|H1(Ω~n)\|\partial_{\bm{n}}u^{n}_{h}\|_{L^{2}(\Gamma^{n}_{\eta})}\lesssim h^{-\frac{1}{2}}|u^{n}_{h}|_{H^{1}(\tilde{\Omega}^{n})}. For any ε(0,1)\varepsilon\in(0,1), by the definitions of 𝒫hn\mathcal{P}^{n}_{h} and the Cauchy-Schwarz inequality, one has

𝒜hn(uhn, 2uhnU~hn1,n)= 2𝒜hn(uhn,uhn)ahn(uhn,Uhn1,n)\displaystyle\mathscr{A}^{n}_{h}(u_{h}^{n},\,2u_{h}^{n}-\tilde{U}^{n-1,n}_{h})=\,2\mathscr{A}^{n}_{h}(u_{h}^{n},u_{h}^{n})-a^{n}_{h}(u_{h}^{n},U^{n-1,n}_{h})
32|uhn|𝒯hn252εhuhnL2(Γηn)2Cε|uhn|H1(Ω~n)212|Uhn1,n|H1(Ωηn)2εh2|Uhn1,n|H1(Γηn)2γ0+ε12hUhn1,nL2(Γηn)2.\displaystyle\,\geq\frac{3}{2}\left\|\left|{u_{h}^{n}}\right|\right\|^{2}_{\mathcal{T}_{h}^{n}}-\frac{5}{2\varepsilon h}\|{u^{n}_{h}}\|^{2}_{L^{2}(\Gamma^{n}_{\eta})}-C\varepsilon|{u^{n}_{h}}|^{2}_{H^{1}(\tilde{\Omega}^{n})}-\frac{1}{2}|{U^{n-1,n}_{h}}|^{2}_{H^{1}(\Omega^{n}_{\eta})}-\frac{\varepsilon h}{2}|{U_{h}^{n-1,n}}|^{2}_{H^{1}(\Gamma_{\eta}^{n})}-\frac{\gamma_{0}+\varepsilon^{-1}}{2h}\|{U^{n-1,n}_{h}}\|^{2}_{L^{2}(\Gamma^{n}_{\eta})}. (74)

Since 𝒘k,hn{\bm{w}}_{k,h}^{n} is bounded with Cw=𝒘k,hn𝑳(Ωηn)C_{w}=\|{\bm{w}}_{k,h}^{n}\|_{{\bm{L}}^{\infty}(\Omega_{\eta}^{n})} and Uhn1,nY(Ωηn)U^{n-1,n}_{h}\in Y(\Omega^{n}_{\eta}), we infer from (72) and the Cauchy-Schwarz inequality that

hn(uhn,vh)\displaystyle\mathscr{B}_{h}^{n}(u_{h}^{n},v_{h})\leq\, |uhn|H1(Ωηn)24+8Cw2(uhnL2(Ωηn)2+Uhn1,nL2(Ωηn)2+Ch2|Uhn1,n|Ωηn2),\displaystyle\frac{|{u_{h}^{n}}|^{2}_{H^{1}(\Omega_{\eta}^{n})}}{4}+8C_{w}^{2}\big{(}\|{u_{h}^{n}}\|^{2}_{L^{2}(\Omega_{\eta}^{n})}+\|{U_{h}^{n-1,n}}\|^{2}_{L^{2}(\Omega_{\eta}^{n})}+Ch^{2}\||{U_{h}^{n-1,n}}\||_{\Omega_{\eta}^{n}}^{2}\big{)},
R1n\displaystyle R^{n}_{1}\leq\, 2fnL2(Ωηn)2+uhnL2(Ωηn)2+Uhn1,nL2(Ωηn)2+Ch2|Uhn1,n|Ωηn2,\displaystyle 2\|{f^{n}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+\|{u^{n}_{h}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+\|{U^{n-1,n}_{h}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+Ch^{2}\||{U^{n-1,n}_{h}}\||_{\Omega^{n}_{\eta}}^{2},
R2n\displaystyle R^{n}_{2}\leq\, Cτε1j=04Uhnj,nL2(Ωηn)2+τε4|Uhn1,n|Ωηn2.\displaystyle C\tau\varepsilon^{-1}\sum_{j=0}^{4}\|{U^{n-j,n}_{h}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+\frac{\tau\varepsilon}{4}\||{U^{n-1,n}_{h}}\||_{\Omega^{n}_{\eta}}^{2}.

Step2: Estimate of Φi4(Uhn¯)\Phi_{i}^{4}(\underline{{\bm{U}}_{h}^{n}}). We deduce from the inverse estimates and Lemma B.3 that

μUhnj,n𝑳2(Ωηn)2\displaystyle\|{\nabla^{\mu}U^{n-j,n}_{h}}\|^{2}_{{\bm{L}}^{2}(\Omega^{n}_{\eta})}\leq\, μuhnj𝑳2(Ωηnj)2+Cτμuhnj𝑳2(Ω~nj)2,μ=0,1,\displaystyle\|{\nabla^{\mu}u^{n-j}_{h}}\|^{2}_{{\bm{L}}^{2}(\Omega^{n-j}_{\eta})}+C\tau\|{\nabla^{\mu}u^{n-j}_{h}}\|^{2}_{{\bm{L}}^{2}(\tilde{\Omega}^{n-j})},\quad\mu=0,1, (75)
Uhn1,nL2(Γηn)2\displaystyle\|{U^{n-1,n}_{h}}\|^{2}_{L^{2}(\Gamma^{n}_{\eta})}\leq\, (1+Cτ)uhn1L2(Γηn1)2+Cτ3uhn1H1(Ω~n1)2,\displaystyle(1+C\tau)\|{u^{n-1}_{h}}\|^{2}_{L^{2}(\Gamma^{n-1}_{\eta})}+C\tau^{3}\|{u^{n-1}_{h}}\|^{2}_{H^{1}(\tilde{\Omega}^{n-1})}, (76)
|Uhn1,n|H1(Γηn)\displaystyle|{U_{h}^{n-1,n}}|_{H^{1}(\Gamma_{\eta}^{n})}\lesssim\, h1/2uhn1H1(Ω~n1).\displaystyle h^{-1/2}\|{u_{h}^{n-1}}\|_{H^{1}(\tilde{\Omega}^{n-1})}. (77)

Note that 𝚽i4(𝑼hn¯)=𝚿i4(𝑼hn1¯)𝑿hn,n1\bm{\Phi}_{i}^{4}\big{(}\underline{{\bm{U}}_{h}^{n}}\big{)}=\bm{\Psi}_{i}^{4}\big{(}\underline{{\bm{U}}_{h}^{n-1}}\big{)}\circ{\bm{X}}_{h}^{n,n-1}. It is easy to see from Lemma B.3 that

𝚽i4(𝑼hn¯)L2(Ωηn)2𝚿i4(𝑼hn1¯)L2(Ωηn1)2+Cτj=1iuhnjL2(Ω~nj)2.\displaystyle\|{\bm{\Phi}_{i}^{4}(\underline{{\bm{U}}_{h}^{n}})}\|^{2}_{L^{2}(\Omega_{\eta}^{n})}\leq\,\|{\bm{\Psi}_{i}^{4}(\underline{{\bm{U}}_{h}^{n-1}})}\|^{2}_{L^{2}(\Omega_{\eta}^{n-1})}+C\tau\sum_{j=1}^{i}\|{u_{h}^{n-j}}\|^{2}_{L^{2}(\tilde{\Omega}^{n-j})}. (78)

Step3: Application of Gronwall’s inequality. Substituting (74)–(78) into (73), using O(h)=τhε1O(h)=\tau\leq h\ll\varepsilon\ll 1, Cw2hεC_{w}^{2}h\leq\varepsilon, applying (28), and taking the sum of the inequalities over 4nm4\leq n\leq m, we end up with

i=14𝚿i4(𝑼hm¯)L2(Ωηm)2+32τn=4m|uhn|𝒯hn2\displaystyle\sum_{i=1}^{4}\|{\bm{\Psi}_{i}^{4}(\underline{{\bm{U}}_{h}^{m}})}\|^{2}_{L^{2}(\Omega_{\eta}^{m})}+\frac{3}{2}\tau\sum_{n=4}^{m}\||{u_{h}^{n}}\||^{2}_{\mathcal{T}_{h}^{n}}
\displaystyle\leq\, n=4mτ(1+Cε)|uhn|𝒯hn2+n=4mτ[CuhnL2(Ωηn)2+2fnL2(Ωηn)2+(6+Cτ2εhγ02h)uhnL2(Γηn)2]+C0,\displaystyle\sum_{n=4}^{m}\tau(1+C\varepsilon)\||u_{h}^{n}\||_{\mathcal{T}_{h}^{n}}^{2}+\sum_{n=4}^{m}\tau\Big{[}C\|{u^{n}_{h}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+2\|{f^{n}}\|^{2}_{L^{2}(\Omega^{n}_{\eta})}+\big{(}\frac{6+C\tau}{2\varepsilon h}-\frac{\gamma_{0}}{2h}\big{)}\|u_{h}^{n}\|_{L^{2}(\Gamma_{\eta}^{n})}^{2}\Big{]}+C_{0},

after careful calculation, where C0C_{0} is related to the initial solutions, C0=Ci=03τ|uhi|𝒯hi2+i=14𝚿i4(𝑼h3¯)L2(Ωη3)2C_{0}=C\sum_{i=0}^{3}\tau\||{u^{i}_{h}}\||^{2}_{\mathcal{T}_{h}^{i}}+\sum_{i=1}^{4}\|{\bm{\Psi}_{i}^{4}(\underline{{\bm{U}}_{h}^{3}})}\|^{2}_{L^{2}(\Omega^{3}_{\eta})}. From [18, Table 2.2], we know 𝚿14(𝑼hm¯)=c1,14uhm\bm{\Psi}_{1}^{4}\big{(}\underline{{\bm{U}}_{h}^{m}}\big{)}=c^{4}_{1,1}u^{m}_{h} (c1,14>0c_{1,1}^{4}>0). Finally, we choose ε\varepsilon small enough such that Cε<1/2C\varepsilon<1/2 and γ0\gamma_{0} large enough such that γ0(6+Cτ)/ε\gamma_{0}\geq(6+C\tau)/\varepsilon. Then the proof is finished by using Gronwall’s inequality. ∎

References

  • [1] S. Adjerid and K. Moon, An immersed discontinuous Galerkin method for acoustic wave propagation in inhomogeneous media, SIAM J. Sci. Comput., 41 (2019), pp. A139–A162.
  • [2] A. M. Andrew, Level set methods and fast marching methods, Cambridge University Press, Cambridge, UK, 2 ed., 1999.
  • [3] U. M. Ascher and S. J. Ruuth, and B. T. Wetton, Implicit-explicit methods for time-dependent partial differential equations, SIAM J. Numer. Anal., 32(1995), pp. 797-823.
  • [4] E. Burman, S. Frei, and A. Massing, Eulerian time-stepping schemes for the non-stationary stokes equations on time-dependent domains, Numer. Math., 150 (2022), pp. 423–478.
  • [5] J. Donea, S. Giuliani, and J. P. Halleux, An arbitrary lagrangian-eulerian finite element method for transient dynamic fluid-structure interactions, Comput. Methods Appl. Mech. Engrg., 33 (1982), pp. 689–723.
  • [6] C. Farhat, P. Geuzaine, and C. Grandmont, The discrete geometric conservation law and the nonlinear stability of ALE schemes for the solution of flow problems on moving grids, J. Comput. Phys., 174 (2001), pp. 669–694.
  • [7] L. Formaggia and F. Nobile, A stability analysis for the arbitrary Lagrangian Eulerian formulation with finite elements, East-West J. Math., 7 (1999), pp. 105–132.
  • [8] T. P. Fries and A. Zilian, On time integration in the XFEM, Internat. J. Numer. Methods Engrg., 79 (2009), pp. 69–93.
  • [9] P. Geuzaine, C. Grandmont, and C. Farhat, Design and analysis of ALE schemes with provable second-order time-accuracy for inviscid and viscous flow simulations, J. Comput. Phys., 191 (2003), pp. 206–227.
  • [10] R. Guo, Solving parabolic moving interface problems with dynamical immersed spaces on unfitted meshes: fully discrete analysis, SIAM J. Numer. Anal., 59 (2021), pp. 797–828.
  • [11] J. Guzmán and M. Olshanskii, Inf-sup stability of geometrically unfitted Stokes finite elements, Math. Comp., 87 (2018), pp. 2091–2112.
  • [12] P. Hansbo, M. G. Larson, and S. Zahedi, Characteristic cut finite element methods for convection–diffusion problems on time dependent surfaces, Comput. Methods Appl. Mech. Engrg., 293 (2015), pp. 431–461.
  • [13] C. W. Hirt, A. A. Amsden, and J. L. Cook, An arbitrary lagrangian-eulerian computing method for all flow speeds, J. Comput. Phys., 14 (1974), pp. 227–253.
  • [14] T. J. Hughes, W. K. Liu, and T. K. Zimmermann, Lagrangian-Eulerian finite element formulation for incompressible viscous flows, Comput. Methods Appl. Mech. Engrg., 29 (1981), pp. 329–349.
  • [15] C. Lehrenfeld and M. Olshanskii, An Eulerian finite element method for PDEs in time-dependent domains, ESAIM Math. Model. Numer. Anal., 53 (2019), pp. 585–614.
  • [16] C. Lehrenfeld and A. Reusken, Analysis of a nitsche XFEM-DG discretization for a class of two-phase mass transport problems, SIAM J.Numer. Anal., 51 (2013), pp. 958–983.
  • [17] C. Y. Li, S. V. Garimella, and J. E. Simpson, Fixed-grid front-tracking algorithm for solidification problems, part I: Method and validation, Numerical Heat Transfer, Part B: Fundamentals, 43 (2003), pp. 117–141.
  • [18] J. Liu, Simple and efficient ALE methods with provable temporal accuracy up to fifth order for the Stokes equations on time varying domains, SIAM J. Numer. Anal., 51 (2013), pp. 743–772.
  • [19] Y. Lou and C. Lehrenfeld, Isoparametric unfitted BDF–finite element method for PDEs on evolving domains, SIAM J. Numer. Anal., 60 (2022), pp. 2069–2098.
  • [20] C. Ma, T. Tian, and W. Zheng, High-order unfitted characteristic finite element methods for moving interface problem of Oseen equations, J. Comput. Appl. Math., 425 (2023),  115028.
  • [21] C. Ma, Q. Zhang, and W. Zheng, A fourth-order unfitted characteristic finite element method for solving the advection-diffusion equation on time-varying domains, SIAM J. Numer. Anal., 60 (2022), pp. 2203–2224.
  • [22] C. Ma and W. Zheng, A fourth-order unfitted characteristic finite element method for free-boundary problems, J. Comput. Phys., 469 (2022),  111552.
  • [23] A. Masud and T. J. Hughes, A space-time Galerkin/least-squares finite element formulation of the Navier-Stokes equations for moving domain problems, Comput. Methods Appl. Mech. Engrg., 146 (1997), pp. 91–126.
  • [24] T. Richter, Fluid-structure interactions: models, analysis and finite elements, Springer, 2017.
  • [25] T. E. Tezduyar, M. Behr, S. Mittal, and J. Liou, A new strategy for finite element computations involving moving boundaries and interfaces—the deforming-spatial-domain/space-time procedure: II. computation of free-surface flows, two-liquid flows, and flows with drifting cylinders, Comput. Methods Appl. Mech. Engrg., 94 (1992), pp. 353–371.
  • [26] H. von Wahl and T. Richter, Error analysis for a parabolic PDE model problem on a coupled moving domain in a fully Eulerian framework, SIAM J. Numer. Anal., 61 (2023), pp. 286–314.
  • [27] H. von Wahl, T. Richter, and C. Lehrenfeld, An unfitted Eulerian finite element method for the time-dependent Stokes problem on moving domains, IMA J. Numer. Anal., 42 (2022), pp. 2505–2544.
  • [28] Q. Zhang, Fourth-and higher-order interface tracking via mapping and adjusting regular semianalytic sets represented by cubic splines, SIAM J. Sci. Comput., 40 (2018), pp. A3755–A3788.
  • [29] P. Zunino, Analysis of backward Euler/Extended finite element discretization of parabolic problems with moving interfaces, Comput. Methods Appl. Mech. Engrg., 258 (2013), pp. 152–165.