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

A numerical domain decomposition method for solving elliptic equations on manifolds

Shuhao Cao Division of Computing, Analytics, and Mathematics, School of Science and Engineering, University of Missouri-Kansas City, Kansas City, MO (). [email protected]    Lizhen Qin Mathematics Department, Nanjing University, Nanjing, Jiangsu, China (). [email protected]
Abstract

A new numerical domain decomposition method is proposed for solving elliptic equations on compact Riemannian manifolds. The advantage of this method is to avoid global triangulations or grids on manifolds. Our method is numerically tested on some 44-dimensional manifolds such as the unit sphere S4S^{4}, the complex projective space 2\mathbb{CP}^{2} and the product manifold S2×S2S^{2}\times S^{2}.

keywords:
Riemannian manifolds, elliptic problems, domain decomposition methods, finite element methods
\headers

DDM on ManifoldsS. Cao and L. Qin

{AMS}

Primary 65N30; Secondary 58J05, 65N55.

1 Introduction

Elliptic partial differential equations on Riemannian manifolds are of fundamental importance both in analysis and geometry (see e.g., [50, 32]). A simple and important example is

(1) Δu+bu=f.-\Delta u+bu=f.

Here Δ\Delta is the Laplace-Beltrami operator, or Laplacian for brevity, defined on a dd-dimensional Riemannian manifold MM. Many manifolds are naturally submanifolds of Euclidean spaces, here the “dimension” of a submanifold is referred to as the topological dimension of the manifold, not the dimension of its ambient Euclidean space. For example, the nn-dimensional unit sphere SnS^{n} is embedded in n+1\mathbb{R}^{n+1}.

When the manifold MM is a two-dimensional Riemannian submanifold of 3\mathbb{R}^{3}, i.e. a surface, the numerical methods to solve PDEs, particularly (1), on MM have been extensively studied for a long history (see e.g. [42, 41, 6, 19, 20, 22, 44]). Over several decades, among many others, the surface finite element method and its variant have had far-reaching developments (see e.g., [2, 4, 15, 16, 27, 30, 43, 47]), and applications to various PDEs (see e.g. [3, 7, 10, 21, 23, 31]); see also e.g. [14, 22, 8] for surveys and bibliographies, and e.g. [1, 5, 26] for software developments. They have also been widely applied to various areas such as computer graphics (e.g. [25]), surface fitting (e.g. [18]), shape analysis (e.g. [17, 48, 49]), isogeometric topology optimization (e.g. [33, 34]), and medical imaging (e.g. [38]). A basic and common feature of these finite element methods is to construct a global triangulation of MM, which provides a necessary grid structure for a global finite element space. This triangulation, following [39, Definition 8.3] in the sense of differential topology, is a bijection between a simplicial complex and the manifold MM, which satisfies certain regularity. In computation, a concrete representation of this triangulation is to approximate MM by a polyhedron in 3\mathbb{R}^{3}. Alternatively, MM is represented implicitly as a level set of a function ϕ\phi, say M=ϕ1(0)M=\phi^{-1}(0), and an equation on MM can be solved by the methods of trace elements or implicit surfaces elements (e.g. [43, 47, 22]).

However, there are several essential difficulties to apply the methods above to general higher dimensional manifolds. First, many important and interesting examples of higher dimensional manifolds are not submanifolds of Euclidean spaces by definition. A notable example is the complex projective spaces n\mathbb{CP}^{n}, which serve as the foundation for algebraic geometry. Though Whitney Embedding Theorem ([54], see also [29, p. 24-27]) and Nash Embedding Theorem ([40]) do reveal that every smooth manifold can be embedded into a Euclidean space k\mathbb{R}^{k} differential-topologically or even geometrically for some large kk. To the best of our knowledge, there is no literature on a universal algorithm to efficiently construct such an explicit embedding for computation on general higher dimensional manifolds. Second, assuming MM is a submanifold of k\mathbb{R}^{k}, if the codimension of the submanifold MM in k\mathbb{R}^{k} is greater than 11, which is often the case, it will be horribly difficult to find an effective polytopal approximation to MM in general due to topological and geometrical complexity, meanwhile, MM also cannot be represent as a level set of a function.

A global triangulation of a manifold is helpful in numerical computation, since it can provide a global discretization of the target problem. We do acknowledge that J. H. C. Whitehead proved ([53], see also [39, Theorem 10.6]) that every smooth manifold can be globally triangulated in an abstract way. However, to the best of our knowledge, in practice, there has been no algorithm to build such a concrete triangulation or a grid over a high-dimensional manifold in general. To circumvent this difficulty, in [46] which serves as our major inspiration, Qin–Zhang–Zhang proposed a new idea to numerically solve elliptic PDEs on manifolds by avoiding global triangulations completely. Since a dd-dimensional manifold MM has local coordinate charts by definition, MM can be decomposed into finitely many subdomains that carry local Cartesian coordinates. Consequently, an elliptic equation on each subdomain can be transformed to one on a domain in d\mathbb{R}^{d}. Thus an elliptic problem on MM can be assembled to either coupled problems on Euclidean domains, or can be solved directly by Domain Decomposition Methods (DDMs). This idea had been numerically verified on the 33-dimensional unit sphere S3S^{3} in [46], where S3S^{3} is decomposed into two subdomains. However, a major drawback of [46] is its lack of flexibility to deal with more general manifolds, whose charts may involve more than two subdomains.

In this paper, we shall develop the idea in [46] further to solve problems on general manifolds. Similar to [46], global triangulations or grids shall be completely avoided. In fact, we shall solve such problems by an overlapping domain decomposition method (DDM). The development of DDM has a long history. It was first invented by H. A. Schwarz [51]. The version of DDM we mainly follow was originally proposed by P. L. Lions in [36, I. 4] for solving continuous problems in Euclidean spaces. It has been well-developed and is later named as Multiplicative Schwarz Method (c.f. [52]). The later development usually takes this DDM as a preconditioner for a globally discretized problem (see e.g. [9, 24], and pure algebraic versions in e.g. [11, 35]). However, a globally discretized problem should be based on a global grid on a manifold MM, which is not accessible in our case. Therefore, we shall more closely follow Lions’ original approach rather than the later development. This original approach is a very simple iteration scheme such that a local problem in a subdomain is solved in each step. We found that this DDM can be well-adapted to solve problems on manifolds. As in [46], a problem in each subdomain can be converted to one in a Euclidean domain. It is simpler to obtain a grid and hence a discretization over each Euclidean domain. To minimize the difficulty of coding, these Euclidean domains can be even chosen as (high-dimensional) rectangles. The transition of information among these Euclidean domains is by virtue of the transition maps of coordinates.

Overall, our approach is a discrete imitation of Lions’ method with additional transition maps of coordinates. It is necessary to point out that the transition maps in the proposed method are not required to preserve nodes or grids. More precisely, in computations, suppose there are two Euclidean domains DiD_{i} and DjD_{j}. Both DiD_{i} and DjD_{j} carry a grid, respectively. If ϕji:DiDj\phi_{ji}:D_{i}\rightarrow D_{j} is such a transition map between DiD_{i} and DjD_{j}, the image of the grid on DiD_{i} under ϕji\phi_{ji} may not match the grid on DjD_{j}. Furthermore, for a node ξDi\xi\in D_{i}, ϕji(ξ)\phi_{ji}(\xi) may not be a node anymore on DjD_{j} (see Fig. 1 for an example). Nevertheless, this incompatibility shows the high flexibility of our approach. In fact, if all transition maps preserved grids, we would obtain a global grid on MM which is practically almost impossible. The transition maps in [46, Section 4] do not preserve grids but do preserve boundary nodes on each subdomain, i.e., the ϕji\phi_{ji} above maps nodes on Di\partial D_{i} to nodes of DjD_{j}. As a result, the methods therein have some flexibility to solve PDEs on S3S^{3}. In order to handle more general high-dimensional manifolds, this paper improves the flexibility further. Particularly, interpolation techniques are employed now to handle nonmatching grids.

The proposed approach is numerically verified on closed manifolds of dimension 44. They are 44-dimensional unit sphere S4S^{4}, the complex projective space 2\mathbb{CP}^{2}, and the product manifold S2×S2S^{2}\times S^{2}. The numerical results show that our method solves equation (1) on these manifolds in a natural manner.

The outline of this paper is as follows. In Section 2, we shall extend P. L. Lions’ method in [36] to a DDM for continuous problems on manifolds and prove its convergence. In Section 3, we shall propose our discrete imitation of Lions’ method. Some specialty of product manifolds will be explained in Section 4. Finally, some numerical results will be presented in Section 5.

2 Theory on Continuous Problems

In this section, we shall first formulate a second order elliptic model problem on general manifolds. Then we introduce a domain decomposition method (DDM) generalized from [36, I. 4] to solve the model problem (see Algorithm 1 below). We shall formulate and prove Theorem 2.1 below on the convergence of this iterative procedure. This procedure also motivates us to propose Algorithm 2 in next section, which gives numerical approximations to the solution to the model problem.

Let MM be a dd-dimensional compact smooth manifold without or with boundary M\partial M. Equipping MM with a Riemannian metric gg, the Laplacian Δ\Delta can be defined on MM. In general, neither gg nor Δ\Delta can be expressed by coordinates globally because MM does not necessarily have a global coordinate chart. In a local chart with coordinates (x1,,xd)(x_{1},\dots,x_{d}), the Riemannian metric tensor gg is expressed as

(2) g=α,β=1dgαβdxαdxβ,g=\sum_{\alpha,\beta=1}^{d}g_{\alpha\beta}\mathrm{d}x_{\alpha}\otimes\mathrm{d}x_{\beta},

where the matrix (gαβ)d×d(g_{\alpha\beta})_{d\times d} is symmetric and positive definite. The Laplacian Δ\Delta can be then expressed in this chart as

Δu=1Gα=1dxα(β=1dgαβGuxβ),\Delta u=\frac{1}{\sqrt{G}}\sum_{\alpha=1}^{d}\frac{\partial}{\partial x_{\alpha}}\left(\sum_{\beta=1}^{d}g^{\alpha\beta}\sqrt{G}\frac{\partial u}{\partial x_{\beta}}\right),

where G=det((gαβ)d×d)G=\det\left((g_{\alpha\beta})_{d\times d}\right) is the determinant of the matrix (gαβ)d×d(g_{\alpha\beta})_{d\times d} and (gαβ)d×d(g^{\alpha\beta})_{d\times d} is the inverse of (gαβ)d×d(g_{\alpha\beta})_{d\times d}. It is well-known that Δ\Delta is an elliptic differential operator of second order.

We consider the following model problem on MM

(3) {Δu+bu=f,u|M=0,\left\{\begin{aligned} -\Delta u+bu&=f,\\ u|_{\partial M}&=0,\end{aligned}\right.

where b0b\geq 0 is a constant and fL2(M)f\in L^{2}(M). A weak solution to (3) is a solution to the following problem: Find a uH01(M)u\in H^{1}_{0}(M) such that, vH01(M)\forall v\in H^{1}_{0}(M),

(4) M(u,v+buv)dvol=Mfvdvol.\int_{M}(\langle\nabla u,\nabla v\rangle+buv)\,\mathrm{dvol}=\int_{M}fv\,\mathrm{dvol}.

Here u\nabla u and v\nabla v are the gradients of uu and vv with respect to gg, u,v\langle\nabla u,\nabla v\rangle is the inner product of u\nabla u and v\nabla v, and dvol\mathrm{dvol} is the volume form. In terms of local coordinates,

(5) u,v=α,β=1dgαβuxαvxβ\langle\nabla u,\nabla v\rangle=\sum_{\alpha,\beta=1}^{d}g^{\alpha\beta}\frac{\partial u}{\partial x_{\alpha}}\frac{\partial v}{\partial x_{\beta}}

and

(6) dvol=Gdx1dxd.\mathrm{dvol}=\sqrt{G}\mathrm{d}x_{1}\cdots\mathrm{d}x_{d}.

For brevity, we shall omit the symbol dvol\mathrm{dvol} in integrals on MM throughout this paper.

Note that it suffices to solve (4) on each component of MM. Therefore, without loss of generality, MM is assumed to be connected. In addition, if M=\partial M=\emptyset, the condition above u|M=0u|_{\partial M}=0 is vacuously satisfied in regards to H01(M)=H1(M)H^{1}_{0}(M)=H^{1}(M). To guarantee (4) is well-posed, we further assume b>0b>0 if M=\partial M=\emptyset. (Actually, if b=0b=0, one may imposed additional conditions such as Mudvol=0\int_{M}u\mathrm{dvol}=0 to guarantee the well-posedness. On the other hand, the numerical algorithm would be more complicated in this situation. We shall study the case later.)

Now we describe a domain decomposition iterative procedure to solve (4). This method was originally proposed by P. L. Lions in [36, I. 4], in which the classical Schwarz Alternating Method is extended to an iterative procedure with many subdomains. The manifold nature of this method is intrinsically adapted to solve PDEs on manifolds. More precisely, suppose MM is decomposed into mm subdomains, i.e.

M=i=1mIntMi.M=\bigcup_{i=1}^{m}\mathrm{Int}M_{i}.

Here MiM_{i} is a closed subdomain (submanifold with codimension 0) of MM with Lipschitz boundary, and IntMi\mathrm{Int}M_{i} is the interior of MiM_{i} in the sense of point-set topology of MM. Clearly, an element in H01(Mi)H^{1}_{0}(M_{i}) can be naturally considered as an element in H01(M)H^{1}_{0}(M) by zero extension. Thus, H01(Mi)H01(M)H^{1}_{0}(M_{i})\subseteq H^{1}_{0}(M). The iterative procedure to solve (4) is the following Algorithm 1.

Algorithm 1 A DDM for the continuous problem.
1:Choose an arbitrary initial guess u0H01(M)u^{0}\in H^{1}_{0}(M) for (4).
2:For each n>0n>0, assuming un1u^{n-1} has been obtained, define un,0=un1u^{n,0}=u^{n-1}. For 1im1\leq i\leq m, assuming un,ju^{n,j} has been obtained for all j<ij<i, find a un,iH01(M)u^{n,i}\in H^{1}_{0}(M) such that
(7) {(vH01(Mi))Miun,i,v+bun,iv=Mifv,un,i|MIntMi=un,i1|MIntMi.\left\{\begin{array}[]{rcl}(\forall v\in H^{1}_{0}(M_{i}))\ \ \ \int_{M_{i}}\langle\nabla u^{n,i},\nabla v\rangle+bu^{n,i}v&=&\int_{M_{i}}fv,\\ u^{n,i}|_{M\setminus\mathrm{Int}M_{i}}&=&u^{n,i-1}|_{M\setminus\mathrm{Int}M_{i}}.\end{array}\right.
3:Let un=un,mu^{n}=u^{n,m}.

To obtain the un,iu^{n,i} in (7), one first solve an elliptic problem on MiM_{i} with Dirichlet boundary condition un,i|Mi=un,i1|Miu^{n,i}|_{\partial M_{i}}=u^{n,i-1}|_{\partial M_{i}}, then extend the solution to a function un,iu^{n,i} on MM by defining un,i|MMi=un,i1|MMiu^{n,i}|_{M\setminus M_{i}}=u^{n,i-1}|_{M\setminus M_{i}}. Thus the un,iu^{n,i} is well-defined and hence Algorithm 1 is well-posed.

We have the following theorem on the geometrical convergence of Algorithm 1.

Theorem 2.1.

There exist constants C0>0C_{0}>0 and L[0,1)L\in[0,1) such that, u0H01(M)\forall u^{0}\in H^{1}_{0}(M), n>0\forall n>0,

uunH01(M)C0Lnuu0H01(M),\|u-u^{n}\|_{H^{1}_{0}(M)}\leq C_{0}L^{n}\|u-u^{0}\|_{H^{1}_{0}(M)},

where uu is the solution to (4) and unu^{n} is the nnth iterated approximation in Algorithm 1 with initial guess u0u^{0}.

Theorem 2.1 was originally proved by P. L. Lions ([36, Theorem. I.2]) in the case that MM is a Euclidean domain. We shall adapt his proof to the case of manifolds.

An elementary proof of the following lemma can be found in [36, p. 17]. It is also an immediate corollary of [56, (1.2)].

Lemma 2.2.

Suppose VV is a Hilbert space and ViV_{i} (1im1\leq i\leq m) are closed subspaces of VV such that V=i=1mViV=\sum_{i=1}^{m}V_{i}. Then

PVmPVm1PV1<1,\|P_{V_{m}^{\bot}}P_{V_{m-1}^{\bot}}\cdots P_{V_{1}^{\bot}}\|<1,

where each ViV_{i}^{\bot} is the orthogonal complement of ViV_{i}, and PViP_{V_{i}^{\bot}} is the orthogonal projection of VV onto ViV_{i}^{\bot}.

Define the energy bilinear form on H01(M)H^{1}_{0}(M) as

a(w,v)=Mw,v+bwv.a(w,v)=\int_{M}\langle\nabla w,\nabla v\rangle+bwv.
Proof 2.3 (Proof of Theorem 2.1).

It’s easy to see that the energy norm a(,)12a(\cdot,\cdot)^{\frac{1}{2}} is equivalent to the original H1H_{1}-norm on H01(M)H^{1}_{0}(M). In other words, there are positive constants C1C_{1} and C2C_{2} such that C1a(,)12H01(M)C2a(,)12C_{1}a(\cdot,\cdot)^{\frac{1}{2}}\leq\|\cdot\|_{H^{1}_{0}(M)}\leq C_{2}a(\cdot,\cdot)^{\frac{1}{2}}. Thus it suffices to show that: there exists L[0,1)L\in[0,1) such that, n>0\forall n>0,

(8) a(uun,uun)12Lna(uu0,uu0)12.a(u-u^{n},u-u^{n})^{\frac{1}{2}}\leq L^{n}a(u-u^{0},u-u^{0})^{\frac{1}{2}}.

Actually, we will have uunH01(M)C0Lnuu0H01(M)\|u-u^{n}\|_{H^{1}_{0}(M)}\leq C_{0}L^{n}\|u-u^{0}\|_{H^{1}_{0}(M)} with C0=C11C2C_{0}=C_{1}^{-1}C_{2}. (See also [45, Lemma 4.4] and its proof.) Let ViV_{i} denote H01(Mi)H^{1}_{0}(M_{i}). By adapting the argument in [36, Theorem. I.2], we obtain

uun,i=PVi(uun,i1),u-u^{n,i}=P_{V_{i}^{\bot}}(u-u^{n,i-1}),

where PViP_{V_{i}^{\bot}} is the orthogonal projection onto ViV_{i}^{\bot} with respect to a(,)a(\cdot,\cdot). Therefore

(9) uun\displaystyle u-u^{n} =\displaystyle= uun,m=PVm(uun,m1)==PVmPV1(uun,0)\displaystyle u-u^{n,m}=P_{V_{m}^{\bot}}(u-u^{n,m-1})=\cdots=P_{V_{m}^{\bot}}\cdots P_{V_{1}^{\bot}}(u-u^{n,0})
=\displaystyle= PVmPV1(uun1)=(PVmPV1)n(uu0).\displaystyle P_{V_{m}^{\bot}}\cdots P_{V_{1}^{\bot}}(u-u^{n-1})=(P_{V_{m}^{\bot}}\cdots P_{V_{1}^{\bot}})^{n}(u-u^{0}).

By Lemma 2.2, we have

PVmPV1<1\|P_{V_{m}^{\bot}}\cdots P_{V_{1}^{\bot}}\|<1

with respect to a(,)a(\cdot,\cdot). Define L=PVmPV1L=\|P_{V_{m}^{\bot}}\cdots P_{V_{1}^{\bot}}\|, then L[0,1)L\in[0,1) and (9) implies (8) which finishes the proof.

3 Numerical Scheme

In this section, we propose a numerical DDM iterative procedure (Algorithm 2 below) to obtain approximations to the solution of (4). The procedure is as follow. First, MM is decomposed into overlapping subdomains MiM_{i} (1im1\leq i\leq m), and each MiM_{i} is in a coordinate chart. Second, a DDM iterative procedure, which serves as a discrete counterpart of Algorithm 1 is applied. Due to MiM_{i} being a coordinate chart, an elliptic problem on MiM_{i} can be naturally converted to one on a domain in a Euclidean space. Then, this problem on MiM_{i} can be solved approximately using conventional finite element methods. The transition of information among subdomains is by interpolation.

For the purpose of presentation, only manifolds without boundary are considered in the numerical examples. A forthcoming work will study in detail the numerical implementation on manifolds with boundaries. In that more general case, we shall have to apply some special technique to deal with the boundary.

3.1 Finite Element Spaces over a d-Rectangle

Suppose a manifold MM has dimension dd. As indicated above, each subdomain MiM_{i} of MM shall be converted to a domain DidD_{i}\subset\mathbb{R}^{d}. This conversion substantially reduces the difficulty of the numerical scheme in consideration. However, when d>3d>3, the construction of a discretized problem on DiD_{i} still remains a difficult task. The main reason is that the geometric intuition used in implementing finite element spaces in 2\mathbb{R}^{2} or 3\mathbb{R}^{3} cannot be simply ported to higher dimensions. To minimize the difficulty of tessellation in higher dimensions, we shall choose DiD_{i} as a dd-rectangle and use the tensor product-type finite element space of dd-rectangles (see e.g., [12, p. 56-64]).

Recall that a dd-rectangle DD is

D=i=1d[ai,bi]={(x1,,xd)i,xi[ai,bi]}.D=\prod_{i=1}^{d}[a_{i},b_{i}]=\{(x_{1},\cdots,x_{d})\mid\forall i,x_{i}\in[a_{i},b_{i}]\}.

We refine each coordinate factor interval [ai,bi][a_{i},b_{i}] by adding points of partition:

ai=ci,0<ci,1<<ci,Ni=bi.a_{i}=c_{i,0}<c_{i,1}<\cdots<c_{i,N_{i}}=b_{i}.

Then [ai,bi][a_{i},b_{i}] is divided into NiN_{i} subintervals. Define a function φi,j\varphi_{i,j} on [ai,bi][a_{i},b_{i}] for 0jNi0\leq j\leq N_{i} as

(10) φi,j(xi)={xici,j1ci,jci,j1,xi[ci,j1,ci,j];xici,j+1ci,jci,j+1,xi[ci,j,ci,j+1];0,otherwise.\varphi_{i,j}(x_{i})=\begin{cases}\frac{x_{i}-c_{i,j-1}}{c_{i,j}-c_{i,j-1}},&x_{i}\in[c_{i,j-1},c_{i,j}];\\ \frac{x_{i}-c_{i,j+1}}{c_{i,j}-c_{i,j+1}},&x_{i}\in[c_{i,j},c_{i,j+1}];\\ 0,&\text{otherwise}.\end{cases}

Here φi,j(xi)\varphi_{i,j}(x_{i}) is undefined for xi<cjx_{i}<c_{j} (resp. xi>cjx_{i}>c_{j}) when j=0j=0 (resp. j=Nij=N_{i}). Clearly, φi,j\varphi_{i,j} is piecewise linear such that φi,j(ci,j)=1\varphi_{i,j}(c_{i,j})=1 and φi,j(ci,t)=0\varphi_{i,j}(c_{i,t})=0 for tjt\neq j.

The refinement of all such [ai,bi][a_{i},b_{i}] provides a grid on DD. This divides DD into i=1dNi\prod_{i=1}^{d}N_{i} many small dd-rectangles

(11) i=1d[ci,ti1,ci,ti],\prod_{i=1}^{d}[c_{i,t_{i}-1},c_{i,t_{i}}],

where 1tiNi1\leq t_{i}\leq N_{i} for all ii. Each small dd-rectangle (11) is an element of the grid. A vertex of (11) is a node of the grid which is of the form

ξ=(c1,j1,c2,j2,,cd,jd),\xi=(c_{1,j_{1}},c_{2,j_{2}},\cdots,c_{d,j_{d}}),

where 0jiNi0\leq j_{i}\leq N_{i} for all ii. Let WhW_{h} be the finite element space of dd-rectangles of type (11) (see [12, p. 57]). A base function in WhW_{h} associated with the node ξ\xi is

φξ(x1,,xd)=i=1dφi,ji(xi),\varphi_{\xi}(x_{1},\cdots,x_{d})=\prod_{i=1}^{d}\varphi_{i,j_{i}}(x_{i}),

where φi,ji\varphi_{i,j_{i}} is the one in (10).

Since DD is a dd-rectangle, it is relatively easy to handle the finite element space of dd-rectangles. This advantage had been indicated in [12, p. 62]

3.2 Discrete Iterative Procedure

Let MM be a dd-dimensional compact Riemannian manifold without boundary. We try to find numerical approximations to the solution to the problem (4) on MM.

Suppose M=i=1mIntMiM=\bigcup_{i=1}^{m}\mathrm{Int}M_{i}, and there is a smooth diffeomorphism ϕi:DiMi\phi_{i}:D_{i}\rightarrow M_{i} for each ii, where DiD_{i} is a dd-rectangle in d\mathbb{R}^{d}. Theoretically, we can always get such triples (Mi,Di,ϕi)(M_{i},D_{i},\phi_{i}). Actually, for each ζM\zeta\in M, there is an open chart neighborhood UζU_{\zeta} of ζ\zeta, i.e. there is a diffeomorphism ϕζ:ΩζUζ\phi_{\zeta}:\Omega_{\zeta}\rightarrow U_{\zeta}, where Ωζ\Omega_{\zeta} is an open subset of d\mathbb{R}^{d}. Since ϕζ1(ζ)\phi_{\zeta}^{-1}(\zeta) is an interior point of Ωζ\Omega_{\zeta}, we can choose a rectangular neighborhood DζD_{\zeta} of ϕζ1(ζ)\phi_{\zeta}^{-1}(\zeta) such that DζΩζD_{\zeta}\subseteq\Omega_{\zeta}. This yields a diffeomorphism ϕζ:DζMζUζ\phi_{\zeta}:D_{\zeta}\rightarrow M_{\zeta}\subset U_{\zeta}, where MζM_{\zeta} is a neighborhood of ζ\zeta. The interiors of all such MζM_{\zeta} provide an open covering of MM. Since it is compact, MM can be covered by the interiors of finitely many such MζM_{\zeta}. These finitely many (Mζ,Dζ,ϕζ)(M_{\zeta},D_{\zeta},\phi_{\zeta}) yield a desired decomposition of MM.

Let (x1,,xd)(x_{1},\dots,x_{d}) denote the coordinates on DiD_{i}. Then the Riemannian metric gg on MiM_{i} can be expressed as (2). Define an energy bilinear form on H1(Di)H^{1}(D_{i}) as

(12) ai(w,v)=Di(α,β=1dgαβwxαvxβ+bwv)Gdx1dxd.a_{i}(w,v)=\int_{D_{i}}\left(\sum_{\alpha,\beta=1}^{d}g^{\alpha\beta}\frac{\partial w}{\partial x_{\alpha}}\frac{\partial v}{\partial x_{\beta}}+bwv\right)\sqrt{G}\mathrm{d}x_{1}\cdots\mathrm{d}x_{d}.

Define a bilinear form (,)i(\cdot,\cdot)_{i} on L2(Di)L^{2}(D_{i}) as

(w,v)i=DiwvGdx1dxd.(w,v)_{i}=\int_{D_{i}}wv\sqrt{G}\mathrm{d}x_{1}\cdots\mathrm{d}x_{d}.

By (5) and (6), the first line of (7) is converted to the following equation: vH01(Di)\forall v\in H^{1}_{0}(D_{i}),

ai(un,iϕi,v)=(fϕi,v)i.a_{i}(u^{n,i}\circ\phi_{i},v)=(f\circ\phi_{i},v)_{i}.

Create a grid of dd-rectangles over DiD_{i}. Let VhiV_{h}^{i} be the finite element space of dd-rectangles of type (1)(1) over DiD_{i}. A discrete imitation of the first line of (7) would be: find a uhn,iVhiu_{h}^{n,i}\in V_{h}^{i} such that, vhVhiH01(Di)\forall v_{h}\in V_{h}^{i}\cap H^{1}_{0}(D_{i}),

ai(uhn,i,vh)=(fϕi,vh)i.a_{i}(u_{h}^{n,i},v_{h})=(f\circ\phi_{i},v_{h})_{i}.

However, this discrete problem is not well-posed because the degrees of freedom of uhn,iu_{h}^{n,i} on Di\partial D_{i} are undetermined. As an imitation of the second line of (7), we should evaluate these degrees of freedom by the data in dd-rectangles DjD_{j} for jij\neq i. So we have to investigate the transitions of coordinates.

For iji\neq j, let Dij=ϕi1(MiMj)DiD_{ij}=\phi_{i}^{-1}(M_{i}\cap M_{j})\subseteq D_{i} and Dji=ϕj1(MiMj)DjD_{ji}=\phi_{j}^{-1}(M_{i}\cap M_{j})\subseteq D_{j} (see Fig. 1 for an illustration). Then

ϕj1ϕi:DijDji\phi_{j}^{-1}\circ\phi_{i}:\ D_{ij}\rightarrow D_{ji}

is a diffeomorphism which is the transition of coordinates on the overlapping between MiM_{i} and MjM_{j}. As pointed out in Section 1, ϕj1ϕi\phi_{j}^{-1}\circ\phi_{i} preserves neither nodes nor grid necessarily. In other words, ϕj1ϕi\phi_{j}^{-1}\circ\phi_{i} may neither map a node in DijD_{ij} to a node in DjiD_{ji}, nor map the grid over DijD_{ij} to the one over DjiD_{ji}. Since ϕi\phi_{i} and ϕj\phi_{j} can be quite arbitrary, ϕj1ϕi\phi_{j}^{-1}\circ\phi_{i} may map the grid over DijD_{ij} to intractable curves in DjiD_{ji}.

Refer to caption
Figure 1: An illustration of a transition of coordinates on SmS^{m}.

However, this incompatibility among the grids over different DiD_{i} is not a bad sign of our method. We wish to emphasize that this actually shows the high flexibility of our approach. In fact, if all the ϕj1ϕi\phi_{j}^{-1}\circ\phi_{i} preserved the grids, one would obtain a global grid on MM. As mentioned in Section 1, it is too difficult to obtain such a grid in practice. In [46, Section 4], the transitions of coordinates do not preserve grid but do preserve boundary nodes, i.e. the transitions map nodes on DijDiD_{ij}\cap\partial D_{i} to nodes of DjiD_{ji}. As a result, the method in [46] have some flexibility. The problem (4) on S3S^{3} was solved numerically by three ways in [46]. However, those methods are not flexible enough to solve problems on more complicated manifolds in practice. The method in this paper improves the work in [46]. Since our transitions of coordinates do not necessarily preserve nodes, we shall evaluate the degrees of freedom on Di\partial D_{i} by interpolation.

Now we are in a position to propose our discrete Algorithm 2. Define

Vh=i=1mVhi.V_{h}=\bigoplus_{i=1}^{m}V_{h}^{i}.
Algorithm 2 A numerical DDM with a known chart decomposition.
1:Choose an arbitrary initial guess uh0=(uh0,1,,uh0,m)Vhu_{h}^{0}=(u_{h}^{0,1},\cdots,u_{h}^{0,m})\in V_{h}.
2:For each n>0n>0 and 1im1\leq i\leq m, assuming uhn1u_{h}^{n-1} and uhn,ju_{h}^{n,j} have been obtained for all j<ij<i, find a uhn,iVhiu_{h}^{n,i}\in V_{h}^{i} as follows. Suppose ξDi\xi\in\partial D_{i} is a node. If {jj<i,ϕi(ξ)Mj}\{j\mid j<i,\phi_{i}(\xi)\in M_{j}\}\neq\emptyset, then let j0j_{0} be the maximum of this set and define
uhn,i(ξ)=uhn,j0(ϕj01ϕi(ξ)).u_{h}^{n,i}(\xi)=u_{h}^{n,j_{0}}(\phi_{j_{0}}^{-1}\circ\phi_{i}(\xi)).
Otherwise, let j0=max{jji,ϕi(ξ)Mj}j_{0}=\max\{j\mid j\neq i,\phi_{i}(\xi)\in M_{j}\} and define
uhn,i(ξ)=uhn1,j0(ϕj01ϕi(ξ)).u_{h}^{n,i}(\xi)=u_{h}^{n-1,j_{0}}(\phi_{j_{0}}^{-1}\circ\phi_{i}(\xi)).
The interior degrees of freedom of uhn,iu_{h}^{n,i} are determined by vhVhiH01(Di)\forall v_{h}\in V_{h}^{i}\cap H^{1}_{0}(D_{i}),
ai(uhn,i,vh)=(fϕi,vh)i.a_{i}(u_{h}^{n,i},v_{h})=(f\circ\phi_{i},v_{h})_{i}.
3:Define uhn=(uhn,1,,uhn,m)Vhu_{h}^{n}=(u_{h}^{n,1},\cdots,u_{h}^{n,m})\in V_{h}.

Note that, in the second step of Algorithm 2, it is possible that

{jj<i,ϕi(ξ)Mj}=,\{j\mid j<i,\phi_{i}(\xi)\in M_{j}\}=\emptyset,

for instance, i=1i=1. However, the following always holds

{jji,ϕi(ξ)Mj}\{j\mid j\neq i,\phi_{i}(\xi)\in M_{j}\}\neq\emptyset

because M=j=1mIntMjM=\bigcup_{j=1}^{m}\mathrm{Int}M_{j} and ϕi(ξ)IntMi\phi_{i}(\xi)\notin\mathrm{Int}M_{i}. The choice of uhn,j0u_{h}^{n,j_{0}} or uhn1,j0u_{h}^{n-1,j_{0}} follows the principle that we use the latest iterate in other subdomains to evaluate the boundary value of uhn,iu_{h}^{n,i}. Also note that ϕj01ϕi(ξ)\phi_{j_{0}}^{-1}\circ\phi_{i}(\xi) is not necessarily a node. However, we can calculate uhn,j0(ϕj01ϕi(ξ))u_{h}^{n,j_{0}}(\phi_{j_{0}}^{-1}\circ\phi_{i}(\xi)) or uhn1,j0(ϕj01ϕi(ξ))u_{h}^{n-1,j_{0}}(\phi_{j_{0}}^{-1}\circ\phi_{i}(\xi)) by virtue of the coordinates of ϕj01ϕi(ξ)\phi_{j_{0}}^{-1}\circ\phi_{i}(\xi) in Dj0D_{j_{0}}. This is essentially by an interpolation of the degrees of freedom of uhn,j0u_{h}^{n,j_{0}} or uhn1,j0u_{h}^{n-1,j_{0}}.

Now the uhnu_{h}^{n} in Algorithm 2 is the nnth iterated discrete approximation to the solution to (4). Unlike the un,iu^{n,i} in Algorithm 1 which is globally defined on MM, the uhn,iu_{h}^{n,i} is a component of uhnu_{h}^{n} and is only defined on DiD_{i}. Furthermore, uhn,iϕi1u_{h}^{n,i}\circ\phi_{i}^{-1} and uhn,jϕj1u_{h}^{n,j}\circ\phi_{j}^{-1} usually disagree on the overlapping MiMjM_{i}\cap M_{j}. However, this disagreement is of no importance at all from the viewpoint of approximation. As far as uhn,iu_{h}^{n,i} approximates uϕiu\circ\phi_{i} well on DiD_{i} for each ii, we know uhn,iϕi1u_{h}^{n,i}\circ\phi_{i}^{-1} approximates uu well on MiM_{i}. Since MM is covered by these MiM_{i}, good numerical data would be obtained everywhere on MM.

Remark 3.1.

Algorithm 2 implicitly defines a discretization. Actually, we “discretize” the iteration in Algorithm 1 rather than the global problem (4). On the other hand, this makes a rigorous theoretical analysis more difficult.

4 Product Manifolds

Suppose MM and MM^{\prime} are compact manifolds with dimensions dd and dd^{\prime} respectively. The Cartesian product M×MM\times M^{\prime} is a compact manifold with dimension d+dd+d^{\prime}. A decomposition of MM and another one of MM^{\prime} canonically result in a decomposition of M×MM\times M^{\prime}. Actually, suppose M=i=1mIntMiM=\bigcup_{i=1}^{m}\mathrm{Int}M_{i} and M=i=1mIntMiM^{\prime}=\bigcup_{i^{\prime}=1}^{m^{\prime}}\mathrm{Int}M^{\prime}_{i^{\prime}}, where MiM_{i} (resp. MiM^{\prime}_{i^{\prime}}) are subdomains of MM (resp. MM^{\prime}). Then

M×M=i=1mi=1mInt(Mi×Mi),M\times M^{\prime}=\bigcup_{i=1}^{m}\bigcup_{i^{\prime}=1}^{m^{\prime}}\mathrm{Int}(M_{i}\times M^{\prime}_{i^{\prime}}),

where M×MM\times M^{\prime} is decomposed into mmmm^{\prime} subdomains Mi×MiM_{i}\times M^{\prime}_{i^{\prime}} for 1im1\leq i\leq m and 1im1\leq i^{\prime}\leq m^{\prime}.

This canonical decomposition of M×MM\times M^{\prime} reflects another advantage of the spaces of rectangular finite elements. More precisely, suppose there are diffeomorphisms ϕi:DiMi\phi_{i}:D_{i}\rightarrow M_{i} and ϕi:DiMi\phi^{\prime}_{i^{\prime}}:D^{\prime}_{i^{\prime}}\rightarrow M^{\prime}_{i^{\prime}}, where each DidD_{i}\subset\mathbb{R}^{d} (resp. DidD^{\prime}_{i^{\prime}}\subset\mathbb{R}^{d^{\prime}}) is a dd-rectangle (resp. dd^{\prime}-rectangle). Then we have the diffeomorphisms

ϕi×ϕi:Di×DiMi×Mi,\phi_{i}\times\phi^{\prime}_{i^{\prime}}:\ D_{i}\times D^{\prime}_{i^{\prime}}\rightarrow M_{i}\times M^{\prime}_{i^{\prime}},

where each Di×Did+dD_{i}\times D^{\prime}_{i^{\prime}}\subset\mathbb{R}^{d+d^{\prime}} is a (d+d)(d+d^{\prime})-rectangle. The transition of coordinates between Di×DiD_{i}\times D^{\prime}_{i^{\prime}} and Dj×DjD_{j}\times D^{\prime}_{j^{\prime}} is

(ϕj×ϕj)1(ϕi×ϕi)=(ϕj1ϕi)×(ϕj1ϕi).(\phi_{j}\times\phi^{\prime}_{j^{\prime}})^{-1}\circ(\phi_{i}\times\phi^{\prime}_{i^{\prime}})=(\phi_{j}^{-1}\circ\phi_{i})\times({\phi^{\prime}}_{j^{\prime}}^{-1}\circ\phi^{\prime}_{i^{\prime}}).

If rectangular grids are created over both DiD_{i} and DiD^{\prime}_{i^{\prime}}, a rectangular grid over Di×DiD_{i}\times D^{\prime}_{i^{\prime}} follows automatically.

In summary, the procedures of the decomposition and discretization of factor manifolds are helpful for those of product manifolds.

5 Numerical Experiments

We perform several numerical tests of the proposed method on manifolds S4S^{4}, 2\mathbb{CP}^{2} and S2×S2S^{2}\times S^{2}. They are 44-dimensional compact manifolds without boundary. While the proposed method applies to problems in all dimensions, the sizes of the linear systems derived from subdomains will increase exponentially with respect to the dimension. On one hand, we would have trouble in the storage of data. On the other hand, we would struggle to find solutions to these linear systems with desired accuracy. This difficulty is so called “the curse of dimensionality”. It is actually a typical phenomenon of Euclidean spaces rather than of general manifolds. For the sake of presentation and due to the constraint of computing resources, the numerical examples in this paper consider manifolds with dimension no more than 44. The numerical challenge in higher dimensions will be tackled in a forthcoming future work.

5.1 Two Problems on S4S^{4}

Let M=S4M=S^{4} be the unit sphere in 5\mathbb{R}^{5}, i.e.

M=S4={(y1,y2,y3,y4,y5)5|i=15yi2=1}.M=S^{4}=\left\{(y_{1},y_{2},y_{3},y_{4},y_{5})\in\mathbb{R}^{5}\middle|\sum_{i=1}^{5}y_{i}^{2}=1\right\}.

We decompose S4S^{4} into two subdomains as follows. By stereographic projections from the south pole (0,0,0,0,1)(0,0,0,0,-1) and north pole (0,0,0,0,1)(0,0,0,0,1), we obtain two subdomains M1M_{1} and M2M_{2} with coordinates whose interiors cover S4S^{4}. For an illustration please refer to Fig. 2, where the vertical direction stands for the direction of the 5th coordinate axis, the rectangle [r,r]4[-r,r]^{4} is a domain in 44×{0}5\mathbb{R}^{4}\cong\mathbb{R}^{4}\times\{0\}\subset\mathbb{R}^{5}, and the intersection of 4×{0}\mathbb{R}^{4}\times\{0\} and S4S^{4} is the equator of S4S^{4}. For each point P=(x1,,x4)[r,r]4P=(x_{1},\dots,x_{4})\in[-r,r]^{4}, the line segment between PP and the north pole (0,0,0,0,1)(0,0,0,0,1) intersects S4S^{4} at a single point Q=(y1,,y5)Q=(y_{1},\dots,y_{5}) other than (0,0,0,0,1)(0,0,0,0,1). The map PQP\mapsto Q provides an embedding ϕ1:[r,r]4S4\phi_{1}:[-r,r]^{4}\rightarrow S^{4}. We obtain another embedding ϕ2\phi_{2} if the north pole is replaced with the south pole. More precisely, we have

(13) D1=D2=[r,r]44.D_{1}=D_{2}=[-r,r]^{4}\subset\mathbb{R}^{4}.
Refer to caption
Refer to caption
Figure 2: An illustration of the stereographic projections used in the case of S4S^{4}.

Let x=(x1,x2,x3,x4)x=(x_{1},x_{2},x_{3},x_{4}) denote the coordinates of 4\mathbb{R}^{4}. Let x=i=14xi2\|x\|=\sqrt{\sum_{i=1}^{4}x_{i}^{2}}. The stereographic projections provide diffeomorphisms ϕi:DiMi\phi_{i}:D_{i}\rightarrow M_{i} as

ϕ1:D1\displaystyle\phi_{1}:\ D_{1} \displaystyle\rightarrow M1S45\displaystyle M_{1}\subset S^{4}\subset\mathbb{R}^{5}
(14) x\displaystyle x \displaystyle\mapsto (2x1+x2,1x21+x2)\displaystyle\left(\frac{2x}{1+\|x\|^{2}},\frac{1-\|x\|^{2}}{1+\|x\|^{2}}\right)

and

ϕ2:D2\displaystyle\phi_{2}:\ D_{2} \displaystyle\rightarrow M2S45\displaystyle M_{2}\subset S^{4}\subset\mathbb{R}^{5}
x\displaystyle x \displaystyle\mapsto (2x1+x2,1+x21+x2).\displaystyle\left(\frac{2x}{1+\|x\|^{2}},\frac{-1+\|x\|^{2}}{1+\|x\|^{2}}\right).

To guarantee S4=i=12IntMiS^{4}=\bigcup_{i=1}^{2}\mathrm{Int}M_{i}, we have to let r>1r>1. The larger rr is, the more overlapping there will be. The transitions of coordinates are given by

ϕ21ϕ1(x)=ϕ11ϕ2(x)=xx2.\phi_{2}^{-1}\circ\phi_{1}(x)=\phi_{1}^{-1}\circ\phi_{2}(x)=\frac{x}{\|x\|^{2}}.

Equip S4S^{4} with the Riemannian metric gg inherited from the standard one on 5\mathbb{R}^{5}. On each DiD_{i},

g=4(1+x2)2α=14dxαdxα,g=4(1+\|x\|^{2})^{-2}\sum_{\alpha=1}^{4}\mathrm{d}x_{\alpha}\otimes\mathrm{d}x_{\alpha},

and

Δv\displaystyle\Delta v =\displaystyle= 41(1+x2)4α=14xα((1+x2)2vxα)\displaystyle 4^{-1}(1+\|x\|^{2})^{4}\sum_{\alpha=1}^{4}\frac{\partial}{\partial x_{\alpha}}\left((1+\|x\|^{2})^{-2}\frac{\partial v}{\partial x_{\alpha}}\right)
=\displaystyle= 41(1+x2)2α=142vxα2(1+x2)α=14xαvxα.\displaystyle 4^{-1}(1+\|x\|^{2})^{2}\sum_{\alpha=1}^{4}\frac{\partial^{2}v}{\partial x_{\alpha}^{2}}-(1+\|x\|^{2})\sum_{\alpha=1}^{4}x_{\alpha}\frac{\partial v}{\partial x_{\alpha}}.

Consider the model problem (3) on S4S^{4} with b>0b>0. Choose the true solution to (3) as

u=y5,u=y_{5},

where y5y_{5} is the 55th coordinate of 5\mathbb{R}^{5}. Then

f=(4+b)uf=(4+b)u

in (3). On DiD_{i}, uu has the expression

uϕ1(x)=1x21+x2,uϕ2(x)=1+x21+x2.u\circ\phi_{1}(x)=\frac{1-\|x\|^{2}}{1+\|x\|^{2}},\qquad u\circ\phi_{2}(x)=\frac{-1+\|x\|^{2}}{1+\|x\|^{2}}.

The weak form of (3) on DiD_{i} is formulated as: vH01(Di)\forall v\in H^{1}_{0}(D_{i}),

Di4(1+x2)2α=14uϕixαvxαdx1dx2dx3dx4\displaystyle\int_{D_{i}}4(1+\|x\|^{2})^{-2}\sum_{\alpha=1}^{4}\frac{\partial u\circ\phi_{i}}{\partial x_{\alpha}}\frac{\partial v}{\partial x_{\alpha}}\mathrm{d}x_{1}\mathrm{d}x_{2}\mathrm{d}x_{3}\mathrm{d}x_{4}
+Di16(1+x2)4buϕivdx1dx2dx3dx4\displaystyle+\int_{D_{i}}16(1+\|x\|^{2})^{-4}b\cdot u\circ\phi_{i}\cdot v\mathrm{d}x_{1}\mathrm{d}x_{2}\mathrm{d}x_{3}\mathrm{d}x_{4}
=\displaystyle= Di16(1+x2)4fϕivdx1dx2dx3dx4.\displaystyle\int_{D_{i}}16(1+\|x\|^{2})^{-4}\cdot f\circ\phi_{i}\cdot v\mathrm{d}x_{1}\mathrm{d}x_{2}\mathrm{d}x_{3}\mathrm{d}x_{4}.

Now we choose bb in (3) as 11. For the discretization, we divide each coordinate interval [r,r][-r,r] into NN equal parts. The scale of the grid is thus h=2r/Nh=2r/N. There are (N+1)4(N+1)^{4} nodes on DiD_{i}, most rows of the stiffness matrix have 34=813^{4}=81 nonzero entries. We keep N80N\leq 80 due to the memory limitation of the hardware.

To get the nn-th discrete approximation uhn=(uhn,1,uhn,2)u_{h}^{n}=(u_{h}^{n,1},u_{h}^{n,2}), we need to solve a linear system AiXn,i=bn,iA_{i}X^{n,i}=b_{n,i} for i=1,2i=1,2, where Xn,iX^{n,i} provides the interior degrees of freedom of uhn,iu_{h}^{n,i}. We use the Conjugate Gradient Method (CG) to find Xn,iX^{n,i}. As a result, the process to generate the sequence {uhn}\{u_{h}^{n}\} is a nested iteration. The outer iteration is the DDM procedure Algorithm 2. The initial guess is chosen as uh0=0u_{h}^{0}=0. For each nn, the inner iteration is the CG iteration to solve AiXn,i=bn,iA_{i}X^{n,i}=b_{n,i} for i=1,2i=1,2. Note that AiA_{i} remains the same when nn changes, whereas bn,ib_{n,i} varies because of the evaluation of uhn,i|Diu_{h}^{n,i}|_{\partial D_{i}}. If {uhn}\{u_{h}^{n}\} does converge, Xn1,iX^{n-1,i} will be close to Xn,iX^{n,i} when nn is large enough. Thus, we choose the initial guess of Xn,iX^{n,i} as Xn1,iX^{n-1,i}. The tolerance for CG is set as

AiXn,ibn,i2/bn,i2108.\|A_{i}X^{n,i}-b_{n,i}\|_{2}/\|b_{n,i}\|_{2}\leq 10^{-8}.

Our numerical results show that uhnu_{h}^{n} becomes stable when n=n0n=n_{0} for some n0n_{0}, i.e. uhn=uhn0u_{h}^{n}=u_{h}^{n_{0}} up to machine precision for all nn0n\geq n_{0}. Actually, if

AiXn,ibn+1,i2/bn+1,i2108\|A_{i}X^{n,i}-b_{n+1,i}\|_{2}/\|b_{n+1,i}\|_{2}\leq 10^{-8}

for all ii, then the inner iteration terminates for n+1n+1 and Xn+1,i=Xn,iX^{n+1,i}=X^{n,i}. We found that inner iteration terminates for all n>n0n>n_{0}. In other words, practically, the sequence {uhn}\{u_{h}^{n}\} reaches its limit

uh=uhn0u_{h}^{\infty}=u_{h}^{n_{0}}

at step n0n_{0}.

In the following tables,

Ihu=(Ihu1,Ihu2)Vh,I_{h}u=(I_{h}u^{1},I_{h}u^{2})\in V_{h},

where IhuiVhiI_{h}u^{i}\in V_{h}^{i} is the interpolation of uϕiu\circ\phi_{i} on DiD_{i}. We define the energy norm of the error as

Ihuuha=max{ai(Ihuiuh,i,Ihuiuh,i)12i=1,2}.\|I_{h}u-u_{h}^{\infty}\|_{a}=\max\{a_{i}(I_{h}u^{i}-u_{h}^{\infty,i},I_{h}u^{i}-u_{h}^{\infty,i})^{\frac{1}{2}}\mid i=1,2\}.

The L2L^{2}-norm IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}}, LL^{\infty}-norm IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} and H1H^{1}-seminorm |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} are defined in similar ways. The numerical results are as follows in Tables 1 and 2, where, for each norm, the data on the left side of each cell are errors and orders of convergence are appended to the right.

hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.240.24 0.03020.0302 0.06900.0690 0.23480.2348 0.18300.1830 2222
0.120.12 0.00950.0095 1.71.7 0.01800.0180 1.91.9 0.07170.0717 1.71.7 0.05010.0501 1.91.9 2323
0.060.06 0.00320.0032 1.61.6 0.00460.0046 2.02.0 0.02390.0239 1.61.6 0.01500.0150 1.71.7 2222
0.030.03 7.2393e47.2393e-4 2.22.2 0.00110.0011 2.02.0 0.00820.0082 1.51.5 0.00480.0048 1.61.6 2222
Table 1: Convergence result on S4S^{4} for (u,r)=(y5,1.2)(u,r)=(y_{5},1.2).
hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.40.4 0.14590.1459 1.25781.2578 0.97820.9782 0.57250.5725 1010
0.20.2 0.04580.0458 1.71.7 0.25460.2546 2.32.3 0.29270.2927 1.71.7 0.14160.1416 2.02.0 1010
0.10.1 0.01100.0110 2.12.1 0.06650.0665 1.91.9 0.11990.1199 1.31.3 0.04270.0427 1.71.7 1010
0.050.05 0.00310.0031 1.81.8 0.01650.0165 2.02.0 0.04320.0432 1.51.5 0.01310.0131 1.71.7 1010
Table 2: Convergence result on S4S^{4} for (u,r)=(y5,2)(u,r)=(y_{5},2).

We see that the error IhuuhI_{h}u-u_{h}^{\infty} decays in the optimal order when hh decreases. Furthermore, n0n_{0} decreases when rr becomes larger, i.e. uhnu_{h}^{n} reaches its limit uhu_{h}^{\infty} fast provides that the overlapping between subdomains is large.

Remark 5.1.

As shown in Tables 1, 2, and other tables below, the convergences under H1H^{1}- and the energy norms are significantly better than the optimal first order in hh. The reason is yet to be explored. On the other hand, since the true solution to our example is CC^{\infty}, the finite element spaces are defined on highly symmetric grids (rectangular), the transition maps are smooth, thus any (or all) from these factors may contribute to superconvergence. But we cannot prove this hypothesis at this stage.

Note that Tables 1 and 2 actually show uh,iϕi1u_{h}^{\infty,i}\circ\phi_{i}^{-1} approximates u|Miu|_{M_{i}} well. Hence we obtain good numerical approximation to uu everywhere on MM because M=i=1mMiM=\bigcup_{i=1}^{m}M_{i}. More precisely, there is a 111-1 and onto correspondence between functions on MiM_{i} and those on DiD_{i}, i.e. a function vv on MiM_{i} correspondence to vϕiv\circ\phi_{i} on DiD_{i}. Via this bijection, L(Mi)L^{\infty}(M_{i}) (resp. L2(Mi)L^{2}(M_{i}) and H1(Mi)H^{1}(M_{i})) is isometrically isomorphic to L(Di)L^{\infty}(D_{i}) (resp. L2(Di;g)L^{2}(D_{i};g) and H1(Di;g)H^{1}(D_{i};g)). Here the notation “gg” stands for the metric tensor in (2), L2(Di;g)L^{2}(D_{i};g) and H1(Di;g)H^{1}(D_{i};g) are the L2L^{2}-space and H1H^{1}-space respectively on DiD_{i} with norms

wL2(Di;g)=(Di|w|2Gdx1dxd)12,\|w\|_{L^{2}(D_{i};g)}=\left(\int_{D_{i}}|w|^{2}\sqrt{G}\mathrm{d}x_{1}\cdots\mathrm{d}x_{d}\right)^{\frac{1}{2}},
|w|H1(Di;g)=(Diα,β=1dgαβwxαwxβGdx1dxd)12,|w|_{H^{1}(D_{i};g)}=\left(\int_{D_{i}}\sum_{\alpha,\beta=1}^{d}g^{\alpha\beta}\frac{\partial w}{\partial x_{\alpha}}\frac{\partial w}{\partial x_{\beta}}\sqrt{G}\mathrm{d}x_{1}\cdots\mathrm{d}x_{d}\right)^{\frac{1}{2}},

and

wH1(Di;g)=(wL2(Di;g)2+|w|H1(Di;g)2)12.\|w\|_{H^{1}(D_{i};g)}=\left(\|w\|_{L^{2}(D_{i};g)}^{2}+|w|_{H^{1}(D_{i};g)}^{2}\right)^{\frac{1}{2}}.

The above “isometrically isomorphic” means the bijection is a linear isomorphism preserving norms (see Definition 1.13 in [13, p. 66]), i.e. vL2(Mi)=vϕiL2(Di;g)\|v\|_{L^{2}(M_{i})}=\|v\circ\phi_{i}\|_{L^{2}(D_{i};g)} and so on. Meanwhile, by (12), we also have a(v,v)Mi=a(vϕi,vϕi)Dia(v,v)_{M_{i}}=a(v\circ\phi_{i},v\circ\phi_{i})_{D_{i}}. Furthermore, (gαβ)d×d(g_{\alpha\beta})_{d\times d} is bounded and uniformly elliptic on DiD_{i} because (gαβ)d×d(g_{\alpha\beta})_{d\times d} is CC^{\infty} and DiD_{i} is compact. Thus the L2(Di;g)\|\cdot\|_{L^{2}(D_{i};g)} and ||H1(Di;g)|\cdot|_{H^{1}(D_{i};g)} are equivalent to the usual L2(Di)\|\cdot\|_{L^{2}(D_{i})} and ||H1(Di)|\cdot|_{H^{1}(D_{i})} respectively. Therefore, uh,iϕi1u_{h}^{\infty,i}\circ\phi_{i}^{-1} approximates Ihuiϕi1I_{h}u^{i}\circ\phi_{i}^{-1} well in L(Mi)\|\cdot\|_{L^{\infty}(M_{i})}, L2(Mi)\|\cdot\|_{L^{2}(M_{i})}, ||H1(Mi)|\cdot|_{H^{1}(M_{i})} and a(,)Mia(\cdot,\cdot)_{M_{i}} as far as uh,iu_{h}^{\infty,i} approximates IhuiI_{h}u^{i} well in L(Di)\|\cdot\|_{L^{\infty}(D_{i})}, L2(Di)\|\cdot\|_{L^{2}(D_{i})}, ||H1(Di)|\cdot|_{H^{1}(D_{i})} and a(,)Dia(\cdot,\cdot)_{D_{i}}. Since Ihuiϕi1I_{h}u^{i}\circ\phi_{i}^{-1} is an interpolation of u|Miu|_{M_{i}}, we infer uh,iϕi1u_{h}^{\infty,i}\circ\phi_{i}^{-1} is a good approximation of u|Miu|_{M_{i}}.

We also investigated the number of iterations required to achieve an approximation of the same order of accuracy as IhuuhI_{h}u-u_{h}^{\infty}. So we set a tolerance for the outer iteration as

IhuuhnL2IhuuhL.\|I_{h}u-u_{h}^{n}\|_{L^{\infty}}\leq 2\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}}.

The numerical results are as follows in Tables 3 and 4. We see that nn are much less than n0n_{0}.

hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.240.24 0.05690.0569 0.20660.2066 0.26040.2604 0.21930.2193 44
0.120.12 0.01420.0142 0.04360.0436 0.07560.0756 0.05540.0554 66
0.060.06 0.00520.0052 0.01580.0158 0.02530.0253 0.01790.0179 77
0.030.03 0.00110.0011 0.00330.0033 0.00840.0084 0.00510.0051 99
Table 3: Convergence result on S4S^{4} for (u,r)=(y5,1.2)(u,r)=(y_{5},1.2).
hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.40.4 0.22310.2231 2.31412.3141 1.05921.0592 0.70990.7099 22
0.20.2 0.05500.0550 0.38060.3806 0.29530.2953 0.15510.1551 33
0.10.1 0.02030.0203 0.19450.1945 0.12810.1281 0.06080.0608 33
0.050.05 0.00420.0042 0.03150.0315 0.04340.0434 0.01440.0144 44
Table 4: Convergence result on S4S^{4} for (u,r)=(y5,2)(u,r)=(y_{5},2).

Now we consider a second problem on S4S^{4} with true solution u=y1y5u=y_{1}y_{5} in (3). Then f=(10+b)uf=(10+b)u. On DiD_{i}, uu has the expression

uϕ1(x)=2x1(1x2)(1+x2)2,uϕ2(x)=2x1(1+x2)(1+x2)2.u\circ\phi_{1}(x)=\frac{2x_{1}(1-\|x\|^{2})}{(1+\|x\|^{2})^{2}},\qquad u\circ\phi_{2}(x)=\frac{2x_{1}(-1+\|x\|^{2})}{(1+\|x\|^{2})^{2}}.

We choose the bb in (3) as 11. The numerical results are in tables 5, 6, 7 and 8. The performance of our algorithm on this problem is similar to that of the first one.

hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.240.24 0.04450.0445 0.07820.0782 0.21420.2142 0.16330.1633 99
0.120.12 0.01210.0121 1.91.9 0.02000.0200 2.02.0 0.06660.0666 1.71.7 0.04500.0450 1.91.9 99
0.060.06 0.00310.0031 2.02.0 0.00510.0051 2.02.0 0.02230.0223 1.61.6 0.01360.0136 1.71.7 99
0.030.03 7.8553e47.8553e-4 2.02.0 0.00130.0013 2.02.0 0.00770.0077 1.51.5 0.00430.0043 1.71.7 99
Table 5: Convergence result on S4S^{4} for (u,r)=(y1y5,1.2)(u,r)=(y_{1}y_{5},1.2).
hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.40.4 0.13890.1389 1.09711.0971 1.13161.1316 0.50170.5017 44
0.20.2 0.04780.0478 1.51.5 0.26580.2658 2.02.0 0.35400.3540 1.71.7 0.14230.1423 1.81.8 44
0.10.1 0.01350.0135 1.81.8 0.07010.0701 1.91.9 0.11410.1141 1.61.6 0.04010.0401 1.81.8 55
0.050.05 0.00340.0034 2.02.0 0.01760.0176 2.02.0 0.03750.0375 1.61.6 0.01170.0117 1.81.8 55
Table 6: Convergence result on S4S^{4} for (u,r)=(y1y5,2)(u,r)=(y_{1}y_{5},2).
hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.240.24 0.05510.0551 0.12010.1201 0.23790.2379 0.19330.1933 22
0.120.12 0.01280.0128 0.02350.0235 0.06760.0676 0.04680.0468 33
0.060.06 0.00400.0040 0.00870.0087 0.02400.0240 0.01610.0161 33
0.030.03 8.3731e48.3731e-4 0.00160.0016 0.00770.0077 0.00440.0044 44
Table 7: Convergence result on S4S^{4} for (u,r)=(y1y5,1.2)(u,r)=(y_{1}y_{5},1.2).
hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.40.4 0.13930.1393 1.10061.1006 1.13491.1349 0.50280.5028 22
0.20.2 0.04840.0484 0.27010.2701 0.35800.3580 0.14380.1438 22
0.10.1 0.01420.0142 0.07470.0747 0.11740.1174 0.04160.0416 22
0.050.05 0.00430.0043 0.02220.0222 0.04030.0403 0.01310.0131 22
Table 8: Convergence result on S4S^{4} for (u,r)=(y1y5,2)(u,r)=(y_{1}y_{5},2).

5.2 A Problem on 2\mathbb{CP}^{2}

Let M=2M=\mathbb{CP}^{2} be the complex projective plane. It is a compact complex manifold with complex dimension 22. Certainly, it can be considered as a real manifold with dimension 44.

Unlike S4S^{4}, the 2\mathbb{CP}^{2} is not a submanifold of any Euclidean space by definition. Furthermore, 2\mathbb{CP}^{2} cannot be embedded differential-topologically into k\mathbb{R}^{k} with k<7k<7 by the theory of characteristic classes ([37, Corollary 11.4]). Whitney constructed an explicit embedding of 2\mathbb{CP}^{2} into 7\mathbb{R}^{7} in an ingenious way ([55, Appendix]). Since the codimension of 2\mathbb{CP}^{2} in 7\mathbb{R}^{7} is 33, it is incredibly difficult to build effective polytopal approximations to 2\mathbb{CP}^{2} in 7\mathbb{R}^{7}. On the other hand, by definition, 2\mathbb{CP}^{2} can be constructed by patching together three coordinate charts, where the transitions of coordinates have explicit and neat formulas. Thus, it is very suitable to apply our method to 2\mathbb{CP}^{2}.

The 2\mathbb{CP}^{2} can be easily defined as a quotient space. Let

3{𝟎}={(w0,w1,w2)𝟎(w0,w1,w2)3}.\mathbb{C}^{3}\setminus\{\mathbf{0}\}=\{(w_{0},w_{1},w_{2})\mid\mathbf{0}\neq(w_{0},w_{1},w_{2})\in\mathbb{C}^{3}\}.

Here 𝟎3\mathbf{0}\in\mathbb{C}^{3} is the origin, each wjw_{j} is a complex number for 0j20\leq j\leq 2, and, following the convention of algebraic geometry, the index jj starts from 0 rather than 11. Define a relation of equivalence on 3{𝟎}\mathbb{C}^{3}\setminus\{\mathbf{0}\} as

(w0,w1,w2)(w0,w1,w2)(w_{0},w_{1},w_{2})\sim(w^{\prime}_{0},w^{\prime}_{1},w^{\prime}_{2})

if and only if

(w0,w1,w2)=λ(w0,w1,w2)(w_{0},w_{1},w_{2})=\lambda(w^{\prime}_{0},w^{\prime}_{1},w^{\prime}_{2})

for some 0λ0\neq\lambda\in\mathbb{C}. Define

2=3{𝟎}/.\mathbb{CP}^{2}=\mathbb{C}^{3}\setminus\{\mathbf{0}\}/\sim.

Thus, every P2P\in\mathbb{CP}^{2} can be represented by a vector (w0,w1,w2)3{𝟎}(w_{0},w_{1},w_{2})\in\mathbb{C}^{3}\setminus\{\mathbf{0}\}. Conventionally, we write

P=[w0,w1,w2],P=[w_{0},w_{1},w_{2}],

where [w0,w1,w2][w_{0},w_{1},w_{2}] are called the homogeneous coordinates of PP. Note that, for λ0\lambda\neq 0,

[w0,w1,w2]=[λw0,λw1,λw2].[w_{0},w_{1},w_{2}]=[\lambda w_{0},\lambda w_{1},\lambda w_{2}].

For more details of general k\mathbb{CP}^{k}, see [28, p. 15].

Now we decompose 2\mathbb{CP}^{2} into three subdomains MjM_{j} for 0j20\leq j\leq 2 (note that the index jj is chosen to start from 0 for the convenience of presentation.) In the following, zj=xj+1yjz_{j}=x_{j}+\sqrt{-1}y_{j}\in\mathbb{C}, xjx_{j}\in\mathbb{R}, and yjy_{j}\in\mathbb{R}. We shall identify the complex number zjz_{j} with the 22-dimensional real vector (xj,yj)(x_{j},y_{j}). Let Dj=[r,r]442D_{j}=[-r,r]^{4}\subset\mathbb{R}^{4}\simeq\mathbb{C}^{2}. For 0j20\leq j\leq 2, we have the following diffeomorphisms

ϕ0:D0\displaystyle\phi_{0}:\ D_{0} \displaystyle\rightarrow M02\displaystyle M_{0}\subset\mathbb{CP}^{2}
(z1,z2)\displaystyle(z_{1},z_{2}) \displaystyle\mapsto [1,z1,z2],\displaystyle[1,z_{1},z_{2}],
ϕ1:D1\displaystyle\phi_{1}:\ D_{1} \displaystyle\rightarrow M12\displaystyle M_{1}\subset\mathbb{CP}^{2}
(z0,z2)\displaystyle(z_{0},z_{2}) \displaystyle\mapsto [z0,1,z2],\displaystyle[z_{0},1,z_{2}],

and

ϕ2:D2\displaystyle\phi_{2}:\ D_{2} \displaystyle\rightarrow M22\displaystyle M_{2}\subset\mathbb{CP}^{2}
(z0,z1)\displaystyle(z_{0},z_{1}) \displaystyle\mapsto [z0,z1,1].\displaystyle[z_{0},z_{1},1].

To guarantee 2=j=02IntMj\mathbb{CP}^{2}=\bigcup_{j=0}^{2}\mathrm{Int}M_{j}, we have to let r>1r>1. The larger rr is, the more overlapping there will be. The transitions of coordinates are given by

ϕ11ϕ0(z1,z2)=(1z1,z2z1),ϕ01ϕ1(z0,z2)=(1z0,z2z0),\phi_{1}^{-1}\circ\phi_{0}(z_{1},z_{2})=\left(\frac{1}{z_{1}},\frac{z_{2}}{z_{1}}\right),\qquad\phi_{0}^{-1}\circ\phi_{1}(z_{0},z_{2})=\left(\frac{1}{z_{0}},\frac{z_{2}}{z_{0}}\right),
ϕ21ϕ0(z1,z2)=(1z2,z1z2),ϕ01ϕ2(z0,z1)=(z1z0,1z0),\phi_{2}^{-1}\circ\phi_{0}(z_{1},z_{2})=\left(\frac{1}{z_{2}},\frac{z_{1}}{z_{2}}\right),\qquad\phi_{0}^{-1}\circ\phi_{2}(z_{0},z_{1})=\left(\frac{z_{1}}{z_{0}},\frac{1}{z_{0}}\right),
ϕ21ϕ1(z0,z2)=(z0z2,1z2),ϕ11ϕ2(z0,z1)=(z0z1,1z1).\phi_{2}^{-1}\circ\phi_{1}(z_{0},z_{2})=\left(\frac{z_{0}}{z_{2}},\frac{1}{z_{2}}\right),\qquad\phi_{1}^{-1}\circ\phi_{2}(z_{0},z_{1})=\left(\frac{z_{0}}{z_{1}},\frac{1}{z_{1}}\right).

Equipping it with the classical Fubini-Study metric (c.f. [28, p. 30]), 2\mathbb{CP}^{2} becomes a Kähler manifold with Kähler form

12¯logj=02|wj|2,\frac{\sqrt{-1}}{2}\partial\bar{\partial}\log\sum_{j=0}^{2}|w_{j}|^{2},

where [w0,w1,w2][w_{0},w_{1},w_{2}] are the homogeneous coordinates of 2\mathbb{CP}^{2}. The Fubini-Study metric, denoted by \mathcal{H}, is a Hermitian metric. On each DjD_{j}, it is expressed as

=(1+z2)1α=0,αj2dzαdz¯α(1+z2)2α=0,αj2β=0,βj2z¯αzβdzαdz¯β,\mathcal{H}=(1+\|z\|^{2})^{-1}\sum_{\alpha=0,\alpha\neq j}^{2}\mathrm{d}z_{\alpha}\otimes\mathrm{d}\bar{z}_{\alpha}-(1+\|z\|^{2})^{-2}\sum_{\alpha=0,\alpha\neq j}^{2}\sum_{\beta=0,\beta\neq j}^{2}\bar{z}_{\alpha}z_{\beta}\mathrm{d}z_{\alpha}\otimes\mathrm{d}\bar{z}_{\beta},

where

(15) z2=α=0,αj2|zα|2=α=0,αj2(xα2+yα2).\|z\|^{2}=\sum_{\alpha=0,\alpha\neq j}^{2}|z_{\alpha}|^{2}=\sum_{\alpha=0,\alpha\neq j}^{2}(x_{\alpha}^{2}+y_{\alpha}^{2}).

We choose the Riemannian metric gg on 2\mathbb{CP}^{2} as the real part of \mathcal{H}. This gg provides the underlying Riemannian structure of the above Kähler structure. The Laplacian is expressed as

Δv=2Δv=2Δ¯v\displaystyle\Delta v=2\Delta_{\partial}v=2\Delta_{\bar{\partial}}v
=\displaystyle= 4(1+z2)3α=0,αj2zα((1+z2)2vz¯α+(1+z2)2zαβ=0,βj2z¯βvz¯β)\displaystyle 4(1+\|z\|^{2})^{3}\sum_{\alpha=0,\alpha\neq j}^{2}\frac{\partial}{\partial z_{\alpha}}\left((1+\|z\|^{2})^{-2}\frac{\partial v}{\partial\bar{z}_{\alpha}}+(1+\|z\|^{2})^{-2}z_{\alpha}\sum_{\beta=0,\beta\neq j}^{2}\bar{z}_{\beta}\frac{\partial v}{\partial\bar{z}_{\beta}}\right)
=\displaystyle= 4(1+z2)(α=0,αj22vzαz¯α+α=0,αj2β=0,βj2zαz¯β2vzαz¯β).\displaystyle 4(1+\|z\|^{2})\left(\sum_{\alpha=0,\alpha\neq j}^{2}\frac{\partial^{2}v}{\partial z_{\alpha}\partial\bar{z}_{\alpha}}+\sum_{\alpha=0,\alpha\neq j}^{2}\sum_{\beta=0,\beta\neq j}^{2}z_{\alpha}\bar{z}_{\beta}\frac{\partial^{2}v}{\partial z_{\alpha}\partial\bar{z}_{\beta}}\right).

We consider the model problem (3) with b>0b>0. Choose constants aja_{j}\in\mathbb{R}, 0j20\leq j\leq 2. Choose the true solution to (3) as

(16) u([w0,w1,w2])=j=02aj|wj|2,u([w_{0},w_{1},w_{2}])=\sum_{j=0}^{2}a_{j}|w_{j}|^{2},

where [w0,w1,w2][w_{0},w_{1},w_{2}] are homogeneous coordinates with normalization j=02|wj|2=1\sum_{j=0}^{2}|w_{j}|^{2}=1. It is easy to see that uu is well-defined. The ff in (3) is then

f=(12+b)u4j=02aj.f=(12+b)u-4\sum_{j=0}^{2}a_{j}.

On DjD_{j}, the true solution uu has the expression

uϕj=aj+β=0,βj2aβ|zβ|21+β=0,βj2|zβ|2=aj+β=0,βj2aβ(xβ2+yβ2)1+β=0,βj2(xβ2+yβ2).u\circ\phi_{j}=\frac{a_{j}+\sum_{\beta=0,\beta\neq j}^{2}a_{\beta}|z_{\beta}|^{2}}{1+\sum_{\beta=0,\beta\neq j}^{2}|z_{\beta}|^{2}}=\frac{a_{j}+\sum_{\beta=0,\beta\neq j}^{2}a_{\beta}(x_{\beta}^{2}+y_{\beta}^{2})}{1+\sum_{\beta=0,\beta\neq j}^{2}(x_{\beta}^{2}+y_{\beta}^{2})}.

The weak form of (3) on DjD_{j} is formulated as: vH01(Dj)\forall v\in H^{1}_{0}(D_{j}),

Dj(1+z2)2α=0,αj2(uϕjxαvxα+uϕjyαvyα)\displaystyle\int_{D_{j}}(1+\|z\|^{2})^{-2}\sum_{\alpha=0,\alpha\neq j}^{2}\left(\frac{\partial u\circ\phi_{j}}{\partial x_{\alpha}}\frac{\partial v}{\partial x_{\alpha}}+\frac{\partial u\circ\phi_{j}}{\partial y_{\alpha}}\frac{\partial v}{\partial y_{\alpha}}\right)
+Dj(1+z2)2[α=0,αj2(xαuϕjxα+yαuϕjyα)][α=0,αj2(xαvxα+yαvyα)]\displaystyle+\int_{D_{j}}(1+\|z\|^{2})^{-2}\left[\sum_{\alpha=0,\alpha\neq j}^{2}\left(x_{\alpha}\frac{\partial u\circ\phi_{j}}{\partial x_{\alpha}}+y_{\alpha}\frac{\partial u\circ\phi_{j}}{\partial y_{\alpha}}\right)\right]\cdot\left[\sum_{\alpha=0,\alpha\neq j}^{2}\left(x_{\alpha}\frac{\partial v}{\partial x_{\alpha}}+y_{\alpha}\frac{\partial v}{\partial y_{\alpha}}\right)\right]
+Dj(1+z2)2[α=0,αj2(yαuϕjxαxαuϕjyα)][α=0,αj2(yαvxαxαvyα)]\displaystyle+\int_{D_{j}}(1+\|z\|^{2})^{-2}\left[\sum_{\alpha=0,\alpha\neq j}^{2}\left(y_{\alpha}\frac{\partial u\circ\phi_{j}}{\partial x_{\alpha}}-x_{\alpha}\frac{\partial u\circ\phi_{j}}{\partial y_{\alpha}}\right)\right]\cdot\left[\sum_{\alpha=0,\alpha\neq j}^{2}\left(y_{\alpha}\frac{\partial v}{\partial x_{\alpha}}-x_{\alpha}\frac{\partial v}{\partial y_{\alpha}}\right)\right]
+Dj(1+z2)3buϕjv\displaystyle+\int_{D_{j}}(1+\|z\|^{2})^{-3}b\cdot u\circ\phi_{j}\cdot v
=\displaystyle= Dj(1+z2)3fϕjv,\displaystyle\int_{D_{j}}(1+\|z\|^{2})^{-3}f\circ\phi_{j}\cdot v,

where z2\|z\|^{2} is defined in (15), and the symbols dxα\mathrm{d}x_{\alpha} and dyα\mathrm{d}y_{\alpha} in the integrals are also omitted.

Now we choose bb in (3) as 44, choose (a0,a1,a2)(a_{0},a_{1},a_{2}) in (16) as (0,1,1)(0,1,-1). The numerical results are as follows in Tables 9, 10, 11 and 12. Tables 9 and 10 indicate that the sequence {uhn}\{u_{h}^{n}\} practiaclly reaches its limit uhu_{h}^{\infty} at step n0n_{0}. The convergence rate improves as the overlapping between subdomains increases. By referring to Tables 11 and 12, it is possible to achieve an approximation with the same order of accuracy as IhuuhI_{h}u-u_{h}^{\infty} with much fewer iterative steps.

hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.240.24 0.03760.0376 0.04540.0454 0.15590.1559 0.07180.0718 3838
0.120.12 0.01030.0103 1.91.9 0.01160.0116 2.02.0 0.04410.0441 1.81.8 0.02040.0204 1.81.8 3636
0.060.06 0.00240.0024 2.12.1 0.00290.0029 2.02.0 0.01270.0127 1.81.8 0.00570.0057 1.81.8 3535
0.030.03 6.1832e46.1832e-4 2.02.0 7.3810e47.3810e-4 2.02.0 0.00390.0039 1.71.7 0.00170.0017 1.71.7 3434
Table 9: Convergence result on 2\mathbb{CP}^{2} for r=1.2r=1.2.
hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.40.4 0.10260.1026 0.37870.3787 0.83380.8338 0.22680.2268 1414
0.20.2 0.03120.0312 1.71.7 0.10500.1050 1.91.9 0.24830.2483 1.71.7 0.06740.0674 1.81.8 1414
0.10.1 0.00940.0094 1.71.7 0.02730.0273 1.91.9 0.07710.0771 1.71.7 0.01980.0198 1.81.8 1414
0.050.05 0.00200.0020 2.22.2 0.00670.0067 2.02.0 0.02450.0245 1.71.7 0.00610.0061 1.71.7 1313
Table 10: Convergence result on 2\mathbb{CP}^{2} for r=2r=2.
hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.240.24 0.06910.0691 0.18320.1832 0.21470.2147 0.13320.1332 33
0.120.12 0.01750.0175 0.04800.0480 0.04620.0462 0.03210.0321 66
0.060.06 0.00430.0043 0.01270.0127 0.01320.0132 0.00870.0087 99
0.030.03 0.00120.0012 0.00340.0034 0.00400.0040 0.00250.0025 1212
Table 11: Convergence result on 2\mathbb{CP}^{2} for r=1.2r=1.2.
hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.40.4 0.13820.1382 0.91930.9193 0.88960.8896 0.27930.2793 22
0.20.2 0.04320.0432 0.23610.2361 0.25160.2516 0.08060.0806 33
0.10.1 0.01230.0123 0.06010.0601 0.07770.0777 0.02270.0227 44
0.050.05 0.00270.0027 0.01480.0148 0.02460.0246 0.00670.0067 55
Table 12: Convergence result on 2\mathbb{CP}^{2} for r=2r=2.

5.3 A Problem on S2×S2S^{2}\times S^{2}

Let S2S^{2} be the unit sphere in 3\mathbb{R}^{3}, i.e.

S2={(y1,y2,y3)3|i=13yi2=1}.S^{2}=\left\{(y_{1},y_{2},y_{3})\in\mathbb{R}^{3}\middle|\sum_{i=1}^{3}y_{i}^{2}=1\right\}.

Let M=S2×S2M=S^{2}\times S^{2}. Similar to S4S^{4}, we can decompose S2S^{2} into two subdomains via stereographic projections. This decomposition results in a product decomposition of S2×S2S^{2}\times S^{2} with 2×2=42\times 2=4 subdomains.

In the following, let x=(x1,x2)x=(x_{1},x_{2}) and x=(x1,x2)x^{\prime}=(x^{\prime}_{1},x^{\prime}_{2}) denote the coordinates of 2\mathbb{R}^{2}. Let x=i=12xi2\|x\|=\sqrt{\sum_{i=1}^{2}x_{i}^{2}} and x=i=12xi2\|x^{\prime}\|=\sqrt{\sum_{i=1}^{2}{x^{\prime}}_{i}^{2}}. For 1i41\leq i\leq 4, let

Di=[r,r]4={(x,x)x[r,r]2,x[r,r]2}.D_{i}=[-r,r]^{4}=\{(x,x^{\prime})\mid x\in[-r,r]^{2},x^{\prime}\in[-r,r]^{2}\}.

The product decomposition of S2×S2S^{2}\times S^{2} is given by diffeomorphisms

ϕ1:D1\displaystyle\phi_{1}:\ D_{1} \displaystyle\rightarrow M1S2×S23×3\displaystyle M_{1}\subset S^{2}\times S^{2}\subset\mathbb{R}^{3}\times\mathbb{R}^{3}
(x,x)\displaystyle(x,x^{\prime}) \displaystyle\mapsto (2x1+x2,1x21+x2,2x1+x2,1x21+x2),\displaystyle\left(\frac{2x}{1+\|x\|^{2}},\frac{1-\|x\|^{2}}{1+\|x\|^{2}},\frac{2x^{\prime}}{1+\|x^{\prime}\|^{2}},\frac{1-\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}}\right),
ϕ2:D2\displaystyle\phi_{2}:\ D_{2} \displaystyle\rightarrow M2S2×S23×3\displaystyle M_{2}\subset S^{2}\times S^{2}\subset\mathbb{R}^{3}\times\mathbb{R}^{3}
(x,x)\displaystyle(x,x^{\prime}) \displaystyle\mapsto (2x1+x2,1x21+x2,2x1+x2,1+x21+x2),\displaystyle\left(\frac{2x}{1+\|x\|^{2}},\frac{1-\|x\|^{2}}{1+\|x\|^{2}},\frac{2x^{\prime}}{1+\|x^{\prime}\|^{2}},\frac{-1+\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}}\right),
ϕ3:D3\displaystyle\phi_{3}:\ D_{3} \displaystyle\rightarrow M3S2×S23×3\displaystyle M_{3}\subset S^{2}\times S^{2}\subset\mathbb{R}^{3}\times\mathbb{R}^{3}
(x,x)\displaystyle(x,x^{\prime}) \displaystyle\mapsto (2x1+x2,1+x21+x2,2x1+x2,1x21+x2),\displaystyle\left(\frac{2x}{1+\|x\|^{2}},\frac{-1+\|x\|^{2}}{1+\|x\|^{2}},\frac{2x^{\prime}}{1+\|x^{\prime}\|^{2}},\frac{1-\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}}\right),

and

ϕ4:D4\displaystyle\phi_{4}:\ D_{4} \displaystyle\rightarrow M4S2×S23×3\displaystyle M_{4}\subset S^{2}\times S^{2}\subset\mathbb{R}^{3}\times\mathbb{R}^{3}
(x,x)\displaystyle(x,x^{\prime}) \displaystyle\mapsto (2x1+x2,1+x21+x2,2x1+x2,1+x21+x2).\displaystyle\left(\frac{2x}{1+\|x\|^{2}},\frac{-1+\|x\|^{2}}{1+\|x\|^{2}},\frac{2x^{\prime}}{1+\|x^{\prime}\|^{2}},\frac{-1+\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}}\right).

To guarantee S2×S2=i=14IntMiS^{2}\times S^{2}=\bigcup_{i=1}^{4}\mathrm{Int}M_{i}, we have to let r>1r>1. The larger rr is, the more overlapping there will be. The transitions of coordinates are given by

ϕ21ϕ1(x,x)=(x,xx2),ϕ31ϕ1(x,x)=(xx2,x),\phi_{2}^{-1}\circ\phi_{1}(x,x^{\prime})=\left(x,\frac{x^{\prime}}{\|x^{\prime}\|^{2}}\right),\qquad\phi_{3}^{-1}\circ\phi_{1}(x,x^{\prime})=\left(\frac{x}{\|x\|^{2}},x^{\prime}\right),
ϕ41ϕ1(x,x)=(xx2,xx2),ϕ31ϕ2(x,x)=(xx2,xx2),\phi_{4}^{-1}\circ\phi_{1}(x,x^{\prime})=\left(\frac{x}{\|x\|^{2}},\frac{x^{\prime}}{\|x^{\prime}\|^{2}}\right),\qquad\phi_{3}^{-1}\circ\phi_{2}(x,x^{\prime})=\left(\frac{x}{\|x\|^{2}},\frac{x^{\prime}}{\|x^{\prime}\|^{2}}\right),
ϕ41ϕ2(x,x)=(xx2,x),ϕ41ϕ3(x,x)=(x,xx2),\phi_{4}^{-1}\circ\phi_{2}(x,x^{\prime})=\left(\frac{x}{\|x\|^{2}},x^{\prime}\right),\qquad\phi_{4}^{-1}\circ\phi_{3}(x,x^{\prime})=\left(x,\frac{x^{\prime}}{\|x^{\prime}\|^{2}}\right),

and ϕi1ϕj=ϕj1ϕi\phi_{i}^{-1}\circ\phi_{j}=\phi_{j}^{-1}\circ\phi_{i} for all ii and jj.

Equip S2S^{2} with the Riemannian metric gg inherited from the standard one on 3\mathbb{R}^{3}. Equip S2×S2S^{2}\times S^{2} with the product metric. On each DiD_{i}, the metric has the form

g=4(1+x2)2α=12dxαdxα+4(1+x2)2α=12dxαdxα,g=4(1+\|x\|^{2})^{-2}\sum_{\alpha=1}^{2}\mathrm{d}x_{\alpha}\otimes\mathrm{d}x_{\alpha}+4(1+\|x^{\prime}\|^{2})^{-2}\sum_{\alpha^{\prime}=1}^{2}\mathrm{d}x^{\prime}_{\alpha^{\prime}}\otimes\mathrm{d}x^{\prime}_{\alpha^{\prime}},

and

Δv=41(1+x2)2α=122vxα2+41(1+x2)2α=122vxα2.\Delta v=4^{-1}(1+\|x\|^{2})^{2}\sum_{\alpha=1}^{2}\frac{\partial^{2}v}{\partial x_{\alpha}^{2}}+4^{-1}(1+\|x^{\prime}\|^{2})^{2}\sum_{\alpha^{\prime}=1}^{2}\frac{\partial^{2}v}{\partial{x^{\prime}}_{\alpha^{\prime}}^{2}}.

Consider the model problem (3) on S2×S2S^{2}\times S^{2} with b>0b>0. Choose the true solution uu to (3) as

u=y3+y3,u=y_{3}+y^{\prime}_{3},

where S2×S23×3S^{2}\times S^{2}\subset\mathbb{R}^{3}\times\mathbb{R}^{3}, and y3y_{3} (resp. y3y^{\prime}_{3}) is the 33rd coordinate of the first (resp. second) factor 3\mathbb{R}^{3}. Then

f=(2+b)uf=(2+b)u

in (3). On DiD_{i}, the true solution uu has the expression

uϕ1=1x21+x2+1x21+x2,uϕ2=1x21+x2+1+x21+x2,u\circ\phi_{1}=\frac{1-\|x\|^{2}}{1+\|x\|^{2}}+\frac{1-\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}},\qquad u\circ\phi_{2}=\frac{1-\|x\|^{2}}{1+\|x\|^{2}}+\frac{-1+\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}},
uϕ3=1+x21+x2+1x21+x2,uϕ4=1+x21+x2+1+x21+x2.u\circ\phi_{3}=\frac{-1+\|x\|^{2}}{1+\|x\|^{2}}+\frac{1-\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}},\qquad u\circ\phi_{4}=\frac{-1+\|x\|^{2}}{1+\|x\|^{2}}+\frac{-1+\|x^{\prime}\|^{2}}{1+\|x^{\prime}\|^{2}}.

The weak form of (3) on DiD_{i} is formulated as: vH01(Di)\forall v\in H^{1}_{0}(D_{i}),

Di[4(1+x2)2α=12uϕixαvxα+4(1+x2)2α=12uϕixαvxα\displaystyle\int_{D_{i}}\left[4(1+\|x^{\prime}\|^{2})^{-2}\sum_{\alpha=1}^{2}\frac{\partial u\circ\phi_{i}}{\partial x_{\alpha}}\frac{\partial v}{\partial x_{\alpha}}+4(1+\|x\|^{2})^{-2}\sum_{\alpha^{\prime}=1}^{2}\frac{\partial u\circ\phi_{i}}{\partial x^{\prime}_{\alpha^{\prime}}}\frac{\partial v}{\partial x^{\prime}_{\alpha^{\prime}}}\right.
+16(1+x2)2(1+x2)2buϕiv]dx1dx2dx1dx2\displaystyle\left.+16(1+\|x\|^{2})^{-2}(1+\|x^{\prime}\|^{2})^{-2}b\cdot u\circ\phi_{i}\cdot v\right]\mathrm{d}x_{1}\mathrm{d}x_{2}\mathrm{d}x^{\prime}_{1}\mathrm{d}x^{\prime}_{2}
=\displaystyle= Di16(1+x2)2(1+x2)2fϕivdx1dx2dx1dx2.\displaystyle\int_{D_{i}}16(1+\|x\|^{2})^{-2}(1+\|x^{\prime}\|^{2})^{-2}f\circ\phi_{i}\cdot v\mathrm{d}x_{1}\mathrm{d}x_{2}\mathrm{d}x^{\prime}_{1}\mathrm{d}x^{\prime}_{2}.

Now we choose bb in (3) as 22. The numerical results are as follows in Tables 13, 14, 15 and 16. The performance of our algorithm on S2×S2S^{2}\times S^{2} is similar to that on S4S^{4} and 2\mathbb{CP}^{2}.

hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.240.24 0.02070.0207 0.05880.0588 0.16710.1671 0.21750.2175 2222
0.120.12 0.00450.0045 2.22.2 0.01440.0144 2.02.0 0.04790.0479 1.81.8 0.06060.0606 1.81.8 2222
0.060.06 0.00120.0012 1.91.9 0.00360.0036 2.02.0 0.01350.0135 1.81.8 0.01640.0164 1.91.9 2222
0.030.03 3.1235e43.1235e-4 1.91.9 8.6478e48.6478e-4 2.12.1 0.00450.0045 1.61.6 0.00530.0053 1.61.6 2121
Table 13: Convergence result on S2×S2S^{2}\times S^{2} for r=1.2r=1.2.
hh IhuuhL\|I_{h}u-u_{h}^{\infty}\|_{L^{\infty}} IhuuhL2\|I_{h}u-u_{h}^{\infty}\|_{L^{2}} |Ihuuh|H1|I_{h}u-u_{h}^{\infty}|_{H^{1}} Ihuuha\|I_{h}u-u_{h}^{\infty}\|_{a} n0n_{0}
0.40.4 0.14520.1452 0.97630.9763 1.19521.1952 1.07661.0766 99
0.20.2 0.02340.0234 2.62.6 0.19850.1985 2.32.3 0.36460.3646 1.71.7 0.30140.3014 1.81.8 99
0.10.1 0.00900.0090 1.41.4 0.05580.0558 1.81.8 0.11760.1176 1.61.6 0.08840.0884 1.81.8 99
0.050.05 0.00160.0016 2.52.5 0.01320.0132 2.12.1 0.03740.0374 1.71.7 0.02680.0268 1.71.7 99
Table 14: Convergence result on S2×S2S^{2}\times S^{2} for r=2r=2.
hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.240.24 0.03340.0334 0.09750.0975 0.15430.1543 0.26400.2640 55
0.120.12 0.00630.0063 0.02150.0215 0.04480.0448 0.06870.0687 77
0.060.06 0.00220.0022 0.00790.0079 0.01280.0128 0.02140.0214 88
0.030.03 4.8270e44.8270e-4 0.00170.0017 0.00440.0044 0.00590.0059 1010
Table 15: Convergence result on S2×S2S^{2}\times S^{2} for r=1.2r=1.2.
hh IhuuhnL\|I_{h}u-u_{h}^{n}\|_{L^{\infty}} IhuuhnL2\|I_{h}u-u_{h}^{n}\|_{L^{2}} |Ihuuhn|H1|I_{h}u-u_{h}^{n}|_{H^{1}} Ihuuhna\|I_{h}u-u_{h}^{n}\|_{a} nn
0.40.4 0.24360.2436 1.15761.1576 1.39371.3937 1.47081.4708 22
0.20.2 0.02960.0296 0.18290.1829 0.37080.3708 0.31860.3186 33
0.10.1 0.00880.0088 0.05460.0546 0.11750.1175 0.08930.0893 44
0.050.05 0.00210.0021 0.01200.0120 0.03750.0375 0.02810.0281 44
Table 16: Convergence result on S2×S2S^{2}\times S^{2} for r=2r=2.

Acknowledgements

We thank Feng Wang, Yuanming Xiao and Xuejun Xu for various discussions. We thank the anonymous reviewers for many constructive comments and suggestions which led to an improved presentation of this paper. S. Cao was partially supported by NSF awards DMS-2136075 and DMS-2309778. L. Qin was partially supported by NSFC11871272.

References

  • [1] R. Anderson, J. Andrej, A. Barker, J. Bramwell, J.-S. Camier, J. Cerveny, V. Dobrev, Y. Dudouit, A. Fisher, T. Kolev, W. Pazner, M. Stowell, V. Tomov, I. Akkerman, J. Dahm, D. Medina, and S. Zampini, Mfem: A modular finite element methods library, Computers & Mathematics with Applications, 81 (2021), pp. 42–74, https://www.sciencedirect.com/science/article/pii/S0898122120302583. Development and Application of Open-source Software for Problems with Numerical PDEs.
  • [2] P. F. Antonietti, A. Dedner, P. Madhavan, S. Stangalino, B. Stinner, and M. Verani, High order discontinuous Galerkin methods for elliptic problems on surfaces, SIAM Journal on Numerical Analysis, 53 (2015), pp. 1145–1171.
  • [3] E. Bachini, M. W. Farthing, and M. Putti, Intrinsic finite element method for advection-diffusion-reaction equations on surfaces, Journal of Computational Physics, 424 (2021), p. 109827.
  • [4] E. Bachini, G. Manzini, and M. Putti, Arbitrary-order intrinsic virtual element method for elliptic equations on surfaces, Calcolo, 58 (2021), pp. 1–28.
  • [5] W. Bangerth, R. Hartmann, and G. Kanschat, deal.II-a general-purpose object-oriented finite element library, ACM Transactions on Mathematical Software (TOMS), 33 (2007), pp. 24–es.
  • [6] J. R. Baumgardner and P. O. Frederickson, Icosahedral discretization of the two-sphere, SIAM Journal on Numerical Analysis, 22 (1985), pp. 1107–1115.
  • [7] A. Bonito, A. Demlow, and M. Licht, A divergence-conforming finite element method for the surface Stokes equation, SIAM Journal on Numerical Analysis, 58 (2020), pp. 2764–2798.
  • [8] A. Bonito, A. Demlow, and R. H. Nochetto, Finite element methods for the Laplace–Beltrami operator, in Handbook of Numerical Analysis, vol. 21, Elsevier, 2020, pp. 1–103.
  • [9] J. H. Bramble, J. E. Pasciak, J. P. Wang, and J. Xu, Convergence estimates for product iterative methods with applications to domain decomposition, Mathematics of Computation, 57 (1991), pp. 1–21.
  • [10] J. Brannick and S. Cao, A bootstrap multigrid eigensolver, SIAM Journal on Matrix Analysis and Applications, 43 (2022), pp. 1627–1657.
  • [11] X.-C. Cai and Y. Saad, Overlapping domain decomposition algorithms for general sparse matrices, Numerical linear algebra with applications, 3 (1996), pp. 221–237.
  • [12] P. G. Ciarlet, The Finite Element Method for Elliptic Problems SIAM, 2002.
  • [13] J. B. Conway, A course in functional analysis, Second edition, Grad. Texts in Math., 96, Springer-Verlag, New York, 1990.
  • [14] K. Deckelnick, G. Dziuk, and C. M. Elliott, Computation of geometric partial differential equations and mean curvature flow, Acta numerica, 14 (2005), pp. 139–232.
  • [15] A. Demlow, Higher-order finite element methods and pointwise error estimates for elliptic problems on surfaces, SIAM Journal on Numerical Analysis, 47 (2009), pp. 805–827.
  • [16] A. Demlow and G. Dziuk, An adaptive finite element method for the Laplace–Beltrami operator on implicitly defined surfaces, SIAM Journal on Numerical Analysis, 45 (2007), pp. 421–442.
  • [17] V. Dobrev, J.-L. Guermond, and B. Popov, Surface reconstruction and image enhancement via L1L^{1}-minimization, SIAM Journal on Scientific Computing, 32 (2010), pp. 1591–1616.
  • [18] V. Dobrev, P. Knupp, T. Kolev, K. Mittal, and V. Tomov, Adaptive tangential relaxation and surface fitting for high-order mesh optimization, in 10th International Conference on Adaptative Modeling and Simulation, CIMNE, 2021, https://doi.org/10.23967/admos.2021.015, https://doi.org/10.23967/admos.2021.015.
  • [19] G. Dziuk, Finite elements for the Beltrami operator on arbitrary surfaces, Partial differential equations and calculus of variations, (1988), pp. 142–155.
  • [20] G. Dziuk, An algorithm for evolutionary surfaces, Numerische Mathematik, 58 (1990), pp. 603–611.
  • [21] G. Dziuk and C. M. Elliott, Finite elements on evolving surfaces, IMA journal of numerical analysis, 27 (2007), pp. 262–292.
  • [22] G. Dziuk and C. M. Elliott, Finite element methods for surface PDEs, Acta Numerica, 22 (2013), pp. 289–396.
  • [23] C. M. Elliott and T. Ranner, Evolving surface finite element method for the Cahn–Hilliard equation, Numerische Mathematik, 129 (2015), pp. 483–534.
  • [24] B. Engquist and H.-K. Zhao, Absorbing boundary conditions for domain decomposition, Applied numerical mathematics, 27 (1998), pp. 341–365.
  • [25] O. Etzmuß, M. Keckeisen, and W. Straßer, A fast finite element solution for cloth modelling, in 11th Pacific Conference onComputer Graphics and Applications, 2003. Proceedings., IEEE, 2003, pp. 244–251.
  • [26] A. Fabri, G.-J. Giezeman, L. Kettner, S. Schirra, and S. Schönherr, On the design of cgal a computational geometry algorithms library, Software: Practice and Experience, 30 (2000), pp. 1167–1202.
  • [27] M. Frittelli and I. Sgura, Virtual element method for the Laplace-Beltrami equation on surfaces, ESAIM: Mathematical Modelling and Numerical Analysis, 52 (2018), pp. 965–993.
  • [28] P. Griffiths and J. Harris, Principles of Algebraic Geometry, John Wiley & Sons, 2014.
  • [29] M. W. Hirsch, Differential Topology, vol. 33, Springer Science & Business Media, 2012.
  • [30] T. Jankuhn, M. A. Olshanskii, A. Reusken, and A. Zhiliakov, Error analysis of higher order trace finite element methods for the surface Stokes equation, Journal of Numerical Mathematics, 29 (2021), pp. 245–267.
  • [31] M. Jin, X. Feng, and K. Wang, Gradient recovery-based adaptive stabilized mixed fem for the convection–diffusion–reaction equation on surfaces, Computer Methods in Applied Mechanics and Engineering, 380 (2021), p. 113798.
  • [32] J. Jost, Riemannian Geometry and Geometric Analysis, vol. 42005, Springer, 2008.
  • [33] B. Jüttler, A. Mantzaflaris, R. Perl, and M. Rumpf, On numerical integration in isogeometric subdivision methods for pdes on surfaces, Computer Methods in Applied Mechanics and Engineering, 302 (2016), pp. 131–146.
  • [34] P. Kang and S.-K. Youn, Isogeometric topology optimization of shell structures using trimmed nurbs surfaces, Finite Elements in Analysis and Design, 120 (2016), pp. 18–40.
  • [35] R. Li and Y. Saad, Low-rank correction methods for algebraic domain decomposition preconditioners, SIAM Journal on Matrix Analysis and Applications, 38 (2017), pp. 807–828.
  • [36] P.-L. Lions, On the Schwarz alternating method. I, in First international symposium on domain decomposition methods for partial differential equations, vol. 1, Paris, France, 1988, p. 42.
  • [37] J. W. Milnor and J. D. Stasheff, Characteristic Classes, no. 76 in Annals of Mathematics Studies, Princeton university press, 1974.
  • [38] A. Mohamed and C. Davatzikos, Finite element modeling of brain tumor mass-effect from 3d medical images, in International conference on medical image computing and computer-assisted intervention, Springer, 2005, pp. 400–408.
  • [39] J. R. Munkres, Elementary Differential Topology.(AM-54), Volume 54, vol. 54, Princeton University Press, 2016.
  • [40] J. Nash, The imbedding problem for Riemannian manifolds, Annals of mathematics, (1956), pp. 20–63.
  • [41] J. Nedelec, Curved finite element methods for the solution of singular integral equations on surfaces in r3, Computer Methods in Applied Mechanics and Engineering, 8 (1976), pp. 61–80.
  • [42] J.-C. Nedelec and J. Planchard, Une méthode variationnelle d’éléments finis pour la résolution numérique d’un problème extérieur dans 𝐑3\mathbf{R}^{3}, Revue française d’automatique, informatique, recherche opérationnelle. Analyse numérique, 7 (1973), pp. 105–129.
  • [43] M. A. Olshanskii and A. Reusken, Trace finite element methods for PDEs on surfaces, in Geometrically unfitted finite element methods and applications, Springer, 2017, pp. 211–258.
  • [44] M. A. Olshanskii, A. Reusken, and J. Grande, A finite element method for elliptic equations on surfaces, SIAM Journal on numerical analysis, 47 (2009), pp. 3339–3358.
  • [45] L. Qin and X. Xu, Optimized Schwarz methods with Robin transmission conditions for parabolic problems, SIAM Journal on Scientific Computing, 31 (2008), pp. 608-623.
  • [46] L. Qin, S. Zhang, and Z. Zhang, Finite element formulation in flat coordinate spaces to solve elliptic problems on general closed riemannian manifolds, SIAM Journal on Scientific Computing, 36 (2014), pp. A2149–A2165.
  • [47] A. Reusken, Analysis of trace finite element methods for surface partial differential equations, IMA Journal of Numerical Analysis, 35 (2015), pp. 1568–1590.
  • [48] M. Reuter, Hierarchical shape segmentation and registration via topological features of Laplace-Beltrami eigenfunctions, International Journal of Computer Vision, 89 (2010), pp. 287–308.
  • [49] M. Reuter, S. Biasotti, D. Giorgi, G. Patanè, and M. Spagnuolo, Discrete Laplace–Beltrami operators for shape analysis and segmentation, Computers & Graphics, 33 (2009), pp. 381–390.
  • [50] R. M. Schoen and S.-T. Yau, Lectures on Differential Geometry, vol. 2, International press Cambridge, MA, 1994.
  • [51] H. A. Schwarz, Ueber einige Abbildungsaufgaben, Journal für die Reine und Angewandte Mathematik, 70 (1869), 105-120.
  • [52] A. Toselli and O. Widlund, Domain Decomposition Methods-Algorithms and Theory, vol. 34, Springer Science & Business Media, 2004.
  • [53] J. H. C. Whitehead, On C1C^{1}-complexes, Annals of Mathematics, (1940), pp. 809–824.
  • [54] H. Whitney, Differentiable manifolds, Annals of Mathematics, (1936), pp. 645–680.
  • [55] H. Whitney, The self-intersections of a smooth nn-manifold in 2n2n-space, Annals of Mathematics, (1944), pp. 220–246.
  • [56] J. Xu and L. Zikatanov, The method of alternating projections and the method of subspace corrections in Hilbert space, Journal of the American Mathematical Society, 15 (2002), pp. 573–597.