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

Multilevel Preconditioner with Stable Coarse Grid Corrections for the Helmholtz Equation

Huangxin Chen, Haijun Wu, and Xuejun Xu School of Mathematical Sciences, Xiamen University, Fujian, 361005, P.R. China (chx@xmu.edu.cn).Department of Mathematics, Nanjing University, Jiangsu, 210093, P.R. China (hjw@nju.edu.cn).Institute of Computational Mathematics and Scientific/Engineering Computing, Academy of Mathematics and Systems Science, Chinese Academy of Sciences, P.O. Box 2719, Beijing, 100190, P.R. China (xxj@lsec.cc.ac.cn).
Abstract

In this paper we consider a class of robust multilevel precontioners for the Helmholtz equation with high wave number. The key idea in this work is to use the continuous interior penalty finite element methods (CIP-FEM) studied in [19, 21] to construct the stable coarse grid correction problems. The multilevel methods, based on GMRES smoothing on coarse grids, are then served as a preconditioner in the outer GMRES iteration. In the one dimensional case, convergence property of the modified multilevel methods is analyzed by the local Fourier analysis. From our numerical results, we find that the proposed methods are efficient for a reasonable range of frequencies. The performance of the algorithms depends relatively mildly on wave number. In particular, only one GMRES smoothing step may guarantee the optimal convergence of our multilevel algorithm, which remedies the shortcoming of the multilevel algorithm in [5].

Keywords. Multilevel method, Helmholtz equation, high wave number, continuous penalty finite element method, GMRES method, local Fourier analysis

1 Introduction

The efficient and accurate numerical approximation of high frequency wave propagation is of fundamental importance in many applications such as acoustic, electromagnetic, elasticity and geophysical surveys. When the problem is linear and time-harmonic, it can be typically modeled by Helmholtz equation. The interest of this paper is to consider the Helmholtz equation with Robin boundary condition which is the first order approximation of the radiation condition:

uκ2u\displaystyle-\triangle u-\kappa^{2}u =finΩ,\displaystyle=f\qquad{\rm in}\ \Omega, (1.1)
un+𝐢κu\displaystyle\frac{\partial u}{\partial{n}}+{\bf i}\kappa u =gonΩ,\displaystyle=g\qquad{\rm on}\ \partial\Omega, (1.2)

where Ωd,d=2,3\Omega\subset\mathbb{R}^{d},d=2,3 is a polygonal/polyhedral domain, κ>0\kappa>0 is known as the wave number, 𝐢=1{\bf i}=\sqrt{-1} denotes the imaginary unit, and n{n} denotes the unit outward normal to Ω\partial\Omega.

For high wave number κ\kappa, the linear system from the discretization of Helmholtz equation is usually strongly indefinite, causing most of iterative methods to converge slowly or diverge. In recent years, there have been many advances in the development of iterative methods and preconditioners for the solution of the Helmholtz equation (cf. [10, 11]).

Due to high efficiency of multilevel methods for positive definite problems, more and more attentions have been received to develop robust multilevel methods for Helmholtz equation. However, a direct application of multilevel methods with standard smoothing and coarse grid correction are ineffective. Some strategies have been proposed to remedy the problem (cf. [3, 5, 15, 16]). For instance, Elman, Ernst and O’Leary [5] proposed GMRES smoothing together with flexible GMRES acceleration. But in order to achieve convergence, a relatively large number of GMRES smoothing steps are needed on the intermediate grids. Another approach, the so-called wave-ray multigrid methods [3, 16], was proposed by Brandt and Livshits through defining a meaningful coarse problem by augmenting the standard V-cycle with ray grids and using coarse grid basis derived from plane waves. This method performs well with increasing wave number, but it does not easily generalize to unstructured grids and complicated Helmholtz problems. Alternatively, instead of applying multigrid iterations directly to the Helmholtz equation, a class of shifted Laplacian preconditioners [8, 9] has recently attracted a lot of attention, which precondition the Helmholtz equation with a complex shifted operator and is shown to be an efficient Krylov method preconditioner. We would like to mention that Engquist and Ying [6, 7] recently proposed two new types of sweeping preconditioners for central difference scheme of the Helmholtz equation based on an approximate LDLtLDL^{t} factorization by sweeping the domain layer by layer starting from an absorbing layer. Similar to the wave-ray multigrid methods, the new preconditioners have a nearly linear computational cost and the number of outer iterations is essentially independent of the number of unknowns and the wave number when combined with the GMRES solver.

Our objective is to develop robust multilevel methods for the Helmholtz problem (1.1)-(1.2). Although the pollution error is inherent in the standard finite element methods (FEM) to solve Helmholtz equation, the FEM can still be used on fine grid whose mesh size is sufficiently small to reduce the pollution error. In a recent work [21], the pre-asymptotic analyses of both the FEM and the continuous interior penalty finite element methods (CIP-FEM) are given. In particular, the well-posedness of standard FEM has been proved under the condition of κhpC0(pκ)1p+1\frac{\kappa h}{p}\leq C_{0}(\frac{p}{\kappa})^{\frac{1}{p+1}}, where hh is mesh size, pp is the polynomial order of approximation space, and C0C_{0} is a constant independent of κ,h,p\kappa,h,p. Thus, without this condition, the well-posedness of standard FEM on coarse grids in the multilevel method can not be guaranteed. Moreover, oscillations on the scale of the wavelength can not be resolved well by standard FEM on the coarse grids. By contrast, the CIP-FEM has been proved in [21] that the associated discrete problem is always well-posed without any mesh condition. Intrinsically, the main technique in the stable CIP-FEM is to add a complex shift in the bilinear form, which is similar to the idea of shifted Laplace preconditioner approach. Comparing to adding a shift to the original problem in shifted Laplace operator, the well-posed CIP-FEM is consistent with the original equation and only changes the discrete bilinear form. Based on these observations, the new approach proposed in this work is to apply the CIP-FEM to construct the stable coarse grid correction problems. Standard Jacobi or Gauss-Seidel smoothers become unstable on the coarse grids, especially on the intermediate grids in multilevel methods. Motivated by the smoothing approach presented in [5], the smoothing in this work is to use GMRES method based on CIP-FEM on the coarse grids, and standard Jacobi or Gauss-Seidel relaxation on the fine enough grids. From our numerical experiments, we find that the number of GMRES smoothing steps in our algorithm can be much smaller than that in [5], even if one GMRES smoothing step may guarantee the optimal convergence of our multilevel algorithm.

Our main tool to analyze the multilevel methods for Helmholtz equation is the Local Fourier analysis (LFA), which has been introduced for multigrid analysis by Achi Brandt in 1977 [2]. Comprehensive surveys can be found in [18] and the references therein. We mainly utilize the LFA to analyze smoothing properties of relaxations and convergence properties of two- and three-level methods in one dimensional case. This may provide quantitative insights into the multilevel methods for Helmholtz problem (1.1)-(1.2).

The remaining part of this paper is organized as follows: In section 2, we introduce some notation, recall the formulation of CIP-FEM, and present the multilevel method for the linear system from CIP-FEM approximation. Section 3 is to present the modified multilevel method for Helmholtz problem. Standard FEM is used on fine enough grids, on coarse grids the CIP-FEM is utilized instead. Section 4 is devoted to the LFA of the multilevel method for one dimensional Helmholtz problem, we focus on the smoothing analysis and two- and three-level analysis. In the last section, we give some numerical results to illustrate the performance of the proposed multilevel methods.

2 Notations and Preliminaries

2.1 Formulation of CIP-FEM

Let 𝒯h\mathcal{T}_{h} be a conforming quasi-uniform triangulation of Ω\Omega, and denote the collection of edges/faces by h\mathcal{E}_{h}, while the set of interior edges/faces by hI\mathcal{E}_{h}^{I} and the set of boundary edges/faces by hB\mathcal{E}_{h}^{B}. For any T𝒯hT\in{\mathcal{T}}_{h}, we define hT:=diam(T)h_{T}:={\rm diam}(T). Similarly, for ehe\in{\mathcal{E}}_{h}, define he:=diam(e)h_{e}:={\rm diam}(e). Let h:=maxT𝒯hhTh:=\max_{T\in{\mathcal{T}}_{h}}h_{T}. Throughout this paper we use the standard notations and definitions for Sobolev spaces (cf. [1]). In particular, (,)Q(\cdot,\cdot)_{Q} and ,Σ\left\langle\cdot,\cdot\right\rangle_{\Sigma} for ΣQ\Sigma\subset\partial Q denote the L2L^{2}-inner product on complex-valued L2(Q)L^{2}(Q) and L2(Σ)L^{2}(\Sigma) spaces, respectively. Denote by (,):=(,)Ω(\cdot,\cdot):=(\cdot,\cdot)_{\Omega} and ,:=,Ω\left\langle\cdot,\cdot\right\rangle:=\left\langle\cdot,\cdot\right\rangle_{\partial\Omega}.

Now we define the energy space V:=H1(Ω)T𝒯hH2(T)V:=H^{1}(\Omega)\cap\prod_{T\in{\mathcal{T}}_{h}}H^{2}(T). For any vVv\in V and an interior edge/face e=T1T2e=T_{1}\cap T_{2}, where T1T_{1} and T2T_{2} are two distinct elements of 𝒯h{\mathcal{T}}_{h} with respective outer normals n1n_{1} and n2n_{2}, we introduce the jump vn|e=vn1|T1+vn2|T2\Lbrack\nabla v\cdot n\Rbrack|_{e}=\nabla v\cdot n_{1}|_{T_{1}}+\nabla v\cdot n_{2}|_{T_{2}}. Define the sesquilinear form bh(,)b_{h}(\cdot,\cdot) on V×VV\times V as follows:

bh(u,v):=(u,v)+J(u,v),u,vV,b_{h}(u,v):=(\nabla u,\nabla v)+J(u,v),\qquad u,v\in V,

where

J(u,v):=ehI𝐢γeheun,vne,\displaystyle J(u,v):=\sum_{e\in{\mathcal{E}}_{h}^{I}}{\bf i}\gamma_{e}h_{e}\left\langle\Lbrack\nabla u\cdot n\Rbrack,\Lbrack\nabla v\cdot n\Rbrack\right\rangle_{e}, (2.1)

where 𝐢γe{\bf i}\gamma_{e} is a complex number with positive imaginary part. The terms in J(u,v)J(u,v) are so-called penalty terms and 𝐢γe{\bf i}\gamma_{e} are penalty parameters (cf. [19, 21]).

Clearly, J(u,v)=0J(u,v)=0 if uH2(Ω)u\in H^{2}(\Omega) and vVv\in V. Thus, if uH2(Ω)u\in H^{2}(\Omega) is the solution of (1.1)-(1.2), then there holds

ah(u,v):=bh(u,v)κ2(u,v)+𝐢κu,v=(f,v)+g,v,vV.a_{h}(u,v):=b_{h}(u,v)-\kappa^{2}(u,v)+{\bf i}\kappa\langle u,v\rangle=(f,v)+\langle g,v\rangle,\qquad v\in V. (2.2)

We define the CIP approximation space VhV_{h} as the standard finite element space of order pp, i.e.,

Vh:={vhH1(Ω):vh|T𝒫p(T),T𝒯h},V_{h}:=\big{\{}v_{h}\in H^{1}(\Omega):v_{h}|_{T}\in\mathcal{P}_{p}(T),\,T\in{\mathcal{T}}_{h}\big{\}},

where 𝒫p(T)\mathcal{P}_{p}(T) is the space of polynomials of degree at most pp on TT. The CIP finite element approximation is to find uhVhu_{h}\in V_{h} such that

ah(uh,vh)=(f,vh)+g,vh,vhVh.a_{h}(u_{h},v_{h})=(f,v_{h})+\langle g,v_{h}\rangle,\qquad v_{h}\in V_{h}. (2.3)

It is clear that if the parameters γe0\gamma_{e}\equiv 0, then the CIP-FEM reduces to the standard FEM. It has been proved that the CIP-FEM is stable for any κ,h,p>0\kappa,h,p>0 [19, 21]. The penalty parameters may be tuned to reduce the pollution errors. The numerical results in [19] show that using about the same total degrees of freedom (DOFs), the CIP-FEM yields the least phase error comparing to the standard FEM and IPDG method [12].

2.2 Multilevel methods for CIP-FEM

Let {𝒯l}l=0L\{\mathcal{T}_{l}\}_{l=0}^{L} be a shape regular family of nested geometrically conforming simplicial triangulations of the computational domain Ω\Omega obtained by successive quasi-uniform refinement of an intentionally chosen coarse grid 𝒯0\mathcal{T}_{0}. We denote by VlV_{l} the CIP approximation space on 𝒯l{\mathcal{T}}_{l}. It is easy to see that the spaces {Vl}l=0L\{V_{l}\}_{l=0}^{L} are nested, i.e., V0V1VLV_{0}\subset V_{1}\subset\cdots\subset V_{L}. But the bilinear forms {al(,)}l=0L\{a_{l}(\cdot,\cdot)\}_{l=0}^{L} defined as in (2.3) on each VlV_{l} are nonnested. Thus in this work we consider the following multilevel methods for CIP-FEM discretizations, which has also been used for solving the linear systems from nonconforming P1 finite element approximations (cf. [17]).

For brevity, we denote by al(,)a_{l}(\cdot,\cdot) the bilinear form ahl(,)a_{h_{l}}(\cdot,\cdot) on VlV_{l}, where hlh_{l} is the mesh size of 𝒯l{\mathcal{T}}_{l}. Define projections PlCP^{C}_{l}, QlQ_{l} : VLVlV_{L}\rightarrow V_{l} according to

al(PlCv,w)=aL(v,w),(Qlv,w)=(v,w),vVL,wVl.a_{l}(P^{C}_{l}v,w)=a_{L}(v,w),\quad(Q_{l}v,w)=(v,w),\quad\forall v\in V_{L},\,w\in V_{l}.

The existence and uniqueness of the discrete problem (2.3) imply the well-posedness of each PlCP^{C}_{l}. For 0lL0\leq l\leq L, we also define AlC:VlVlA^{C}_{l}:V_{l}\rightarrow V_{l} by means of

(AlCv,w)=al(v,w),v,wVl.\displaystyle(A^{C}_{l}v,w)=a_{l}(v,w),\quad\forall v,w\in V_{l}. (2.4)

Define FlVlF_{l}\in V_{l} by

(Fl,v)=(f,v)+g,vvVl.(F_{l},v)=(f,v)+\langle g,v\rangle\quad\forall v\in V_{l}.

Then the CIP-FEM on level ll (cf. (2.3)) can be rewritten as:

AlCulC=Fl.A^{C}_{l}u^{C}_{l}=F_{l}. (2.5)

For the smoothing strategy, in fact, when κhl/p\kappa h_{l}/p is small enough, either of weighted Jacobi and Gauss-Seidel relaxation can be applied. Otherwise, GMRES relaxation can be used as a smoother and this will be explained in the following sections. To be precise, we describe the smoothing strategy as follows:

Definition 2.1.

Let R0C=(A0C)1R_{0}^{C}=(A_{0}^{C})^{-1}. Given α>0\alpha>0 and 0<lL0<l\leq L, let SL={l:κhl/p<α, 1lL}S_{L}=\{l:\ \kappa h_{l}/p<\alpha,\ 1\leq l\leq L\} and GL={l:1lL}SLG_{L}=\{l:1\leq l\leq L\}\setminus S_{L}. If lSLl\in S_{L}, let RlCR_{l}^{C} be the weighted Jacobi relaxation Rl,CJR^{J}_{l,C} or Gauss-Seidel relaxation Rl,CGSR^{GS}_{l,C} based on AlCA^{C}_{l}. Otherwise, when lGLl\in G_{L} we use the GMRES relaxation based on AlCA^{C}_{l}.

We remark that we choose α=0.5\alpha=0.5 in this paper. This choice is motivated by the efficient relaxation of Jacobi and Gauss-Seidel smoothers, and it ensures that the amplification factor for the Jacobi and Gauss-Seidel smoothers will not become too large (cf. section 4.3 in the following and section 2.1 in [5]). When lSLl\in S_{L}, we also define the operator (RlC)t(R_{l}^{C})^{t} by

((RlC)tw¯,v¯)=(RlCv,w),v,wVl.\displaystyle\big{(}(R_{l}^{C})^{t}\bar{w},\bar{v}\big{)}=\big{(}R_{l}^{C}v,w\big{)},\quad\forall v,w\in V_{l}. (2.6)

We can now state the multilevel method for solving the CIP-FEM discretization system on level LL which is non-recursive version of multilevel method.

Algorithm 2.1.

Given an arbitrarily chosen initial iterate u0VLu^{0}\in V_{L}, we seek unVLu^{n}\in V_{L} as follows:

Let v0=un1v_{0}=u^{n-1}.

  • 1)1)

    For l=0,1,,Ll=0,1,\cdots,L. When l=0l=0 or lSLl\in S_{L},

    vl+1=vl+μlRlCQl(FLALCvl).v_{l+1}=v_{l}+\mu_{l}R_{l}^{C}Q_{l}(F_{L}-A^{C}_{L}v_{l}).

    Otherwise, perform GMRES relaxation for the correction problem AlCwl=Ql(FLALCvl)A^{C}_{l}w_{l}=Q_{l}(F_{L}-A^{C}_{L}v_{l}) and set vl+1=vl+μlwlv_{l+1}=v_{l}+\mu_{l}w_{l}. Here μl>0\mu_{l}>0 is a scaling parameter to weaken the influence of the error during prolongations. In this paper, we always set μl0.5\mu_{l}\equiv 0.5.

  • 2)2)

    For l=L,,1,0l=L,\cdots,1,0. When l=0l=0 or lSLl\in S_{L},

    v2L+2l=v2L+1l+μl(RlC)tQl(FLALCv2L+1l).\displaystyle v_{2L+2-l}=v_{2L+1-l}+\mu_{l}(R_{l}^{C})^{t}Q_{l}(F_{L}-A^{C}_{L}v_{2L+1-l}).

    Otherwise, perform GMRES relaxation based on v2L+1lv_{2L+1-l} as in step 1 to get v2L+2lv_{2L+2-l}.

  • 3)3)

    Set un=v2L+2u^{n}=v_{2L+2}.

Here we note that the GMRES relaxation is not linear in the starting value, and the error operator of the above algorithm can not be written directly in a product form which holds only for the particular case GL=G_{L}=\emptyset. When this particular case is considered reasonably, we set

Tl:=μlRlCAlCPlC and Tl:=μl(RlC)tAlCPlC,l=0,1,,L,GL=,\displaystyle T_{l}:=\mu_{l}R_{l}^{C}A^{C}_{l}P^{C}_{l}\text{ and }T_{l}^{*}:=\mu_{l}(R_{l}^{C})^{t}A^{C}_{l}P^{C}_{l},\quad l=0,1,\ldots,L,\ G_{L}=\emptyset,

Then the error operator of Algorithm 2.1 for the case GL=G_{L}=\emptyset can be derived as

EMEM, where EM:=(ITL)(IT1)(IT0),EM:=(IT0)(IT1)(ITL),\displaystyle E_{M}^{*}E_{M},\text{ where }E_{M}:=(I-T_{L})\cdots(I-T_{1})(I-T_{0}),E_{M}^{*}:=(I-T_{0}^{*})(I-T_{1}^{*})\cdots(I-T_{L}^{*}), (2.7)

where II is the identity operator in VLV_{L}.

3 Modified multilevel methods for standard FEM

In general, in order to reduce the pollution error, 6-10 grid points per wavelength are typically chosen to yield reasonable accuracy. In a unit square domain [0,1]2[0,1]^{2}, it is well know that κh=2π/nw\kappa h=2\pi/n_{w}, where nwn_{w} is the number of points per wavelength (cf. [14]), implies the deterioration of κh\kappa h for increasing grid points per wavelength. In theory, the well-posedness of discrete solution by CIP-FEM holds for any κ,h\kappa,h and pp, but it is not guaranteed for discrete solution by standard FEM on coarser grids with κh/pC\kappa h/p\geq C particularly [19, 21], where the constant CC is independent of κ,h\kappa,h and pp. Therefore, in order to establish stable coarse grid correction problems in multilevel methods for Helmholtz equation, the CIP-FEM will be applied on the coarse grids. Besides, for standard FEM, eigenvalues of discrete system close to the origin may undergo a sign change after discretization on a coarser grid. If a sign change occurs, the coarse grid correction does not give a convergence acceleration to the finer grid problem but gives a severe convergence degradation instead. This is analyzed in [5] and a remedy combining GMRES method as a smoother on coarse grids is proposed. This idea has been applied in Algorithm 2.1, and we will also utilize this strategy in the following modified multilevel methods to get more efficient smoother for indefinite discrete systems.

For brevity, let AlFA^{F}_{l} denote the linear operator AlCA^{C}_{l} with the parameters γe0\gamma_{e}\equiv 0, i.e., the linear operator for standard FEM, and energy operator PlF=PlC|γe0P^{F}_{l}=P^{C}_{l}|_{\gamma_{e}\equiv 0}. The discrete problem on level ll is AlCulC=FlA^{C}_{l}u^{C}_{l}=F_{l} for CIP-FEM (cf. (2.5)) or AlFulF=FlA^{F}_{l}u^{F}_{l}=F_{l} for standard FEM. When the mesh size is sufficiently small to reduce the pollution error and satisfy the accuracy requirement, both CIP-FEM and standard FEM can be utilized. However, the nonzero elements of linear system from standard FEM may be less than that from CIP-FEM, and when the grid is fine enough, the pollution error by the standard FEM is also small. Thus, we may still apply the standard FEM on the fine grids.

Similar to Definition 2.1, we use the smoothing strategy on each VlV_{l} as follows.

Definition 3.1.

Let R0=(A0C)1R_{0}=(A_{0}^{C})^{-1}. Given α>0\alpha>0 and 0<lL0<l\leq L. If lSLl\in S_{L}, let RlR_{l} be the weighted Jacobi relaxation Rl,FJR^{J}_{l,F} or Gauss-Seidel relaxation Rl,FGSR^{GS}_{l,F} based on AlFA^{F}_{l}. Otherwise, when lGLl\in G_{L} we use the GMRES relaxation based on AlCA^{C}_{l}.

When lSLl\in S_{L}, the operator RltR_{l}^{t} is defined similarly as in (2.6). Now we state the modified multilevel method for the discrete system from standard FEM for the Helmholtz problem (1.1)-(1.2) as follows.

Algorithm 3.1.

Given an arbitrarily chosen initial iterate u0VLu^{0}\in V_{L} and integers m1,m20m_{1},m_{2}\geq 0, we seek unVLu^{n}\in V_{L} as follows:

Let v0=un1v_{0}=u^{n-1}.

  • 1)1)

    For l=0,1,,Ll=0,1,\cdots,L. Given μl>0\mu_{l}>0 is chosen as in Algorithm 2.1, when l=0l=0 or lSLl\in S_{L},

    vl+1=vl+μl(Rl)mQl(FLALFvl),v_{l+1}=v_{l}+\mu_{l}(R_{l})^{m}Q_{l}(F_{L}-A^{F}_{L}v_{l}),

    where m=m1m=m_{1} if l>0l>0, m=1m=1 if l=0l=0. Otherwise, perform m1m_{1} steps of GMRES relaxation for the correction problem AlCwl=Ql(FLALFvl)A^{C}_{l}w_{l}=Q_{l}(F_{L}-A^{F}_{L}v_{l}) and set vl+1=vl+μlwlv_{l+1}=v_{l}+\mu_{l}w_{l}.

  • 2)2)

    For l=L,,1,0l=L,\cdots,1,0. When l=0l=0 or lSLl\in S_{L},

    v2L+2l=v2L+1l+μl(Rlt)mQl(FLALFv2L+1l),\displaystyle v_{2L+2-l}=v_{2L+1-l}+\mu_{l}(R_{l}^{t})^{m}Q_{l}(F_{L}-A^{F}_{L}v_{2L+1-l}),

    where m=m1m=m_{1} if l>0l>0, m=1m=1 if l=0l=0. Otherwise, perform m2m_{2} steps of GMRES relaxation for the correction problem AlCwl=Ql(FLALFv2L+1l)A^{C}_{l}w_{l}=Q_{l}(F_{L}-A^{F}_{L}v_{2L+1-l}) and set v2L+2l=v2L+1l+μlwlv_{2L+2-l}=v_{2L+1-l}+\mu_{l}w_{l}.

  • 3)3)

    Set un=v2L+2u^{n}=v_{2L+2}.

Remark 3.1.

In [5], in order to prevent unnecessary damping of smoothing modes which should be handled by the coarse grid correction, the postsmoothing is always favored over presmoothing (cf. section 3.2 in [5]). This is also true in our work. However, due to the utilization of stable coarse grid correction method, the above algorithm will converge even when the smoothing is chosen as one step in both post- and presmoothing.

Remark 3.2.

Comparing to Algorithm 2.1, the CIP-FEM is only applied in the coarse grid correction when lGLl\in G_{L} in the above algorithm. From the first numerical example in this paper, we can see that the convergence property of this two algorithms is similar. Actually, this phenomena can also be observed in the following LFA. In order to reduce computations, one may prefer Algorithm 3.1 in practice.

4 The one dimensional local Fourier analysis

In this section, we aim to analyze different approaches based on Algorithm 2.1 for the discrete system from one dimensional Helmholtz equation, where standard FEM is utilized on the finest grid. We focus on the analysis for the discretization from linear continuous interior penalty finite element method (CIP-P1) by the important tool LFA in multigrid analysis. Here we mainly focus on the analysis of two-level methods. The LFA of three-level methods is also mentioned. The analysis will imply the efficiency of Algorithm 3.1. The following presentation is related to the notation and philosophy from [18].

4.1 Basic tools in one dimensional local Fourier analysis

The LFA is based on certain idealized assumptions and simplifications: the boundary conditions are neglected and the problem is considered on regular indefinite grids 𝒢h={x:x=xj=jh,j}{\mathcal{G}}_{h}=\{x:x=x_{j}=jh,j\in\mathbb{Z}\}. Although the Robin boundary condition (1.2) and other absorbing boundary conditions are often applied in realistic Helmholtz problem, the neglect of boundary condition does usually not affect the validity of LFA (cf. [18]). The LFA in this section can be considered as simplification for the analysis of multilevel method for Helmholtz problem (1.1-1.2) in one dimension.

Let 𝑳h{\boldsymbol{L}}_{h} be a discrete operator with a stencil representation 𝑳h=[lj]h,j{\boldsymbol{L}}_{h}=[l_{j}]_{h},j\in\mathbb{Z}. For any uhu_{h} defined on 𝒢h{\mathcal{G}}_{h} and a fixed point x𝒢hx\in{\mathcal{G}}_{h}, 𝑳huh{\boldsymbol{L}}_{h}u_{h} reads in stencil notation as

𝑳huh(x)=jJljuh(x+jh),\displaystyle{\boldsymbol{L}}_{h}u_{h}(x)=\sum_{j\in J}l_{j}u_{h}(x+jh),

where JJ\subset\mathbb{Z} is a certain finite index defining the so-called stencil. The basic quantities in the LFA are the Fourier modes φh(θ,x)=e𝐢θx/h\varphi_{h}(\theta,x)=e^{{\bf i}\theta x/h} with x𝒢hx\in{\mathcal{G}}_{h} and Fourier frequency θ\theta\in\mathbb{R}. In fact, the frequency θ\theta can be restricted to the interval (π,π](-\pi,\pi] as a fact that φh(θ+2π,x)=φh(θ,x)\varphi_{h}(\theta+2\pi,x)=\varphi_{h}(\theta,x). It is easy to see that the Fourier modes are all the formal eigenfunctions of 𝑳h{\boldsymbol{L}}_{h}:

𝑳hφh(θ,x)=𝑳~h(θ)φh(θ,x),x𝒢h,θ(π,π],\displaystyle{\boldsymbol{L}}_{h}\varphi_{h}(\theta,x)=\widetilde{{\boldsymbol{L}}}_{h}(\theta)\varphi_{h}(\theta,x),\qquad x\in{\mathcal{G}}_{h},\theta\in(-\pi,\pi], (4.1)

where the eigenvalues of 𝑳h{\boldsymbol{L}}_{h} can be presented as 𝑳~h(θ)=jJlje𝐢θj\widetilde{{\boldsymbol{L}}}_{h}(\theta)=\sum_{j\in J}l_{j}e^{{\bf i}\theta j}, which is called the Fourier symbol of the operator 𝑳h{\boldsymbol{L}}_{h}. Given a so-called low frequency θ0Θlow=(π/2,π/2]\theta^{0}\in\Theta_{\rm low}=(-\pi/2,\pi/2], its complementary frequency θ1\theta^{1} is defined as

θ1=θ0sign(θ0)π.\displaystyle\theta^{1}=\theta^{0}-{\rm sign}(\theta^{0})\pi. (4.2)

Interpreting the Fourier modes as coarse grid functions yields

φh(θ0,x)=φ2h(2θ0,x)=φ2h(2θ1,x)=φh(θ1,x),θ0Θlow,x𝒢2h.\varphi_{h}(\theta^{0},x)=\varphi_{2h}(2\theta^{0},x)=\varphi_{2h}(2\theta^{1},x)=\varphi_{h}(\theta^{1},x),\qquad\theta^{0}\in\Theta_{\rm low},\ x\in{\mathcal{G}}_{2h}.

The Fourier modes φh(θ0,x)\varphi_{h}(\theta^{0},x) and φh(θ1,x)\varphi_{h}(\theta^{1},x) are called 2h2h-harmonics. These Fourier modes coincide on the coarse grid with mesh size H=2hH=2h, and they can be represented by a single coarse grid mode φ2h(2θ0,x)\varphi_{2h}(2\theta^{0},x). Hence, each low frequency mode is associated with a high frequency mode. For a given θ0Θlow\theta^{0}\in\Theta_{\rm low}, define the two dimensional subspace of 2h2h-harmonics by

E2hθ0:=span{φh(θ0,x),φh(θ1,x)},\displaystyle E_{2h}^{\theta^{0}}:={\rm span}\{\varphi_{h}(\theta^{0},x),\varphi_{h}(\theta^{1},x)\}, (4.3)

where θ1\theta^{1} is defined as in (4.2). A crucial observation is that the space E2hθ0E_{2h}^{\theta^{0}} is invariant under both smoothing operators and correction schemes for general cases by two-level method. The invariance property holds for many well-known smoothing methods (cf. [18]), such as Jacobi relaxation, lexicographical Gauss-Seidel relaxation, et al.

Let 𝑴h{\boldsymbol{M}}_{h} be a discrete two-level operator. In the following, we will show that a block-diagonal representation for 𝑴h{\boldsymbol{M}}_{h} consists of 2×22\times 2 blocks 𝑴~h(θ)\widetilde{{\boldsymbol{M}}}_{h}(\theta) (cf. [18]), which denotes the representation of 𝑴h{\boldsymbol{M}}_{h} on E2hθ0E_{2h}^{\theta^{0}}. Then the convergence factor of 𝑴h{\boldsymbol{M}}_{h} by the LFA is defined as follows:

ρ(𝑴h)=sup{ρ(𝑴~h(θ)):θΘlow},\rho({\boldsymbol{M}}_{h})={\rm sup}\{\rho(\widetilde{{\boldsymbol{M}}}_{h}(\theta)):\ \theta\in\Theta_{\rm low}\},

where ρ(𝑴~h(θ))\rho(\widetilde{{\boldsymbol{M}}}_{h}(\theta)) is the spectral radius of the matrix 𝑴~h(θ)\widetilde{{\boldsymbol{M}}}_{h}(\theta). We can refer to [18] for generalizations to kk-level analysis.

4.2 One dimensional Fourier symbols

Now we give the Fourier symbols of different operators in multilevel method for CIP-P1 discretization of one dimensional Helmholtz equation (1.1). We always assume γeγ\gamma_{e}\equiv\gamma for some constant γ\gamma in (2.1). Denote by t=κht=\kappa h, R=1𝐢4γt2/6R=-1-{\bf i}4\gamma-t^{2}/6, S=1+𝐢3γt2/3S=1+{\bf i}3\gamma-t^{2}/3. Since the boundary condition is neglected in the LFA, the stencil presentation of discretization operator 𝑨hC{\boldsymbol{A}}^{C}_{h} from (2.4) can be derived as (cf. [20])

𝑨hC=1h[𝐢γR 2SR𝐢γ]h.\displaystyle{\boldsymbol{A}}^{C}_{h}=\frac{1}{h}[{\bf i}\gamma\ \ R\ \ 2S\ \ R\ \ {\bf i}\gamma]_{h}.

Combining the above expression and (4.1) yields the Fourier symbol of 𝑨hC{\boldsymbol{A}}^{C}_{h} as

𝑨~hC(θ)=1h(𝐢2γcos2θ+2Rcosθ+2S).\displaystyle\widetilde{{\boldsymbol{A}}}^{C}_{h}(\theta)=\frac{1}{h}({\bf i}2\gamma\cos{2\theta}+2R\cos{\theta}+2S). (4.4)

Obviously, the Fourier symbol of standard FEM with piecewise P1 approximation (FEM-P1) is 𝑨~hF(θ)=𝑨~hC(θ)|γ=0\widetilde{{\boldsymbol{A}}}^{F}_{h}(\theta)=\widetilde{{\boldsymbol{A}}}^{C}_{h}(\theta)|_{\gamma=0}.

For simplicity, we use standard weighted Jacobi (ω\omega-JAC) and lexicographical Gauss-Seidel (GS-LEX) relaxations as the smoothers in the LFA. It is easy to derive the weighted Jacobi relaxation matrix as 𝑺hJ=𝑰hω𝑫h1𝑨hC{\boldsymbol{S}}^{J}_{h}={\boldsymbol{I}}_{h}-\omega{\boldsymbol{D}}^{-1}_{h}{\boldsymbol{A}}^{C}_{h}, where 𝑰h{\boldsymbol{I}}_{h} is indentity matrix, 𝑫h{\boldsymbol{D}}_{h} consists of the diagonal of 𝑨hC{\boldsymbol{A}}^{C}_{h} and ω\omega is a weighted parameter. Due to the fact that 𝑫h=2Sh𝑰h{\boldsymbol{D}}_{h}=\frac{2S}{h}{\boldsymbol{I}}_{h}, one can easily deduce the Fourier symbol of weighted Jacobi relaxation as follows:

𝑺~hJ(θ)=1ωS(𝐢γcos2θ+Rcosθ+S).\displaystyle\widetilde{{\boldsymbol{S}}}^{J}_{h}(\theta)=1-\frac{\omega}{S}({\bf i}\gamma\cos{2\theta}+R\cos{\theta}+S). (4.5)

The GS-LEX relaxation matrix is 𝑺hGS=(𝑫h𝑳h)1𝑼h{\boldsymbol{S}}^{GS}_{h}=({\boldsymbol{D}}_{h}-{\boldsymbol{L}}_{h})^{-1}{\boldsymbol{U}}_{h}, where 𝑳h-{\boldsymbol{L}}_{h} is the strictly lower triangular part of 𝑨hC{\boldsymbol{A}}^{C}_{h} and 𝑼h-{\boldsymbol{U}}_{h} is the strictly upper triangular part of 𝑨hC{\boldsymbol{A}}^{C}_{h}. The Fourier symbol of 𝑺hGS{\boldsymbol{S}}^{GS}_{h} can also be directly derived that

𝑺~hGS(θ)=Re𝐢θ+𝐢γe𝐢2θRe𝐢θ+𝐢γe𝐢2θ+2S.\displaystyle\widetilde{{\boldsymbol{S}}}^{GS}_{h}(\theta)=-\frac{Re^{{\bf i}\theta}+{\bf i}\gamma e^{{\bf i}2\theta}}{Re^{{\bf-i}\theta}+{\bf i}\gamma e^{{\bf-i}2\theta}+2S}. (4.6)

Note that for the restriction matrix 𝑰h2h=[rj]h2h{\boldsymbol{I}}^{2h}_{h}=[r_{j}]^{2h}_{h} and x𝒢2hx\in{\mathcal{G}}_{2h}, there holds

(𝑰h2hφh(θα,))(x)=jJrje𝐢jθαφh(θα,x)=jJrje𝐢jθαφ2h(2θ0,x),α=0,1.({\boldsymbol{I}}^{2h}_{h}\varphi_{h}(\theta^{\alpha},\cdot))(x)=\sum_{j\in J}r_{j}e^{{\bf i}j\theta^{\alpha}}\varphi_{h}(\theta^{\alpha},x)=\sum_{j\in J}r_{j}e^{{\bf i}j\theta^{\alpha}}\varphi_{2h}(2\theta^{0},x),\qquad\alpha=0,1.

By an analogous stencil argument, the stencil presentation of full weighting restriction matrix can be derived to be 𝑰h2h=[1/4,1/2,1/4]h2h{\boldsymbol{I}}^{2h}_{h}=[1/4,1/2,1/4]_{h}^{2h}. Then one can obtain the Fourier symbol of 𝑰h2h{\boldsymbol{I}}^{2h}_{h} is

𝑰~h2h(θ)=12(1+cosθ).\widetilde{{\boldsymbol{I}}}^{2h}_{h}(\theta)=\frac{1}{2}(1+\cos{\theta}).

For the linear prolongation matrix 𝑰2hh{\boldsymbol{I}}^{h}_{2h}, one can also derive its Fourier symbol as follows (cf. [18]):

𝑰~2hh(θ)=12(1+cosθ).\widetilde{{\boldsymbol{I}}}^{h}_{2h}(\theta)=\frac{1}{2}(1+\cos{\theta}).

4.3 Smoothing analysis

Weighted Jacobi and lexicographical Gauss-Seidel relaxations are the general smoothing operators. It is well-known that such two smoothers are unstable especially for linear systems from indefinite Helmholtz equation by standard FEM approximations. This is caused by negative eigenvalues of the associated linear system and divergence occurs under such two smoothers. To make up the problem, an improvement is introduced by the shifted Laplacian preconditioner [8, 9, 10], which preconditions the Helmholtz equation with a complex shifted operator

Δ(1+𝐢β)k2,\displaystyle-\Delta-(1+{\bf i}\beta)k^{2}, (4.7)

where β\beta are free parameter. The use of a shift is important to guarantee multilevel convergence, whereas the multilevel method for the shifted operator only converges for a sufficiently large shift. But this contradicts the fact that the outer Krylov acceleration prefers a small shift. Based on the idea of adding a shift to the original problem, we consider the stable CIP-FEM which is consistent with original equation and adds a complex shift in the bilinear form.

Recalling that every two dimensional subspace of 2h2h-harmonics E2hθ0E^{\theta^{0}}_{2h} with θ0Θlow\theta^{0}\in\Theta_{\rm low} is left invariant under the ω\omega-JAC and GS-LEX relaxations, then the Fourier representation of smoother 𝑺h=𝑺hJ{\boldsymbol{S}}_{h}={\boldsymbol{S}}^{J}_{h} or 𝑺hGS{\boldsymbol{S}}^{GS}_{h} with respect to E2hθ0E^{\theta^{0}}_{2h} can be written as

[𝑺~h(θ0)00𝑺~h(θ1)],\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{S}}}_{h}(\theta^{0})&0\\ 0&\widetilde{{\boldsymbol{S}}}_{h}(\theta^{1})\\ \end{array}\right], (4.8)

where 𝑺~h(θ)\widetilde{{\boldsymbol{S}}}_{h}(\theta) is the smoother symbol derived in (4.5) and (4.6). The spectral radius of the smoother operator can be easily calculated since the above matrix is diagonal.

For simplicity, we concern on the weighted Jacobi relaxation. The frequency θ\theta maximizing |𝑺~hJ(θ)||\widetilde{{\boldsymbol{S}}}^{J}_{h}(\theta)| over (π,π](-\pi,\pi] can be calculated by its first and second derivatives, which reveal θ=0\theta=0 or θ=π\theta=\pi maximizing |𝑺~hJ(θ)||\widetilde{{\boldsymbol{S}}}^{J}_{h}(\theta)|. Hence, the spectral radius of 𝑺hJ{\boldsymbol{S}}^{J}_{h} can be deduced as follows: ρ(𝑺hJ)=max{|𝑺~hJ(0)|,|𝑺~hJ(π)}|,\rho({\boldsymbol{S}}^{J}_{h})=\max\{|\widetilde{{\boldsymbol{S}}}^{J}_{h}(0)|,|\widetilde{{\boldsymbol{S}}}^{J}_{h}(\pi)\}|, where

|𝑺~hJ(0)|=|1ωωS(𝐢γ+R)|,|𝑺~hJ(π)|=|1ωωS(𝐢γR)|.\displaystyle|\widetilde{{\boldsymbol{S}}}^{J}_{h}(0)|=|1-\omega-\frac{\omega}{S}({\bf i}\gamma+R)|,\quad|\widetilde{{\boldsymbol{S}}}^{J}_{h}(\pi)|=|1-\omega-\frac{\omega}{S}({\bf i}\gamma-R)|.

Since the parameter γ\gamma in CIP-FEM may influence the pollution error, it is critical to make a suitable choice. For the case t=κh1t=\kappa h\leq 1, it has been derived in [20] that for one dimensional problem there exists an optimal choice 𝐢γo=6cost6+t2cost+2t212(1cost)2{\bf i}\gamma_{o}=\frac{6\cos{t}-6+t^{2}\cos{t}+2t^{2}}{12(1-\cos{t})^{2}} such that the pollution error vanishes when the penalty parameter is chosen as |𝐢γ𝐢γo|Cκ2h|{\bf i}\gamma-{\bf i}\gamma_{o}|\leq\frac{C}{\kappa^{2}h}, where the constant CC is independent of κ\kappa and hh. The left graph of Figure 1 shows the Fourier symbols 𝑺~hJ(θ)\widetilde{{\boldsymbol{S}}}^{J}_{h}(\theta) for ω\omega-JAC smoother with γ=γo\gamma=\gamma_{o} and ω=0.6\omega=0.6. We find that 𝑺~hJ(θ)1\widetilde{{\boldsymbol{S}}}^{J}_{h}(\theta)\geq 1 always occur at the low frequencies, and small tt leads to a better relaxation. Similar phenomenon is observed for GS-LEX smoother in the right graph of Figure 1. Thus, for fixed wave number κ\kappa, both ω\omega-JAC and GS-LEX relaxations can be used as smoother on fine grid. Actually, the relaxation properties of these two smoothers are similar when 𝐢γ{\bf i}\gamma is a complex number.

Refer to caption
Refer to caption
Figure 1: |𝑺~hJ(θ)||\widetilde{{\boldsymbol{S}}}^{J}_{h}(\theta)| with ω=0.6\omega=0.6 (left) and |𝑺~hGS(θ)||\widetilde{{\boldsymbol{S}}}^{GS}_{h}(\theta)| (right) over (π,π](-\pi,\pi] for t=0.1,0.5,1t=0.1,0.5,1.

When ω\omega-JAC smoother is utilized in the standard P1 FEM approximation, then

ρ(𝑺hJ)|γ=0=max{1ω(12t2)62t2,1+3ωt262t2}.\displaystyle\rho({\boldsymbol{S}}^{J}_{h})|_{\gamma=0}=\max\big{\{}\mid 1-\frac{\omega(12-t^{2})}{6-2t^{2}}\mid,\mid 1+\frac{3\omega t^{2}}{6-2t^{2}}\mid\big{\}}. (4.9)

Thus, for t=κht=\kappa h near 3\sqrt{3}, the error is amplified by the smoothing. For CIP-FEM, there is a (complex) shift in the bilinear form, therefore the amplification factor is reduced. This may permit the use of ω\omega-JAC smoother again. In fact, in our modified multilevel method, GMRES iteration is used on the intermediate grids (cf. [5]). In contrast with ω\omega-JAC and GS-LEX relaxations, the Fourier symbol can not be derived for GMRES smoothing. In the following, we will give some explanations for the performance of GMRES smoothing from the numerical approach.

Refer to caption
Figure 2: Amplification factor of GS, Jacobi and GMRES smoothing for t=0.8t=0.8.

For simplification, we consider the one dimensional Helmholtz equation on an interval (0,10)(0,10) with homogeneous Dirichlet boundary conditions. The piecewise P1 CIP-FEM discretization leads to a linear system with N×NN\times N coefficient matrix (cf. [20])

𝑨hC=1h(2S𝐢γR𝐢γR2SR𝐢γ𝐢γR2SR𝐢γ𝐢γR2SR𝐢γ𝐢γR2SR𝐢γR2S𝐢γ)N×N,\displaystyle{\boldsymbol{A}}^{C}_{h}=\frac{1}{h}\left(\begin{array}[]{ccccccc}2S-{\bf i}\gamma&R&{\bf i}\gamma&&\\ R&2S&R&{\bf i}\gamma&\\ {\bf i}\gamma&R&2S&R&{\bf i}\gamma\\ &\ddots&\ddots&\ddots&\ddots&\ddots&\\ &&{\bf i}\gamma&R&2S&R&{\bf i}\gamma\\ &&&{\bf i}\gamma&R&2S&R\\ &&&&{\bf i}\gamma&R&2S-{\bf i}\gamma\\ \end{array}\right)_{N\times N}, (4.17)

where N=10/h1N=10/h-1. For κ=200\kappa=200, we apply the grid with mesh size h=0.004h=0.004, i.e., t=0.8t=0.8, and choose γ=γo\gamma=\gamma_{o}. Let the vector u0=e𝐢θ𝒙/hu_{0}=e^{{\bf i}\theta{\boldsymbol{x}}/h} be an initial choice for smoothing, where 𝒙=[x1,,xN1],xk=xk1+h,x0=0,k=1,,N1{\boldsymbol{x}}=[x_{1},\cdots,x_{N-1}],x_{k}=x_{k-1}+h,x_{0}=0,k=1,\cdots,N-1, θ(π,π]\theta\in(-\pi,\pi]. We assume that 𝑺h{\boldsymbol{S}}_{h} is the relaxation iteration matrix and u1=𝑺hu0u_{1}={\boldsymbol{S}}_{h}u_{0} is the new vector after one step of smoothing. Then for fixed θ\theta, we obtain the amplification factor for one step of smoothing ρs(θ)=u1/u0\rho_{s}(\theta)=\|u_{1}\|/\|u_{0}\|, where \|\cdot\| stands for the Euclidean norm. Figure 2 shows the amplification factors of three different kinds of smoothing strategies: GS, Jacobi and GMRES relaxations. We can see that for high and low frequencies θ\theta, the amplification factor of GMRES relaxation is always smaller than that of GS and Jacobi relaxations. The GMRES relaxation can still be convergent even when the other two smoothers lead to a divergent relaxation. The similar phenomenon can also be observed for other choices of tt and γ\gamma. Therefore, when standard Jacobi or GS relaxation fails as a smoother, we can replace this with the GMRES smoothing.

From the above analysis, we can see that for the relaxation of linear system for Helmholtz problem, the standard Jacobi or GS relaxation can be performed on the fine grids, but they fail on the coarse grids. In particular, the GMRES relaxation is efficient for smoothing on the intermediate coarse grids. In the next section, the correction scheme will be taken into account to obtain a more realistic convergence analysis of the multilevel method.

4.4 Two- and three-level local Fourier analysis

For more details about convergence property of multilevel method in Algorithm 3.1, we will focus on the LFA of two-level method. The LFA of three-level method will also be mentioned. For simplicity, we consider the case without postsmoothing. Since the Fourier symbol can not be obtained for GMRES smoothing, we only consider the two and three level methods with ω\omega-JAC or GS-LEX relaxation. Then the iteration operator of Algorithm 3.1 in this case is the nonsymmetric version and can be derived as in (2.7). Three cases of iteration operator will be analyzed in the following: multilevel methods by shifted Laplace approach, multilevel methods with stable CIP-FEM coarse grid corrections, and multilevel methods with stable CIP-FEM for fine grid smoothing and coarse grid corrections.

The iteration operator of two-level method can be written as (IT1)(IT0)(I-T_{1})(I-T_{0}), and the operator TlT_{l} is chosen for the corresponding algorithm. Since standard FEM is applied on the finest grid, we have Tl=μlRlAlFPlFT_{l}=\mu_{l}R_{l}A^{F}_{l}P^{F}_{l}, where RlR_{l} is smoother determined by the algorithm. When CIP-FEM is used for smoothing on the grids 𝒯1{\mathcal{T}}_{1} and 𝒯0{\mathcal{T}}_{0} respectively, the iteration matrix is given by

𝑴2C=(𝑰1μ1(𝑰1𝑺1C)(𝑨1C)1𝑨1F)(𝑰1μ0𝑰01(𝑨0C)1𝑰10𝑨1F).{\boldsymbol{M}}^{C}_{2}=\Big{(}{\boldsymbol{I}}_{1}-\mu_{1}({\boldsymbol{I}}_{1}-{\boldsymbol{S}}^{C}_{1})({\boldsymbol{A}}^{C}_{1})^{-1}{\boldsymbol{A}}^{F}_{1}\Big{)}\Big{(}{\boldsymbol{I}}_{1}-\mu_{0}{\boldsymbol{I}}^{1}_{0}({\boldsymbol{A}}^{C}_{0})^{-1}{\boldsymbol{I}}^{0}_{1}{\boldsymbol{A}}^{F}_{1}\Big{)}.

Here, for l0l\geq 0, 𝑺lC=𝑰l𝑹lC𝑨lC{\boldsymbol{S}}^{C}_{l}={\boldsymbol{I}}_{l}-{\boldsymbol{R}}^{C}_{l}{\boldsymbol{A}}^{C}_{l} is smoothing relaxation matrix, 𝑰l{\boldsymbol{I}}_{l} with the same size as 𝑨lF{\boldsymbol{A}}^{F}_{l} is identity matrix, 𝑰ls(s>l){\boldsymbol{I}}^{s}_{l}(s>l) is prolongation matrix from level ll to ss, 𝑰ls(s<l){\boldsymbol{I}}^{s}_{l}(s<l) is restriction matrix from level ll to ss, and 𝑹lC{\boldsymbol{R}}^{C}_{l} stands for matrix representation of smoother RlCR_{l}^{C} with CIP-FEM approximation. Let 𝑴2SL{\boldsymbol{M}}^{SL}_{2} (cf. [4]) stand for two-level iteration matrix, taking the smoothing based on shifted Laplace operator (4.7) with standard FEM approximation. Similarly, when applying standard FEM for smoothing on 𝒯1{\mathcal{T}}_{1} and CIP-FEM for correction on 𝒯0{\mathcal{T}}_{0}, the associated iteration matrix is derived as

𝑴2FC=(𝑰1μ1(𝑰1𝑺1F))(𝑰1μ0𝑰01(𝑨0C)1𝑰10𝑨1F),{\boldsymbol{M}}^{FC}_{2}=\Big{(}{\boldsymbol{I}}_{1}-\mu_{1}({\boldsymbol{I}}_{1}-{\boldsymbol{S}}^{F}_{1})\Big{)}\Big{(}{\boldsymbol{I}}_{1}-\mu_{0}{\boldsymbol{I}}^{1}_{0}({\boldsymbol{A}}^{C}_{0})^{-1}{\boldsymbol{I}}^{0}_{1}{\boldsymbol{A}}^{F}_{1}\Big{)},

where 𝑺1F=𝑰1𝑹1F𝑨1F{\boldsymbol{S}}^{F}_{1}={\boldsymbol{I}}_{1}-{\boldsymbol{R}}^{F}_{1}{\boldsymbol{A}}^{F}_{1}, and 𝑹1F{\boldsymbol{R}}^{F}_{1} stands for smoothing matrix representation of R1R_{1} with standard FEM approximation.

Since every two dimensional subspace (4.3) of 2h2h-harmonics E2h1θ0E^{\theta^{0}}_{2h_{1}} with θ0(π/2,π/2]\theta^{0}\in(-\pi/2,\pi/2] is left invariant under ω\omega-JAC or GS-LEX smoothing operator and correction operator, the representation of two-level iteration matrix of 𝑴2C{\boldsymbol{M}}^{C}_{2} on E2h1θ0E^{\theta^{0}}_{2h_{1}} is given by a 2×22\times 2 matrix as follows:

𝑴~2C\displaystyle\widetilde{{\boldsymbol{M}}}^{C}_{2} =\displaystyle= [𝑰~1μ1(𝑰~1[𝑺~1C(θ0)𝑺~1C(θ1)]D)[𝑨~1C(θ0)𝑨~1C(θ1)]D1[𝑨~1F(θ0)𝑨~1F(θ1)]D]\displaystyle\left[\widetilde{{\boldsymbol{I}}}_{1}-\mu_{1}\left(\widetilde{{\boldsymbol{I}}}_{1}-\left[\begin{array}[]{c}\widetilde{{\boldsymbol{S}}}_{1}^{C}(\theta^{0})\\ \widetilde{{\boldsymbol{S}}}_{1}^{C}(\theta^{1})\\ \end{array}\right]_{D}\right)\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{A}}}_{1}^{C}(\theta^{0})\\ \widetilde{{\boldsymbol{A}}}_{1}^{C}(\theta^{1})\\ \end{array}\right]_{D}^{-1}\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{A}}}_{1}^{F}(\theta^{0})\\ \widetilde{{\boldsymbol{A}}}_{1}^{F}(\theta^{1})\\ \end{array}\right]_{D}\right] (4.31)
[𝑰~1μ0[𝑰~01(θ0)𝑰~01(θ1)]𝑨~0C(2θ0)1[𝑰~10(θ0)𝑰~10(θ1)]t[𝑨~1F(θ0)𝑨~1F(θ1)]D],\displaystyle\cdot\left[\widetilde{{\boldsymbol{I}}}_{1}-\mu_{0}\left[\begin{array}[]{c}\widetilde{{\boldsymbol{I}}}_{0}^{1}(\theta^{0})\\ \widetilde{{\boldsymbol{I}}}_{0}^{1}(\theta^{1})\\ \end{array}\right]\widetilde{{\boldsymbol{A}}}^{C}_{0}(2\theta^{0})^{-1}\left[\begin{array}[]{c}\widetilde{{\boldsymbol{I}}}_{1}^{0}(\theta^{0})\\ \widetilde{{\boldsymbol{I}}}_{1}^{0}(\theta^{1})\\ \end{array}\right]^{t}\left[\begin{array}[]{c}\widetilde{{\boldsymbol{A}}}_{1}^{F}(\theta^{0})\\ \widetilde{{\boldsymbol{A}}}_{1}^{F}(\theta^{1})\\ \end{array}\right]_{D}\right],

where 𝑰~1\widetilde{{\boldsymbol{I}}}_{1} is 2×22\times 2 identity matrix and the subscript-D denotes the transformation of a vector into a diagonal matrix. Similarly, the representation 𝑴~2SL\widetilde{{\boldsymbol{M}}}^{SL}_{2} of two-level iteration matrix based on shifted Laplace operator, can be easily obtained (cf. [4]). For the iteration operator 𝑴2FC{\boldsymbol{M}}^{FC}_{2}, its representation on E2h1θ0E^{\theta^{0}}_{2h_{1}} is given by

𝑴~2FC\displaystyle\widetilde{{\boldsymbol{M}}}^{FC}_{2} =\displaystyle= [𝑰~1μ1(𝑰~1[𝑺~1F(θ0)𝑺~1F(θ1)]D)]\displaystyle\left[\widetilde{{\boldsymbol{I}}}_{1}-\mu_{1}\left(\widetilde{{\boldsymbol{I}}}_{1}-\left[\begin{array}[]{c}\widetilde{{\boldsymbol{S}}}_{1}^{F}(\theta^{0})\\ \widetilde{{\boldsymbol{S}}}_{1}^{F}(\theta^{1})\\ \end{array}\right]_{D}\right)\right] (4.41)
[𝑰~1μ0[𝑰~01(θ0)𝑰~01(θ1)]𝑨~0C(2θ0)1[𝑰~10(θ0)𝑰~10(θ1)]t[𝑨~1F(θ0)𝑨~1F(θ1)]D].\displaystyle\cdot\left[\widetilde{{\boldsymbol{I}}}_{1}-\mu_{0}\left[\begin{array}[]{c}\widetilde{{\boldsymbol{I}}}_{0}^{1}(\theta^{0})\\ \widetilde{{\boldsymbol{I}}}_{0}^{1}(\theta^{1})\\ \end{array}\right]\widetilde{{\boldsymbol{A}}}^{C}_{0}(2\theta^{0})^{-1}\left[\begin{array}[]{c}\widetilde{{\boldsymbol{I}}}_{1}^{0}(\theta^{0})\\ \widetilde{{\boldsymbol{I}}}_{1}^{0}(\theta^{1})\\ \end{array}\right]^{t}\left[\begin{array}[]{c}\widetilde{{\boldsymbol{A}}}_{1}^{F}(\theta^{0})\\ \widetilde{{\boldsymbol{A}}}_{1}^{F}(\theta^{1})\\ \end{array}\right]_{D}\right].

Then the spectral radius of 𝑴~2SL,𝑴~2FC\widetilde{{\boldsymbol{M}}}^{SL}_{2},\widetilde{{\boldsymbol{M}}}^{FC}_{2} and 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} for different θ0Θlow\theta^{0}\in\Theta_{\rm low} can be obtained analytically and numerically.

Refer to caption
Refer to caption
Figure 3: Left: Case 1-Case 3: ρ(𝑴~2SL)\rho(\widetilde{{\boldsymbol{M}}}^{SL}_{2}) with β=0.5,0.3\beta=0.5,0.3 and 0.10.1 over θ0Θlow\theta^{0}\in\Theta_{\rm low}, Case 4 and Case 5 denote for ρ(𝑴~2FC)\rho(\widetilde{{\boldsymbol{M}}}^{FC}_{2}) and ρ(𝑴~2C)\rho(\widetilde{{\boldsymbol{M}}}^{C}_{2}) with 𝐢γ=𝐢γo+0.01𝐢{\bf i}\gamma={\bf i}\gamma_{o}+0.01{\bf i} (𝐢γo0.085{\bf i}\gamma_{o}\thickapprox-0.085 when t=0.8t=0.8) on the finest grid. Right: ρ(𝑴~2C)\rho(\widetilde{{\boldsymbol{M}}}^{C}_{2}) with different 𝐢γ{\bf i}\gamma. Since t=kh>1t=kh>1 on the coarse grid, we formally choose the corresponding parameter 𝐢γ=𝐢γo+0.05𝐢{\bf i}\gamma={\bf i}\gamma_{o}+0.05{\bf i} for the cases in this figure.

The left graph of Figure 3 shows the spectral radius of 𝑴~2SL\widetilde{{\boldsymbol{M}}}^{SL}_{2} with β=0.5,0.3\beta=0.5,0.3 and 0.10.1 over θ0Θlow\theta^{0}\in\Theta_{\rm low}, and 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2}, 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} under ω\omega-JAC relaxation for t=0.8t=0.8 on the finest grid. In order to make some comparisons between different algorithms, the penalty parameter 𝐢γ{\bf i}\gamma in the CIP-FEM is always chosen as a complex number in the following if there is no special notation. If no confusion is possible, we will always denote t=𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡t={\it constant} for t=κhLt=\kappa h_{L} on the finest grid. The parameter in ω\omega-JAC relaxation is always chosen as ω=0.6\omega=0.6 in this section. We observe that for most of the frequencies θ0Θlow\theta^{0}\in\Theta_{\rm low} the spectral radius is smaller than one. The spectral radius tends to larger than one only under a few frequencies. Actually, the appearance of such a resonance is caused by the coarse grid correction and originates from the inversion of the coarse grid discretization symbol, 𝑨~0C(2θ0)\widetilde{{\boldsymbol{A}}}^{C}_{0}(2\theta^{0}) in (4.31) and (4.41) for instance. The minimum of |𝑨~0C(2θ0)||\widetilde{{\boldsymbol{A}}}^{C}_{0}(2\theta^{0})| maximizes the spectral radius of 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2} and 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} for fixed tt and 𝐢γ{\bf i}\gamma.

Moreover, for different choices of parameter β\beta in shifted Laplace operator, the properties of 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2} and 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} with 𝐢γ=𝐢γo+0.01𝐢{\bf i}\gamma={\bf i}\gamma_{o}+0.01{\bf i} on the finest grid always perform better than that of 𝑴~2SL\widetilde{{\boldsymbol{M}}}^{SL}_{2}. Although the spectral radius of 𝑴~2SL\widetilde{{\boldsymbol{M}}}^{SL}_{2} with β=0.1\beta=0.1 is almost equivalent to that of 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2} over the frequencies spreading near zero, there is a relatively large resonance causing by the coarse grid correction. Small β\beta in (4.7) deteriorates the coarse grid correction. However, large β\beta amplifies the corresponding spectral radius over the frequencies spreading near zero. We can refer to a recent work [4] about the choice of minimal complex shift parameter in shifted Laplace preconditioner. The choice of 𝐢γ{\bf i}\gamma in 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2} or 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} would also influence its spectral radius. The right graph of Figure 3 shows the spectral radius of 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} with different 𝐢γ{\bf i}\gamma under ω\omega-JAC relaxation for t=0.8t=0.8. The influence of 𝐢γ{\bf i}\gamma in 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} is similar to β\beta in shifted Laplace preconditioner. For other small t<1t<1, the properties of 𝑴~2SL\widetilde{{\boldsymbol{M}}}^{SL}_{2}, 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2} and 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} can also be observed as above. Moreover, when t=κht=\kappa h on the finest grid is small enough, the spectral radiuses of 𝑴~2SL\widetilde{{\boldsymbol{M}}}^{SL}_{2}, 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2} and 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} can all be smaller than one over all θ0Θlow\theta^{0}\in\Theta_{\rm low}. Due to the fact that our aim is to study the influences of smoothing corrections in two- and three-level methods, in the following we do not always assume tt to be sufficiently small.

Refer to caption
Figure 4: Case 1 (t=3t=\sqrt{3}) and Case 3 (t=4t=4) denote for ρ(𝑴~2SL)\rho(\widetilde{{\boldsymbol{M}}}^{SL}_{2}) with β=0.8\beta=0.8 over θ0Θlow\theta^{0}\in\Theta_{\rm low}, Case 2 (t=3t=\sqrt{3}) and Case 4 (t=4t=4) denote for ρ(𝑴~2C)\rho(\widetilde{{\boldsymbol{M}}}^{C}_{2}) with γ=0.8\gamma=0.8.

Alternatively, for large tt, the maximum of spectral radiuses of 𝑴~2SL\widetilde{{\boldsymbol{M}}}^{SL}_{2}, 𝑴~2FC\widetilde{{\boldsymbol{M}}}^{FC}_{2} and 𝑴~2C\widetilde{{\boldsymbol{M}}}^{C}_{2} appear to be spread around θ0=0\theta^{0}=0. Actually, for large tt the main factor determining the spectral radius is the smoothing operator rather than the coarse grid correction. We assume that ω\omega-JAC relaxation is used as smoother. It can be observed from Figure 4 that ρ(𝑴~2SL)\rho(\widetilde{{\boldsymbol{M}}}^{SL}_{2}) and ρ(𝑴~2C)\rho(\widetilde{{\boldsymbol{M}}}^{C}_{2}) reach a maximum in θ0=0\theta^{0}=0 for t=3t=\sqrt{3} and t=4t=4. Indeed, it has been analyzed in previous section and [4] that ρ(𝑺hJ)\rho({\boldsymbol{S}}^{J}_{h}) for CIP-FEM and shifted Laplace preconditioner (4.7) with continuous piecewise P1 approximation reaches a maximum at θ=0\theta=0 or θ=π\theta=\pi. From (4.9), it is easy to see that the ω\omega-JAC relaxation for discrete system from standard P1 FEM approximation is inefficient when tt is near 3\sqrt{3}. Thus, two-level method 𝑴2FC{\boldsymbol{M}}^{FC}_{2} is inefficient in this case. However, 𝑴2SL{{\boldsymbol{M}}}^{SL}_{2} and 𝑴2C{{\boldsymbol{M}}}^{C}_{2} can still be applied due to the adding of complex parts on the fine and coarse grids. One can observe from Figure 4 that the spectral radiuses of ρ(𝑴~2SL)\rho(\widetilde{{\boldsymbol{M}}}^{SL}_{2}) and ρ(𝑴~2C)\rho(\widetilde{{\boldsymbol{M}}}^{C}_{2}) is always larger than one when t=3t=\sqrt{3}, which implies the divergence of the associated two-level methods when applying ω\omega-JAC smoother. However, the spectral radius is smaller than one when t=4t=4, which may permit the use of ω\omega-JAC as a smoother on very coarse grids, but we shall not use this and utilize GMRES smoothing on coarse grids instead.

Furthermore, to get a comprehensive insight into the influence of multiple coarse grid corrections, we next carry out a three-level local Fourier analysis. The iteration operator of three-level method by nonsymmetric version of Algorithm 3.1 is derived as (IT2)(IT1)(IT0)(I-T_{2})(I-T_{1})(I-T_{0}). Similar to the two-level method, when CIP-FEM is used for smoothing on 𝒯2,𝒯1{\mathcal{T}}_{2},{\mathcal{T}}_{1} and 𝒯0{\mathcal{T}}_{0}, the iteration matrix can be deduced to be

𝑴3C=(𝑰2μ2(𝑰2𝑺2C)(𝑨2C)1𝑨2F)(𝑰2μ1𝑰12(𝑰1𝑺1C)(𝑨1C)1𝑰21𝑨2F)(𝑰2μ0𝑰02(𝑨0C)1𝑰20𝑨2F).{\boldsymbol{M}}^{C}_{3}=\Big{(}{\boldsymbol{I}}_{2}-\mu_{2}({\boldsymbol{I}}_{2}-{\boldsymbol{S}}^{C}_{2})({\boldsymbol{A}}^{C}_{2})^{-1}{\boldsymbol{A}}^{F}_{2}\Big{)}\Big{(}{\boldsymbol{I}}_{2}-\mu_{1}{\boldsymbol{I}}^{2}_{1}({\boldsymbol{I}}_{1}-{\boldsymbol{S}}^{C}_{1})({\boldsymbol{A}}^{C}_{1})^{-1}{\boldsymbol{I}}^{1}_{2}{\boldsymbol{A}}^{F}_{2}\Big{)}\Big{(}{\boldsymbol{I}}_{2}-\mu_{0}{\boldsymbol{I}}^{2}_{0}({\boldsymbol{A}}^{C}_{0})^{-1}{\boldsymbol{I}}^{0}_{2}{\boldsymbol{A}}^{F}_{2}\Big{)}.

Define the four dimensional 4h4h-harmonics by

E4h2θ0:=span{φh2(θ00,x),φh2(θ01,x),φh2(θ10,x),φh2(θ11,x)},E_{4h_{2}}^{\theta^{0}}:={\rm span}\{\varphi_{h_{2}}(\theta^{00},x),\varphi_{h_{2}}(\theta^{01},x),\varphi_{h_{2}}(\theta^{10},x),\varphi_{h_{2}}(\theta^{11},x)\},

where θα0=θα2,θα1=θα2sign(θα2)π,α=0,1\theta^{\alpha 0}=\frac{\theta^{\alpha}}{2},\theta^{\alpha 1}=\frac{\theta^{\alpha}}{2}-{\rm sign}(\frac{\theta^{\alpha}}{2})\pi,\alpha=0,1. For any θ0Θlow\theta^{0}\in\Theta_{\rm low}, the three-level operators leaves the space of 4h4h-harmonics E4h2θ0E_{4h_{2}}^{\theta^{0}} invariant (cf. [18]). This yields a block diagonal representation of 𝑴3C{\boldsymbol{M}}^{C}_{3} with the following 4×44\times 4 matrix 𝑴~3C\widetilde{{\boldsymbol{M}}}^{C}_{3}:

𝑴~3C\displaystyle\widetilde{{\boldsymbol{M}}}^{C}_{3} =\displaystyle= [𝑰~2μ2(𝑰~2𝑺~2C)(𝑨~2C)1𝑨~2F]\displaystyle\left[\widetilde{{\boldsymbol{I}}}_{2}-\mu_{2}\left(\widetilde{{\boldsymbol{I}}}_{2}-\widetilde{{\boldsymbol{S}}}^{C}_{2}\right)(\widetilde{{\boldsymbol{A}}}^{C}_{2})^{-1}\widetilde{{\boldsymbol{A}}}^{F}_{2}\right]
[𝑰~2μ1𝑰~12(𝑰~1[𝑺~1C(θ0)𝑺~1C(θ1)]D)[𝑨~1C(θ0)𝑨~1C(θ1)]D1(𝑰~21)t𝑨~2F]\displaystyle\cdot\left[\widetilde{{\boldsymbol{I}}}_{2}-\mu_{1}\widetilde{{\boldsymbol{I}}}_{1}^{2}\left(\widetilde{{\boldsymbol{I}}}_{1}-\left[\begin{array}[]{c}\widetilde{{\boldsymbol{S}}}_{1}^{C}(\theta^{0})\\ \widetilde{{\boldsymbol{S}}}_{1}^{C}(\theta^{1})\\ \end{array}\right]_{D}\right)\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{A}}}_{1}^{C}(\theta^{0})\\ \widetilde{{\boldsymbol{A}}}_{1}^{C}(\theta^{1})\\ \end{array}\right]_{D}^{-1}(\widetilde{{\boldsymbol{I}}}_{2}^{1})^{t}\widetilde{{\boldsymbol{A}}}^{F}_{2}\right]
[𝑰~2μ0𝑰~12𝑰~01𝑨~0C(2θ0)1(𝑰~10)t(𝑰~21)t𝑨~2F],\displaystyle\cdot\left[\widetilde{{\boldsymbol{I}}}_{2}-\mu_{0}\widetilde{{\boldsymbol{I}}}_{1}^{2}\widetilde{{\boldsymbol{I}}}_{0}^{1}\widetilde{{\boldsymbol{A}}}^{C}_{0}(2\theta^{0})^{-1}(\widetilde{{\boldsymbol{I}}}_{1}^{0})^{t}(\widetilde{{\boldsymbol{I}}}_{2}^{1})^{t}\widetilde{{\boldsymbol{A}}}^{F}_{2}\right],

where 𝑰~2\widetilde{{\boldsymbol{I}}}_{2} is 4×44\times 4 identity matrix, 𝑺~2C=[𝑺~2C(θ00)𝑺~2C(θ01)𝑺~2C(θ10)𝑺~2C(θ11)]D\widetilde{{\boldsymbol{S}}}^{C}_{2}=\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{S}}}_{2}^{C}(\theta^{00})\\ \widetilde{{\boldsymbol{S}}}_{2}^{C}(\theta^{01})\\ \widetilde{{\boldsymbol{S}}}_{2}^{C}(\theta^{10})\\ \widetilde{{\boldsymbol{S}}}_{2}^{C}(\theta^{11})\\ \end{array}\right]_{D}, 𝑨~2C=[𝑨~2C(θ00)𝑨~2C(θ01)𝑨~2C(θ10)𝑨~2C(θ11)]D\widetilde{{\boldsymbol{A}}}^{C}_{2}=\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{A}}}_{2}^{C}(\theta^{00})\\ \widetilde{{\boldsymbol{A}}}_{2}^{C}(\theta^{01})\\ \widetilde{{\boldsymbol{A}}}_{2}^{C}(\theta^{10})\\ \widetilde{{\boldsymbol{A}}}_{2}^{C}(\theta^{11})\\ \end{array}\right]_{D}, 𝑨~2F=[𝑨~2F(θ00)𝑨~2F(θ01)𝑨~2F(θ10)𝑨~2F(θ11)]D\widetilde{{\boldsymbol{A}}}^{F}_{2}=\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{A}}}_{2}^{F}(\theta^{00})\\ \widetilde{{\boldsymbol{A}}}_{2}^{F}(\theta^{01})\\ \widetilde{{\boldsymbol{A}}}_{2}^{F}(\theta^{10})\\ \widetilde{{\boldsymbol{A}}}_{2}^{F}(\theta^{11})\\ \end{array}\right]_{D}, 𝑰~12=[𝑰~12(θ00)0𝑰~12(θ01)00𝑰~12(θ10)0𝑰~12(θ11)]\widetilde{{\boldsymbol{I}}}_{1}^{2}=\left[\begin{array}[]{cc}\widetilde{{\boldsymbol{I}}}_{1}^{2}(\theta^{00})\quad 0\\ \widetilde{{\boldsymbol{I}}}_{1}^{2}(\theta^{01})\quad 0\\ 0\quad\widetilde{{\boldsymbol{I}}}_{1}^{2}(\theta^{10})\\ 0\quad\widetilde{{\boldsymbol{I}}}_{1}^{2}(\theta^{11})\\ \end{array}\right], 𝑰~01=[𝑰~10(θ0)𝑰~10(θ1)]\widetilde{{\boldsymbol{I}}}_{0}^{1}=\left[\begin{array}[]{c}\widetilde{{\boldsymbol{I}}}_{1}^{0}(\theta^{0})\\ \widetilde{{\boldsymbol{I}}}_{1}^{0}(\theta^{1})\\ \end{array}\right], and 𝑰~21,𝑰~10\widetilde{{\boldsymbol{I}}}_{2}^{1},\widetilde{{\boldsymbol{I}}}_{1}^{0} are defined similarly. For another approach, we assume that standard FEM is applied on 𝒯2{\mathcal{T}}_{2} for smoothing and CIP-FEM is used for correction on 𝒯1{\mathcal{T}}_{1} and 𝒯0{\mathcal{T}}_{0}, then the associated iteration matrix is given by

𝑴3FC=(𝑰2μ2(𝑰2𝑺2F))(𝑰2μ1𝑰12(𝑰1𝑺1C)(𝑨1C)1𝑰21𝑨2F)(𝑰2μ0𝑰02(𝑨0C)1𝑰20𝑨2F).{\boldsymbol{M}}^{FC}_{3}=\Big{(}{\boldsymbol{I}}_{2}-\mu_{2}({\boldsymbol{I}}_{2}-{\boldsymbol{S}}^{F}_{2})\Big{)}\Big{(}{\boldsymbol{I}}_{2}-\mu_{1}{\boldsymbol{I}}^{2}_{1}({\boldsymbol{I}}_{1}-{\boldsymbol{S}}^{C}_{1})({\boldsymbol{A}}^{C}_{1})^{-1}{\boldsymbol{I}}^{1}_{2}{\boldsymbol{A}}^{F}_{2}\Big{)}\Big{(}{\boldsymbol{I}}_{2}-\mu_{0}{\boldsymbol{I}}^{2}_{0}({\boldsymbol{A}}^{C}_{0})^{-1}{\boldsymbol{I}}^{0}_{2}{\boldsymbol{A}}^{F}_{2}\Big{)}.

Moreover, similar to 𝑴2SL{\boldsymbol{M}}^{SL}_{2}, we denote by 𝑴3SL{\boldsymbol{M}}^{SL}_{3} the three-level iteration matrix, which takes the smoothing based on shifted Laplace operator (4.7) with standard FEM approximation. Then the representations 𝑴~3FC\widetilde{{\boldsymbol{M}}}^{FC}_{3}, 𝑴~3SL\widetilde{{\boldsymbol{M}}}^{SL}_{3} with respect to 𝑴3FC{\boldsymbol{M}}^{FC}_{3} and 𝑴3SL{\boldsymbol{M}}^{SL}_{3} respectively on the 4h4h-harmonics E4h2θ0E_{4h_{2}}^{\theta^{0}} can be derived directly.

Refer to caption
Refer to caption
Figure 5: Left: ρ(𝑴~3C)\rho(\widetilde{{\boldsymbol{M}}}^{C}_{3}) over θ0Θlow\theta^{0}\in\Theta_{\rm low} for the cases t=0.6,0.4,0.2t=0.6,0.4,0.2. Right (t=0.4t=0.4): Case 1-Case 2: ρ(𝑴~3SL)\rho(\widetilde{{\boldsymbol{M}}}^{SL}_{3}), ρ(𝑴~3C)\rho(\widetilde{{\boldsymbol{M}}}^{C}_{3}), Case 3-Case 4: ρ(𝑴~3FC)\rho(\widetilde{{\boldsymbol{M}}}^{FC}_{3}) with one step and two step intermediate coarse grid (the 2nd level) correction. ω\omega-JAC relaxation is applied for smoothing. The parameters in shifted Laplace operator and CIP-FEM are chosen as β=0.5\beta=0.5, 𝐢γ=𝐢γo+0.01𝐢{\bf i}\gamma={\bf i}\gamma_{o}+0.01{\bf i} when t<1t<1 and 𝐢γ=𝐢γo+0.05𝐢{\bf i}\gamma={\bf i}\gamma_{o}+0.05{\bf i} when t>1t>1.

For fixed wave number κ\kappa, the performance of three-level methods is shown in Figure 5. For instance, one can observe from the left graph of Figure 5 that a coarser initial grid used for 𝑴3C{\boldsymbol{M}}^{C}_{3} will deteriorate its convergence. The right graph of Figure 5 shows the spectral radiuses for three kinds of approaches: 𝑴3SL{\boldsymbol{M}}^{SL}_{3}, 𝑴3C{\boldsymbol{M}}^{C}_{3} and 𝑴3FC{\boldsymbol{M}}^{FC}_{3}. For t=0.4t=0.4, both of 𝑴3FC{\boldsymbol{M}}^{FC}_{3} and 𝑴3C{\boldsymbol{M}}^{C}_{3} perform better than 𝑴3SL{\boldsymbol{M}}^{SL}_{3} over most of the frequencies, and the spectral radius of 𝑴3FC{\boldsymbol{M}}^{FC}_{3} is similar to that of 𝑴3C{\boldsymbol{M}}^{C}_{3} over most of frequencies in this case. Actually, the performance of 𝑴3FC{\boldsymbol{M}}^{FC}_{3} is always comparable with 𝑴3C{\boldsymbol{M}}^{C}_{3}. This is also true for two-level method, which can be observed from the left graph of Figure 3. Thus, in order to reduce the computational cost, the modified multilevel methods (Algorithm 3.1) with stable CIP-FEM corrections on coarser grids and standard FEM corrections on finer grids can be usually utilized in practical. From the right graph of Figure 5, we can also see that the convergence of 𝑴3FC{\boldsymbol{M}}^{FC}_{3} does not always perform better with more steps of intermediate coarse grid correction.

5 Numerical results

In this section, we present some numerical results to illustrate the performance of Algorithm 2.1 and modified multilevel method (Algorithm 3.1) for two Helmholtz problems in two dimension. The multilevel method is used as a preconditioner in outer GMRES iterations (PGMRES). Gauss-Seidel relaxation is always used as smoother when lSLl\in S_{L}, and the smoothing steps are always chosen as one step if there is no any annotation. At the ll-th level, the discrete problem reads 𝐀l𝐮l=𝐅l{\bf A}_{l}{\bf u}_{l}={\bf F}_{l}. We denote by 𝐫ln=𝐅l𝐀l𝐮ln{\bf r}^{n}_{l}={\bf F}_{l}-{\bf A}_{l}{\bf u}^{n}_{l} the residual with respect to the nn-th iteration. The PGMRES algorithm terminates when

𝐫ln/𝐫l0106.\|{\bf r}^{n}_{l}\|/\|{\bf r}^{0}_{l}\|\leq 10^{-6}.

The number of iteration steps required to achieve the desired accuracy is denoted by iter.

Example 5.1.

We consider a two dimensional Helmholtz equation with the first order absorbing boundary condition (cf. [12, 19]):

uκ2u\displaystyle-\triangle u-\kappa^{2}u =f:=sin(κr)rinΩ,\displaystyle=f:=\frac{\sin(\kappa r)}{r}\quad{\rm in}\ \Omega,
un+iκu\displaystyle\frac{\partial u}{\partial n}+{\rm i}\kappa u =gonΩ.\displaystyle=g\quad{\rm on}\ \partial\Omega.

Here Ω\Omega is a unit square with center (0,0)(0,0) and gg is chosen such that the exact solution is

u=cos(κr)κcosκ+isinκκ(J0(κ)+iJ1(κ))J0(κr),u=\frac{\cos(\kappa r)}{\kappa}-\frac{\cos\kappa+\rm i\sin\kappa}{\kappa(J_{0}(\kappa)+{\rm i}J_{1}(\kappa))}J_{0}(\kappa r),

where Jν(z)J_{\nu}(z) are Bessel functions of the first kind.

Table 1: Iteration counts of PGMRES for discrete system from CIP-P1 and CIP-P2 (κ=100\kappa=100). The smoothing relaxations on fine and coarse grids are all based on CIP-FEM.
Level 2 3 4 5
DOFs 16641 66049 263169 1050625
iter (P1) 27 24 21 20
iter (P2) 26 22 19 18

In this example, we assume that the coarsest level of multilevel method satisfies κh0/p2\kappa h_{0}/p\thickapprox 2 for κ360\kappa\leq 360. For 400κ600400\leq\kappa\leq 600, we choose the coarsest grid condition such that κh0/p1.11.7\kappa h_{0}/p\approx 1.1\thicksim 1.7. The parameters in (2.1) are chosen as γe0.01+0.07𝐢\gamma_{e}\equiv 0.01+0.07{\bf i} (cf. [12]) for CIP-P1, and γe0.005+0.035𝐢\gamma_{e}\equiv 0.005+0.035{\bf i} for piecewise P2 CIP-FEM (CIP-P2). We first test the algorithms for the case with wave number κ=100\kappa=100. When Algorithm 2.1 is applied to the discrete system ALCuL=FLA^{C}_{L}u_{L}=F_{L}, the smoothing relaxations are all based on the CIP-FEM approximation on fine and coarse grids. Table 1 shows the corresponding iteration counts of PGMRES for discrete system from CIP-P1 and CIP-P2 approximations. We can observe that for fixed κ\kappa the algorithm is robust on different levels.

Table 2: Iteration counts of PGMRES for discrete system from FEM-P1. The smoothing relaxations on fine and coarse grids are all based on FEM-P1 for κ=100,200\kappa=100,200.
κ=100\kappa=100 Level 2 3 4 5
DOFs 16641 66049 263169 1050625
iter (m2=1m_{2}=1) 58 87 94 90
iter (m2=20m_{2}=20) 46 53 49 47
κ=200\kappa=200 Level 2 3 4 5
DOFs 66049 263169 1050625 4198401
iter (m2m_{2}=1) 177 >200>200 >200>200 >200>200

For fixed κ\kappa, when the grid is fine enough to get the accuracy, standard FEM can be utilized again to discretize the problem. But standard multigrid method fails to solve the corresponding discrete system for the case with large wave number. In the following, we mainly apply the multilevel method presented in Algorithm 3.1 with different smoothing strategies. Table 2 shows the iteration counts of this multilevel-preconditioned GMRES method with GMRES smoothing on coarse grids for κ=100,200\kappa=100,200, and the smoothing relaxations on fine and coarse grids are all based on standard P1 FEM (FEM-P1) approximation. We find that the iteration count is mesh independent for fixed κ\kappa, but it increases rapidly with larger wave number. For instance, for the case κ=200\kappa=200 with m2=1m_{2}=1, the iteration counts of PGMRES will be more than 200, even when the GMRES smoothing is performed by m2=20m_{2}=20 steps, the iteration counts are still more than 100. Although more steps of GMRES smoothing can reduce the total iteration counts, it requires more memory to store data in the computation. Actually, the slow convergence of the algorithm mainly lies in the bad approximation on coarse grids. Next, we apply CIP-FEM to construct stable coarse grid corrections.

Refer to caption
Refer to caption
Figure 6: Surface plot of imaginary part of discrete FEM-P2 solutions for κ=100\kappa=100 (left) and κ=200\kappa=200 (right) on the grid with mesh condition κh/p0.55\kappa h/p\approx 0.55.
Table 3: Iteration counts of PGMRES for discrete system from FEM-P1 approximation (κ=100\kappa=100). Case 1: the smoothing relaxations on fine and coarse grids are both based on shifted Laplace operator (4.7) with β=0.2\beta=0.2 by FEM-P1, Case 2: the smoothing relaxations on fine and coarse grids are both based on CIP-P1.
Level 2 3 4 5
DOFs 16641 66049 263169 1050625
iter (Case 1) 52 64 66 63
iter (Case 2) 27 24 22 21

Figure 6 displays the surface plot of imaginary part of discrete FEM-P2 solutions for κ=100,200\kappa=100,200 on the grid with mesh condition κh/p0.55\kappa h/p\approx 0.55. Indeed, the discrete solutions have correct shapes and amplitudes as the exact solutions. Table 3 presents the comparing of shifted Laplace operator with FEM-P1 and the original Helmholtz operator with CIP-P1. For the case κ=100\kappa=100, we can see that the second approach has the minimum iteration counts. The iteration counts shown in Table 4 examine the performance of PGMRES based on Algorithm 3.1 with different steps of GMRES smoothing. The CIP-P1 and CIP-P2 are only used for approximations on coarse grids. From Tables 1, 3 and 4, it suggests that Algorithm 3.1 is also very efficient as Algorithm 2.1 and the second approach in Table 3 which uses CIP-FEM on both fine and coarse grids, and the growth in GMRES smoothing steps does not always reduce iteration counts obviously. This supports the similar phenomena shown in the right graph of Figure 5. Hence, in the following we will only use one step (m2=1m_{2}=1) of GMRES smoothing in Algorithm 3.1.

Table 4: Iteration counts of PGMRES based on Algorithm 3.1 for discrete system from FEM-P1 and FEM-P2 with different steps of post GMRES smoothing (κ=100\kappa=100).
Level 3 4 5
DOFs 66049 263169 1050625
iter (P1, m2=1m_{2}=1) 24 23 22
iter (P1, m2=10m_{2}=10) 21 23 23
iter (P2, m2=1m_{2}=1) 27 21 19
iter (P2, m2=10m_{2}=10) 32 21 18
Table 5: Iteration counts of PGMRES based on Algorithm 3.1 for discrete system from FEM-P1 and FEM-P2 for the cases κ=50,200,360\kappa=50,200,360 with the coarsest grid condition κh0/p2\kappa h_{0}/p\thickapprox 2.
κ=50\kappa=50 Level 3 4 5
DOFs 16641 66049 263169
iter (P1) 15 15 15
iter (P2) 23 14 13
κ=200\kappa=200 Level 3 4 5
DOFs 263169 1050625 4198401
iter (P1) 49 57 55
iter (P2) 44 41 36
κ=360\kappa=360 Level 2 3 4
DOFs 263169 1050625 4198401
iter (P1) 111 115 112
iter (P2) 41 44 39
Table 6: Iteration counts of PGMRES based on Algorithm 3.1 for discrete system from FEM-P1 and FEM-P2 for the cases κ=400,500,600\kappa=400,500,600. The coarsest grid condition is chosen with mesh size h00.00276h_{0}\thickapprox 0.00276 such that κh0/p1.11.7\kappa h_{0}/p\thickapprox 1.1\thicksim 1.7.
FEM-P1 Level 2 3 FEM-P2 Level 2 3
DOFs 1050625 4198401 DOFs 1050625 4198401
iter (κ=400\kappa=400) 30 26 iter (κ=400\kappa=400) 21 14
iter (κ=500\kappa=500) 71 50 iter (κ=500\kappa=500) 29 25
iter (κ=600\kappa=600) 156 141 iter (κ=600\kappa=600) 43 47

Tables 5 and 6 show the iteration counts of PGMRES based on Algorithm 3.1 with CIP-FEM used for correction only on coarse grids for discrete system from FEM-P1 and FEM-P2. For fixed κ\kappa, one can observe that the iteration counts are robust on different levels. Due to the same coarsest grid used for κ=400,500,600\kappa=400,500,600, the growth of iteration counts is a little faster than linear when κ400\kappa\geq 400. But when the FEM-P2 is applied in particular, the growth of iteration counts with increasing wave number is more stable than FEM-P1.


Example 5.2.

In a unit square domain Ω\Omega with center (0,0)(0,0), we consider the Helmholtz problem (1.1)-(1.2) with discontinuous wave number, which is defined by

κ={κ1,if(x1,x2)(0.5,0)×(0,0.5)(0,0.5)×(0.5,0),κ2,elsewhere,\kappa=\begin{cases}\kappa_{1},&\text{if}\;\;(x_{1},x_{2})\in(-0.5,0)\times(0,0.5)\bigcup(0,0.5)\times(-0.5,0),\\ \kappa_{2},&\text{elsewhere},\end{cases}

where κ2=qκ1,q>1\kappa_{2}=q\kappa_{1},q>1. We set the Robin boundary condition (1.2) with g=0g=0 and the external force f(x)f(x) to be a narrow Gaussian point source (cf. [7]) located at (r1,r2)=(0.25,0.25)(r_{1},r_{2})=(-0.25,-0.25):

f(x1,x2)=e(4κπ)2((x1r1)2+(x2r2)2).f(x_{1},x_{2})=e^{-(\frac{4\kappa}{\pi})^{2}((x_{1}-r_{1})^{2}+(x_{2}-r_{2})^{2})}.
Refer to caption
Refer to caption
Figure 7: Surface plot of imaginary part of discrete FEM-P2 solutions for κ2=300,q=3\kappa_{2}=300,q=3 (left) and κ2=300,q=10\kappa_{2}=300,q=10 (right) on the grid with mesh condition κ2h/p0.8\kappa_{2}h/p\approx 0.8.
Table 7: Iteration counts of PGMRES based on Algorithm 3.1 for discrete system from FEM-P1 and FEM-P2 for the cases κ2=180(q=3,10),κ2=300(q=3,10)\kappa_{2}=180\ (q=3,10),\ \kappa_{2}=300\ (q=3,10).
κ2=180\kappa_{2}=180 Level 3 4
DOFs 263169 1050625
iter (P1,q=3q=3) 26 27
iter (P1,q=10q=10) 28 29
iter (P2,q=3q=3) 16 15
iter (P2,q=10q=10) 17 15
κ2=300\kappa_{2}=300 Level 3 4
DOFs 1050625 4198401
iter (P1,q=3q=3) 30 30
iter (P1,q=10q=10) 32 33
iter (P2,q=3q=3) 16 15
iter (P2,q=10q=10) 16 15

In this example, we test Algorithm 3.1 for the Helmholtz problem with discontinuous wave number. Due to the fact that κ2=max{κ1,κ2}\kappa_{2}=\max\{\kappa_{1},\kappa_{2}\}, the coarsest mesh condition is according to the choice of κ2h0/p\kappa_{2}h_{0}/p. Figure 7 displays the surface plot of imaginary part of discrete FEM-P2 solutions for κ2=300\kappa_{2}=300 with q=3q=3 and q=10q=10 on the grid with mesh condition κ2h/p0.8\kappa_{2}h/p\approx 0.8. The iteration counts of PGMRES based on Algorithm 3.1 for different κ2\kappa_{2} and qq are listed in Table 7. For fixed κ2\kappa_{2} and polynomial order, we note that the iteration counts are robust on different levels. Similar to the first example, the PGMRES iteration for FEM-P2 is more stable than FEM-P1.

References

  • [1] R. Adams, Sobolev Spaces, Academic Press, New York, 1975.
  • [2] A. Brandt, Multi-level adaptive solutions to boundary-value problems, Math. Comp., 31 (1977), pp. 333–390.
  • [3] A. Brandt and I. Livshits, Wave-ray multigrid method for standing wave equations, Electron. Trans. Numer. Anal., 6 (1997), pp. 162–181.
  • [4] S. Cools and W. Vanroose, Local Fourier Analysis of the Complex Shifted Laplacian preconditioner for Helmholtz problems, preprint, 2011.
  • [5] H.C. Elman, O.G. Ernst, and D.P. O’Leary, A multigrid method enhanced by Krylov subspace iteration for discrete Helmholtz equations, SIAM J. Sci. Comput., 23 (2001), pp. 1291–1315.
  • [6] B. Engquist and L. Ying, Sweeping preconditioner for the Helmholtz equation: hierarchical matrix representation, Comm. Pure Appl. Math., 64 (2011), pp. 697–735.
  • [7] B. Engquist and L. Ying, Sweeping preconditioner for the Helmholtz equation: moving perfectly matched layers, Multiscale Model. Simul., 9 (2011), pp. 686–710.
  • [8] Y. A. Erlangga, C. Vuik, and C. W. Oosterlee, On a class of preconditioners for solving the Helmholtz equation, Appl. Numer. Math., 50 (2004), pp. 409–425.
  • [9] Y. A. Erlangga, C. W. Oosterlee, and C. Vuik, A novel multigrid based preconditioner for heterogeneous Helmholtz problems, SIAM J. Sci. Comput., 27 (2006), pp. 1471–1492.
  • [10] Y.A. Erlangga, Advances in iterative methods and preconditioners for the Helmholtz equation, Arch. Comput. Methods Eng., 15 (2008), pp. 37–66.
  • [11] O.G. Ernst and M.J. Gander, Why it is difficult to solve Helmholtz problems with classical iterative methods?, in: I. Graham, T. Hou, O. Lakkis, R. Scheichl (Eds.), Numerical Analysis of Multiscale Problems, Springer, 2011.
  • [12] X. Feng and H. Wu, Discontinuous Galerkin methods for the Helmholtz equation with large wave number, SIAM J. Numer. Anal, 47 (2009), pp. 2872–2896.
  • [13] X. Feng and H. Wu, hp-discontinuous Galerkin methods for the Helmholtz equation with large wave number, Math. Comp, 80 (2011), pp. 1997–2024.
  • [14] F. Ihlenburg, Finite Element Analysis of Acoustic Scattering, Springer-Verlag, New York, 1998.
  • [15] S. Kim and S. Kim, Multigrid simulations for high-frequency solutions of the Helmholtz problem in heterogeneous media, SIAM J. Sci. Comput., 24 (2002), pp. 684–701.
  • [16] I. Livshits and A. Brandt, Accuracy properties of the wave-ray multigrid algorithm for Helmholtz equations, SIAM J. Sci. Comput., 28 (2006), pp. 1228–1251.
  • [17] P.S. Vassilevski and J. Wang, An application of the abstract multilevel theory to nonconforming finite element methods, SIAM J. Numer. Anal., 32 (1995), 235–248.
  • [18] R. Wienands and W. Joppich, Practical Fourier Analysis for Multigrid Methods, Chapman & Hall/CRC, London, 2004.
  • [19] H. Wu, Pre-asymptotic error analysis of CIP-FEM and FEM for Helmholtz equation with high wave number. Part I: Linear version, to appear, 2013.
  • [20] L. Zhu, E. Burman, and H. Wu, Continuous Interior Penalty Finite Element Method for Helmholtz Equation with Large Wave Number: One Dimension Analysis, to appear, 2013.
  • [21] L. Zhu and H. Wu, Pre-asymptotic error analysis of CIP-FEM and FEM for Helmholtz equation with high wave number. Part II: hphp version, SIAM J. Numer. Anal., to appear, 2013. (See also arXiv: http://arxiv.org/pdf /1204.5061v1).