Skip to main content
Log in

A Semi Matrix-Free Twogrid Preconditioner for the Helmholtz Equation with Near Optimal Shifts

  • Original Article
  • Published:
Journal of Scientific Computing Aims and scope Submit manuscript

Abstract

Due to its significance in terms of wave phenomena a considerable effort has been put into the design of preconditioners for the Helmholtz equation. One option to design a preconditioner is to apply a multigrid method on a shifted operator. In such an approach, the wavenumber is shifted by some imaginary value. This step is motivated by the observation that the shifted problem can be more efficiently handled by iterative solvers when compared to the standard Helmholtz equation. However, up to now, it is not obvious what the best strategy for the choice of the shift parameter is. It is well known that a good shift parameter depends sensitively on the wavenumber and the discretization parameters such as the order and the mesh size. Therefore, we study the choice of a near optimal complex shift such that a flexible generalized minimal residual (FGMRES) solver converges with fewer iterations. Our goal is to provide a map which returns the near optimal shift for the preconditioner depending on the wavenumber and the mesh size. In order to compute this map, a data driven approach is considered: We first generate many samples, and in a second step, we perform a nonlinear regression on this data. With this representative map, the near optimal shift can be obtained by a simple evaluation. Our preconditioner is based on a twogrid V-cycle applied to the shifted problem, allowing us to implement a semi matrix-free method in which only the coarse grid and boundary matrices need to be stored in memory. On the fine grid, only the action of the matrix applied to a vector is computed, without assembling the global matrix. This enables efficient use of computing resources and allows problems to be solved at scales that were previously limited by the available memory. The performance of our preconditioned FGMRES solver is illustrated by several benchmark problems with heterogeneous wavenumbers in two and three space dimensions.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. https://doi.org/10.5281/zenodo.4607330.

References

  1. Amoco statics test. https://software.seg.org/datasets/2D/Statics_1994/ (1994). [Online]. Accessed 26 Jan 2021

  2. Alvarado, A., Castillo, P.: Computational performance of LDG methods applied to time harmonic Maxwell equation in polyhedral domains. J. Sci. Comput. 67(2), 453–474 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  3. Amestoy, P., Buttari, A., L’Excellent, J.-Y., Mary, T.: Performance and scalability of the block low-rank multifrontal factorization on multicore architectures. ACM Trans. Math. Softw. 45, 1–26 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  4. Anderson, R., Andrej, J., Barker, A., Bramwell, J., Camier, J.-S., Dobrev, J.C.V., Dudouit, Y., Fisher, A., Kolev, T., Pazner, W., Stowell, M., Tomov, V., Akkerman, I., Dahm, J., Medina, D., Zampini, S.: MFEM: a modular finite element library. Comput. Math. Appl. 81, 42–74 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  5. Balay, S., Abhyankar, S., Adams, M.F., Brown, J., Brune, P., Buschelman, K., Dalcin, L., Dener, A., Eijkhout, V., Gropp, W.D., Karpeyev, D., Kaushik, D., Knepley, M.G., May, D.A.,McInnes, L.C., Mills, R.T., Munson, T., Rupp, K., Sanan, P., Smith, B.F., Zampini, S., Zhang, H., Zhang, H.: PETSc Web page. https://www.mcs.anl.gov/petsc (2019)

  6. Bayliss, A., Goldstein, C.I., Turkel, E.: An iterative method for the Helmholtz equation. J. Comput. Phys. 49(3), 443–457 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  7. Bolten, M., Rittich, H.: Fourier analysis of periodic stencils in multigrid methods. SIAM J. Sci. Comput. 40(3), A1642–A1668 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  8. Brandt, A., Ta’asan, S.: Multigrid method for nearly singular and slightly indefinite problems. In: Multigrid Methods II, pp. 99–121. Springer (1986)

  9. Briggs, W.L., Henson, V.E., McCormick, S.F.: A multigrid tutorial. SIAM (2000)

  10. Cocquet, P.-H., Gander, M.J.: How large a shift is needed in the shifted Helmholtz preconditioner for its effective inversion by multigrid? SIAM J. Sci. Comput. 39(2), A438–A478 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  11. Cocquet, P.-H., Gander, M.J., Xiang, X.: Closed form dispersion corrections including a real shifted wavenumber for finite difference discretizations of 2D constant coefficient Helmholtz problems. SIAM J. Sci. Comput. 43(1), A278–A308 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  12. Cools, S., Vanroose, W.: Local Fourier analysis of the complex shifted Laplacian preconditioner for Helmholtz problems. Numer. Linear Algebra Appl. 20(4), 575–597 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  13. Drzisga, D., Rüde, U., Wohlmuth, B.: Stencil scaling for vector-valued PDEs on hybrid grids with applications to generalized Newtonian fluids. SIAM J. Sci. Comput. 42(6), B1429–B1461 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  14. Elman, H.C., Ernst, O.G., O’leary, D.P.: A multigrid method enhanced by Krylov subspace iteration for discrete Helmholtz equations. SIAM J. Sci. Comput. 23(4), 1291–1315 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  15. Erlangga, Y.A., Oosterlee, C.W., Vuik, C.: A novel multigrid based preconditioner for heterogeneous Helmholtz problems. SIAM J. Sci. Comput. 27(4), 1471–1492 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  16. Erlangga, Y.A., Vuik, C., Oosterlee, C.W.: On a class of preconditioners for solving the Helmholtz equation. Appl. Numer. Math. 50(3–4), 409–425 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  17. Ernst, O.G., Gander, M.J.: Why it is difficult to solve Helmholtz problems with classical iterative methods. In: Lecture Notes in Computational Science and Engineering, pp. 325–363. Springer, Berlin (2011)

  18. Ernst, O.G., Gander, M.J.: Multigrid methods for Helmholtz problems: a convergent scheme in 1D using standard components. In: Direct and Inverse Problems in Wave Propagation and Applications. De Gruyer, pp. 135–186 (2013)

  19. Esterhazy, S., Melenk, J.M.: An analysis of discretizations of the Helmholtz equation in L2 and in negative norms. Comput. Math. Appl. 67(4), 830–853 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  20. Gander, M.J., Graham, I.G., Spence, E.A.: Applying GMRES to the Helmholtz equation with shifted Laplacian preconditioning: What is the largest shift for which wavenumber-independent convergence is guaranteed? Numer. Math. 131(3), 567–614 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  21. Gander, M.J., Zhang, H.: A class of iterative solvers for the Helmholtz equation: factorizations, sweeping preconditioners, source transfer, single layer potentials, polarized traces, and optimized Schwarz methods. SIAM Rev. 61(1), 3–76 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  22. Gray, S.H., Marfurt, K.J.: Migration from topography: improving the near-surface image. Can. J. Explor. Geophys. 31(1–2), 18–24 (1995)

    Google Scholar 

  23. Greenfeld, D., Galun, M., Basri, R., Yavneh, I., Kimmel, R.: Learning to optimize multigrid PDE solvers. In: Proceedings of Machine Learning Research, vol. 97, pp. 2415–2423, Long Beach, California, USA (2019). PMLR

  24. Heinlein, A., Klawonn, A., Lanser, M., Weber, J.: Machine learning in adaptive domain decomposition methods–predicting the geometric location of constraints. SIAM J. Sci. Comput. 41(6), A3887–A3912 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  25. Kiefer, J.: Sequential minimax search for a maximum. Proc. Am. Math. Soc. 4(3), 502 (1953)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kronbichler, M., Ljungkvist, K.: Multigrid for matrix-free high-order finite element computations on graphics processors. ACM Trans. Parallel Comput. 6(1), 1–32 (2019)

    Article  Google Scholar 

  27. Kumar, P., Rodrigo, C., Gaspar, F.J., Oosterlee, C.W.: On local Fourier analysis of multigrid methods for PDEs with jumping and random coefficients. SIAM J. Sci. Comput. 41(3), A1385–A1413 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  28. Li, C., Qiao, Z.: A fast preconditioned iterative algorithm for the electromagnetic scattering from a large cavity. J. Sci. Comput. 53(2), 435–450 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  29. Liu, F., Ying, L.: Recursive sweeping preconditioner for the three-dimensional Helmholtz equation. SIAM J. Sci. Comput. 38(2), A814–A832 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  30. LRZ. Hardware of SuperMUC-NG. https://doku.lrz.de/display/PUBLIC/Hardware+of+SuperMUC-NG. Accessed 25 Feb 2020

  31. Lu, P., Xu, X.: A robust multilevel preconditioner based on a domain decomposition method for the Helmholtz equation. J. Sci. Comput. 81(1), 291–311 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  32. Luz, I., Galun, M., Maron, H., Basri, R., Yavneh, I.: Learning algebraic multigrid using graph neural networks. In: International Conference on Machine Learning, pp. 6489–6499. PMLR (2020)

  33. Melenk, J.M.: On generalized finite element methods. PhD thesis, research directed by Dept. of Mathematics. University of Maryland at College Park (1995)

  34. Paszke, A. et al.: PyTorch: an imperative style, high-performance deep learning library. In: Wallach, H., Larochelle, H., Beygelzimer, A., Buc, F., Fox, E., Garnett, R. (eds.) Advances in Neural Information Processing Systems, vol. 32, pp. 8024–8035. Curran Associates, Inc. (2019)

  35. Ramos, L.G., Nabben, R.: A two-level shifted Laplace preconditioner for Helmholtz problems: field-of-values analysis and wavenumber-independent convergence. arXiv preprint arXiv:2006.08750 (2020)

  36. Saad, Y.: A flexible inner-outer preconditioned GMRES algorithm. SIAM J. Sci. Comput. 14(2), 461–469 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  37. Saad, Y., Schultz, M.H.: GMRES: a generalized minimal residual algorithm for solving nonsymmetric linear systems. SIAM J. Sci. Stat. Comput. 7(3), 856–869 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  38. Stolk, C.C., Ahmed, M., Bhowmik, S.K.: A multigrid method for the Helmholtz equation with optimized coarse grid corrections. SIAM J. Sci. Comput. 36(6), A2819–A2841 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  39. Sun, Y., Wu, S., Xu, Y.: A parallel-in-time implementation of the Numerov method for wave equations. J. Sci. Comput. 90(1), 1–31 (2022)

    Article  MathSciNet  Google Scholar 

  40. Taus, M., Zepeda-Núñez, L., Hewett, R.J., Demanet, L.: L-Sweeps: a scalable, parallel preconditioner for the high-frequency Helmholtz equation. J. Comput. Phys. 420, 109706 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  41. Trottenberg, U., Oosterlee, C.W., Schuller, A.: Multigrid. Elsevier (2000)

  42. Versteeg, R.: The Marmousi experience: velocity model determination on a synthetic complex data set. Lead. Edge 13(9), 927–936 (1994)

    Article  Google Scholar 

  43. Wienands, R., Joppich, W.: Practical Fourier Analysis for Multigrid Methods. Chapman and Hall/CRC (2004)

Download references

Acknowledgements

This work was partly supported by the German Research Foundation by Grant WO671/11-1. The authors gratefully acknowledge the Gauss Centre for Supercomputing e.V. (www.gauss-centre.eu) for funding this project by providing computing time on the GCS Supercomputer SuperMUC at Leibniz Supercomputing Centre (www.lrz.de).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tobias Köppl.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

A Local Fourier Analysis for the Twogrid Operator

A Local Fourier Analysis for the Twogrid Operator

The convergence analysis by means of LFA is based on the assumption that the error function \(e_h^l\) on the fine grid \(\textbf{G}_h\) can be represented by a linear combination of Fourier modes [41]:

$$\begin{aligned} \varphi _h( {\theta },\textbf{x} ) = e^{\textrm{i}{\theta } \cdot \textbf{x}/h} = e^{\textrm{i}\theta _1 x_1/h} e^{\textrm{i}\theta _2 x_2/h},\;\textbf{x} \in \textbf{G}_h. \end{aligned}$$

\(\theta = ( \theta _1,\theta _2 ) \in \mathbb {R}^2\) denote the Fourier frequencies, which may be restricted to the domain \(\varTheta = \left( -\pi ,\pi \right] ^2 \subset \mathbb {R}^2\), see [12], as a consequence of the fact that

$$\begin{aligned} \varphi _h( {\theta } + 2\pi ,\textbf{x} ) = \varphi _h( {\theta },\textbf{x} ). \end{aligned}$$

Thus the space of Fourier modes is given by:

$$\begin{aligned} E_h = \text {span}\left\{ \left. \varphi _h( {\theta },\textbf{x} ) = e^{\textrm{i}{\theta } \cdot \textbf{x}/h} \right| \textbf{x} \in \textbf{G}_h,\; {\theta } \in \varTheta \right\} . \end{aligned}$$

According to [41], the Fourier modes are the formal eigenfunctions of an operator

$$\begin{aligned} H \in \left\{ T_h^{2h},\;I_h,\;I_h^{2h},\;I_{2h}^h,\;L_{h},\;L_{2h},\;S_h \right\} . \end{aligned}$$

This means that:

$$\begin{aligned} H \varphi _h( {\theta },\textbf{x} ) = \tilde{H}( {\theta } ) \varphi _h( {\theta },\textbf{x} ), \end{aligned}$$

where \(\tilde{H}( {\theta } )\) is called the Fourier symbol of the operator H. In a next step, we decompose the frequency space into a low frequency space \(T^{low} = \left( -\frac{\pi }{2},\frac{\pi }{2}\right] ^2\) and a high frequency space \(T^{high} = \varTheta \setminus T^{low}\). For each low frequency \({\theta } \in T^{low} = \left( -\frac{\pi }{2}, \frac{\pi }{2} \right] ^2,\) we define the following four frequencies:

$$\begin{aligned} {\theta }^{(0,0)} = ( \theta _1,\theta _2 ),\; {\theta }^{(1,1)} = ( \overline{\theta }_1,\overline{\theta }_2 ),\; {\theta }^{(1,0)} = ( \overline{\theta }_1,\theta _2 ),\; {\theta }^{(0,1)} = ( \theta _1,\overline{\theta }_2 ), \end{aligned}$$

where

$$\begin{aligned} \overline{{\theta }}_i = {\left\{ \begin{array}{ll} {\theta }_i + \pi &{}\quad \text { if } \theta _i<0,\\ {\theta }_i - \pi &{}\quad \text { if } \theta _i \ge 0. \end{array}\right. } \end{aligned}$$

Obviously, the four frequencies fulfil

$$\begin{aligned} \varphi _h( {\theta }^\alpha ,\textbf{x} ) = \varphi _{2h}( 2{\theta }^{(0,0)},\textbf{x} ),\; \textbf{x} \in \textbf{G}_{2h}, \;\mathbf {\alpha } \in I = \left\{ (0,0),\;(1,1),\;(1,0),\;(0,1)\right\} . \end{aligned}$$

Fourier modes with this relationship are called harmonic to each other. Considering a given low frequency \({\theta } = {\theta }^{(0,0)}\), we define a four dimensional space of harmonics by

$$\begin{aligned} E_h^{{\theta }} = \text {span}\left\{ \left. \varphi ( {\theta }^{\alpha }, \cdot ) \right| \mathbf {\alpha } \in I \right\} . \end{aligned}$$

An important feature of these spaces is that they are invariant under the twogrid operator \(T_h^{2h}\). Applying \(T_h^{2h}\) to an arbitrary element \(\varPsi \in E_h^{{\theta }}\) which is represented by some coefficients \(A^{\mathbf {\alpha }}\):

$$\begin{aligned} \varPsi = \sum _{\mathbf {\alpha } \in I} A^{\mathbf {\alpha }} \varphi ( {\theta }^{\alpha }, \cdot ), \end{aligned}$$

yields new coefficients \(B^{\mathbf {\alpha }}\):

$$\begin{aligned} \begin{pmatrix} B^{(0,0)} \\ B^{(1,1)} \\ B^{(1,0)} \\ B^{(0,1)} \end{pmatrix} = \hat{T}_h^{2h} ( {\theta },\sigma ) \begin{pmatrix} A^{(0,0)} \\ A^{(1,1)} \\ A^{(1,0)} \\ A^{(0,1)} \end{pmatrix}. \end{aligned}$$

\(\hat{T}_h^{2\,h} ( {\theta },\sigma ) \in \mathbb {C}^{4 \times 4}\) is a matrix depending on the frequency \({\theta }\) and the exponent \(\sigma \). It represents the operator \(T_h^{2h}\) with respect to \(E_h^{{\theta }}\). In order to determine the convergence behavior of the twogrid method, we study how the coefficients \(A^{\mathbf {\alpha }}\) are transformed by \(\hat{T}_h^{2h} ( {\theta },\sigma )\) into new coordinates \(B^{\mathbf {\alpha }}\). Therefore, the local convergence factor \(\rho _{loc}( \sigma )\) depending on the complex shift exponent in the twogrid operator is taken into account:

$$\begin{aligned} \rho _{loc}( \sigma ) = \sup \left\{ \left. \rho ( \hat{T}_h^{2h}( {\theta },\sigma ) ) \right| {\theta } \in T^{low}, {\theta } \notin \varLambda \right\} . \end{aligned}$$

Thereby \(\varLambda \subset \left( -\pi ,\pi \right] ^2\) denotes the set of frequencies for which \(\hat{L}_{2\,h}\) and \(\hat{L}_{h}\) are not invertible and \(\rho ( \hat{T}_h^{2h}( {\theta },\sigma ) )\) is the spectral radius of \(\hat{T}_h^{2h}( {\theta },\sigma )\). This means that finding the convergence factor for a fixed frequency is reduced to finding the spectral radius of a \(4\times 4\) matrix. To obtain the convergence rate of the twogrid solver, one has to search for the maximum of this quantity in the low frequency space \(T^{low}\). The search for the maximal spectral radius is restricted to \(T^{low}\), since the pattern of \(\rho ( \hat{T}_h^{2h} )\) in \(T^{low}\) is extended periodically to the whole frequency domain \(\varTheta \). Furthermore, an important feature of a multigrid solver is the damping of the low frequency error components [41].

To compute \(\hat{T}_h^{2h} ( {\theta },\sigma )\), the operators occurring in the definition of \(T_h^{2h}\) are represented with respect to the harmonic space \(E_h^{{\theta }}\) by means of \(4\times 4\) matrices. Exceptions are the prolongation, restriction and \(L^{-1}_{2h}\) operator whose matrices are given by \(1\times 4\), \(4\times 1\) and \(1\times 1\) matrices, since these are mappings related to \(E_h^{{\theta }}\) and \(E_{2h}^{{\theta }}\). The latter space is the space of harmonics with respect to the coarse grid \(\textbf{G}_{2h}\):

$$\begin{aligned} E_{2h}^{{\theta }} = \text {span}\left\{ \left. \varphi _{2h}( {\theta }^{\alpha }, \cdot ) \right| \alpha \in \left\{ (0,0) \right\} \right\} . \end{aligned}$$

The matrices defined with respect to the harmonic spaces are indicated by a hat:

$$\begin{aligned} \hat{T}_h^{2h}( {\theta },\sigma )&= \hat{S}_h^{\nu _2}( {\theta },\sigma ) \hat{K}_h^{2h}( {\theta },\sigma ) \hat{S}_h^{\nu _1}( {\theta },\sigma ), \\ \ \\ \hat{K}_h^{2h}( {\theta },\sigma )&= \hat{I}_h - \hat{I}_{2h}^h( {\theta } ) \hat{L}_{2h}^{-1}( 2{\theta },\sigma ) \hat{I}_h^{2h}( {\theta } ) \hat{L}_{h}( {\theta },\sigma ). \end{aligned}$$

In the following, we list the matrices with respect to the harmonic spaces. Obviously, the matrix for the identity operator \(\hat{I}_h\) is given by the identity matrix. The discretization operator \(L_h\) can also be represented by a diagonal matrix [12] [41, Chapter 2]:

$$\begin{aligned} \hat{L}_h( {\theta },\sigma ) = \text {diag} \begin{pmatrix} \tilde{L}_h( {\theta }^{(0,0)}, \sigma ),&\tilde{L}_h( {\theta }^{(1,1)}, \sigma ),&\tilde{L}_h( {\theta }^{(1,0)}, \sigma ),&\tilde{L}_h( {\theta }^{(0,1)}, \sigma ) \end{pmatrix}. \end{aligned}$$

The symbol of \(L_h\) is given by:

$$\begin{aligned} \tilde{L}_h( {\theta }, \sigma )&= \frac{1}{3} ( 8- 2 \cos (\theta _1) - 2 \cos (\theta _2) - 4 \cos (\theta _1) \cos (\theta _2) ) \\&\quad -\frac{\lambda h^2}{36} (16+8\cos (\theta _1) +8\cos (\theta _2) + 4 \cos (\theta _1) \cos (\theta _2) ). \end{aligned}$$

The matrix for the coarse grid operator and \({\theta } = {\theta }^{(0,0)}\) is given by:

$$\begin{aligned} \hat{L}_{2h}({\theta },\sigma ) = \tilde{L}_{2h}({\theta },\sigma ),\; L_{2h}(\sigma )\varphi _{2h}( 2{\theta },\textbf{x} ) = \tilde{L}_{2h}({\theta },\sigma ) \varphi _{2h}( 2{\theta },\textbf{x} ). \end{aligned}$$

The exact formula for \(\tilde{L}_{2h}({\theta },\sigma )\) reads as follows:

$$\begin{aligned} \tilde{L}_{2h}(2{\theta },\sigma )&= \frac{1}{3} ( 8- 2 \cos (2\theta _1) - 2 \cos (2\theta _2) - 4 \cos (2\theta _1) \cos (2\theta _2) ) \\&\quad -\frac{\lambda h^2}{9} (16+8\cos (2\theta _1) +8\cos (2\theta _2) + 4 \cos (2\theta _1) \cos (2\theta _2) ). \end{aligned}$$

In a next step the matrices for the restriction operator and the prolongation operator are derived:

$$\begin{aligned} \hat{I}_{2h}^h ({\theta } ) = \begin{pmatrix} \tilde{I}_{2h}^h( {\theta }^{(0,0)} ) \\ \tilde{I}_{2h}^h( {\theta }^{(1,1)} ) \\ \tilde{I}_{2h}^h( {\theta }^{(1,0)} ) \\ \tilde{I}_{2h}^h( {\theta }^{(0,1)} ) \end{pmatrix} = \begin{pmatrix} (1+\cos (\theta _1) ) (1+\cos (\theta _2) ) \\ (1-\cos (\theta _1) ) (1-\cos (\theta _2) ) \\ (1-\cos (\theta _1) ) (1+\cos (\theta _2) ) \\ (1+\cos (\theta _1) ) (1-\cos (\theta _2) ) \end{pmatrix} \text { and } \hat{I}_h^{2h}({\theta } ) = \frac{1}{4} \cdot (\hat{I}_{2h}^h ({\theta } ) )^T. \end{aligned}$$

Thereby the matrix entries are determined by the following equation:

$$\begin{aligned} I_{2h}^h \varphi _{2h}( 2{\theta },\textbf{x} ) = \sum _{\alpha } \tilde{I}_{2h}^h( {\theta }^\alpha ) \varphi _{h}( {\theta }^\alpha ,\textbf{x} ) \text { and } I_h^{2h} \varphi _{h}( {\theta }^\alpha ,\textbf{x} ) = \tilde{I}_h^{2h}({\theta }^\alpha ) \varphi _{2h}( 2{\theta }^{(0,0)},\textbf{x} ),\; \alpha \in I. \end{aligned}$$
Fig. 11
figure 11

Convergence factors \(\rho \) in the frequency space \(T^{low}\) in case of four different shifts \(\sigma \in \left\{ 1.0,1.25,1.75,2.0 \right\} \)

Fig. 12
figure 12

Convergence factor \(\rho \) for \(\theta _2=0\) and \(\theta _1 \in \left[ -\frac{\pi }{2},\frac{\pi }{2} \right] \).The exponents for the shift are chosen as follows: \(\sigma \in \left\{ 1.00,1.25,1.50,1.75,2.00 \right\} \)

We point out that the restriction and prolongation operator are independent of k and \(\sigma \). Summarizing the previous considerations, one obtains the matrix \(\hat{K}_h^{2h}({\theta },\sigma )\). Finally, it remains to specify the matrix for the smoother. According to [41, Chapter 4], it is given by a diagonal matrix:

$$\begin{aligned} \hat{S}_{h}( {\theta },\sigma ) = \text {diag} \begin{pmatrix} \tilde{S}_{h}( {\theta }^{(0,0)},\sigma ),&\tilde{S}_{h}( {\theta }^{(1,1)},\sigma ),&\tilde{S}_{h}( {\theta }^{(1,0)},\sigma ),&\tilde{S}_{h}( {\theta }^{(0,1)},\sigma ) \end{pmatrix}. \end{aligned}$$

A straightforward calculation yields for \({\theta } = {\theta }^{(0,0)}\):

$$\begin{aligned} \tilde{S}_{h}( {\theta }, \sigma )&= (1-\omega ) + \omega \frac{\frac{2}{3} + \lambda h^2 \frac{8}{36}}{\frac{8}{3} - \lambda h^2 \frac{16}{36}} \cos ( \theta _1 ) \\&\quad + \omega \frac{\frac{2}{3} + \lambda h^2 \frac{8}{36}}{\frac{8}{3} - \lambda h^2 \frac{16}{36}} \cos ( \theta _2 ) + \omega \frac{\frac{4}{3} + \lambda h^2 \frac{4}{36}}{\frac{8}{3} - \lambda h^2 \frac{16}{36}} \cos ( \theta _1 ) \cos ( \theta _2 ). \end{aligned}$$

Finally, we have all ingredients at hand to assemble the matrix \(\hat{T}_h^{2h}\) so that \(\rho _{loc}( \sigma )\) can be computed numerically.

In Fig. 11, we show some typical surfaces for \(\rho \) in the space consisting of low frequencies for \(\nu _1=\nu _2=3\), \(\omega = 2/3\), \(h=2^{-5}\) and \(k=0.35/h\). It can be observed that the maximum of the spectral radius \(\rho ( \hat{T}_h^{2h} )\), i.e., \(\rho _{loc}\) increases as \(\sigma \) is decreased from \(\sigma = 2\) towards \(\sigma = 1\). In addition to that, there is a symmetry with respect to the axes of the coordinate system. This can be explained by the fact that the entries of the Fourier symbols are composed of cosine functions (see “Appendix A”). We further observe numerically that the maxima are located on the axes for \(\theta _1 = 0\) or \(\theta _2 =0\) (see Fig. 1). Motivated by these observations, we restrict our search for the maximal spectral radius of \(\hat{T}_h^{2h}\) to the line with \(\theta _2 =0\) (see Fig. 12).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Drzisga, D., Köppl, T. & Wohlmuth, B. A Semi Matrix-Free Twogrid Preconditioner for the Helmholtz Equation with Near Optimal Shifts. J Sci Comput 95, 82 (2023). https://doi.org/10.1007/s10915-023-02195-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10915-023-02195-5

Keywords

Mathematics Subject Classification

Navigation