Skip to main content
Log in

A Symmetry-Based Decomposition Approach to Eigenvalue Problems

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

Abstract

In this paper, we propose a decomposition approach to differential eigenvalue problems with Abelian or non-Abelian symmetries. In the approach, we divide the original differential problem into eigenvalue subproblems which require less eigenpairs and can be solved independently. Our approach can be seamlessly incorporated with grid-based discretizations such as finite difference, finite element, or finite volume methods. We place the approach into a two-level parallelization setting, which saves the CPU time remarkably. For illustration and application, we implement our approach with finite elements and carry out electronic structure calculations of some symmetric cluster systems, in which we solve thousands of eigenpairs with millions of degrees of freedom and demonstrate the effectiveness of the approach.

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

Similar content being viewed by others

Notes

  1. In [6], \(\varSigma _g\) is the “internal” boundary \(\partial \varOmega _0\setminus \partial \varOmega \) of \(\varOmega _0\), and irreducible subdomain \(\varOmega _0\) is called symmetry cell.

  2. Cubic crystals are crystals where the unit cell is a cube. All irreducible representations of the associated symmetry group are one-, two-, or three-dimensional.

  3. Symmetry element of operation \(R\) is a point of reference about which \(R\) is carried out, such as a point to do inversion, a rotation axis, or a reflection plane. Symmetry element is invariant under the associated symmetry operation.

  4. We count the average CPU time of a single iteration for each subproblem and then accumulate them. Taking Table 4 for example, we have that \(4.73=17.95/32 + 15.53/26 + 14.88/25 + 14.59/24 + 15.95/28 + 15.78/27 + 14.17/23 + 14.92/25\).

  5. The configuration is visualized by PyMOL.

  6. The charge density is visualized by JADLIB [40] and JAVIS [28] developed by HPCC, Institute of Applied Physics and Computational Mathematics.

References

  1. Ackermann, J., Erdmann, B., Roitzsch, R.: A self-adaptive multilevel finite element method for the stationary Schrödinger equation in three space dimensions. J. Chem. Phys. 101, 7643–7650 (1994)

    Article  Google Scholar 

  2. Agmon, S.: Lectures on the Exponential Decay of Solutions of Second-Order Elliptic Operators. Princeton University Press, Princeton (1981)

    Google Scholar 

  3. Banjai, L.: Eigenfrequencies of fractal drums. J. Comput. Appl. Math. 198, 1–18 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  4. Beck, T.L.: Real-space mesh techniques in density-functional theory. Rev. Mod. Phys. 72, 1041–1080 (2000)

    Article  Google Scholar 

  5. Bossavit, A.: Symmetry, groups, and boundary value problems. A progressive introduction to noncommutative harmonic analysis of partial differential equations in domains with geometrical symmetry. Comput. Methods Appl. Mech. Eng. 56, 167–215 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  6. Bossavit, A.: Boundary value problems with symmetry, and their approximation by finite elements. SIAM J. Appl. Math. 53, 1352–80 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  7. Chelikowsky, J.R., Troullier, N., Saad, Y.: Finite-difference-pseudopotential method: electronic structure calculations without a basis. Phys. Rev. Lett. 72, 1240–1243 (1994)

    Article  Google Scholar 

  8. Cornwell, J.F.: Group Theory in Physics: An Introduction. Academic Press, California (1997)

    MATH  Google Scholar 

  9. Cotton, F.A.: Chemical Applications of Group Theory, 3rd edn. Wiley-Interscience, New York (1990)

  10. Dai, X., Gong, X., Yang, Z., Zhang, D., Zhou, A.: Finite volume discretizations for eigenvalue problems with applications to electronic structure calculations. Multiscale Model. Simul. 9, 208–240 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  11. Fang, J., Gao, X., Zhou, A.: A Kohn–Sham equation solver based on hexahedral finite elements. J. Comput. Phys. 231, 3166–3180 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  12. Fattebert, J.-L., Hornung, R.D., Wissink, A.M.: Finite element approach for density functional theory calculations on locally-refined meshes. J. Comput. Phys. 223, 759–773 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  13. Friesecke, G., Goddard, B.D.: Asymptotics-based CI models for atoms: properties, exact solution of a minimal model for Li to Ne, and application to atomic spectra. Multiscale Model. Simul. 7, 1876–1897 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  14. Gårding, L.: On the essential spectrum of Schrödinger operators. J. Funct. Anal. 52, 1–10 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  15. Genovese, L., Neelov, A., Goedecker, S., Deutsch, T., Ghasemi, S.A., Willand, A., Caliste, D., Zilberberg, O., Rayson, M., Bergman, A., Schneider, R.: Daubechies wavelets as a basis set for density functional pseudopotential calculations. J. Chem. Phys. 129, 014109 (2008)

    Article  Google Scholar 

  16. Golub, G.H., van Loan, C.F.: Matrix Computations. Johns Hopkins University Press, Baltimore (1996)

    MATH  Google Scholar 

  17. Gull, E., Millis, A.J., Lichtenstein, A.I., Rubtsov, A.N., Troyer, M., Werner, P.: Continuous-time Monte Carlo methods for quantum impurity models. Rev. Mod. Phys. 83, 349–404 (2011)

    Article  Google Scholar 

  18. Gong, X., Shen, L., Zhang, D., Zhou, A.: Finite element approximations for Schrödinger equations with applications to electronic structure computations. J. Comput. Math. 23, 310–327 (2008)

    MathSciNet  Google Scholar 

  19. Hasegawa, Y., Iwata, J.-I., Tsuji, M., Takahashi, D., Oshiyama, A., Minami, K., Boku, T., Shoji, F., Uno, A., Kurokawa, M., Inoue, H., Miyoshi, I., Yokokawa M.: First-principles calculations of electron states of a silicon nanowire with 100,000 atoms on the K computer. In: Proceedings of 2011 International Conference for High Performance Computing, Networking, Storage and Analysis (SC2011), pp. 1–11 (2011)

  20. Haule, K.: Quantum Monte Carlo impurity solver for cluster dynamical mean-field theory and electronic structure calculations with adjustable cluster base. Phys. Rev. B 75, 155113 (2007)

    Article  Google Scholar 

  21. Hohenberg, P., Kohn, W.: Inhomogeneous electron gas. Phys. Rev. B 136((3B)), B864–B871 (1964)

    Article  MathSciNet  Google Scholar 

  22. Kohn, W., Sham, L.J.: Self-consistent equations including exchange and correlation effects. Phys. Rev. 140((4A)), A1133–A1138 (1965)

    Article  MathSciNet  Google Scholar 

  23. Kronik, L., Makmal, A., Tiago, M.L., Alemany, M.M.G., Jain, M., Huang, X., Saad, Y., Chelikowsky, J.R.: Parsec—the pseudopotential algorithm for real-space electronic structure calculations: recent advances and novel applications to nano-structures. Phys. Stat. Sol. B. 243, 1063–1079 (2006)

    Article  Google Scholar 

  24. Kuttler, J.R., Sigillito, V.G.: Eigenvalues of the Laplacian in two dimensions. SIAM Rev. 26, 163–193 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  25. Lehoucq, R.B., Sorensen, D.C., Yang, C.: ARPACK Users’ Guide: Solution of Large-scale Eigenvalue Problems with Implicitly Restarted Arnoldi Methods. SIAM, Philadelphia (1998)

    Book  Google Scholar 

  26. Martin, R.M.: Electronic Structure: Basic Theory and Practical Methods. Cambridge University Press, Cambridge (2004)

    Book  Google Scholar 

  27. Mendl, C.B., Friesecke, G.: Efficient algorithm for asymptotics-based configuration-interaction methods and electronic structure of transition metal atoms. J. Chem. Phys. 133, 184101 (2010)

    Article  Google Scholar 

  28. Mo, Z., Zhang, A., (eds.): User’s guide for JASMIN Technical Report No. T09-JMJL-01. http://www.iapcm.ac.cn/jasmin (2009)

  29. Neuberger, J.M., Sieben, N., Swift, J.W.: Computing eigenfunctions on the Koch Snowflake: a new grid and symmetry. J. Comput. Appl. Math. 191, 126–142 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  30. Ono, T., Hirose, K.: Real-space electronic-structure calculations with a time-saving double-grid technique. Phys. Rev. B. 72, 085115 (2005)

    Article  Google Scholar 

  31. Pask, J.E., Sterne, P.A.: Finite element methods in ab initio electronic structure calculations. Model. Simul. Mater. Sci. Eng. 13, 71–96 (2005)

    Article  Google Scholar 

  32. Simon, B.: Schrödinger operators in the twentieth century. J. Math. Phys. 41, 3523–3555 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  33. Suryanarayana, P., Gavini, V., Blesgen, T.: Non-periodic finite-element formulation of Kohn–Sham density functional theory. J. Mech. Phys. Solids 58, 256–280 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  34. Tinkham, M.: Group Theory and Quantum Mechanics. McGraw-Hill, New York (1964)

    MATH  Google Scholar 

  35. Torsti, T., Eirola, T., Enkovaara, J., Hakala, T., Havu, P., Havu, V., Höynälänmaa, T., Ignatius, J., Lyly, M., Makkonen, I., Rantala, T.T., Ruokolainen, J., Ruotsalainen, K., Räsänen, E., Saarikoski, H., Puska, M.J.: Three real-space discretization techniques in electronic structure calculations. Phys. Stat. Sol. B243, 1016–1053 (2006)

    Article  Google Scholar 

  36. Trefethen, L.N., Betcke, T., Computed eigenmodes of planar regions. In: Recent advances in differential equations and mathematical physics, volume 412 of Contemp. Math., pp. 297–314. Providence, RI, Amer. Math. Soc (2006)

  37. Tsuchida, E., Tsukada, M.: Electronic-structure calculations based on the finite-element method. Phys. Rev. B. 52, 5573–5578 (1995)

    Article  Google Scholar 

  38. White, S.R., Wilkins, J.W., Teter, M.P.: Finite-element method for electronic structure. Phys. Rev. B. 39, 5819–5833 (1989)

    Article  Google Scholar 

  39. Wohlever, J.C.: Some computational aspects of a group theoretic finite element approach to the buckling and postbuckling analyses of plates and shells-of-revolution. Comput. Methods Appl. Mech. Eng. 170, 373–406 (1999)

    Article  MATH  Google Scholar 

  40. Xia, F., Chen, H., Song, L., Shen, W.: Design and implementation of numerical simulation mesh data model. J. Comput. Res. Develop. 46((Supp. I)), 258–264 (2009). (in Chinese)

    Google Scholar 

  41. Zhang, D., Shen, L., Zhou, A., Gong, X.: Finite element method for solving Kohn-Sham equations based on self-adaptive tetrahedral mesh. Phys. Lett. A. 372, 5071–5076 (2008)

    Article  MATH  Google Scholar 

  42. Zienkiewicz, O.C., Taylor, R.L.: The Finite Element Method for Solid and Structural Mechanics, 6th edn. Elsevier, London (2005)

    MATH  Google Scholar 

  43. Zingoni, A.: Group-theoretic exploitations of symmetry in computational solid and structural mechanics. Int. J. Numer. Methods Eng. 79, 253–289 (2009)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The authors would like to thank Prof. Xiaoying Dai, Prof. Xingao Gong, Prof. Lihua Shen, Dr. Zhang Yang, and Mr. Jinwei Zhu for their stimulating discussions on electronic structure calculations. The second author is grateful to Prof. Zeyao Mo for his encouragement. The authors would also like to thank the referee for his/her constructive comments and suggestions that improved the presentation of this paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Aihui Zhou.

Additional information

This work was partially supported by the National Science Foundation of China under Grant 61170310, the Funds for Creative Research Groups of China under Grant 11021101, the National Basic Research Program of China under Grants 2011CB309702 and 2011CB309703, and the National Center for Mathematics and Interdisciplinary Sciences, Chinese Academy of Sciences.

Appendices

Appendix 1: Basic Concept of Group Theory

In this appendix, we include some basic concepts of group theory for a more self-contained exposition. They could be found in standard textbooks like [8, 9, 34].

A group \(G\) is a set of elements \(\{R\}\) with a well-defined multiplication operation which satisfy the following requirements:

  1. 1.

    The set is closed under the multiplication.

  2. 2.

    The associative law holds.

  3. 3.

    There exists a unit element \(E\) such that \(ER=RE=R\) for any \(R\in G\).

  4. 4.

    There is an inverse \(R^{-1}\) in \(G\) to each element \(R\) such that \(RR^{-1} = R^{-1}R = E\).

If the commutative law of multiplication also holds, \(G\) is called an Abelian group. Group \(G\) is called a finite group if it contains a finite number of elements. And this number, denoted by \(g\), is said to be the order of the group. The rearrangement theorem tells that the elements of \(G\) are only rearranged by multiplying each by any \(R\in G\), i.e., \(RG=G\) for any \(R\in G\).

An element \(R_1\in G\) is called to be conjugate to \(R_2\) if \(R_2=SR_1S^{-1}\), where \(S\) is some element in the group. All the mutually conjugate elements form a class of elements. It can be proved that group \(G\) can be divided into distinct classes. Denote the number of classes as \(n_c\). In an Abelian group, any two elements are commutative, so each element forms a class by itself, and \(n_c\) equals the order of the group.

Groups \(G=\{R\}\) and \(G^{\prime }=\{R^{\prime }\}\) are called to be homomorphic, if there exists a correspondence between the elements of \(G\) and \(G^{\prime }\) as \(R\leftrightarrow R^{\prime }_1,R^{\prime }_2,\ldots \), which means that if \(RS=T\) then the product of any \(R^{\prime }_i\) with any \(S^{\prime }_j\) will belong to the set \(\{T^{\prime }_1,T^{\prime }_2,\ldots \}\). This is a many-to-one correspondence in general. If the correspondence specializes to be one-to-one, the two groups are called to be isomorphic.

A matrix representation of group \(G\) means a group of matrices which is homomorphic to \(G\). Two representations are said to be equivalent if they are associated by a similarity transformation. If a representation can not be equivalent to representations of lower dimensionality, it is called irreducible. Any matrix representation with nonzero determinants is equivalent to a unitary representation, i.e., a representation by unitary matrices.

The number of all the inequivalent, irreducible, unitary representations is equal to \(n_c\), which is the number of classes in \(G\). The Celebrated Theorem tells that

$$\begin{aligned} \sum _{\nu =1}^{n_c}d_{\nu }^2=g, \end{aligned}$$

where \(d_{\nu }\) denotes the dimensionality of the \(\nu \)th representation. Since the number of classes of an Abelian group equals the number of elements, an Abelian group of order \(g\) has \(g\) one-dimensional irreducible representations.

The groups used in this paper are all crystallographic point groups. Groups \(D_2,\;D_{2h},\;D_{2d}\) and \(D_4\) are four dihedral groups; the first two groups are Abelian and the other two are non-Abelian. In Tables 2 and 7, \(C_{nj}\) denotes a rotation about axis \(Oj\) by \(2\pi /n\) in the right-hand screw sense and \(I\) is the inversion operation [8]. The \(Oj\) axes are illustrated in Fig. 6. We refer to textbooks like [8, 9, 34] for more details about crystallographic point groups.

Fig. 6
figure 6

Rotation axes in Tables 2 and 7

Appendix 2: Proof of Proposition 2

Proof

(a) Since \(\{P_R\}\) are unitary operators, we have

$$\begin{aligned} {{\fancyscript{P}}^{(\nu )}_{ml}}^* = \frac{d_{\nu }}{g} \sum _{R\in G} \varGamma ^{(\nu )}(R)_{ml} P_{R^{-1}} = \frac{d_{\nu }}{g} \sum _{S\in G} \varGamma ^{(\nu )}(S^{-1})_{ml} P_S, \end{aligned}$$

which together with the fact that \(\varGamma ^{(\nu )}\) is a unitary representation derives

$$\begin{aligned} {{\fancyscript{P}}^{(\nu )}_{ml}}^* = \frac{d_{\nu }}{g} \sum _{S\in G} {\varGamma ^{(\nu )}(S)}^*_{lm} P_S = {\fancyscript{P}}^{(\nu )}_{lm}. \end{aligned}$$

(b) It follows from the definition that

$$\begin{aligned} {\fancyscript{P}}^{(\nu )}_{ml} {\fancyscript{P}}^{(\nu ^{\prime })}_{m^{\prime }l^{\prime }}&= \left( \frac{d_{\nu }}{g}\sum _{R\in G} {\varGamma ^{(\nu )}(R)}^*_{ml}P_R\right) \left( \frac{d_{\nu ^{\prime }}}{g} \sum _{S\in G} {\varGamma ^{(\nu ^{\prime })}(S)}^*_{m^{\prime }l^{\prime }}P_S\right) \\&= \frac{d_{\nu }d_{\nu ^{\prime }}}{g^2} \sum _{R\in G}{\varGamma ^{(\nu )}(R)}^*_{ml} \left( \sum _{S\in G} {\varGamma ^{(\nu ^{\prime })}(S)}^*_{m^{\prime }l^{\prime }}P_{RS} \right) . \end{aligned}$$

Note that the rearrangement theorem implies that, when \(S\) runs over all the group elements, \(S^{\prime }=RS\) for any \(R\) also runs over all the elements. Hence we get

$$\begin{aligned} {\fancyscript{P}}^{(\nu )}_{ml} {\fancyscript{P}}^{(\nu ^{\prime })}_{m^{\prime }l^{\prime }}&= \frac{d_{\nu }d_{\nu ^{\prime }}}{g^2} \sum _{R\in G}{\varGamma ^{(\nu )}(R)}^*_{ml} \left( \sum _{S^{\prime }\in G} {\varGamma ^{(\nu ^{\prime })}(R^{-1}S^{\prime })}^*_{m^{\prime }l^{\prime }}P_{S^{\prime }} \right) \\&= \frac{d_{\nu }d_{\nu ^{\prime }}}{g^2} \sum _{S^{\prime }\in G} \left( \sum _{R\in G} {\varGamma ^{(\nu )}(R)}^*_{ml}~{\varGamma ^{(\nu ^{\prime })}(R^{-1}S^{\prime })}^*_{m^{\prime }l^{\prime }} \right) P_{S^{\prime }}. \end{aligned}$$

We may calculate as follows

$$\begin{aligned} \sum _{R\in G} {\varGamma ^{(\nu )}(R)}^*_{ml} {\varGamma ^{(\nu ^{\prime })}(R^{-1}S^{\prime })}^*_{m^{\prime }l^{\prime }}&= \sum _{R\in G} {\varGamma ^{(\nu )}(R)}^*_{ml} \left( \sum _{n=1}^{d_{\nu ^{\prime }}} {\varGamma ^{(\nu ^{\prime })}(R^{-1})}^*_{m^{\prime }n}~{\varGamma ^{(\nu ^{\prime })}(S^{\prime })}^*_{nl^{\prime }} \right) \\&= \sum _{R\in G} {\varGamma ^{(\nu )}(R)}^*_{ml} \left( \sum _{n=1}^{d_{\nu ^{\prime }}} {\varGamma ^{(\nu ^{\prime })}(R)}_{nm^{\prime }}~{\varGamma ^{(\nu ^{\prime })}(S^{\prime })}^*_{nl^{\prime }} \right) \\&= \sum _{n=1}^{d_{\nu ^{\prime }}} {\varGamma ^{(\nu ^{\prime })}(S^{\prime })}^*_{nl^{\prime }} \left( \sum _{R\in G} {\varGamma ^{(\nu )}(R)}^*_{ml}~{\varGamma ^{(\nu ^{\prime })}(R)}_{nm^{\prime }} \right) , \end{aligned}$$

which together with the great orthogonality theorem yields

$$\begin{aligned} \sum _{R\in G} {\varGamma ^{(\nu )}(R)}^*_{ml} {\varGamma ^{(\nu ^{\prime })}(R^{-1}S^{\prime })}^*_{m^{\prime }l^{\prime }} = \delta _{\nu \nu ^{\prime }} \delta _{lm^{\prime }}\frac{g}{d_{\nu ^{\prime }}} {\varGamma ^{(\nu )}(S^{\prime })}^*_{ml^{\prime }}. \end{aligned}$$

Thus we arrive at

$$\begin{aligned} {\fancyscript{P}}^{(\nu )}_{ml} {\fancyscript{P}}^{(\nu ^{\prime })}_{m^{\prime }l^{\prime }} = \delta _{\nu \nu ^{\prime }} \delta _{lm^{\prime }} \frac{d_{\nu }}{g} \sum _{S^{\prime }\in G} {\varGamma ^{(\nu )}(S^{\prime })}^*_{ml^{\prime }} P_{S^{\prime }} = \delta _{\nu \nu ^{\prime }} \delta _{lm^{\prime }} {\fancyscript{P}}^{(\nu )}_{ml^{\prime }}. \end{aligned}$$

\(\square \)

Appendix 3: An Example

We take the Laplacian in square \((-1,1)^2\) as an example to illustrate the subproblem formulation in Corollary 2. Namely, we consider the decomposition of the following eigenvalue problem

$$\begin{aligned} \left\{ \begin{array}{rcll} -\Delta u &{} = &{}\lambda u &{}\quad \text{ in }~\varOmega =(-1,1)^2,\\ u &{} = &{}0 &{}\quad \text{ on }~\partial \varOmega . \end{array} \right. \end{aligned}$$
(27)

Note that \(G=\{E,\sigma _x,\sigma _y,I\}\) is a symmetry group associated with (27), where \(E\) represents the identity operation, \(\sigma _x\) a reflection about \(x\)-axis, \(\sigma _y\) a reflection about \(y\)-axis, and \(I\) the inversion operation. We see that \(G\) is an Abelian group of order 4, and has 4 one-dimensional irreducible representations as shown in Table 9.

Table 9 Representation matrices of example group \(G\)

According to Theorem 1 and Corollary 2, eigenvalue problem (27) can be decomposed into 4 subproblems (due to \(\sum _{\nu =1}^{n_c}d_{\nu }=4\)). And the symmetry characteristic conditions, the third equation in (13), for the 4 subproblems are

$$\begin{aligned} \{u^{(1)}(R_1x),u^{(1)}(R_2x),u^{(1)}(R_3x),u^{(1)}(R_4x)\}&= \{1, 1, 1, 1\} ~u^{(1)}(x), \end{aligned}$$
(28)
$$\begin{aligned} \{u^{(2)}(R_1x),u^{(2)}(R_2x),u^{(2)}(R_3x),u^{(2)}(R_4x)\}&= \{1, 1,-1,-1\} ~u^{(2)}(x), \end{aligned}$$
(29)
$$\begin{aligned} \{u^{(3)}(R_1x),u^{(3)}(R_2x),u^{(3)}(R_3x),u^{(3)}(R_4x)\}&= \{1,-1,-1, 1\} ~u^{(3)}(x), \end{aligned}$$
(30)
$$\begin{aligned} \{u^{(4)}(R_1x),u^{(4)}(R_2x),u^{(4)}(R_3x),u^{(4)}(R_4x)\}&= \{1,-1, 1,-1\} ~u^{(4)}(x), \end{aligned}$$
(31)

where \(x\in \varOmega \) is an arbitrary point and subscripts of \(\{u_1^{(\nu )}:\nu =1,2,3,4\}\) are omitted.

In Fig. 7, we illustrate four eigenfunctions of (27) belonging to different subproblems. We see that \(u_2\) and \(u_3\) are degenerate eigenfunctions corresponding to \(\lambda =\frac{5}{4}\pi ^2\) with double degeneracy. In other words, a doubly-degenerate eigenvalue of the original problem becomes nondegenerate for subproblems. This implies a relation between symmetry and degeneracy [24, 29, 36]. Moreover, the first subproblem does not have this eigenvalue, which shows that the decomposition approach has improved the spectral separation.

Fig. 7
figure 7

Four eigenfunctions of problem (27): \(u_1\) keeps invariant under \(\{E,\sigma _x,\sigma _y,I\}\) and satisfies Eq. (28), and \(u_2, u_3\) and \(u_4\) satisfy (29), (30) and (31), respectively

Appendix 4: Spatial Symmetry in Reciprocal Space

Plane wave method is widely used for solving the Kohn–Sham equations of crystals. Actually, plane waves may be regarded as grid-based discretizations in reciprocal space. We will show that the symmetry relation in real space is kept in reciprocal space. The solution domain \(\varOmega \) of crystals can be spanned by three lattice vectors in real space. We denote them as \(\mathbf{a}_i (i=1,2,3)\). If function \(f\) is invariant with integer multiple translations of the lattice vectors, we then present the function in reciprocal space as like:

$$\begin{aligned} \hat{f}(\mathbf{q})=\frac{1}{N}\sum _{ \mathbf {r}}f(\mathbf{r}) e^{-\imath \mathbf{q}\cdot \mathbf{r}}, \end{aligned}$$

where \(\mathbf{q}\) is any vector in reciprocal space satisfying \(\mathbf{q}\cdot \mathbf{a}_i=\frac{2\pi n}{N_i}\) with \(n\) an integer, \(N_i\) the number of degrees of freedom along direction \(\mathbf{a}_i~(i=1,2,3)\), and \(N=N_1N_2N_3\) the total number of degrees of freedom. Assume that \(f\) is kept invariant under coordinate transformation \(R\) in \(\varOmega \). We obtain from

$$\begin{aligned} \hat{f}(R\mathbf{q}) = \frac{1}{N}\sum _{ \mathbf{r}}f(\mathbf{r}) e^{-\imath (R\mathbf{q})\cdot \mathbf{r}} \end{aligned}$$

and the coordinate transformation \(R\) can be represented as an orthogonal matrix that

$$\begin{aligned} \hat{f}(R\mathbf{q}) = \frac{1}{N}\sum _{ \mathbf{r}}f( \mathbf{r}) e^{-\imath \mathbf{q}\cdot \left( R^{-1}\mathbf{r}\right) }. \end{aligned}$$

Since

$$\begin{aligned} f(R^{-1}\mathbf{r})=f(\mathbf{r})\quad \forall \mathbf{r}\in \varOmega , \end{aligned}$$

we have

$$\begin{aligned} \hat{f}(R\mathbf{q}) = \frac{1}{N}\sum _{R^{-1}\mathbf{r}}f(R^{-1}\mathbf{r}) e^{-\imath \mathbf{q}\cdot \left( R^{-1}\mathbf{r}\right) } = \hat{f}(\mathbf{q}). \end{aligned}$$

Hence the decomposition approach is probably applicable to plane waves.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Fang, J., Gao, X. & Zhou, A. A Symmetry-Based Decomposition Approach to Eigenvalue Problems. J Sci Comput 57, 638–669 (2013). https://doi.org/10.1007/s10915-013-9719-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10915-013-9719-7

Keywords

Mathematics Subject Classification (2000)

Navigation