A skew-symmetric Lanczos bidiagonalization method for computing several largest eigenpairs of a large skew-symmetric matrix
Abstract
The spectral decomposition of a real skew-symmetric matrix can be mathematically transformed into a specific structured singular value decomposition (SVD) of . Based on such equivalence, a skew-symmetric Lanczos bidiagonalization (SSLBD) method is proposed for the specific SVD problem that computes extreme singular values and the corresponding singular vectors of , from which the eigenpairs of corresponding to the extreme conjugate eigenvalues in magnitude are recovered pairwise in real arithmetic. A number of convergence results on the method are established, and accuracy estimates for approximate singular triplets are given. In finite precision arithmetic, it is proven that the semi-orthogonality of each set of basis vectors and the semi-biorthogonality of two sets of basis vectors suffice to compute the singular values accurately. A commonly used efficient partial reorthogonalization strategy is adapted to maintaining the needed semi-orthogonality and semi-biorthogonality. For a practical purpose, an implicitly restarted SSLBD algorithm is developed with partial reorthogonalization. Numerical experiments illustrate the effectiveness and overall efficiency of the algorithm.
keywords:
Skew-symmetric matrix, spectral decomposition, eigenvalue, eigenvector, singular value decomposition, singular value, singular vector, Lanczos bidiagonalization, partial reorthogonalizationAMS:
65F15, 15A18, 65F10, 65F251 Introduction
Let be a large scale and possibly sparse skew-symmetric matrix, i.e., , where the superscript denotes the transpose of a matrix or vector. We consider the eigenvalue problem
(1.1) |
where is an eigenvalue of and with the 2-norm is an corresponding eigenvector. It is well known that the nonzero eigenvalues of are purely imaginary and come in conjugate pairs. The real skew-symmetric eigenvalue problem arises in a variety of applications, such as matrix function computations [4, 7], the solution of linear quadratic optimal control problems [19, 39], model reductions [22, 37], wave propagation solution for repetitive structures [12, 38], crack following in anisotropic materials [1, 21], and some others [20, 26, 36].
For small to medium sized skew-symmetric matrices, several numerical algorithms have been developed to compute their spectral decompositions in real arithmetic [8, 23, 26, 35]. For a large scale skew-symmetric , however, it still lacks a high-performance numerical algorithm that can make use of its skew-symmetric structure to compute several eigenvalues and/or the corresponding eigenvectors only in real arithmetic. In this paper, we close this gap by proposing a Krylov subspace type method to compute several extreme conjugate eigenvalues in magnitude and the corresponding eigenvectors of that only uses real arithmetic.
Our work originates from a basic fact that, as will be shown in Section 2, the spectral decomposition of the skew-symmetric has a close relationship with its specifically structured singular value decomposition (SVD), where each conjugate pair of purely imaginary eigenvalues of corresponds to a double singular value of and the mutually orthogonal real and imaginary parts of associated eigenvectors are the corresponding left and right singular vectors of . Therefore, the solution of eigenvalue problem of the skew-symmetric is equivalent to the computation of its structured SVD, and the desired eigenpairs can be recovered from the converged singular triplets by exploiting this equivalence.
It has been well realized in [35] that the spectral decomposition of the skew-symmetric can be reduced to the SVD of a certain bidiagonal matrix whose size is for even. Exploiting such property, Ward and Gray [35] have designed an eigensolver for dense skew-symmetric matrices. For large scale, however, the computation of using orthogonal similarity transformations is prohibitive due to the requirement of excessive storage and computations. Nevertheless, with the help of Lanczos bidiagonalization (LBD) [24], whose complete bidiagonal reduction is originally due to [9], we are able to compute a sequence of leading principal matrices of , based on which, approximations to the extreme singular values of and/or the corresponding singular vectors can be computed. Along this line, a number of LBD based methods have been proposed and intensively studied, and several practical explicitly and implicitly restarted LBD type algorithms have been well developed [2, 3, 5, 11, 14, 15, 17, 18, 34].
For the skew-symmetric matrix , LBD owns some unique properties in both mathematics and numerics. However, direct applications of the aforementioned algorithms do not make use of these special properties, and will thus encounter some severe numerical difficulties. We will fully exploit these properties to propose a skew-symmetric LBD (SSLBD) method for computing several extreme singular triplets of , from which its extreme conjugate eigenvalues in magnitude and the associate eigenvectors are recovered. Attractively, our algorithm is performed in real arithmetic when computing complex eigenpairs of . Particularly, one conjugate pair of approximations to two conjugate pairwise eigenpairs can be recovered from one converged approximate singular triplet. We estimate the distance between the subspace generated by the real and imaginary parts of a pair of desired conjugate eigenvectors and the Krylov subspaces generated by SSLBD, and prove how it tends to zero as the subspace dimension increases. With these estimates, we establish a priori error bounds for the approximate eigenvalues and approximate eigenspaces obtained by the SSLBD method. One of them is the error bound for the two dimensional approximate eigenspace, which extends [29, pp.103, Theorem 4.6], a classic error bound for the Ritz vector in the real symmetric (complex Hermitian) case. These results show how fast our method converges when computing several extreme singular triplets.
As the subspace dimension increases, the SSLBD method will eventually become impractical due to the excessive memory and computational cost. For a practical purpose, it is generally necessary to restart the method by selecting an increasingly better starting vector when the maximum iterations allowed are attained, until the restarted algorithm converges ultimately. We will develop an implicitly restarted SSLBD algorithm. Initially, implicit restart was proposed by Sorensen [31] for large eigenvalue problems, and then it has been developed by Larsen [18] and Jia and Niu [14, 15] for large SVD computations. Our implicit restart is based on them but specialized to the skew-symmetric . We will show that two sets of left and right Lanczos vectors generated by SSLBD are meanwhile biorthogonal for the skew-symmetric . However, in finite precision arithmetic, this numerical biorthogonality loses gradually even if they themselves are kept numerically orthonormal. As a matter of fact, we have found that the numerical orthogonality of the computed left and right Lanczos vectors in an available state-of-the-art Lanczos bidiagonalization type algorithm does not suffice for the SVD problem of the skew-symmetric , and unfortunate ghosts appear when the numerical biorthogonality loses severely; that is, a singular value of may be recomputed a few times. Therefore, certain numerical biorthogonality is crucial to make the method work regularly.
As has been proved by Simon [30] for the symmetric Lanczos method on the eigenvalue problem and extended by Larsen [17] to the LBD method for the SVD problem, as far as the accurate computation of singular values is concerned, the numerical semi-orthogonality suffices for a general , and the numerical orthogonality to the level of machine precision is not necessary. Here semi-orthogonality means that two vectors are numerically orthogonal to the level of the square root of machine precision. To maintain the numerical stability and make the SSLBD method behave like it in exact arithmetic, we shall prove that the semi-orthogonality of each set of Lanczos vectors and the semi-biorthogonality of left and right Lanczos vectors suffice for the accurate computation of singular values of the skew-symmetric . We will seek for an effective and efficient reorthogonalization strategy for this purpose. Specifically, besides the commonly used partial reorthogonalization to maintain the desired semi-orthogonality of each set of Lanczos vectors, we will propose a partial reorthogonalization strategy to maintain the numerical semi-biorthogonality of two sets of Lanczos vectors in order to avoid the ghost phenomena and make the algorithm converge regularly. We will introduce such kind of reorthogonalization strategy into the implicitly restarted SSLBD algorithm and design a simple scheme to recover the desired conjugate eigenpairs of pairwise from the converged singular triplets.
The rest of this paper is organized as follows. In Section 2, we describe some properties of the spectral decomposition of , and show its mathematical equivalence with the SVD of . Then we establish a close relationship between the SVD of and that of an upper bidiagonal matrix whose order is half of that of when the order of is even. In Section 3, we present some properties of the SSLBD process, and propose the SSLBD method for computing several extreme singular values and the corresponding singular vectors of , from which conjugate approximate eigenpairs of are pairwise reconstructed. In Section 4, we establish convergence results on the SSLBD method. In Section 5, we design a partial reorthogonalization scheme for the SSLBD process, and develop an implicitly restarted SSLBD algorithm to compute several extreme singular triplets of . Numerical experiments are reported in Section 6. We conclude the paper in Section 7.
Throughout this paper, we denote by the imaginary unit, by and the conjugate transpose and pseudoinverse of , by the range space of , and by and the - and Frobenius norms of , respectively. We denote by and the real and complex spaces of dimension , by the identity matrix of order , and by and the zero matrices of orders and , respectively. The subscripts of identity and zero matrices are omitted when they are clear from the context.
2 Preliminaries
For the skew-symmetric matrix , the following five properties are well known and easily justified: Property (i): is diagonalizable by a unitary similarity transformation; Property (ii): the eigenvalues of are either purely imaginary or zero; Property (iii): for an eigenpair of , the conjugate is also an eigenpair of ; Property (iv): the real and imaginary parts of the eigenvector corresponding to a purely imaginary eigenvalue of have the same length, and are mutually orthogonal; Property (v): of odd order must be singular, and has at least one zero eigenvalue. It is easily deduced that a nonsingular must have even order, i.e., for some integer , and its eigenvalues are purely imaginary and come in conjugate, i.e., plus-minus, pairs.
We formulate the following basic results as a theorem, which establishes close structure relationships between the spectral decomposition of and its SVD and forms the basis of our method and algorithm to be proposed. For completeness and our frequent use in this paper, we will give a rigorous proof by exploiting the above five properties.
Theorem 2.1.
The spectral decomposition of the nonsingular skew-symmetric with is
(2.1) |
with
(2.2) |
where and are orthonormal and biorthogonal:
(2.3) |
and with . The SVD of is
(2.4) |
By Properties (i) and (ii) and the assumption that is nonsingular, we denote by the diagonal matrix consisting of all the eigenvalues of with positive imaginary parts, and by the corresponding orthonormal eigenvector matrix. By Property (iv), we can write as
(2.5) |
where the columns of and have unit-length, and and . By these and Property (iii), the conjugate of is
(2.6) |
and and . As a result, and defined by (2.2) are the eigenvector and eigenvalue matrices of , respectively, and
(2.7) |
Premultiplying the two sides of by delivers , which, by the skew-symmetry of , shows that
Since and is a positive diagonal matrix, it follows from the above equation that
By definition (2.2) of and the fact that and are column orthonormal, is unitary. Postmultiplying (2.7) by yields (2.1), from which it follows that (2.4) holds.
Inserting into and and solving the resulting equations for and , we obtain (2.3). This shows that both and are orthogonal. Therefore, (2.4) is the SVD of .
Remark 1.
For a singular skew-symmetric with where is the multiplicity of zero eigenvalues of , we can still write the spectral decomposition of as (2.1) with some slight rearrangements for the eigenvalue and eigenvector matrices and :
(2.8) |
where and are as in Theorem 2.1 and the columns of form an orthonormal basis of the null space of . The proof is analogous to that of Theorem 2.1 and thus omitted. In this case, relation (2.4) gives the thin SVD of .
For and in (2.4), write
On the basis of Theorem 2.1, in the sequel, we denote by
(2.9) |
The SVD (2.4) indicates that and are two singular triplets of corresponding to the multiple singular value . Therefore, in order to obtain the conjugate eigenpairs of , one can compute only the singular triplet or , and then recovers the desired eigenpairs from the singular triplet by (2.9).
Next we present a bidiagonal decomposition of the nonsingular skew-symmetric with order , which essentially appears in [35]. For completeness, we include a proof.
Theorem 2.2.
Assume that the skew-symmetric with is nonsingular. Then there exist two orthonormal matrices and satisfying and an upper bidiagonal matrix such that
(2.10) |
It is known from, e.g., [35] that is orthogonally similar to a skew-symmetric tridiagonal matrix, that is, there exists an orthogonal matrix such that
(2.11) |
Denote and with being the th column of , . Then it is easy to verify that
(2.12) |
Combining these two relations with (2.11), we obtain
(2.13) | |||||
which proves (2.10), where we have denoted and . Note that is orthogonal and , are not only orthonormal but also biorthogonal: and . Therefore, and .
Remark 2.
For a singular skew-symmetric with where is the multiplicity of zero eigenvalues of , the lower diagonal elements of in (2.11) are zeros. Taking and as in (2.12) and , we obtain
(2.14) |
with defined by (2.12). Using the same derivations as in the proof of Theorem 2.2, we have
where and are as in (2.13) and is the orthonormal basis matrix of the null space of . This decomposition is a precise extension of (2.10) to with any order.
Combining Remark 2 with Remark 1 after Theorem 2.1, for ease of presentation, in the sequel we will always assume that has even order and is nonsingular. However, our method and algorithm to be proposed apply to a singular skew-symmetric with only some slight modifications.
Decomposition (2.10) is a bidiagonal decomposition of that reduces to using the structured orthogonal matrices and from the left and right, respectively. The singular values of are the union of those of and but the SVD of alone delivers the whole SVD of : Let be the SVD of . Then Theorem 2.2 indicates that is an -dimensional partial SVD of with and being orthonormal and . By Theorem 2.1, the spectral decomposition (2.1) of can be restored directly from . This is exactly the working mechanism of the eigensolver proposed in [35] for small to medium sized skew-symmetric matrices.
For large, using orthogonal transformations to construct , and in (2.10) and computing the SVD of are unaffordable. Fortunately, based on decomposition (2.10), we can perform the LBD process on the skew-symmetric , i.e., the SSLBD process, to successively generate the columns of and and the leading principal matrices of , so that the Rayleigh–Ritz projection comes into the computation of a partial SVD of .
3 The SSLBD method
For ease of presentation, from now on, we always assume that the eigenvalues of in (2.9) are simple, and that are labeled in decreasing order:
(3.1) |
Our goal is to compute the pairs of extreme, e.g., largest conjugate eigenvalues and/or the corresponding eigenvectors defined by (2.9). By Theorem 2.1, this amounts to computing the following partial SVD of :
3.1 The -step SSLBD process
Algorithm 1 sketches the -step SSLBD process on the skew-symmetric , which is a variant of the lower LBD process proposed by Paige and Saunders [24].
Assume that the -step SSLBD process does not break down for , and note that and . Then the process computes the orthonormal base and of the Krylov subspaces
(3.2) |
generated by and the starting vectors and with , respectively. Denote for and for , whose columns are called left and right Lanczos vectors, respectively. Then the -step SSLBD process can be written in the matrix form:
(3.3) |
where is the th column of . It is well known that the singular values of are simple whenever the -step SSLBD process does not break down. For later use, denote . Then we can write the second relation in (3.3) as .
Theorem 3.1.
Let and be generated by Algorithm 1 without breakdown for . Then and are biorthogonal:
(3.4) |
We use induction on . Since and , by the skew-symmetry of and , , we have
which means that and thus proves (3.4) for .
Suppose that (3.4) holds for , i.e., . For , partition
(3.5) |
It follows from (3.3) and the inductive hypothesis that
which means that and . Since , , is nonsingular, from (3.3) and the above relations we have
(3.6) |
by noting that . Particularly, relation (3.6) shows that and . Making use of them and , and noticing that , we obtain
(3.7) | |||||
Applying (3.6) and (3.7) to (3.5) yields (3.4) for . By induction, we have proved that relation (3.4) holds for all .
Theorem 3.1 indicates that any vectors from and , which are called left and right subspaces, are biorthogonal. Therefore, any left and right approximate singular vectors extracted from and are biorthogonal, a desired property as the left and right singular vectors of are so, too.
The SSLBD process must break down no later than steps since has distinct singular values; see [13] for a detailed analysis on the multiple singular value case. Once breakdown occurs for some , all the singular values of are exact ones of , and exact singular triplets of have been found, as we shall see in Section 4. Furthermore, by the assumption that is nonsingular, breakdown can occur only for ; for if , then and thus would have a zero singular value.
3.2 The SSLBD method
With the left and right searching subspaces and , we use the standard Rayleigh–Ritz projection, i.e., the standard extraction approach [2, 3, 5, 14, 17, 18], to compute the approximate singular triplets of with the unit-length vectors and that satisfy the requirement
(3.8) |
Set and for . Then (3.3) indicates that (3.8) becomes
(3.9) |
that is, , are the singular triplets of . Therefore, the standard extraction amounts to computing the SVD of with the singular values ordered as and takes as approximations to some singular triplets of . The resulting method is called the SSLBD method, and the are called the Ritz approximations of with respect to the left and right searching subspaces and ; particularly, the are the Ritz values, and the and are the left and right Ritz vectors, respectively.
Since , by the Cauchy interlace theorem of eigenvalues (cf. [25, Theorem 10.1.1, pp.203]), for fixed, as increases, and monotonically converge to and from below and above, respectively. Therefore, we can take as approximations to the largest singular triplets , of with . Likewise, we may use , to approximate the smallest singular triplets , of .
4 A convergence analysis
For later use, for each , we denote
(4.1) |
Then , and is the two dimensional eigenspace of associated with the conjugate eigenvalues , . We call the pair a block eigenpair of , . From (2.3), is orthogonal, and (2.1) can be written as
(4.2) |
To make a convergence analysis on the SSLBD method, we review some preliminaries. Let the columns of , and form orthonormal base of , and the orthogonal complement of , respectively. Denote by the canonical angles between and [10, pp.329–330]. The 2-norm distance between and is defined by
which is the sine of the largest canonical angle between and (cf. Jia and Stewart [16]). It is worthwhile to point out that if the dimensions of and are not equal then this measure is not symmetric in its arguments. In fact, if the dimension of is greater than that of , then , although generally . In this paper, we will use the -norm distance
(4.3) |
which equals the square root of the squares sum of sines of all the canonical angles between the two subspaces. Correspondingly, we define to be the square root of the squares sum of the tangents of all the canonical angles between the subspaces. These tangents are the generalized singular values of the matrix pair .
Notice that for and with equal dimensions, if is nonsingular then the tangents of canonical angles are the singular values of , so that
(4.4) |
which holds when the -norm is replaced by the 2-norm. We remark that the inverse in the above cannot be replaced by the pseudoinverse when is singular. In fact, by definition, is infinite when is singular, and the generalized singular values of are not the singular values of in this case.
Note that the real and imaginary parts of the eigenvectors of can be interchanged since are the eigenvectors of and its left and right singular vectors and are the right and left ones, too (cf. (2.4)). As we have pointed out, any pair of approximate left and right singular vectors extracted from the biorthogonal and are mutually orthogonal, using which we construct the real and imaginary parts of an approximate eigenvector of . As a consequence, in the SVD context of the skew-symmetric , because of these properties, when analyzing the convergence of the SSLBD method, we consider the orthogonal direct sum as a whole, and estimate the distance between a desired two dimensional eigenspace of and the -dimensional subspace for small. We remind that their dimensions are unequal for . Using these distances and their estimates, we can establish a priori error bounds for approximate eigenvalues and approximate eigenspaces obtained by the SSLBD method, showing how fast they converge when computing largest eigenvalues in magnitudes and the associated eigenvectors.
In terms of the definition of , we present the following estimate for .
Theorem 4.1.
Let and be defined by (3.2), and suppose that with the initial is nonsingular. Then the following estimate holds for any integer :
(4.5) |
where is the degree Chebyshev polynomial of the first kind and
(4.6) |
For a fixed , denote by and which delete and from and in (4.2), respectively. Then relation (4.2) can be written as
(4.7) |
Notice that is orthonormal by Theorem 3.1, and . Since in (4.2) is orthogonal, there exists an orthonormal matrix with such that
(4.8) |
For an arbitrary polynomial in , the set of polynomials of degree no more than that satisfies , write and . Then
(4.9) |
From (4.1), it is known that , .
Since is nonsingular, , and the columns of form an orthonormal basis of , by definition (4.4) and relation (4.9), we obtain
(4.10) | |||||
Note that for any . By definition and (4.10), it holds that
where the last inequality comes from the proof of Theorem 1 in [27], is the degree Chebyshev polynomial of the first kind, and and are defined by (4.6).
Theorem 4.1 establishes accuracy estimates for approximating the searching subspaces . But one must be aware that they are only significant for . The scalars and defined by (4.6) are constants depending only on the eigenvalue or singular value distribution of . For a fixed integer , as long as the initial searching subspace contains some information on , that is, is finite in (4.5), then the larger is, the smaller is and the closer is to . Moreover, the better the singular value is separated from the other singular values of , the smaller and the larger are, meaning that approaches more quickly as increases. Generally, decays faster for smaller. Two extreme cases are that and . In the first case, Algorithm 1 breaks down at step , and we already have the exact eigenspace . The second case indicates that the initial is deficient in , such that does not contain any information on ; consequently, one cannot find any meaningful approximations to and from and for any .
Theorem 4.1 shows that the bounds tend to zero for small as increases. Using a similar proof, we can establish the following analogue of (4.5) for small:
(4.11) |
where
Similar arguments show that generally tends to zero faster for smaller as increases. It indicates that also favors the eigenvectors of corresponding to several smallest eigenvalues in magnitude if the smallest singular values are not clustered. In the sequel, for brevity, we only discuss the computation of the largest conjugate eigenvalues in magnitude and the associated eigenvectors of .
In term of , we next present a priori accuracy estimates for the Ritz approximations computed by the SSLBD method. To this end, we first establish the following two lemmas.
Lemma 4.2.
Let be the orthogonal projector onto the subspace . Then
(4.12) |
are the block eigenpairs of the linear operator restricted to :
(4.13) |
Since the columns of form an orthonormal basis of , the orthogonal projector onto is . By and , we have
It is known from the proof of Theorem 3.1 that . Making use of that and , as well as (3.9), by , we obtain
Lemma 4.3.
Since and with , by the fact that and , we have
(4.15) | |||||
Denote . Then the trace
applying which to (4.15) gives
(4.16) |
Making use of Lemmas 4.2–4.3, we are now in a position to establish a priori accuracy estimates for the Ritz approximations.
Theorem 4.4.
For , assume that is a Ritz approximation to the desired where is the closest to among the Ritz values. Denote and . Then
(4.17) |
where is the orthogonal projector onto the subspace .
Notice that the two dimensional subspace is the orthogonal projection of onto with respect to the -norm. Therefore, by the definition of , we have
(4.18) |
Let and be the orthonormal basis matrices of and the orthogonal complement of with respect to , respectively. The orthonormal basis matrix of can be expressed as
(4.19) |
where and . Since , from the above relation we have
Premultiplying this relation by and taking the -norms in the two hand sides, from , and , we obtain
(4.20) |
Recall the definition of in (4.12), and denote . Note that the columns of form an orthonormal basis of . Since , we can decompose its orthonormal basis matrix into
(4.21) |
where and . Lemma 4.2 shows that and with . Therefore, by (4.21), we obtain
(4.22) | |||||
where the last two relations hold since is orthonormal. Write
with each . Then by Lemma 4.3 we obtain
(4.23) | |||||
Therefore, combining (4.20), (4.22) and (4.23), we obtain
(4.24) |
Since both and are column orthonormal, decomposition (4.19) shows that
(4.25) |
Substituting (4.21) into (4.19) yields
(4.26) |
which, by the fact that the columns of are orthonormal, means that
Then (4.17) follows by in turn applying (4.24), (4.25) and (4.18) to the above relation.
Remark 3.
This theorem extends a classic result (cf. [29, pp.103, Theorem 4.6]) on the a priori error bound for the Ritz vector in the real symmetric (complex Hermitian) case to the skew-symmetric case for the Ritz block .
Theorem 4.5.
Decompose the orthonormal matrix defined by (4.12) into the orthogonal direct sum
(4.30) |
where and . Then
(4.31) |
By (4.13), (4.30) and , , we obtain
(4.32) |
where the last equality holds since with . Notice that and . Taking the -norms in the two hand sides of the above relation and exploiting and yield
(4.33) | |||||
Write , and notice that and . Then by (4.27) we have
(4.34) |
since, from (4.31), the 2-norms of columns and of are no more than one. As a consequence, we obtain
Let be the singular values of . Therefore, by definition we have and
Then it is straightforward to verify that
Applying the above inequality to the right hand side of (4.33), we obtain
Similarly, we get bound (4.29) from (4.32) by using and
We can amplify bound (4.28) as
Applying (4.17) to , we are able to estimate how fast tends to zero as increases. Since approaches for a fixed small as increases, if is well separated from the Ritz values of other than the approximate singular value , that is, is not small, then converges to as fast as tends to . In this case, Theorem 4.5 shows that the convergence of to is quadratic relative to that of . We should point out that assumption (4.27) is weak and is met very soon as increases.
For brevity, we now omit the subscript and denote by a Ritz approximation at the th step. We can then recover two conjugate approximate eigenpairs from in the form of . Note that their residuals are
(4.35) |
Obviously, if and only if is an exact singular triplet of . We claim that has converged if its residual norm satisfies
(4.36) |
where is a prescribed tolerance. In practical computations, one can replace by the largest Ritz value . Notice that
is nothing but the residual norm of the Ritz approximation of . Therefore, the eigenvalue problem and SVD problem of essentially share the same general-purpose stopping criterion.
By inserting and into (4.35) and making use of (3.3) and (3.9), it is easily justified that
(4.37) |
This indicates that we can calculate the residual norms of the Ritz approximations cheaply without explicitly forming the approximate singular vectors. We only compute the approximate singular vectors of until the corresponding residual norms defined by (4.37) drop below a prescribed tolerance. Moreover, (4.37) shows that once the -step SSLBD process breaks down, i.e., , all the computed approximate singular triplets are the exact ones of , as we have mentioned at the end of Section 3.1.
5 An implicitly restarted SSLBD algorithm with partial reorthogonalization
This section is devoted to efficient partial reorthogonalization and the development of an implicitly restarted SSLBD algorithm, which are crucial for a practical purpose and the numerical reliability so as to avoid ghost phenomena of computed Ritz approximations and make the algorithm behave as if it is in exact arithmetic.
5.1 Partial reorthogonalization
We adopt a commonly used two-sided partial reorthogonalization strategy, as done in [14, 15, 18], to make the columns of and numerically orthogonal to the level with the machine precision. In finite precision arithmetic, however, such semi-orthogonality cannot automatically guarantee the numerical semi-biorthogonality of and , causing that computed Ritz values have duplicates or ghosts and the convergence is delayed. To avoid this severe deficiency and make the method as if it does in exact arithmetic, one of our particular concerns is to design an effective and efficient partial reorthogonalization strategy, so that the entries of are in size; that is, the columns of and are numerically semi-biorthogonal. As it will turn out, how to achieve this goal efficiently is involved. Why the semi-orthogonality and semi-biorthogonality suffice is based on the following fundamental result.
Theorem 5.1.
Let the bidiagonal and and be produced by the -step SSLBD process. Assume that and defined in (3.3) are not small, and
(5.1) |
Let be the Q-factor in the QR decomposition of with the diagonals of the upper triangular matrix being positive, and define the orthonormal matrices and . Then
(5.2) |
where the entries of are in size.
Recall the equivalence of decompositions (2.10) and (2.11), and notice that decomposition (2.11) can be realized by the skew-symmetric Lanczos process, and (2.10) can be computed by our SSLBD process. The result then follows from a proof analogous to that of Theorem 4 in [30], and we thus omit details.
Theorem 5.1 shows that if the computed Lanczos vectors are semi-orthogonal and semi-biorthogonal then the upper bidiagonal matrix is, up to roundoff error, equal to the projection matrix of with respect to the left and right searching subspaces and . This means that the singular values of , the computed Ritz values, are as accurate as those of , i.e., the exact Ritz values, within the level . In other words, the semi-orthogonality and semi-biorthogonality suffice to guarantee that the computed and true Ritz values are equally accurate in finite precision arithmetic. Therefore, the full numerical orthogonality and biorthogonality at the level do not help and instead cause unnecessary waste for the accurate computation of singular values.
To achieve the desired numerical semi-orthogonality and semi-biorthogonality, we first need to efficiently estimate the levels of orthogonality and biorthogonality among all the Lanczos vectors to monitor if (5.1) is guaranteed. Whenever (5.1) is violated, we will use an efficient partial reorthogonalization to restore (5.1). To this end, we adapt the recurrences in Larsen’s PhD thesis [17] to the effective estimations of the desired semi-orthogonality and semi-biorthogonality, as is shown below.
Denote by and the estimate matrices of orthogonality between the columns of and , respectively, and by the estimate matrix of biorthogonality of and . That is, their -elements , and , respectively. Adopting the recurrences in [17], we set for , and at the th step use the following recurrences to compute
(5.3) | |||||
(5.4) |
where is the sign function, and is an estimate for . The terms in (5.3) and (5.4) take into account the rounding errors in LBD, which makes the calculated items as large as possible in magnitude so that it is safe to take them as estimates for the levels of orthogonality among the left and right Lanczos vectors. Based on the sizes of and , one can decide to which previous Lanczos vectors the newly computed one should be reorthogonalized.
Making use of the skew-symmetry of , we can derive analogous recurrences to those in [17] to compute the elements of . The derivations are rigorous but tedious, so we omit details here, and present the recurrences directly. Initially, set and for and . At the th step, we compute the elements of by
(5.6) | |||||
Having computed , and , at the th step of the SSLBD process, we determine the index sets:
(5.7) | |||||
(5.8) |
These four sets consist of the indices that correspond to the left and right Lanczos vectors and that have lost the semi-orthogonality and semi-biorthogonality with and , respectively. We then make use of the modified Gram–Schmidt (MGS) orthogonalization procedure [10, 28, 32] to reorthogonalize the newly computed left Lanczos vector first against the left Lanczos vector(s) with and then the right one(s) with . Having done these, we reorthogonalize the right Lanczos vector , which has newly been computed using the updated , first against the right Lanczos vector(s) for and then the left one(s) for .
- (3.1)
-
(3.2)
for do partially reorthogonalize to
-
Calculate , and update .
-
Overwrite for and for .
-
-
end for
-
(3.3)
for do partially reorthogonalize to
-
Compute , and update .
-
Overwrite for and for .
-
-
end for
- (3.4)
- (5.1)
-
(5.2)
for do partially reorthogonalize to
-
Calculate , and update .
-
Overwrite and for .
-
-
end for
-
(5.3)
for do partially reorthogonalize to
-
Compute , and update .
-
Overwrite and for .
-
-
end for
- (5.4)
Remarkably, once a Lanczos vector is reorthogonalized to a previous one, the relevant elements in , and change correspondingly, which may no longer be reliable estimates for the levels of orthogonality and biorthogonality among the updated Lanczos vector and all the previous ones. To this end, we need to update those changed values of , and by the MGS procedure111In his PhD thesis [17], Larsen resets those quantities as and includes the parts in of (5.3)–(5.6). But he does not explain why those quantities can automatically remain as the procedure is going on. In our procedure, we guarantee those vectors to be numerically orthogonal and biorthogonal by explicitly performing the MGS procedure.. Since and are symmetric matrices with diagonals being ones, it suffices to compute their strictly upper triangular parts. Algorithm 2 describes the whole partial reorthogonalization process in SSLBD, where we insert steps (3.1)–(3.4) and (5.1)–(5.4) between steps 3–4 and steps 5–6 of Algorithm 1, respectively. In such a way, we have developed an -step SSLBD process with partial reorthogonalization.
We see from Algorithm 2 and relations (5.3)–(5.6) that the th step takes flops to compute the new elements of , and for the first time, which is negligible relative to the cost of performing matrix-vector multiplications with at that step. Also, it takes extra flops to update the elements of , and after one step of reorthogonalization. Therefore, we can estimate the levels of orthogonality and biorthogonality very efficiently.
5.2 An implicitly restarted algorithm
When the dimension of the searching subspaces reaches the maximum number allowed, implicit restart needs to select certain shifts. The shifts can particularly be taken as the unwanted Ritz values , called the exact shifts [18, 31]. For those so-called bad shifts, i.e., those close to some of the desired singular values, we replace them with certain more appropriate values. As suggested in [17] and adopted in [14, 15] with necessary modifications when computing the smallest singular triplets, we regard a shift as a bad one and reset it to be zero if
(5.9) |
where is the residual norm of the approximate singular triplet with and defined by (4.35).
Implicit restart formally performs implicit QR iterations [29, 33] on with the shifts by equivalently acting Givens rotations directly on and accumulating the orthogonal matrices and with order such that
Using these matrices, we first compute
(5.10) |
where is the th leading principal submatrix of , and and consist of the first columns of and , respectively. We then compute
(5.11) |
with the st column of and , the - and -elements of and , respectively. As a result, when the above implicit restart is run, we have already obtained a -step SSLBD process rather than over scratch. We then perform -step SSLBD process from step onwards to obtain a new -step SSLBD process.
Once the SSLBD process is restarted, the matrices , and must be updated. This can be done efficiently by making use of (5.10) and (5.11). Note that and are symmetric matrices with orders and , respectively, and is an flat matrix. The corresponding updated matrices are with orders , and size , respectively, which are updated by
and
Here we have denoted by and the leading principal submatrix and the first entries in the th column of the estimate matrices of orthogonality and biorthogonality, respectively. The updating process needs extra flops, which is negligible relative to the -step implicit restart for .
As we have seen previously, the convergence criterion (4.36) and the efficient partial reorthogonalization scheme (cf. (5.3)–(5.6)) depend on a rough estimate for , which itself is approximated by the largest approximate singular value in our algorithm. Therefore, we simply set to replace in (4.36). Notice that, before the end of the first cycle of the implicitly restarted algorithm, there is no approximate singular value available. We adopt a strategy analogous to that in [17] to estimate : Set for , and update it for by
where .
Algorithm 3 sketches the implicitly restarted SSLBD algorithm with partial reorthogonalization for computing the largest singular triplets of . It requires the device to compute for an arbitrary -dimensional vector and the number of the desired conjugate eigenpairs of . The number is the maximum subspace dimension, is the maximum restarts allowed, is the stopping tolerance in (4.36), and the unit-length vector is an initial right Lanczos vector. The defaults of these parameters are set as , , and , respectively.
6 Numerical examples
In this section, we report numerical experiments on several matrices to illustrate the performance of our implicitly restarted SSLBD algorithm with partial reorthogonalization for the eigenvalue problem of a large skew-symmetric . The algorithm will be abbreviated as IRSSLBD, and has been coded in the Matlab language. All the experiments were performed on an Intel (R) core (TM) i9-10885H CPU 2.40 GHz with the main memory 64 GB and 16 cores using the Matlab R2021a with the machine precision under the Microsoft Windows 10 64-bit system.
plsk1919 | 1919 | 9662 | 1.71 | 3.79e-17 | 1.34e-1 | 5.17e-3 | 5.17e-3 |
tols4000 | 4000 | 9154 | 1.17e+7 | 9.83e-56 | 4.97e-3 | 4.97e-3 | 4.97e-3 |
M80PI | 4028 | 8066 | 2.84 | 6.60e-18 | 7.03e-3 | 3.68e-3 | 3.68e-3 |
visco1 | 4326 | 71572 | 3.70e+1 | 5.42e-21 | 1.59e-2 | 3.50e-4 | 3.50e-4 |
utm5940 | 5940 | 114590 | 1.21 | 1.83e-7 | 5.14e-2 | 9.57e-4 | 9.57e-4 |
coater2 | 9540 | 264448 | 3.11e+2 | 1.35e-13 | 9.89e-3 | 7.38e-3 | 4.51e-4 |
flowm | 9669 | 54130 | 3.32e-2 | 7.56e-36 | 2.87e-3 | 8.27e-4 | 6.88e-4 |
inlet | 11730 | 440562 | 3.35 | 6.17e-7 | 1.61e-2 | 5.63e-4 | 5.63e-4 |
e40r0 | 17281 | 906340 | 1.04e+1 | 1.54e-16 | 9.64e-2 | 3.99e-3 | 2.00e-3 |
ns3Da | 20414 | 1660392 | 5.12e-1 | 2.35e-6 | 2.59e-2 | 2.38e-3 | 8.37e-5 |
epb2 | 25228 | 199152 | 1.09 | 2.85e-12 | 1.15e-2 | 8.89e-4 | 1.86e-4 |
barth | 6691 | 39496 | 2.23 | 5.46e-18 | 5.98e-3 | 1.85e-5 | 1.85e-5 |
Hamrle2 | 5952 | 44282 | 5.55e+1 | 6.41e-5 | 5.27e-6 | 5.27e-6 | 2.56e-6 |
delf | 9824 | 30794 | 3.92e+3 | 0 | 3.12e-3 | 3.12e-3 | 3.12e-3 |
large | 12899 | 41270 | 4.04e+3 | 0 | 2.21e-3 | 2.21e-3 | 2.21e-3 |
Marag | 31407 | 1075388 | 1.33e+1 | 0 | 2.78e-1 | 1.66e-2 | 7.89e-3 |
r05 | 14880 | 208290 | 1.82e+1 | 0 | 2.07e-3 | 2.07e-3 | 5.73e-4 |
nl | 22364 | 94070 | 2.87e+2 | 0 | 2.27e-2 | 1.38e-3 | 7.94e-4 |
deter7 | 24528 | 74262 | 9.67 | 0 | 3.42e-1 | 5.56e-4 | 5.56e-4 |
Kemel | 38145 | 201750 | 2.41e+2 | 0 | 2.17e-2 | 6.96e-3 | 5.50e-4 |
p05 | 14680 | 118090 | 1.28e+1 | 0 | 5.08e-3 | 5.08e-3 | 1.93e-4 |
storm | 51926 | 188548 | 5.45e+2 | 0 | 1.64e-2 | 2.28e-5 | 2.28e-5 |
south31 | 54746 | 224796 | 1.29e+4 | 0 | 6.86e+1 | 3.96e-1 | 2.02e-2 |
dano3 | 19053 | 163266 | 1.82e+3 | 0 | 1.17e-3 | 2.31e-6 | 2.31e-6 |
Table 1 lists some basic properties of the test skew-symmetric matrices and with being the square and rectangular real matrices selected from the University of Florida Sparse Matrix Collection [6], respectively, where denotes the number of nonzero elements of , and are the largest and smallest singular values of , respectively, and . All the singular values of are computed, for the experimental purpose, by using the Matlab built-in function svd. As Theorem 4.1 shows, the size of critically affects the performance of IRSSLBD for computing the largest conjugate eigenpairs of ; the bigger is, the faster IRSSLBD converges. Otherwise, IRSSLBD may converge slowly.
Unless mentioned otherwise, we always run the IRSSLBD algorithm with the default parameters described in the end of Section 5. We also test the Matlab built-in functions eigs and svds in order to illustrate the efficiency and reliability of IRSSLBD. Concretely, with the same parameters as those in IRSSLBD, we use eigs to compute eigenpairs corresponding to the largest conjugate eigenvalues in magnitude and the corresponding eigenvectors of , and use svds to compute the or largest singular triplets of to obtain the largest eigenpairs of . We record the number of matrix-vector multiplications, abbreviated as , and the CPU time in second, denoted by , that the three algorithms use to achieve the same stopping tolerance. We mention that, for most of the test matrices, the CPU time used by each of the three algorithms is too short, e.g., fewer than 0.1s, so that it is hard to make a convincing and reliable comparison using CPU time. Therefore, we will not report the CPU time in most of the experiments. We should remind that a comparison of CPU time used by our algorithm and eigs, svds may also be misleading because our IRSSLBD code is the pure Matlab language while eigs and svds were programmed using advanced and much higher efficient C/C++ language. Nevertheless, we aim to illustrate that IRSSLBD is indeed fast and not slower than eigs and svds even in terms of CPU time.
Experiment 6.1.
We compute pairs of the largest conjugate eigenvalues in magnitude of and the corresponding eigenvectors.
Algorithm | ||||||
---|---|---|---|---|---|---|
Mv | Mv | Mv | ||||
eigs | 58 | 4.52e-3 | 122 | 1.02e-2 | 186 | 3.06e-2 |
SVDS() | 118 | 4.11e-3 | 258 | 1.22e-2 | 416 | 1.61e-2 |
SVDS() | 266 | 8.35e-3 | 416 | 1.61e-2 | 448 | 2.04e-2 |
IRSSLBD | 50 | 8.26e-3 | 106 | 4.03e-2 | 134 | 5.20e-2 |
Table 2 displays the Mv and used by the three algorithms, where SVDS() and SVDS() indicate that we use svds to compute the and largest singular triplets of , respectively. We have found that IRSSLBD and eigs solved the SVD problems successfully and computed the desired eigenpairs of correctly. In terms of Mv, we can see from the table that IRSSLBD is slightly more efficient than eigs for but outperforms eigs substantially for . IRSSLBD is thus superior to eigs for this matrix. We also see from Table 2 that the CPU time used by IRSSLBD is very short and competitive with that used by eigs. The same phenomena have been observed for most of the test matrices.
Remarkably, we have found that, without using explicit partial reorthogonalization to maintain the semi-biorthogonality of the left and right Lanczos vectors, IRSSLBD and svds encounter some severe troubles, and they behave similarly. Indeed, for , SVDS() succeeded in computing the largest singular triplet of , from which the desired largest conjugate eigenpairs are recovered as described in Section 3.2. For and , however, SVDS() computes duplicate approximations to the singular triplets associated with each conjugate eigenvalue pair of , and ghosts appeared, causing that SVDS() computed only half of the desired eigenpairs, i.e., the first to third and the first to fifth conjugate eigenpairs, respectively. The ghost phenomena not only delay the convergence of SVDS() considerably but also make it converge irregularly and fail to find all the desired singular triplets. SVDS() also uses much more Mv than IRSSLBD does for . Likewise, for each , SVDS() succeeds to compute only approximations to the desired largest singular triplets of , from which the desired approximate eigenpairs can be recovered, but consumes times Mv as many as those used by IRSSLBD. This shows that direct application of an LBD algorithm to skew-symmetric matrix eigenvalue problems does not work well in finite precision arithmetic. In contrast, IRSSLBD works well when the semi-orthogonality and semi-biorthogonality are maintained. Therefore, the semi-biorthogonality of left and right Lanczos vectors is crucial.
Experiment 6.2.
We compute pairs of largest conjugate eigenvalues in magnitude and the corresponding eigenvectors of the twelve square matrices with , , , , , , , , , , and , respectively. We will report the results on IRSSLBD and eigs only since svds behaves irregularly and computes only partial desired singular triplets because of the severe loss of numerical biorthogonality of left and right Lanczos vectors.
Algorithm: eigs | Algorithm: IRSSLBD | |||||
---|---|---|---|---|---|---|
tols4000 | 338 | 278 | 460 | 174 (51.48) | 232 (83.45) | 292 (63.48) |
M80PI | 58 | 110 | 254 | 44 (75.86) | 100 (90.91) | 158 (62.20) |
visco1 | 198 | 302 | 912 | 132 (66.67) | 282 (93.38) | 370 (40.57) |
utm5940 | 86 | 336 | 788 | 76 (88.37) | 278 (82.74) | 306 (38.83) |
coater2 | 142 | 162 | 282 | 84 (59.15) | 130 (80.25) | 152 (53.90) |
flowm | 478 | 384 | 702 | 228 (47.70) | 346 (90.10) | 394 (56.13) |
inlet | 170 | 412 | 572 | 132 (77.65) | 316 (76.70) | 310 (54.20) |
e40r0 | 86 | 168 | 576 | 70 (81.40) | 148 (88.10) | 284 (49.31) |
ns3Da | 114 | 226 | 960 | 92 (80.70) | 204 (90.27) | 376 (39.17) |
epb2 | 198 | 328 | 828 | 138 (69.70) | 278 (84.76) | 384 (46.38) |
barth | 310 | 392 | 510 | 172 (55.48) | 318 (81.12) | 322 (63.14) |
Hamrle2 | 3558 | 300 | 16178 | 136 ( 3.82) | 190 (63.33) | 354 ( 2.19) |
IRSSLBD and eigs successfully compute the desired eigenpairs of all the test matrices. Table 3 displays the Mv used by them, where each quantity in the parenthesis is the percentage of Mv used by IRSSLBD over that by eigs and we drop to save the space.
For , we see from Table 3 that IRSSLBD is always more efficient than eigs, and it often outperforms the latter considerably for most of the test matrices. For instance, IRSSLBD uses fewer than Mv of eigs for tols4000, coater2, flowm and barth. Strikingly, IRSSLBD consumes only of the Mv used by eigs for . For , IRSSLBD is often considerably more efficient than eigs for most of the test matrices, especially Hamrle2, where IRSSLBD uses fewer Mv. For , IRSSLBD uses – of the Mv consumed by eigs for tols4000, M80PI and barth, and it uses – of the Mv for coater2, flowm, and inlet. Therefore, for these six matrices, the advantage of IRSSLBD over eigs is considerable. For the five matrices visco1, utm5940, e40r0, ns3Da and epb2, IRSSLBD is even more efficient than as it consumes no more than half of the Mv used by eigs. For Hamrle2, IRSSLBD only uses of the Mv cost by eigs, and the save is huge.
In summary, for these twelve test matrices, our IRSSLBD algorithm outperforms eigs and is often considerably more efficient than the latter for the given three .
Experiment 6.3.
We compute pairs of the largest conjugate eigenvalues of and the corresponding eigenvectors with the eleven matrices delf, large, Marag, r05, nl, deter7, Kemel, p05, storm, south31 and dano3.
Algorithm: eigs | Algorithm: IRSSLBD | |||||
---|---|---|---|---|---|---|
delf | 478 | 214 | 192 | 142 (29.71) | 142 (66.36) | 138 (71.88) |
large | 590 | 284 | 268 | 188 (31.86) | 172 (60.56) | 172 (64.18) |
Marag | 58 | 84 | 178 | 38 (65.52) | 82 (97.62) | 130 (73.03) |
r05 | 86 | 70 | 668 | 54 (62.79) | 56 (80.00) | 222 (33.23) |
nl | 114 | 242 | 222 | 76 (66.67) | 172 (71.07) | 126 (56.76) |
deter7 | 58 | 346 | 144 | 36 (62.07) | 148 (42.77) | 122 (84.72) |
Kemel | 142 | 220 | 636 | 118 (83.10) | 198 (90.00) | 306 (48.11) |
p05 | 114 | 90 | 692 | 58 (50.88) | 62 (68.89) | 234 (32.82) |
storm | 58 | 84 | 78 | 46 (79.31) | 64 (76.19) | 66 (84.62) |
south31 | 30 | 42 | 146 | 8 (26.67) | 26 (61.90) | 92 (63.01) |
dano3 | 114 | 70 | 5922 | 58 (50.88) | 70 (100.0) | 196 ( 3.31) |
These test matrices are all singular with large dimensional null spaces. In order to purge the null space from the searching subspaces, we take the initial vector in IRSSLBD and eigs to be the unit-length vector normalized from . All the other parameters are the same for the two algorithms with the default values as described previously. Both IRSSLBD and eigs succeed in computing all the desired eigenpairs of for the three ’s. Table 4 displays the #Mv, where the quantities in the parentheses are as in Table 3.
We have also observed an interesting phenomenon from this table and the previous experiments. As is seen from Table 4, for some test matrices, e.g., delf, large, nl, deter7, p05 and dano, when more conjugate eigenpairs of are desired, IRSSLBD or eigs may use even fewer Mv. This occurs mainly due to the implicit restart technique. Specifically, when more conjugate eigenpairs are desired, larger dimensional new initial searching subspaces and are retained in the next restart, so that the resulting restarted searching subspaces may contain more information on the desired eigenvectors, causing that the largest approximate eigenpairs may converge faster.
As for an overall efficiency comparison of IRSSLBD and eigs, we have also observed the advantage of the former over the latter, similar to that in Experiment 6.2. Therefore, we now make some comments on them together. Before doing so, let us regard each test matrix with each selected as one independent problem, so in total we have 33 problems, which will be distinguished by the name-value pairs .
As is seen from Table 4, for and , IRSSLBD and eigs use almost the same Mv. For , , and , IRSSLBD is slightly better than eigs and uses – of the Mv. For , , , , , , , , , , , , , , , and , IRSSLBD outperforms eigs considerably as it costs – of the Mv used by eigs. For , , , , , , , , and , IRSSLBD is substantially more efficient than eigs as it uses only approximately half of the Mv or even fewer. Particularly, for , the advantage of IRSSLBD over eigs is very substantial since, Mv used by the former is only of that by the latter.
In summary, by choosing an appropriate starting vector, IRSSLBD suits well for computing several largest conjugate eigenpairs of a singular skew-symmetric matrix, and it is more and can be much more efficient than eigs.
7 Conclusions
We have shown that the eigendecomposition of a real skew-symmetric matrix has a close relationship with its structured SVD, by which the computation of its several extreme conjugate eigenpairs can be equivalently transformed into the computation of the largest singular triplets in real arithmetic. For large, as a key step toward the efficient computation of the desired eigenpairs, by means of the equivalence result on the tridiagonal decomposition of and a half sized bidiagonal decomposition, the SSLBD process is exploited to successively compute a sequence of partial bidiagonal decompositions. The process provides us with a sequence of left and right searching subspaces and bidiagonal matrices, making the computation of a partial SVD of possible, from which the desired eigenpairs of is obtained without involving complex arithmetic.
We have made a detailed theoretical analysis on the SSLBD process. Based on the results obtained, we have proposed a SSLBD method for computing several extreme singular triplets of , from which the desired extreme conjugate eigenvalues in magnitude and the corresponding eigenvectors can be recovered. We have established estimates for the distance between a desired eigenspace and the underlying subspaces that the SSLBD process generates, showing how it converges to zero. In terms of the distance between the desired eigenspace and the searching subspaces, we have derived a priori error bounds for the recovered conjugate eigenvalues and the associated eigenspaces. The results indicate that the SSLBD method generally favors extreme conjugate eigenpairs of .
We have made a numerical analysis on the SSLBD process and proposed an efficient and reliable approach to track the orthogonality and biorthogonality among the computed left and right Lanczos vectors. Unlike the standard LBD process, for its skew-symmetric variant SSLBD, we have proved that only the semi-orthogonality of the left and right Lanczos vectors does not suffice and that the semi-biorthogonality of these two sets of vectors is absolutely necessary, which has been numerically confirmed. Based on these, we have designed an effective partial reorthogonalization strategy for the SSLBD process to maintain the desired semi-orthogonality and semi-biorthogonality. Combining the implicit restart with the partial reorthogonalization proposed, we have developed an implicitly restarted SSLBD algorithm to compute several largest singular triplets of a large real skew-symmetric matrix.
A lot of numerical experiments have confirmed the effectiveness and high efficiency of the implicitly restarted SSLBD algorithm. They have indicated that it is at least competitive with eigs and often outperforms the latter considerably.
References
- [1] T. Apel, V. Mehrmann, and D. Watkins, Structured eigenvalue methods for the computation of corner singularities in 3d anisotropic elastic structures, Comput. Methods Appl. Mech. Eng., 191 (2002), pp. 4459–4473.
- [2] J. Baglama and L. Reichel, Augmented implicitly restarted Lanczos bidiagonalization methods, SIAM J. Sci. Comput., 27 (2005), pp. 19–42.
- [3] M. W. Berry, Large-scale sparse singular value decomposition, Int. J. Supercomput. Appl., 6 (1992), pp. 13–49.
- [4] J. R. Cardoso and F. S. Leite, Exponentials of skew-symmetric matrices and logarithms of orthogonal matrices, J. Comput. Appl. Math., 233 (2010), pp. 2867–2875.
- [5] J. Cullum, R. A. Willoughby, and M. Lake, A Lanczos algorithm for computing singular values and vectors of large matrices, SIAM J. Sci. Statist. Comput., 4 (1983), pp. 197–215.
- [6] T. A. Davis and Y. Hu, The University of Florida sparse matrix collection, ACM Trans. Math. Software, 38 (2011), pp. 1–25. Data available online at http://www.cise.ufl.edu/research/sparse/matrices/.
- [7] N. Del Buono, L. Lopez, and R. Peluso, Computation of the exponential of large sparse skew-symmetric matrices, SIAM J. Sci. Comput., 27 (2005), pp. 278–293.
- [8] K. V. Fernando, Accurately counting singular values of bidiagonal matrices and eigenvalues of skew-symmetric tridiagonal matrices, SIAM J. Matrix Anal. Appl., 20 (1998), pp. 373–399.
- [9] G. Golub and W. Kahan, Calculating the singular values and pseudo-inverse of a matrix, J. Soc. Indust. Appl. Math. Ser. B Numer. Anal., 2 (1965), pp. 205–224.
- [10] G. Golub and C. F. van Loan, Matrix Computations, 4th ed., The John Hopkins University Press, Baltimore, 2012.
- [11] M. E. Hochstenbach, Harmonic and refined extraction methods for the singular value problem, with applications in least squares problems, BIT Numer. Math., 44 (2004), pp. 721–754.
- [12] M. Hǎrǎguş and T. Kapitula, On the spectra of periodic waves for infinite-dimensional Hamiltonian systems, Physica D, 237 (2008), pp. 2649–2671.
- [13] Z. Jia, Regularization properties of LSQR for linear discrete ill-posed problems in the multiple singular value case and best, near best and general low rank approximations, Inverse Probl., 36 (2020), 085009 (38pp).
- [14] Z. Jia and D. Niu, An implicitly restarted refined bidiagonalization Lanczos method for computing a partial singular value decomposition, SIAM J. Matrix Anal. Appl., 25 (2003), pp. 246–265.
- [15] Z. Jia and D. Niu, A refined harmonic Lanczos bidiagonalization method and an implicitly restarted algorithm for computing the smallest singular triplets of large matrices, SIAM J. Sci. Comput., 32 (2010), pp. 714–744.
- [16] Z. Jia and G. W. Stewart, An analysis of the Rayleigh–Ritz method for approximating eigenspaces, Math. Comput., 70 (2001), pp. 637–647.
- [17] R. M. Larsen, Lanczos bidiagonalization with partial reorthogonalization, PhD thesis, University of Aarhus, 1998.
- [18] R. M. Larsen, Combining implicit restarts and partial reorthogonalization in Lanczos bidiagonalization, SCCM, Stanford University, (2001).
- [19] V. Mehrmann, The Autonomous Linear Quadratic Control Problem: Theory and Numerical Solution, vol. 163, Springer, Heidelberg, 1991.
- [20] V. Mehrmann, C. Schröder, and V. Simoncini, An implicitly-restarted Krylov subspace method for real symmetric/skew-symmetric eigenproblems, Linear Algebra Appl., 436 (2012), pp. 4070–4087.
- [21] V. Mehrmann and D. Watkins, Polynomial eigenvalue problems with Hamiltonian structure, Electron. Trans. Numer. Anal., 13 (2002), pp. 106–118.
- [22] J.-M. Mencik and D. Duhamel, A wave-based model reduction technique for the description of the dynamic behavior of periodic structures involving arbitrary-shaped substructures and large-sized finite element models, Finite Elem. Anal. Des., 101 (2015), pp. 1–14.
- [23] M. Paardekooper, An eigenvalue algorithm for skew-symmetric matrices, Numer. Math., 17 (1971), pp. 189–202.
- [24] C. C. Paige and M. A. Saunders, LSQR: an algorithm for sparse linear equations and sparse least squares, ACM Trans. Math. Soft., 8 (1982), pp. 43–71.
- [25] B. N. Parlett, The Symmetric Eigenvalue Problem, SIAM, Philadelphia, 1998.
- [26] C. Penke, A. Marek, C. Vorwerk, C. Draxl, and P. Benner, High performance solution of skew-symmetric eigenvalue problems with applications in solving the Bethe-Salpeter eigenvalue problem, Parallel Comput., 96 (2020), p. 102639.
- [27] Y. Saad, On the rates of convergence of the Lanczos and the block-Lanczos methods, SIAM J. Numer. Anal., 17 (1980), pp. 687–706.
- [28] Y. Saad, Iterative Methods for Sparse Linear Systems, SIAM, Philadelphia, 2003.
- [29] Y. Saad, Numerical Methods for Large Eigenvalue Problems: revised edition, SIAM, Philadelphia, 2011.
- [30] H. D. Simon, Analysis of the symmetric Lanczos algorithm with reorthogonalization methods, Linear Algebra Appl., 61 (1984), pp. 101–131.
- [31] D. C. Sorensen, Implicit application of polynomial filters in a k-step Arnoldi method, SIAM J. Matrix Anal. Appl., 13 (1992), pp. 357–385.
- [32] G. W. Stewart, Matrix Algorithms I: Basic Decompositions, SIAM, Philadelphia, 1998.
- [33] G. W. Stewart, Matrix Algorithms II: Eigensystems, SIAM, Philadelphia, 2001.
- [34] M. Stoll, A Krylov–Schur approach to the truncated SVD, Linear Algebra Appl., 436 (2012), pp. 2795–2806.
- [35] R. C. Ward and L. J. Gray, Eigensystem computation for skew-symmetric and a class of symmetric matrices, ACM Trans. Math. Software, 4 (1978), pp. 278–285.
- [36] M. Wimmer, Algorithm 923: Efficient numerical computation of the pfaffian for dense and banded skew-symmetric matrices, ACM Trans. Math. Software, 38 (2012), pp. 1–17.
- [37] W.-Y. Yan and J. Lam, An approximate approach to optimal model reduction, IEEE Trans. Automat. Control, 44 (1999), pp. 1341–1358.
- [38] W. X. Zhong and F. W. Williams, On the direct solution of wave propagation for repetitive structures, J. Sound Vibr., 181 (1995), pp. 485–501.
- [39] K. Zhou, J. C. Doyle, K. Glover, et al., Robust and Optimal Control, Prentice Hall, New Jersey, 1996.