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

Backward Stability of Explicit External Deflation for the Symmetric Eigenvalue Problem

Chao-Ping Lin Department of Mathematics, University of California, Davis, CA 95616, USA, cplin@ucdavis.edu    Ding Lu Department of Mathematics, University of Kentucky, Lexington, KY 40506, USA, Ding.Lu@uky.edu    Zhaojun Bai Department of Mathematics and Department of Computer Science, University of California, Davis, CA 95616, USA, zbai@ucdavis.edu
Abstract

A thorough backward stability analysis of Hotelling’s deflation, an explicit external deflation procedure through low-rank updates for computing many eigenpairs of a symmetric matrix, is presented. Computable upper bounds of the loss of the orthogonality of the computed eigenvectors and the symmetric backward error norm of the computed eigenpairs are derived. Sufficient conditions for the backward stability of the explicit external deflation procedure are revealed. Based on these theoretical results, the strategy for achieving numerical backward stability by dynamically selecting the shifts is proposed. Numerical results are presented to corroborate the theoretical analysis and to demonstrate the stability of the procedure for computing many eigenpairs of large symmetric matrices arising from applications.

1 Introduction

Let AA be an n×nn\times n real symmetric matrix with ordered eigenvalues λ1λ2λn\lambda_{1}\leq\lambda_{2}\leq\cdots\leq\lambda_{n}, and the corresponding eigenvectors v1,v2,,vnv_{1},v_{2},\ldots,v_{n}. Let =[λlow,λupper]\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}] be an interval containing the eigenvalues λ1,λ2,,λne\lambda_{1},\lambda_{2},\ldots,\lambda_{n_{e}} at the lower end of the spectrum of AA. We are interested in computing the partial eigenvalue decomposition

AVne=VneΛne,AV_{n_{e}}=V_{n_{e}}\Lambda_{n_{e}}, (1.1)

where Λne=diag(λ1,λ2,,λne)\Lambda_{n_{e}}=\mbox{diag}(\lambda_{1},\lambda_{2},\ldots,\lambda_{n_{e}}), Vne=[v1,v2,,vne]V_{n_{e}}=[v_{1},v_{2},\ldots,v_{n_{e}}] and VneTVne=IneV^{\mathrm{T}}_{n_{e}}V_{n_{e}}=I_{n_{e}}. We are particularly interested in the case where the interval is of the width ||=λupperλlow12A2|\mathcal{I}|=\lambda_{\rm upper}-\lambda_{\rm low}\leq\frac{1}{2}\|A\|_{2}, and may contain a large number of eigenvalues of AA.

A majority of subspace projection methods for computing the partial eigenvalue decomposition (1.1), such as the Lanczos methods [13, 1], typically converge first to the eigenvalues on the periphery of the spectrum. A projection subspace of a large dimension is necessary to compute the eigenvalues located deep inside the spectrum. Maintaining the orthogonality of a large projection subspace is computationally expensive. To reduce the dimensions of projection subspaces and accelerate the convergence, a number of techniques such as the restarting [17, 8, 20] and the filtering [11] have been developed.

This paper revisits a classical deflation technique known as Hotelling’s deflation [6, 19, 13]. Hotelling’s deflation computes the partial eigenvalue decomposition (1.1) through a sequence of low-rank updates. In the simplest case, supposing that the lowest eigenpair (λ1,v1)(\lambda_{1},v_{1}) of AA is computed by an eigensolver, by choosing a real shift σ1\sigma_{1} and defining the rank-one update matrix

A1A+σ1v1v1T,\displaystyle A_{1}\equiv A+\sigma_{1}v_{1}v_{1}^{\mathrm{T}},

the eigenpair (λ1,v1)(\lambda_{1},v_{1}) of AA is displaced to an eigenpair (λ1+σ1,v1)(\lambda_{1}+\sigma_{1},v_{1}) of A1A_{1}, while all the other eigenpairs of AA and A1A_{1} are the same. If the shift σ1>λupperλ1\sigma_{1}>\lambda_{\rm upper}-\lambda_{1}, then the eigenpair (λ1,v1)(\lambda_{1},v_{1}) is shifted outside the interval \mathcal{I} and the eigenpair (λ2,v2)(\lambda_{2},v_{2}) of AA becomes the lowest eigenpair of A1A_{1}. Subsequently, we can use the eigensolver again to compute the lowest eigenpair (λ2,v2)(\lambda_{2},v_{2}) of A1A_{1}. The procedure is repeated until all the eigenvalues in the interval \mathcal{I} are found. Since the low-rank update of the original AA can be done outside of the eigensolver, we refer to this strategy as an explicit external deflation (EED) procedure, to distinguish from the deflation techniques used inside an eigensolver, such as TRLan [20] and ARPACK [9]. Recently the EED has been combined with Krylov subspace methods to progressively compute the eigenpairs toward the interior of the spectrum of symmetric eigenvalue problems [12, 21], and extended to structured eigenvalue problems [2, 3].

In the presence of finite precision arithmetic, the EED procedure is susceptible to numerical instability due to the accumulation of numerical errors from previous computations; see, e.g., [19, p. 585] and [15, p. 96]. In [13, Chap. 5.1], Parlett showed that the change to the smallest eigenvalue in magnitude by deflating out the largest computed eigenvalue would be at the same order of the round-off error incurred. In [14], for nonsymmetric matrices, Saad derived an upper bound on the backward error norm and showed that the stability is determined by the angle between the computed eigenvectors and the deflated subspace. Saad concluded that “if the angle between the computed eigenvector and the previous invariant subspace is small at every step, the process may quickly become unstable. On the other hand, if this is not the case then the process is quite safe, for small number of requested eigenvalues.”

However, none of the existing studies had examined the effect of the choice of shifts on the numerical stability. In this paper, we present a thorough backward stability analysis of the EED procedure. We first derive governing equations of the EED procedure in the presence of finite precision arithmetic, and then derive upper bounds of the loss of the orthogonality of the computed eigenvectors and the symmetric backward error norm of the computed eigenpairs. From these upper bounds, under mild assumptions, we conclude that the EED procedure is backward stable if the shifts σj\sigma_{j} are dynamically chosen such that (i) the spectral gaps, defined as the separation between the computed eigenvalues and shifted ones, are maintained at the same order of the norm of the matrix AA, and (ii) the shift-gap ratios, defined as the ratios between the largest shift in magnitude and the spectral gaps, are at the order of one. Our analysis implies that the re-orthogonalization of computed eigenvectors and the large angles between the computed eigenvectors and the previous invariant subspaces are unnecessary. Based on theoretical analysis, we provide a practical shift selection scheme to guarantee the backward stability. For a set of test matrices from applications, we demonstrate that the EED in combination with an established eigensolver can compute a large number of eigenvalues with backward stability.

The rest of the paper is organized as follows. In Section 2, we present the governing equations of the EED in finite precision arithmetic. In Section 3, we derive the upper bounds on the loss of orthogonality of computed eigenvectors and the symmetric backward error norm of computed eigenpairs. From these upper bounds, we derive conditions for the backward stability of the EED procedure. In Section 4, we discuss how to properly choose the shifts to satisfy the stability conditions and outline an algorithm for combining the EED procedure with an eigensolver. Numerical results to verify the sharpness of the upper bounds and to demonstrate the backward stability of the EED for symmetric eigenvalue problems from applications are presented in Section 5. Concluding remarks are given in section 6.

Following the convention of matrix computations, we use the upper case letters for matrices and the lower case letters for vectors. In particular, we use InI_{n} for the identity matrix of dimension nn. If not specified, the dimensions of matrices and vectors conform to the dimensions used in the context. We use T\cdot^{\mathrm{T}} for transpose, and 2\|\cdot\|_{2} and F\|\cdot\|_{\rm F} for 22-norm and the Frobenius norm, respectively. The range of a matrix AA is denoted by (A)\mathcal{R}(A). We also use σmin()\sigma_{\min}(\cdot) to denote the minimal singular values of a matrix. Other notations will be explained as used.

2 EED in finite precision arithmetic

Suppose that we use an established eigensolver called EIGSOL, such as TRLan [20] and ARPACK [9], to successfully compute the lowest eigenpair (λ^,v^)(\widehat{\lambda},\widehat{v}) of AA with

Av^=λ^v^+η,A\widehat{v}=\widehat{\lambda}\widehat{v}+\eta,

where v^2=1\|\widehat{v}\|_{2}=1 and the residual vector η\eta satisfies

η2tolA2\|\eta\|_{2}\leq tol\cdot\|A\|_{2} (2.1)

for a prescribed convergence tolerance toltol. The EED procedure starts with computing the lowest eigenpair (λ^1,v^1)(\widehat{\lambda}_{1},\widehat{v}_{1}) of AA by EIGSOL satisfying

Av^1=λ^1v^1+η1,\displaystyle A\widehat{v}_{1}=\widehat{\lambda}_{1}\widehat{v}_{1}+\eta_{1},

where λ^1=[λlow,λupper]\widehat{\lambda}_{1}\in\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}], and the residual vector η1\eta_{1} satisfies (2.1). At the first EED step, we choose a shift σ1\sigma_{1} and define

A^1A+σ1v^1v^1T.\displaystyle\widehat{A}_{1}\equiv A+\sigma_{1}\widehat{v}_{1}\widehat{v}_{1}^{\mathrm{T}}.

By choosing the shift σ1>λupperλ^1\sigma_{1}>\lambda_{\rm upper}-\widehat{\lambda}_{1}, the lowest eigenpair of A^1\widehat{A}_{1} is an approximation of the second eigenpair (λ2,v2)({\lambda}_{2},{v}_{2}) of AA. Subsequently, we use EIGSOL to compute the lowest eigenpair (λ^2,v^2)(\widehat{\lambda}_{2},\widehat{v}_{2}) of A^1\widehat{A}_{1} satisfying

A^1v^2=λ^2v^2+η2,\displaystyle\widehat{A}_{1}\widehat{v}_{2}=\widehat{\lambda}_{2}\widehat{v}_{2}+\eta_{2},

where the residual vector η2\eta_{2} satisfies (2.1). Meanwhile, expressing the computed eigenpair (λ^1,v^1)(\widehat{\lambda}_{1},\widehat{v}_{1}) in terms of A^1\widehat{A}_{1}, we have

A^1v^1=(λ^1+σ1)v^1+η1.\widehat{A}_{1}\widehat{v}_{1}=(\widehat{\lambda}_{1}+\sigma_{1})\widehat{v}_{1}+\eta_{1}.

Proceeding to the second EED step, we choose a shift σ2\sigma_{2} and define

A^2A^1+σ2v^2v^2T=A+V^2Σ2V^2T,\widehat{A}_{2}\equiv\widehat{A}_{1}+\sigma_{2}\widehat{v}_{2}\widehat{v}_{2}^{\mathrm{T}}=A+\widehat{V}_{2}\Sigma_{2}\widehat{V}_{2}^{\mathrm{T}},

where V^2=[v^1,v^2]\widehat{V}_{2}=[\widehat{v}_{1},\widehat{v}_{2}] and Σ2=diag(σ1,σ2)\Sigma_{2}=\operatorname{diag}(\sigma_{1},\sigma_{2}). By choosing the shift σ2>λupperλ^2\sigma_{2}>\lambda_{\rm upper}-\widehat{\lambda}_{2}, the lowest eigenpair of A^2\widehat{A}_{2} is an approximation of the third eigenpair (λ3,v3)({\lambda}_{3},{v}_{3}) of AA. Then we use EIGSOL again to compute the lowest eigenpair (λ^3,v^3)(\widehat{\lambda}_{3},\widehat{v}_{3}) of A^2\widehat{A}_{2} satisfying

A^2v^3=λ^3v^3+η3,\displaystyle\widehat{A}_{2}\widehat{v}_{3}=\widehat{\lambda}_{3}\widehat{v}_{3}+\eta_{3},

where the residual vector η3\eta_{3} satisfies (2.1). Meanwhile, expressing the computed eigenpairs (λ^1,v^1)(\widehat{\lambda}_{1},\widehat{v}_{1}) and (λ^2,v^2)(\widehat{\lambda}_{2},\widehat{v}_{2}) in terms of A^2\widehat{A}_{2}, we have

A^2V^2=V^2(Λ^2+Σ2)+V^2Σ2Φ2+E2,\displaystyle\widehat{A}_{2}\widehat{V}_{2}=\widehat{V}_{2}(\widehat{\Lambda}_{2}+\Sigma_{2})+\widehat{V}_{2}\Sigma_{2}\Phi_{2}+E_{2},

where Λ^2=diag(λ^1,λ^2)\widehat{\Lambda}_{2}=\operatorname{diag}(\widehat{\lambda}_{1},\widehat{\lambda}_{2}), E2=[η1,η2]E_{2}=[\eta_{1},\eta_{2}], and Φ22×2\Phi_{2}\in{\mathbb{R}}^{2\times 2} is the strictly lower triangular part of the matrix V^2TV^2I2\widehat{V}_{2}^{\mathrm{T}}\widehat{V}_{2}-I_{2}, i.e., Φ2+Φ2T=V^2TV^2I2\Phi_{2}+\Phi^{\mathrm{T}}_{2}=\widehat{V}_{2}^{\mathrm{T}}\widehat{V}_{2}-I_{2}.

In general, at the jj-th EED step, we choose a shift σj\sigma_{j} and define

A^jA^j1+σjv^jv^jT=A+V^jΣjV^jT,\displaystyle\widehat{A}_{j}\equiv\widehat{A}_{j-1}+\sigma_{j}\widehat{v}_{j}\widehat{v}_{j}^{\mathrm{T}}=A+\widehat{V}_{j}\Sigma_{j}\widehat{V}_{j}^{\mathrm{T}}, (2.2)

where V^j[v^1,,v^j]\widehat{V}_{j}\equiv[\widehat{v}_{1},\ldots,\widehat{v}_{j}] and Σj=diag(σ1,,σj)\Sigma_{j}=\operatorname{diag}(\sigma_{1},\ldots,\sigma_{j}) with A^0A\widehat{A}_{0}\equiv A. Then by choosing the shift σj>λupperλ^j\sigma_{j}>\lambda_{\rm upper}-\widehat{\lambda}_{j}, the lowest eigenpair of A^j\widehat{A}_{j} is an approximation of the (j+1)(j+1)-th eigenpair (λj+1,vj+1)({\lambda}_{j+1},{v}_{j+1}) of AA. We use EIGSOL to compute the lowest eigenpair (λ^j+1,v^j+1)(\widehat{\lambda}_{j+1},\widehat{v}_{j+1}) of A^j\widehat{A}_{j} satisfying

A^jv^j+1=λ^j+1v^j+1+ηj+1,\displaystyle\widehat{A}_{j}\widehat{v}_{j+1}=\widehat{\lambda}_{j+1}\widehat{v}_{j+1}+\eta_{j+1}, (2.3)

where v^j+12=1\|\widehat{v}_{j+1}\|_{2}=1 and the residual vector ηj+1\eta_{j+1} satisfies (2.1), i.e.,

ηj+12tolA2.\displaystyle\|\eta_{j+1}\|_{2}\leq tol\cdot\|A\|_{2}. (2.4)

Meanwhile, for the computed eigenpairs (λ^j,v^j)(\widehat{\lambda}_{j},\widehat{v}_{j}) in terms of A^j\widehat{A}_{j}, we have

A^jv^j=A^j1v^j+σjv^j=(λj+σj)v^j+ηj,\displaystyle\widehat{A}_{j}\widehat{v}_{j}=\widehat{A}_{j-1}\widehat{v}_{j}+\sigma_{j}\widehat{v}_{j}=(\lambda_{j}+\sigma_{j})\widehat{v}_{j}+\eta_{j}, (2.5)

and for the computed eigenpairs (λ^i,v^i)(\widehat{\lambda}_{i},\widehat{v}_{i}) with 1ij11\leq i\leq j-1 in terms of A^j\widehat{A}_{j}, we have

A^jv^i\displaystyle\widehat{A}_{j}\widehat{v}_{i} =(A^j1+σjv^jv^jT)v^i\displaystyle=\left(\widehat{A}_{j-1}+\sigma_{j}\widehat{v}_{j}\widehat{v}_{j}^{\mathrm{T}}\right)\widehat{v}_{i}
=(A^i1+σiv^iv^iT+V^i+1:jΣi+1:jV^i+1:jT)v^i\displaystyle=\left(\widehat{A}_{i-1}+\sigma_{i}\widehat{v}_{i}\widehat{v}_{i}^{\mathrm{T}}+\widehat{V}_{i+1:j}\Sigma_{i+1:j}\widehat{V}_{i+1:j}^{\mathrm{T}}\right)\widehat{v}_{i}
=(A^i1+σiv^iv^iT)v^i+V^i+1:jΣi+1:jV^i+1:jTv^i\displaystyle=\left(\widehat{A}_{i-1}+\sigma_{i}\widehat{v}_{i}\widehat{v}_{i}^{\mathrm{T}}\right)\widehat{v}_{i}+\widehat{V}_{i+1:j}\Sigma_{i+1:j}\widehat{V}_{i+1:j}^{\mathrm{T}}\widehat{v}_{i}
=λ^iv^i+ηi+σiv^i+V^i+1:jΣi+1:jV^i+1:jTv^i\displaystyle=\widehat{\lambda}_{i}\widehat{v}_{i}+\eta_{i}+\sigma_{i}\widehat{v}_{i}+\widehat{V}_{i+1:j}\Sigma_{i+1:j}\widehat{V}_{i+1:j}^{\mathrm{T}}\widehat{v}_{i}
=(λ^i+σi)v^i+V^i+1:jΣi+1:jV^i+1:jTv^i+ηi\displaystyle=(\widehat{\lambda}_{i}+\sigma_{i})\widehat{v}_{i}+\widehat{V}_{i+1:j}\Sigma_{i+1:j}\widehat{V}_{i+1:j}^{\mathrm{T}}\widehat{v}_{i}+\eta_{i}
=(λ^i+σi)v^i+V^jΣj[0V^i+1:jTv^i]+ηi,\displaystyle=(\widehat{\lambda}_{i}+\sigma_{i})\widehat{v}_{i}+\widehat{V}_{j}\Sigma_{j}\left[\begin{array}[]{c}0\\ \widehat{V}_{i+1:j}^{\mathrm{T}}\widehat{v}_{i}\end{array}\right]+\eta_{i}, (2.8)

where V^i+1:j[v^i+1,,v^j]\widehat{V}_{i+1:j}\equiv[\widehat{v}_{i+1},\ldots,\widehat{v}_{j}] and Σi+1:jdiag(σi+1,,σj)\Sigma_{i+1:j}\equiv\operatorname{diag}(\sigma_{i+1},\ldots,\sigma_{j}).

Combining (2.5) and (2.8), we have

A^jV^j=V^j(Λ^j+Σj)+V^jΣjΦj+Ej,\displaystyle\widehat{A}_{j}\widehat{V}_{j}=\widehat{V}_{j}(\widehat{\Lambda}_{j}+\Sigma_{j})+\widehat{V}_{j}\Sigma_{j}\Phi_{j}+E_{j}, (2.9)

where Λ^j=diag(λ^1,,λ^j)\widehat{\Lambda}_{j}=\operatorname{diag}(\widehat{\lambda}_{1},\ldots,\widehat{\lambda}_{j}), Ej=[η1,,ηj]E_{j}=[\eta_{1},\ldots,\eta_{j}], Φj\Phi_{j} is the strictly lower triangular part of the matrix V^jTV^jIj\widehat{V}_{j}^{\mathrm{T}}\widehat{V}_{j}-I_{j} and Φj+ΦjT=V^jTV^jIj\Phi_{j}+\Phi^{\mathrm{T}}_{j}=\widehat{V}_{j}^{\mathrm{T}}\widehat{V}_{j}-I_{j}. Eqs. (2.3) and (2.9) are referred to as the governing equations of jj steps of (inexact) EED procedure.

For the backward stability analysis of the EED procedure, we introduce the following two quantities associated with the shifts σ1,,σj\sigma_{1},\dots,\sigma_{j} for a jj-step EED:

  • the spectral gap of A^j\widehat{A}_{j}, defined as the separation between the computed eigenvalues and the shifted ones:

    γjminλj+1,θ𝒥j|λθ|>0,\gamma_{j}\equiv\min_{\lambda\in\mathcal{I}_{j+1},\theta\in\mathcal{J}_{j}}|\lambda-\theta|>0, (2.10)

    where j+1{λ^1,,λ^j,λ^j+1}\mathcal{I}_{j+1}\equiv\{\widehat{\lambda}_{1},\ldots,\widehat{\lambda}_{j},\widehat{\lambda}_{j+1}\}, the set of computed eigenvalues, and 𝒥j{λ^1+σ1,,λ^j+σj}\mathcal{J}_{j}\equiv\{\widehat{\lambda}_{1}+\sigma_{1},\ldots,\widehat{\lambda}_{j}+\sigma_{j}\}, the set of computed eigenvalues after shifting (see Figure 1 for an illustration);

    λ^1\widehat{\lambda}_{1}λ^i\widehat{\lambda}_{i}λ^j+1\widehat{\lambda}_{j+1}λ^1+σ1\widehat{\lambda}_{1}+\sigma_{1}λ^i+σi\widehat{\lambda}_{i}+\sigma_{i}λ^j+σj\widehat{\lambda}_{j}+\sigma_{j}γj\gamma_{j}
    Figure 1: Illustration of the spectral gap γj\gamma_{j}
  • the shift-gap ratio, defined as the ratio of the largest shift and the spectral gap γj\gamma_{j}:

    τj1γjmax1ij|σi|.\tau_{j}\equiv\frac{1}{\gamma_{j}}\cdot\max_{1\leq i\leq j}|\sigma_{i}|. (2.11)

We will see that γj\gamma_{j} and τj\tau_{j} are crucial quantities to characterize the backward stability of the EED procedure.

3 Backward stability analysis of EED

In this section, we first review the notion of backward stability for the computed eigenpairs. Then we derive upper bounds on the loss of orthogonality of computed eigenvectors and the backward error norm of computed eigenpairs by the EED procedure, and reveal conditions for the backward stability of the procedure.

3.1 Notion of backward stability

By the well-established notion of backward stability analysis in numerical linear algebra [5, 18], the following two quantities measure the accuracy of the approximate eigenpairs (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}) of AA computed by the EED procedure:

  • the loss of orthogonality of the computed eigenvectors V^j+1\widehat{V}_{j+1},

    ωj+1V^j+1TV^j+1Ij+1F,\displaystyle\omega_{j+1}\equiv\|\widehat{V}_{j+1}^{\mathrm{T}}\widehat{V}_{j+1}-I_{j+1}\|_{\rm F}, (3.1)
  • the symmetric backward error norm of the computed eigenpairs (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}),

    δj+1minΔQj+1ΔF,\displaystyle\delta_{j+1}\equiv\min_{\Delta\in\mathcal{H}_{Q_{j+1}}}\|\Delta\|_{\rm F}, (3.2)

    where Qj+1\mathcal{H}_{Q_{j+1}} is the set of the symmetric backward errors for the orthonormal basis Qj+1Q_{j+1} from the polar decomposition111The polar decomposition of a matrix XX is given by X=UPX=UP where UU is orthonormal and PP is symmetric positive semi-definite. of V^j+1\widehat{V}_{j+1}, namely,

    Qj+1{Δ(A+Δ)Qj+1=Qj+1Λ^j+1,Δ=ΔTn×n}.\displaystyle\mathcal{H}_{Q_{j+1}}\equiv\left\{\Delta\mid(A+\Delta)Q_{j+1}=Q_{j+1}\widehat{\Lambda}_{j+1},\ \Delta=\Delta^{\mathrm{T}}\in{\mathbb{R}}^{n\times n}\right\}. (3.3)

For a prescribed tolerance toltol of the stopping criterion (2.4) for an eigensolver EIGSOL, the EED procedure is considered to be backward stable if

ωj+1=O(tol)\displaystyle\omega_{j+1}=O(tol) (3.4)

and

δj+1=O(tolA2),\displaystyle\delta_{j+1}=O(tol\cdot\|A\|_{2}), (3.5)

where the constants in the big-O notations are low-degree polynomials in the number jj of the EED steps.

3.2 Loss of orthogonality

We first prove the following lemma to reveal the structure of the orthogonality between the computed eigenvectors.

Lemma 3.1.

By the governing equations (2.3) and (2.9) of jj steps of EED, if τjωj<2\tau_{j}\omega_{j}<\sqrt{2}, then for i=1,2,,ji=1,2,\dots,j, the matrices ΓiΛ^i+Σiλ^i+1Ii\Gamma_{i}\equiv\widehat{\Lambda}_{i}+\Sigma_{i}-\widehat{\lambda}_{i+1}I_{i} and Ii+ΦiTΣiΓi1I_{i}+\Phi_{i}^{\mathrm{T}}\Sigma_{i}\Gamma_{i}^{-1} are non-singular, and

V^iTv^i+1=Γi1(Ii+ΦiTΣiΓi1)1[V^iTηi+1EiTv^i+1].\displaystyle\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1}=\Gamma_{i}^{-1}\left(I_{i}+\Phi_{i}^{\mathrm{T}}\Sigma_{i}\Gamma_{i}^{-1}\right)^{-1}\left[\widehat{V}_{i}^{\mathrm{T}}\eta_{i+1}-E_{i}^{\mathrm{T}}\widehat{v}_{i+1}\right]. (3.6)

Furthermore,

  1. (i)

    Γi12γj1\|\Gamma_{i}^{-1}\|_{2}\leq\gamma_{j}^{-1},

  2. (ii)

    (Ii+ΦiTΣiΓi1)12(1τjωj/2)1\|(I_{i}+\Phi_{i}^{\mathrm{T}}\Sigma_{i}\Gamma_{i}^{-1})^{-1}\|_{2}\leq(1-\tau_{j}\omega_{j}/\sqrt{2})^{-1},

where γj\gamma_{j} and τj\tau_{j} are the spectral gap and the shift-gap ratio defined in (2.10) and (2.11), respectively, and ωj\omega_{j} is the loss of the orthogonality defined in (3.1).

Proof.

By the governing equations (2.3) and (2.9) of jj steps of EED, for 1ij1\leq i\leq j, we have

V^iTA^iv^i+1=λ^i+1V^iTv^i+1+V^iTηi+1\displaystyle\widehat{V}_{i}^{\mathrm{T}}\widehat{A}_{i}\widehat{v}_{i+1}=\widehat{\lambda}_{i+1}\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1}+\widehat{V}_{i}^{\mathrm{T}}\eta_{i+1}

and

V^iTA^iv^i+1=(Λ^i+Σi)V^iTv^i+1+ΦiTΣiV^iTv^i+1+EiTv^i+1.\displaystyle\widehat{V}_{i}^{\mathrm{T}}\widehat{A}_{i}\widehat{v}_{i+1}=(\widehat{\Lambda}_{i}+\Sigma_{i})\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1}+\Phi_{i}^{\mathrm{T}}\Sigma_{i}\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1}+E_{i}^{\mathrm{T}}\widehat{v}_{i+1}.

Consequently,

(Γi+ΦiTΣi)V^iTv^i+1=V^iTηi+1EiTv^i+1,\displaystyle(\Gamma_{i}+\Phi_{i}^{\mathrm{T}}\Sigma_{i})\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1}=\widehat{V}_{i}^{\mathrm{T}}\eta_{i+1}-E_{i}^{\mathrm{T}}\widehat{v}_{i+1}, (3.7)

where Γi=Λ^i+Σiλ^i+1Ii\Gamma_{i}=\widehat{\Lambda}_{i}+\Sigma_{i}-\widehat{\lambda}_{i+1}I_{i} is a diagonal matrix.

By the definition (2.10) of the spectral gap γj\gamma_{j},

σmin(Γi)=min1ki|λ^k+σkλ^i+1|γj>0.\displaystyle\sigma_{\min}(\Gamma_{i})=\min_{1\leq k\leq i}|\widehat{\lambda}_{k}+\sigma_{k}-\widehat{\lambda}_{i+1}|\geq\gamma_{j}>0. (3.8)

Hence the matrix Γi\Gamma_{i} is non-singular and the bound (i) holds.

Since Φi\Phi_{i} is the strictly lower triangular part of the matrix V^iTV^iIi\widehat{V}_{i}^{\mathrm{T}}\widehat{V}_{i}-I_{i},

ΦiT2ΦiTF=ωi2ωj2.\displaystyle\|\Phi_{i}^{\mathrm{T}}\|_{2}\leq\|\Phi_{i}^{\mathrm{T}}\|_{\rm F}=\frac{\omega_{i}}{\sqrt{2}}\leq\frac{\omega_{j}}{\sqrt{2}}.

Consequently,

ΦiTΣiΓi12ΦiT2Γi12Σi2ωj2γj1Σj2=ωj2τj<1,\displaystyle\|\Phi_{i}^{\mathrm{T}}\Sigma_{i}\Gamma_{i}^{-1}\|_{2}\leq\|\Phi_{i}^{\mathrm{T}}\|_{2}\|\Gamma_{i}^{-1}\|_{2}\|\Sigma_{i}\|_{2}\leq\frac{\omega_{j}}{\sqrt{2}}\cdot\gamma_{j}^{-1}\cdot{\|\Sigma_{j}\|_{2}}=\frac{\omega_{j}}{\sqrt{2}}\cdot\tau_{j}<1, (3.9)

where for the last inequality, we use the assumption τjωj<2\tau_{j}\omega_{j}<\sqrt{2}. By (3.9), the matrix Ii+ΦiTΣiΓi1I_{i}+\Phi_{i}^{\mathrm{T}}\Sigma_{i}\Gamma_{i}^{-1} is non-singular and the bound (ii) holds due to (I+X)12(1X2)1\|(I+X)^{-1}\|_{2}\leq(1-\|X\|_{2})^{-1} for any XX with X2<1\|X\|_{2}<1.

Since both matrices Γi\Gamma_{i} and Ii+ΦiTΣiΓi1I_{i}+\Phi_{i}^{\mathrm{T}}\Sigma_{i}\Gamma_{i}^{-1} are invertible, the identity (3.6) follows from (3.7). ∎

Next we exploit the structure of the product V^iTv^i+1\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1} to derive a computable upper bound on the loss of orthogonality ωj+1\omega_{j+1} of computed eigenvectors V^j+1\widehat{V}_{j+1}.

Theorem 3.1.

By the governing equations (2.3) and (2.9) of jj steps of EED, if τjωj<2\tau_{j}\omega_{j}<\sqrt{2}, then the loss of orthogonality ωj+1\omega_{j+1} of the computed eigenvectors V^j+1\widehat{V}_{j+1} defined in (3.1) satisfies

ωj+12cjγj(1+2cjγjEj+1F)Ej+1F,\displaystyle\omega_{j+1}\leq 2\,\frac{c_{j}}{\gamma_{j}}\left(1+2\frac{c_{j}}{\gamma_{j}}\|E_{j+1}\|_{\rm F}\right)\|E_{j+1}\|_{\rm F}, (3.10)

where cj=(1τjωj/2)1c_{j}=(1-\tau_{j}\omega_{j}/\sqrt{2})^{-1}, and γj\gamma_{j} and τj\tau_{j} are the spectral gap and the shift-gap ratio defined in (2.10) and (2.11), respectively.

Proof.

By the definition (3.1), we have

ωj+12=2Φj+1TF2=2i=1jV^iTv^i+122.\displaystyle\omega_{j+1}^{2}=2\cdot\left\|\Phi^{\mathrm{T}}_{j+1}\right\|^{2}_{\rm F}=2\cdot\sum_{i=1}^{j}\|\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1}\|_{2}^{2}. (3.11)

Recall Lemma 3.1 that, for any 1ij1\leq i\leq j,

V^iTv^i+12cjγjV^iTηi+1EiTv^i+12.\displaystyle\|\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1}\|_{2}\leq\frac{c_{j}}{\gamma_{j}}\cdot\|\widehat{V}_{i}^{\mathrm{T}}\eta_{i+1}-E_{i}^{\mathrm{T}}\widehat{v}_{i+1}\|_{2}.

Hence we can derive from (3.11) that

ωj+12\displaystyle\omega_{j+1}^{2} 2cj2γj2i=1jV^iTηi+1EiTv^i+122\displaystyle\leq\frac{2c^{2}_{j}}{\gamma_{j}^{2}}\cdot\sum_{i=1}^{j}\|\widehat{V}_{i}^{\mathrm{T}}\eta_{i+1}-E_{i}^{\mathrm{T}}\widehat{v}_{i+1}\|_{2}^{2}
=2cj2γj212V^j+1TEj+1Ej+1TV^j+1F2\displaystyle=\frac{2c^{2}_{j}}{\gamma_{j}^{2}}\cdot\frac{1}{2}\|\widehat{V}_{j+1}^{\mathrm{T}}E_{j+1}-E_{j+1}^{\mathrm{T}}\widehat{V}_{j+1}\|_{\rm F}^{2}
2cj2γj22V^j+1TEj+1F24cj2γj2V^j+1T22Ej+1F2.\displaystyle\leq\frac{2c^{2}_{j}}{\gamma_{j}^{2}}\cdot 2\|\widehat{V}_{j+1}^{\mathrm{T}}E_{j+1}\|^{2}_{{\rm F}}\ \leq\ \frac{4c^{2}_{j}}{\gamma_{j}^{2}}\cdot\|\widehat{V}_{j+1}^{\mathrm{T}}\|^{2}_{2}\|E_{j+1}\|^{2}_{{\rm F}}. (3.12)

Since

V^j+1T22=V^j+1TV^j+12Ij+12+V^j+1TV^j+1Ij+121+ωj+1,\displaystyle\|\widehat{V}_{j+1}^{\mathrm{T}}\|^{2}_{2}=\|\widehat{V}_{j+1}^{\mathrm{T}}\widehat{V}_{j+1}\|_{2}\leq\|I_{j+1}\|_{2}+\|\widehat{V}_{j+1}^{\mathrm{T}}\widehat{V}_{j+1}-I_{j+1}\|_{2}\leq 1+\omega_{j+1},

we arrive at

ωj+124cj2γj2(1+ωj+1)Ej+1F2.\displaystyle\omega_{j+1}^{2}\leq\frac{4c_{j}^{2}}{\gamma_{j}^{2}}\cdot(1+\omega_{j+1})\cdot\|E_{j+1}\|_{\rm F}^{2}. (3.13)

Letting t=ωj+1/χj+1t=\omega_{j+1}/{\chi}_{j+1}, where χj+1=2cjEj+1F/γj{\chi}_{j+1}=2c_{j}\|E_{j+1}\|_{\rm F}/\gamma_{j}, then the inequality (3.13) is recast as

t2χj+1t10.\displaystyle t^{2}-{\chi}_{j+1}t-1\leq 0. (3.14)

By the fact that the quadratic polynomial in (3.14) is concave, we conclude that

t12(χj+1+4+χj+12)12(χj+1+2+χj+1)1+χj+1.\displaystyle t\leq\frac{1}{2}\cdot\left({\chi}_{j+1}+\sqrt{4+{\chi}_{j+1}^{2}}\right)\leq\frac{1}{2}\cdot\left({\chi}_{j+1}+2+{\chi}_{j+1}\right)\leq 1+{\chi}_{j+1}.

This proves the upper bound in (3.10). ∎

Remark 3.1.

A straightforward way to derive an upper bound of ωj+1\omega_{j+1} is to use the relation

ωj+12=ωj2+2V^jTv^j+122,\omega_{j+1}^{2}=\omega_{j}^{2}+2\cdot\|\widehat{V}_{j}^{\mathrm{T}}\widehat{v}_{j+1}\|_{2}^{2},

and the following bound from Lemma 3.1

V^jTv^j+122cj2γj2(V^j2ηj+12+EjF)2cj2γj2(Ej+1F2+2ηj+12EjF).\displaystyle\|\widehat{V}_{j}^{\mathrm{T}}\widehat{v}_{j+1}\|_{2}^{2}\leq\frac{c_{j}^{2}}{\gamma_{j}^{2}}\cdot\left(\|\widehat{V}_{j}\|_{2}\|\eta_{j+1}\|_{2}+\|E_{j}\|_{\rm F}\right)^{2}\approx\frac{c_{j}^{2}}{\gamma_{j}^{2}}(\|E_{j+1}\|_{\rm F}^{2}+2\|\eta_{j+1}\|_{2}\|E_{j}\|_{\rm F}).

However, this would lead to a pessimistic upper bound of ωj+1\omega_{j+1}. In contrast, in Theorem 3.1, we exploit the structure of the orthogonality V^iTv^i+1\widehat{V}_{i}^{\mathrm{T}}\widehat{v}_{i+1} to obtain a tighter bound (3.10) of ωj+1\omega_{j+1}. Eq. (3.12) in the proof is the key step. In Example 5.1 in Section 5, we will demonstrate numerically that the bound (3.10) is indeed a tight upper bound on the loss of orthogonality ωj+1\omega_{j+1}.

3.3 Symmetric backward error norm

In this section, we derive a computable upper bound on the symmetric backward error norm δj+1\delta_{j+1} of computed eigenpairs (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}) of AA defined in (3.2). First, the following lemma gives an upper bound of the norm of the residual for (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}):

Rj+1AV^j+1V^j+1Λ^j+1.\displaystyle R_{j+1}\equiv A\widehat{V}_{j+1}-\widehat{V}_{j+1}\widehat{\Lambda}_{j+1}. (3.15)
Lemma 3.2.

By the governing equations (2.3) and (2.9) of jj steps of EED, if τjωj<2\tau_{j}\omega_{j}<\sqrt{2}, then for the computed eigenpairs (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}) of AA, the Frobenius norm of the residual Rj+1R_{j+1} defined in (3.15) satisfies

Rj+1F(1+2cjτj(1+ωj+1))Ej+1F,\displaystyle\|R_{j+1}\|_{\rm F}\leq\left(1+\sqrt{2}c_{j}\tau_{j}(1+\omega_{j+1})\right)\|E_{j+1}\|_{\rm F}, (3.16)

where cj=(1τjωj/2)1c_{j}=(1-\tau_{j}\omega_{j}/\sqrt{2})^{-1}, and γj\gamma_{j} and τj\tau_{j} are the spectral gap and the shift-gap ratio defined in (2.10) and (2.11), respectively.

Proof.

From the governing equation (2.9) of the EED procedure after j+1j+1 steps, we have

A^j+1V^j+1=V^j+1(Λ^j+1+Σj+1)+V^j+1Σj+1Φj+1+Ej+1.\displaystyle\widehat{A}_{j+1}\widehat{V}_{j+1}=\widehat{V}_{j+1}(\widehat{\Lambda}_{j+1}+\Sigma_{j+1})+\widehat{V}_{j+1}\Sigma_{j+1}\Phi_{j+1}+E_{j+1}. (3.17)

On the other hand, by the definition (2.2) of A^j+1\widehat{A}_{j+1}, we have

A^j+1V^j+1\displaystyle\widehat{A}_{j+1}\widehat{V}_{j+1} =AV^j+1+V^j+1Σj+1V^j+1TV^j+1\displaystyle=A\widehat{V}_{j+1}+\widehat{V}_{j+1}\Sigma_{j+1}\widehat{V}_{j+1}^{\mathrm{T}}\widehat{V}_{j+1}
=AV^j+1+V^j+1Σj+1(Φj+1+Ij+1+Φj+1T).\displaystyle=A\widehat{V}_{j+1}+\widehat{V}_{j+1}\Sigma_{j+1}(\Phi_{j+1}+I_{j+1}+\Phi_{j+1}^{\mathrm{T}}). (3.18)

Combining (3.17) and (3.18), we obtain the residual

Rj+1=AV^j+1V^j+1Λ^j+1=Ej+1V^j+1Σj+1Φj+1T.\displaystyle R_{j+1}=A\widehat{V}_{j+1}-\widehat{V}_{j+1}\widehat{\Lambda}_{j+1}=E_{j+1}-\widehat{V}_{j+1}\Sigma_{j+1}\Phi_{j+1}^{\mathrm{T}}. (3.19)

Consequently, the norm of the residual Rj+1R_{j+1} is bounded by

Rj+1FEj+1F+V^j+12Σj+1Φj+1TF.\displaystyle\|R_{j+1}\|_{\rm F}\leq\|E_{j+1}\|_{\rm F}+\|\widehat{V}_{j+1}\|_{2}\|\Sigma_{j+1}\Phi_{j+1}^{\mathrm{T}}\|_{\rm F}. (3.20)

Note that Σj+1=diag(σ1,,σj+1)\Sigma_{j+1}=\operatorname{diag}(\sigma_{1},\ldots,\sigma_{j+1}) and Φj+1T\Phi_{j+1}^{\mathrm{T}} is the strictly upper triangular part of the matrix V^j+1TV^j+1Ij+1\widehat{V}_{j+1}^{\mathrm{T}}\widehat{V}_{j+1}-I_{j+1}, and we have

V^j+12Σj+1Φj+1TF\displaystyle\|\widehat{V}_{j+1}\|_{2}\|\Sigma_{j+1}\Phi_{j+1}^{\mathrm{T}}\|_{\rm F} V^j+12Σj2Φj+1TF\displaystyle\leq\|\widehat{V}_{j+1}\|_{2}\|\Sigma_{j}\|_{2}\|\Phi_{j+1}^{\mathrm{T}}\|_{\rm F}
V^j+12Σj212ωj+1\displaystyle\leq\|\widehat{V}_{j+1}\|_{2}\|\Sigma_{j}\|_{2}\cdot\frac{1}{\sqrt{2}}\omega_{j+1}
12Σj2(1+ωj+1)ωj+12,\displaystyle\leq\frac{1}{\sqrt{2}}\|\Sigma_{j}\|_{2}\sqrt{(1+\omega_{j+1})\omega_{j+1}^{2}}, (3.21)

where, for the third inequality, we again use the fact that V^j+121+ωj+1\|\widehat{V}_{j+1}\|_{2}\leq\sqrt{1+\omega_{j+1}} by the definition of the loss of orthogonality ωj+1\omega_{j+1}. Left-multiplying (3.13) by 1+ωj+11+\omega_{j+1}, we know that

(1+ωj+1)ωj+124cj2γj2(1+ωj+1)2Ej+1F2.\displaystyle(1+\omega_{j+1})\omega_{j+1}^{2}\leq\frac{4c_{j}^{2}}{\gamma_{j}^{2}}\cdot(1+\omega_{j+1})^{2}\cdot\|E_{j+1}\|_{\rm F}^{2}.

Plugging into (3.21), we obtain

V^j+12Σj+1Φj+1TF\displaystyle\|\widehat{V}_{j+1}\|_{2}\|\Sigma_{j+1}\Phi_{j+1}^{\mathrm{T}}\|_{\rm F} 2Σj2cjγj1(1+ωj+1)Ej+1F\displaystyle\leq\sqrt{2}\|\Sigma_{j}\|_{2}c_{j}\gamma_{j}^{-1}\cdot(1+\omega_{j+1})\cdot\|E_{j+1}\|_{\rm F}
=2cjτj(1+ωj+1)Ej+1F.\displaystyle=\sqrt{2}c_{j}\tau_{j}\cdot(1+\omega_{j+1})\cdot\|E_{j+1}\|_{\rm F}.

Combine with (3.20) and we arrive at the upper bound (3.16) of Rj+1F\|R_{j+1}\|_{\rm F}. ∎

Now we recall the following theorem by Sun [18].

Theorem 3.2 (Thm.3.1 in [18]).

Let (Λ,X)(\Lambda,X) be approximate eigenpairs of a symmetric matrix AA. Let QQ be the orthonormal basis from the polar decomposition of XX. Define the set

Q={Δ(A+Δ)Q=QΛ,Δ=ΔTn×n}.\displaystyle\mathcal{H}_{Q}=\{\Delta\mid(A+\Delta)Q=Q\Lambda,\ \Delta=\Delta^{\mathrm{T}}\in{\mathbb{R}}^{n\times n}\}. (3.22)

Then Q\mathcal{H}_{Q} is non-empty and there exists a unique ΔQQ\Delta_{Q}\in\mathcal{H}_{Q} such that

minΔQΔF=ΔQFRF2+𝒫XRF2/σmin(X),\displaystyle\min_{\Delta\in\mathcal{H}_{Q}}\|\Delta\|_{\rm F}=\|\Delta_{Q}\|_{\rm F}\leq\sqrt{\|R\|^{2}_{\rm F}+\|\mathcal{P}_{X}^{\perp}R\|^{2}_{\rm F}}\Big{/}{\sigma_{\min}(X)}, (3.23)

where R=AXXΛR=AX-X\Lambda is the residual and 𝒫X\mathcal{P}_{X}^{\perp} is the orthogonal projection onto the orthogonal complement of (X)\mathcal{R}(X).

From Lemma 3.2 and Theorem 3.2, we have the following computable upper bound on the symmetric backward error norm δj+1\delta_{j+1} of the computed eigenpairs (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}) of AA.

Theorem 3.3.

By the governing equations (2.3) and (2.9) of jj steps of EED, if τjωj<2\tau_{j}\omega_{j}<\sqrt{2} and ωj+1<1\omega_{j+1}<1, then the symmetric backward error norm δj+1\delta_{j+1} of the computed eigenpairs (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}) of AA defined in (3.2) has the following upper bound

δj+12(1+cjτj(1+ωj+1)1ωj+1)Ej+1F,\displaystyle\delta_{j+1}\leq\sqrt{2}\left(\frac{1+c_{j}\tau_{j}(1+\omega_{j+1})}{\sqrt{1-\omega_{j+1}}}\right)\|E_{j+1}\|_{\rm F}, (3.24)

where cj=(1τjωj/2)1c_{j}=(1-\tau_{j}\omega_{j}/\sqrt{2})^{-1}, and γj\gamma_{j} and τj\tau_{j} are the spectral gap and the shift-gap ratio defined in (2.10) and (2.11), respectively.

Proof.

For the computed eigenvectors V^j+1\widehat{V}_{j+1} of AA, let Qj+1Q_{j+1} be the orthonormal basis from the polar decomposition of V^j+1\widehat{V}_{j+1}, and the set Qj+1\mathcal{H}_{Q_{j+1}} be defined in (3.3). It follows from the definition (3.2) and Theorem 3.2 that

δj+1\displaystyle\delta_{j+1} =minΔQj+1ΔF1σmin(V^j+1)Rj+1F2+𝒫Rj+1F2,\displaystyle=\min_{\Delta\in\mathcal{H}_{Q_{j+1}}}\|\Delta\|_{\rm F}\leq\frac{1}{\sigma_{\min}(\widehat{V}_{j+1})}\sqrt{\|R_{j+1}\|^{2}_{\rm F}+\|\mathcal{P}^{\perp}R_{j+1}\|^{2}_{\rm F}}, (3.25)

where Rj+1R_{j+1} is the residual of (Λ^j+1,V^j+1)(\widehat{\Lambda}_{j+1},\widehat{V}_{j+1}) defined in (3.15), and 𝒫\mathcal{P}^{\perp} is the orthogonal projection onto the orthogonal complement of the subspace (V^j+1)\mathcal{R}(\widehat{V}_{j+1}), i.e.,

𝒫=IV^j+1(V^j+1TV^j+1)1V^j+1T.\mathcal{P}^{\perp}=I-\widehat{V}_{j+1}(\widehat{V}^{T}_{j+1}\widehat{V}_{j+1})^{-1}\widehat{V}^{T}_{j+1}.

By the equation (3.19), 𝒫Rj+1=𝒫Ej+1\mathcal{P}^{\perp}R_{j+1}=\mathcal{P}^{\perp}E_{j+1}. Hence we have

𝒫Rj+1F=𝒫Ej+1FEj+1F.\displaystyle\|\mathcal{P}^{\perp}R_{j+1}\|_{\rm F}=\|\mathcal{P}^{\perp}E_{j+1}\|_{\rm F}\leq\|E_{j+1}\|_{\rm F}. (3.26)

On the other hand, by the definition (3.1) of ωj+1\omega_{j+1} and the assumption ωj+1<1\omega_{j+1}<1, we have

|σmin2(V^j+1)1|V^j+1TV^j+1Ij+12ωj+1<1,\displaystyle|\sigma^{2}_{\min}(\widehat{V}_{j+1})-1|\leq\|\widehat{V}_{j+1}^{\mathrm{T}}\widehat{V}_{j+1}-I_{j+1}\|_{2}\leq\omega_{j+1}<1,

which implies the following lower bound of the singular value

σmin(V^j+1)1ωj+1.\displaystyle\sigma_{\min}(\widehat{V}_{j+1})\geq\sqrt{1-\omega_{j+1}}. (3.27)

Plug (3.26) and (3.27) into (3.25) and recall the upper bound of Rj+1F\|R_{j+1}\|_{\rm F} in Lemma 3.2, and then we obtain

δj+1\displaystyle\delta_{j+1} 1+(1+2cjτj(1+ωj+1))2Ej+1F1ωj+1\displaystyle\leq\sqrt{1+\left(1+\sqrt{2}c_{j}\tau_{j}\cdot(1+\omega_{j+1})\right)^{2}}\cdot\frac{\|E_{j+1}\|_{\rm F}}{\sqrt{1-\omega_{j+1}}}
2(1+cjτj(1+ωj+1)1ωj+1)Ej+1F,\displaystyle\leq\sqrt{2}\left(\frac{1+c_{j}\tau_{j}(1+\omega_{j+1})}{\sqrt{1-\omega_{j+1}}}\right)\|E_{j+1}\|_{\rm F},

where the second inequality is due to 1+(1+2a)22(1+2a+a2)=2(1+a)21+(1+\sqrt{2}a)^{2}\leq 2(1+2a+a^{2})=2(1+a)^{2}. This completes the proof. ∎

Remark 3.2.

We note from the inequalities (3.25) and (3.27) that, when ωj+11\omega_{j+1}\ll 1, the residual norm Rj+1F\|R_{j+1}\|_{\rm F} can be used as an easy-to-compute estimate of δj+1\delta_{j+1}. We will use this observation in numerical examples in Section 5.

3.4 Backward stability of EED

In Theorems 3.1 and 3.3, the upper bounds (3.10) and (3.24) for δj+1\delta_{j+1} and ωj+1\omega_{j+1} involve the quantity ωj\omega_{j} from the previous EED step. In this section, under a mild assumption, we derive explicit upper bounds for ωj+1\omega_{j+1} and δj+1\delta_{j+1}, and then reveal conditions for the backward stability of the EED procedure.

Lemma 3.3.

Consider jj steps of EED governed by Eqs. (2.3) and (2.9). Assume

τjA2γj4j+1tol<0.1.\displaystyle\tau_{j}\frac{\|A\|_{2}}{\gamma_{j}}\cdot 4\sqrt{j+1}\cdot tol<0.1. (3.28)

Then

  1. (i)

    it holds that

    τiωi<0.11andci=(1τiωi/2)1<2for i=1,2,,j;\tau_{i}\omega_{i}<0.11\quad\text{and}\quad c_{i}=(1-\tau_{i}\omega_{i}/\sqrt{2})^{-1}<2\quad\text{for $i=1,2,\dots,j$;} (3.29)
  2. (ii)

    the loss of orthogonality ωj+1\omega_{j+1} is bounded by

    ωj+1(A2γj5j+1)tol;\displaystyle\omega_{j+1}\leq\left(\frac{\|A\|_{2}}{\gamma_{j}}\cdot 5\sqrt{j+1}\right)\cdot tol; (3.30)
  3. (iii)

    the backward error norm δj+1\delta_{j+1} is bounded by

    δj+1(τj5j+1)tolA2.\delta_{j+1}\leq\left(\tau_{j}\cdot 5\sqrt{j+1}\right)\cdot tol\cdot\|A\|_{2}. (3.31)
Proof.

First observe that by the definitions (2.10) and (2.11), γi\gamma_{i} is monotonically decreasing with the index ii and τi1\tau_{i}\geq 1 is monotonically increasing with ii. Therefore, the assumption (3.28) implies the inequalities

A2γi4i+1tol<0.1andτiA2γi14itol<0.1for all ij.\displaystyle\frac{\|A\|_{2}}{\gamma_{i}}\cdot 4\sqrt{i+1}\cdot tol<0.1\quad\text{and}\quad\tau_{i}\frac{\|A\|_{2}}{\gamma_{i-1}}\cdot 4\sqrt{i}\cdot tol<0.1\quad\text{for all $i\leq j$}. (3.32)

Since the stopping criterion (2.4) of EIGSOL implies

EiF=[η1,,ηi]FitolA2,\|E_{i}\|_{\rm F}=\|[\eta_{1},\dots,\eta_{i}]\|_{\rm F}\leq\sqrt{i}\cdot tol\cdot\|A\|_{2}, (3.33)

inequalities (3.32) leads to

4γiEi+1F<0.1andτi4γi1EiF<0.1for all ij.\displaystyle\frac{4}{\gamma_{i}}\cdot\|E_{i+1}\|_{\rm F}<0.1\quad\text{and}\quad\tau_{i}\cdot\frac{4}{\gamma_{i-1}}\cdot\|E_{i}\|_{\rm F}<0.1\quad\text{for all $i\leq j$}. (3.34)

(i) We prove the inequality (3.29) by induction. To begin with, recall that v^12=1\|\widehat{v}_{1}\|_{2}=1, which implies ω1=v^1Tv^11F=0\omega_{1}=\|\widehat{v}_{1}^{T}\widehat{v}_{1}-1\|_{F}=0, τ1ω1=0<0.11\tau_{1}\omega_{1}=0<0.11, and c1=1<2c_{1}=1<2. Hence (3.29) holds for i=1i=1. Now, for 2ij2\leq i\leq j, assume that τi1ωi1<0.11\tau_{i-1}\omega_{i-1}<0.11 and ci1<2c_{i-1}<2. Since τi1ωi1<0.11\tau_{i-1}\omega_{i-1}<0.11, we can apply Theorem 3.1 and derive from (3.10) that

τiωiτi2ci1γi1EiF(1+2ci1γi1EiF)<0.1(1+0.1)=0.11,\tau_{i}\omega_{i}\leq\tau_{i}\cdot\frac{2c_{i-1}}{\gamma_{i-1}}\|E_{i}\|_{\rm F}\cdot\left(1+\frac{2c_{i-1}}{\gamma_{i-1}}\|E_{i}\|_{\rm F}\right)<0.1\cdot(1+0.1)=0.11, (3.35)

where the last inequality of (3.35) is by 2ci1<42c_{i-1}<4 and (3.34). This implies immediately

ci=(1τiωi/2)1(10.11/2)1<2.c_{i}=(1-\tau_{i}\omega_{i}/\sqrt{2})^{-1}\leq(1-0.11/\sqrt{2})^{-1}<2.

Therefore, (3.29) follows by induction.

(ii) Since we have τjωj<0.11\tau_{j}\omega_{j}<0.11 and cj<2c_{j}<2 by (3.29), we can apply Theorem 3.1 and derive from (3.10) that

ωj+12cjγjEj+1F(1+2cjγjEj+1F)4γjEj+1F(1+0.1),\omega_{j+1}\leq\frac{2c_{j}}{\gamma_{j}}\cdot\|E_{j+1}\|_{\rm F}\cdot\left(1+\frac{2c_{j}}{\gamma_{j}}\|E_{j+1}\|_{\rm F}\right)\leq\frac{4}{\gamma_{j}}\cdot\|E_{j+1}\|_{\rm F}\cdot\left(1+0.1\right), (3.36)

where in the second inequality we used 2cj<42c_{j}<4 and the first inequality in (3.34). Recall the error bound of Ej+1F\|E_{j+1}\|_{\rm F} from (3.33) and we obtain (3.30).

(iii) We have τjωj<0.11\tau_{j}\omega_{j}<0.11 and cj<2c_{j}<2 by (3.29). It also follows from (3.36) and (3.34) that ωj+1<0.11\omega_{j+1}<0.11. Therefore, we can apply Theorem 3.3 and derive from (3.24) that

δj+12(1+cjτj(1+ωj+1)1ωj+1)Ej+1F2(1+2τj(1+0.11)10.11)Ej+1F,\delta_{j+1}\leq\sqrt{2}\left(\frac{1+c_{j}\tau_{j}(1+\omega_{j+1})}{\sqrt{1-\omega_{j+1}}}\right)\|E_{j+1}\|_{{\rm F}}\leq\sqrt{2}\left(\frac{1+2\tau_{j}(1+0.11)}{\sqrt{1-0.11}}\right)\cdot\|E_{j+1}\|_{{\rm F}},

where in second inequality we used 0ωj+1<0.110\leq\omega_{j+1}<0.11. Since τj1\tau_{j}\geq 1 by definition (2.11), we can relax the leading constant as 2(1.06+2.36τj)2(3.42τj)<5τj\sqrt{2}(1.06+2.36\cdot\tau_{j})\leq\sqrt{2}(3.42\cdot\tau_{j})<5\tau_{j}. Recall the error bound of Ej+1F\|E_{j+1}\|_{\rm F} from (3.33) and we prove (3.31). ∎

By the error bounds (3.30) and (3.31) in Lemma 3.3, we can see that the quantities γj1A2\gamma^{-1}_{j}\|A\|_{2} and τj\tau_{j} play important roles for the stability of the EED procedure. A sufficient condition to achieve the backward stability (3.4) and (3.5) is given by γj1A2=O(1)\gamma^{-1}_{j}\|A\|_{2}=O(1) and τj=O(1)\tau_{j}=O(1). In summary, we have the following theorem for the backward stability of the EED procedure.

Theorem 3.4.

Under the assumptions of the residual norm ηi2\|\eta_{i}\|_{2} of EIGSOL satisfying (2.1) and the inequality (3.28), the backward stability of the EED procedure, in the sense of (3.4) and (3.5), is guaranteed if the shifts σ1,,σj\sigma_{1},\dots,\sigma_{j} are dynamically chosen such that

γj1A2=O(1)andτj=O(1).\gamma_{j}^{-1}\|A\|_{2}=O(1)\quad\text{and}\quad\tau_{j}=O(1). (3.37)

We note that when the shifts σj\sigma_{j} are dynamically chosen such that the conditions (3.37) are satisfied, the assumption of the inequality  (3.28) is indeed mild.

Remark 3.3.

From the upper bound (3.30) of the loss of orthogonality ωj+1\omega_{j+1}, we see that if the spectral gap γj\gamma_{j} is too small, i.e., γjA2\gamma_{j}\ll\|A\|_{2}, then ωj+1\omega_{j+1} could be amplified by a factor of γj1A2\gamma_{j}^{-1}\|A\|_{2}. On the other hand, from the upper bound (3.31) of the symmetric backward error norm δj+1\delta_{j+1}, we see that when τj\tau_{j} is too large, i.e., τj1\tau_{j}\gg 1, δj+1\delta_{j+1} could be amplified by a factor of τj\tau_{j}. We will demonstrate these observations numerically in Example 5.2 in Section 5.

4 EED in practice

In Section 3.4, we derived the required conditions (3.37) of the spectral gap γj\gamma_{j} and the shift-gap ratio τj\tau_{j} for the backward stability of the EED procedure. The quantities γj\gamma_{j} and τj\tau_{j} are controlled by the selection of the shifts σ1,,σj\sigma_{1},\dots,\sigma_{j}. In this section, we discuss a practical selection scheme of the shifts to satisfy the conditions (3.37), and then present an algorithm that combines an eigensolver EIGSOL with the EED procedure to compute eigenvalues in a prescribed interval =[λlow,λupper]\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}].

4.1 Choice of the shifts

We consider the following choice of the shift at the jj-th EED step,

σj=μλ^j,\displaystyle\sigma_{j}=\mu-\widehat{\lambda}_{j}, (4.1)

where μ\mu\in{\mathbb{R}} is a parameter with μ>λupper\mu>\lambda_{\rm upper}. The shifting scheme (4.1) has been used in several previous works, although without elaboration on the choice of parameter μ\mu; see [13, Chap.5.1] and [12, 21]. In the following, we will discuss how to choose the parameter μ\mu such that the conditions (3.37) can hold.

To begin with, the choice of the shifts in (4.1) implies that the spectral gap γj\gamma_{j} in (2.10) satisfies

γj=minθj+1,λ𝒥j|λθ|=min1ij+1|μλ^i|,\displaystyle\gamma_{j}=\min_{\theta\in\mathcal{I}_{j+1},\ \lambda\in\mathcal{J}_{j}}|\lambda-\theta|=\min_{1\leq i\leq j+1}|\mu-\widehat{\lambda}_{i}|, (4.2)

where j+1={λ^1,,λ^j,λ^j+1}\mathcal{I}_{j+1}=\{\widehat{\lambda}_{1},\ldots,\widehat{\lambda}_{j},\widehat{\lambda}_{j+1}\} and 𝒥j={λ^1+σ1,,λ^j+σj}={μ}\mathcal{J}_{j}=\{\widehat{\lambda}_{1}+\sigma_{1},\ldots,\widehat{\lambda}_{j}+\sigma_{j}\}=\{\mu\}. On the other hand, it also implies that the shift-gap ratio τj\tau_{j} in (2.11) satisfies

τj=1γjmax1ij|σi|=max1ij|μλ^i|min1ij+1|μλ^i|.\displaystyle\tau_{j}=\frac{1}{\gamma_{j}}\cdot\max_{1\leq i\leq j}|\sigma_{i}|=\frac{\max_{1\leq i\leq j}|\mu-\widehat{\lambda}_{i}|}{\min_{1\leq i\leq j+1}|\mu-\widehat{\lambda}_{i}|}. (4.3)

Now recall that μ>λupper\mu>\lambda_{\rm upper} and the computed eigenvalues λ^i[λlow,λupper]\widehat{\lambda}_{i}\in[\lambda_{\rm low},\lambda_{\rm upper}], for i=1,2,,j+1i=1,2,\dots,j+1, so we have

μλuppermin1ij+1|μλ^i|max1ij+1|μλ^i|μλlow.\displaystyle\mu-\lambda_{\rm upper}\leq\min_{1\leq i\leq j+1}|\mu-\widehat{\lambda}_{i}|\leq\max_{1\leq i\leq j+1}|\mu-\widehat{\lambda}_{i}|\leq\mu-\lambda_{\rm low}.

Hence, (4.2) and (4.3) lead to

γgγjγgτgandτjτg,\displaystyle\gamma_{g}\leq\gamma_{j}\leq\gamma_{g}\,\tau_{g}\quad\mbox{and}\quad\tau_{j}\leq\tau_{g}, (4.4)

where

γgμλupperandτgμλlowμλupper.\gamma_{g}\equiv\mu-\lambda_{\rm upper}\quad\mbox{and}\quad\tau_{g}\equiv\frac{\mu-\lambda_{\rm low}}{\mu-\lambda_{\rm upper}}.

The problem is then turned to the choice of parameter μ\mu such that the quantities γg\gamma_{g} and τg\tau_{g} to satisfy the conditions (3.37).

Let us consider a frequently encountered case in practice where the width of the interval =[λlow,λupper]\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}] satisfies

λupperλlow12A2.\lambda_{\rm upper}-\lambda_{\rm low}\leq\frac{1}{2}\|A\|_{2}.

Then by setting

μ=λ^1+A2,\mu=\widehat{\lambda}_{1}+\|A\|_{2},

we have

12A2γg=(1λupperλ^1A2)A2A2\frac{1}{2}\|A\|_{2}\leq\gamma_{g}=\left(1-\frac{\lambda_{\rm upper}-\widehat{\lambda}_{1}}{\|A\|_{2}}\right)\|A\|_{2}\leq\|A\|_{2}

and

τg=1+λupperλlowγg2.\tau_{g}=1+\frac{\lambda_{\rm upper}-\lambda_{\rm low}}{\gamma_{g}}\leq 2.

Consequently, by (4.4), the spectral gap γj\gamma_{j} and the shift-gap ratio τj\tau_{j} satisfy the desired conditions (3.37).

In summary, for the EED procedure in practice, we recommend the use of the following strategy for the choice of the shift σj\sigma_{j} at the jj-th EED,

σj=μλ^jwithμ=λ^1+A2,\sigma_{j}=\mu-\widehat{\lambda}_{j}\quad\mbox{with}\quad\mu=\widehat{\lambda}_{1}+\|A\|_{2}, (4.5)

to compute the eigenvalues in an interval =[λlow,λupper]\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}], where λupperλlow12A2\lambda_{\rm upper}-\lambda_{\rm low}\leq\frac{1}{2}\|A\|_{2}.

4.2 Algorithm

An eigensolver EIGSOL combined with the EED procedure to compute the eigenvalues in an interval =[λlow,λupper]\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}] at the lower end of the spectrum of AA is summarized in Algorithm 1.

Algorithm 1 An eigensolver EIGSOL combined with the EED procedure
0:   the interval =[λlow,λupper]\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}] at the lower end of the spectrum of AA; the relative tolerance toltol in (2.1) for EIGSOL.
0:   the approximate eigenpairs (λ^i,v^i)(\widehat{\lambda}_{i},\widehat{v}_{i}) in the interval \mathcal{I}.
1:  A^0=A\widehat{A}_{0}=A;
2:  use EIGSOL to compute the lowest eigenpair (λ^1,v^1)(\widehat{\lambda}_{1},\widehat{v}_{1}) of A^0\widehat{A}_{0} and an estimate Anorm of A2\|A\|_{2};
3:  μ=λ^1+𝙰𝚗𝚘𝚛𝚖\mu=\widehat{\lambda}_{1}+{\tt Anorm};
4:  for j=1,2,j=1,2,\ldots do
5:     σj=μλ^j\sigma_{j}=\mu-\widehat{\lambda}_{j};
6:     A^j=A^j1+σjv^jv^jT=A+V^jΣjV^jT\widehat{A}_{j}=\widehat{A}_{j-1}+\sigma_{j}\widehat{v}_{j}\widehat{v}_{j}^{\mathrm{T}}=A+\widehat{V}_{j}\Sigma_{j}\widehat{V}_{j}^{\mathrm{T}};
7:      compute the lowest eigenpair (λ^j+1,v^j+1)(\widehat{\lambda}_{j+1},\widehat{v}_{j+1}) of A^j\widehat{A}_{j} by EIGSOL;
8:      check if all the eigenpairs in the interval \mathcal{I} have been computed;
9:  end for
10:  return the approximate eigenpairs (λ^i,v^i)(\widehat{\lambda}_{i},\widehat{v}_{i}) in the interval \mathcal{I};

A few remarks are in order:

  • In practice, we never need to form the matrix A^j\widehat{A}_{j} at step 6 explicitly. We can assume that the only operation that is required by EIGSOL is the matrix-vector product y:=A^jxy:=\widehat{A}_{j}x.

  • At step 7, the computation of the lowest eigenpair (λ^j+1,v^j+1)(\widehat{\lambda}_{j+1},\widehat{v}_{j+1}) of A^j\widehat{A}_{j} can be accelerated by warm starting the EIGSOL with the lowest unconverged eigenvectors of A^j1\widehat{A}_{j-1}. This is possible for most iterative eigensolvers, such as TRLan [20] and ARPACK [9].

  • At step 8, an ideal validation method is to use the inertias of the shifted matrix AλupperIA-\lambda_{\rm upper}I. However, computation of the inertias is a prohibitive cost for large matrices. An empirical validation is to monitor the lowest eigenvalue λ^j+1\widehat{\lambda}_{j+1} of A^j\widehat{A}_{j}. All eigenpairs in the interval \mathcal{I} are considered to be found when λ^j+1\widehat{\lambda}_{j+1} is outside the interval \mathcal{I}.

5 Numerical experiments

In this section, we first use synthetic examples to verify the sharpness of the upper bounds (3.10) and (3.24) on the loss of orthogonality and the symmetric backward error norm of the EED procedure under the choice (4.5) of the shifts σj\sigma_{j}. We present the cases where improper choices of the shifts σj\sigma_{j} may lead to numerical instability of the EED procedure. Then we demonstrate the numerical stability of the EED procedure for a set of large sparse symmetric matrices arising from applications.

We use TRLan as the eigensolver. TRLan is a C implementation of the thick-restart Lanczos method with adaptive sizes of the projection subspace [20, 22, 23]. The convergence criterion of an approximate eigenpair (λ^j+1,v^j+1)(\widehat{\lambda}_{j+1},\widehat{v}_{j+1}) is the residual norm satisfying

ηj+12=A^jv^j+1λ^j+1v^j+12<tol𝙰𝚗𝚘𝚛𝚖,\|\eta_{j+1}\|_{2}=\|\widehat{A}_{j}\widehat{v}_{j+1}-\widehat{\lambda}_{j+1}\widehat{v}_{j+1}\|_{2}<tol\cdot{\tt Anorm},

where toltol is a user-specified tolerance and Anorm is a 22-norm estimate of AA computed by TRLan. The starting vector is a random vector.

Example 5.1.

In this example, we demonstrate the sharpness of the upper bounds (3.10) and (3.24) on the loss of orthogonality and the symmetric backward error norm with the choice (4.5) of the shifts σj\sigma_{j}.

We consider a diagonal matrix AA with diagonal elements

akk={12dk,if 1kn/2,12(1+dkn/2),if n/2<kn,\displaystyle a_{kk}=\begin{cases}\frac{1}{2}d_{k},\quad\text{if $1\leq k\leq n/2$},\\ \frac{1}{2}(1+d_{k-n/2}),\quad\text{if $n/2<k\leq n$},\end{cases}

where dk=105(1k1n/21)d_{k}=10^{-5(1-\frac{k-1}{n/2-1})} and the matrix size n=500n=500. The spectrum range of AA is (0,1](0,1]. The eigenvalues of AA are clustered around 0 and 0.50.5. We are interested in computing the ne=65n_{e}=65 eigenvalues in the interval =[0, 104]\mathcal{I}=[0,\,10^{-4}]. The computed 22-norm of AA is 𝙰𝚗𝚘𝚛𝚖=1.00{\tt Anorm}=1.00.

To closely observe the convergence, TRLan is slightly modified so that the convergence test is performed at each Lanczos iteration. The maximal dimension mm of the projection subspace is set to be 4040.

Numerical results of TRLan with the EED procedure for computing all the nen_{e} eigenvalues in the interval \mathcal{I} are summarized in Table 1, where the 4th column is the loss of orthogonality ωne\omega_{n_{e}}, the 5th column is the upper bound (3.10) of ωne\omega_{n_{e}}, the 6th column is the residual norm RneF\|R_{n_{e}}\|_{\rm F} and the 7th column is the upper bound (3.24) of δne\delta_{n_{e}}. Note that here we use the quantities δne\delta_{n_{e}} and RneF\|R_{n_{e}}\|_{\rm F} interchangeably as discussed in Remark 3.2.

From Table 1, we observe that with the choice (4.5) of the shifts σj\sigma_{j}, γg1𝙰𝚗𝚘𝚛𝚖1\gamma^{-1}_{g}{\tt Anorm}\approx 1 and τg1\tau_{g}\approx 1. Therefore, the conditions (3.37) of the spectral gap γj\gamma_{j} and the shift-gap ratio τj\tau_{j} for the backward stability are satisfied. Consequently, the loss of orthogonality of the computed eigenvectors is ωne=O(tol)\omega_{n_{e}}=O(tol) and the symmetric backward error norm of the computed eigenpairs (Λ^ne,V^ne)(\widehat{\Lambda}_{n_{e}},\widehat{V}_{n_{e}}) is δne=O(tol𝙰𝚗𝚘𝚛𝚖)\delta_{n_{e}}=O(tol\cdot{\tt Anorm}). In addition, we observe that the upper bounds (3.10) and (3.24) of ωne\omega_{n_{e}} and δne\delta_{n_{e}} are tight within an order of magnitude.

Table 1: Numerical stability of TRLan with EED for different tolerances toltol (Example 5.1)
toltol μ\mu γg\gamma_{g} ωne\omega_{n_{e}} bound (3.10) RneF\|R_{n_{e}}\|_{\rm F} bound (3.24)
10610^{-6} 1.00 1.001.00 2.371062.37\cdot 10^{-6} 1.591051.59\cdot 10^{-5} 7.871067.87\cdot 10^{-6} 2.241052.24\cdot 10^{-5}
10810^{-8} 1.00 1.001.00 1.781081.78\cdot 10^{-8} 1.581071.58\cdot 10^{-7} 7.951087.95\cdot 10^{-8} 2.241072.24\cdot 10^{-7}
101010^{-10} 1.00 1.001.00 1.8210101.82\cdot 10^{-10} 1.581091.58\cdot 10^{-9} 7.9410107.94\cdot 10^{-10} 2.241092.24\cdot 10^{-9}
Refer to caption
Figure 2: The loss of orthogonality ωj+1\omega_{j+1} and the upper bound (3.10) of ωj+1\omega_{j+1} against the computed eigenvalues λ^j+1\widehat{\lambda}_{j+1} for 2j+1ne2\leq j+1\leq n_{e}, tol=108tol=10^{-8} (Example 5.1).
Example 5.2.

In this example we illustrate that improperly chosen shifts σj\sigma_{j} may lead to instability of the EED procedure.

In Remark 3.3, we stated that if the shifts σj\sigma_{j} are chosen such that the spectral gaps γj\gamma_{j} are too small, i.e., γj𝙰𝚗𝚘𝚛𝚖\gamma_{j}\ll{\tt Anorm}, then the loss of orthogonality of the computed eigenvectors could be amplified by a factor of γj1𝙰𝚗𝚘𝚛𝚖\gamma_{j}^{-1}\cdot{\tt Anorm}. Here is an example using the same diagonal matrix AA in Example 5.1. The combination of TRLan and EED is used to compute the ne=65n_{e}=65 eigenvalues in the interval =[0, 104]\mathcal{I}=[0,\,10^{-4}]. Let us set the shifts σj=μλ^j\sigma_{j}=\mu-\widehat{\lambda}_{j} with μ=2104\mu=2\cdot 10^{-4}, which is much smaller than the recommended value of μ=λ^1+A21.00\mu=\widehat{\lambda}_{1}+\|A\|_{2}\approx 1.00. Numerical results are summarized in Table 2, where the tolerance tol=108tol=10^{-8} for TRLan. We observe that γj=O(γg)𝙰𝚗𝚘𝚛𝚖\gamma_{j}=O(\gamma_{g})\ll{\tt Anorm}, and the loss of orthogonality of the computed eigenvectors is indeed amplified by a factor of γg1𝙰𝚗𝚘𝚛𝚖\gamma_{g}^{-1}\cdot{\tt Anorm}. We note that since τj=O(1)\tau_{j}=O(1), the symmetric backward error norms δne=O(tol𝙰𝚗𝚘𝚛𝚖)\delta_{n_{e}}=O(tol\cdot{\tt Anorm}).

Table 2: Instability of TRLan with EED when the spectral gaps γj=O(γg)\gamma_{j}=O(\gamma_{g}) are too small.
toltol μ\mu γg\gamma_{g} ωne\omega_{n_{e}} bound (3.10) RneF\|R_{n_{e}}\|_{\rm F} bound (3.24)
10610^{-6} 21042\cdot 10^{-4} 10410^{-4} 8.261038.26\cdot 10^{-3} 1.791011.79\cdot 10^{-1} 8.001068.00\cdot 10^{-6} 3.291053.29\cdot 10^{-5}
10810^{-8} 21042\cdot 10^{-4} 10410^{-4} 8.281058.28\cdot 10^{-5} 1.531031.53\cdot 10^{-3} 7.961087.96\cdot 10^{-8} 3.221073.22\cdot 10^{-7}
101010^{-10} 21042\cdot 10^{-4} 10410^{-4} 8.271078.27\cdot 10^{-7} 1.521051.52\cdot 10^{-5} 7.9510107.95\cdot 10^{-10} 3.221093.22\cdot 10^{-9}

In Remark 3.3, we also stated that if the shifts σj\sigma_{j} are chosen such that the shift-gap ratios τj\tau_{j} are too large, then the symmetric backward error norm δj\delta_{j} could be amplified by a factor of τj\tau_{j}. Here is an example. We flip the sign of the diagonal elements of AA defined in Example 5.1, and set n=200n=200. We compute ne=74n_{e}=74 eigenvalues in the interval =[1.0,0.5001]\mathcal{I}=[-1.0,-0.5001] using TRLan with EED procedure. The computed 22-norm of AA is 𝙰𝚗𝚘𝚛𝚖=1.00{\tt Anorm}=1.00.

Refer to caption
Refer to caption
Figure 3: The loss of orthogonality ωj+1\omega_{j+1} (left) and the residual norm Rj+1F\|R_{j+1}\|_{\rm F} (right) against the computed eigenvalues λ^j+1\widehat{\lambda}_{j+1} for 2j+1ne2\leq j+1\leq n_{e}. The red lines are toltol (left) and tol𝙰𝚗𝚘𝚛𝚖tol\cdot{\tt Anorm} (right).

Instead of the choice (4.5) for the shifts σj\sigma_{j}, we set σj=μλ^j\sigma_{j}=\mu-\widehat{\lambda}_{j} with μ=0.5\mu=-0.5. The blue ×\times-lines in Figure 3 are the loss of orthogonality and the residual norms for the computed eigenpairs (λ^j+1,v^j+1)(\widehat{\lambda}_{j+1},\widehat{v}_{j+1}) for 2j+1ne2\leq j+1\leq n_{e}. As we observe that for the first 66 computed eigenvalues in the subinterval [1.0,0.75][-1.0,-0.75] of \mathcal{I}, since the spectral gap γj0.25\gamma_{j}\geq 0.25 and the shift-gap ratio τj2\tau_{j}\leq 2, the computed eigenpairs are backward stable with ω6=2.48109=O(tol)\omega_{6}=2.48\cdot 10^{-9}=O(tol) and R6F=9.05109=O(tol𝙰𝚗𝚘𝚛𝚖)\|R_{6}\|_{\rm F}=9.05\cdot 10^{-9}=O(tol\cdot{\tt Anorm}). However, for the computed eigenvalues in the subinterval [0.75,0.5001][-0.75,-0.5001] of \mathcal{I}, the computed eigenpairs are not backward stable due to the facts that the spectral gaps γj\gamma_{j} become small, γj1.03104\gamma_{j}\approx 1.03\cdot 10^{-4}, and the shift-gap ratios τj\tau_{j} grows up to τj4.86103\tau_{j}\approx 4.86\cdot 10^{3}. Consequently, the loss of orthogonality ωne\omega_{n_{e}} and the residual norm RneF\|R_{n_{e}}\|_{\rm F} are increased by a factor of up to 10310^{3}, respectively. The stability are restored if the shifts are chosen according to the recommendation (4.5) as shown by the green ++-lines in Figure 3.

Example 5.3.

In this example, we demonstrate the numerical stability of TRLan with the EED procedure for a set of large sparse symmetric matrices from applications.

The statistics of the matrices are summarized in Table 3, where nn is the size of the matrix, nnz is the number of nonzero entries of the matrix, [λmin,λmax][\lambda_{\min},\lambda_{\max}] is the spectrum range, and nen_{e} is the number of eigenvalues in the interval =[λlow,λupper]\mathcal{I}=[\lambda_{\rm low},\lambda_{\rm upper}]. The quantities nen_{e} are calculated by computing the inertias of the shifted matrices AλupperIA-\lambda_{\rm upper}I. The Laplacian is the negative 2D Laplacian on a 200200-by-200200 grids with Dirichlet boundary condition [16]. The worms20 is the graph Laplacian worms20_10NN in machine learning datasets [4]. The SiO, Si34H36, Ge87H76 and Ge99H100 are Hamiltonian matrices from PARSEC collection [4].

Table 3: Statistics of the test matrices.
matrix nn nnz [λmin,λmax][\lambda_{\min},\lambda_{\max}] [λlow,λupper][\lambda_{\rm low},\lambda_{\rm upper}] nen_{e}
Laplacian 40,00040,000 199,200199,200 [0,7.9995][0,7.9995] [0,0.07][0,0.07] 205205
worms20 20,05520,055 260,881260,881 [0,6.0450][0,6.0450] [0,0.05][0,0.05] 289289
SiO 33,40133,401 1,317,6551,317,655 [1.6745,84.3139][-1.6745,84.3139] [1.7,2.0][-1.7,2.0] 182182
Si34H36 97,56997,569 5,156,3795,156,379 [1.1586,42.9396][-1.1586,42.9396] [1.2,0.4][-1.2,0.4] 310310
Ge87H76 112,985112,985 7,892,1957,892,195 [1.214,32.764][-1.214,32.764] [1.3,0.0053][-1.3,-0.0053] 318318
Ge99H100 112,985112,985 8,451,3958,451,395 [1.226,32.703][-1.226,32.703] [1.3,0.0096][-1.3,-0.0096] 372372

We run TRLan with a maximal number mm of Lanczos vectors to compute the lowest eigenpairs of the matrix A^j\widehat{A}_{j}. The convergence test is performed at each restart of TRLan. All the converged eigenvalues in the interval \mathcal{I} are shifted by EED. Meanwhile, we also keep a maximal number m0m_{0} of the lowest unconverged eigenvectors as the starting vectors of TRLan for the matrix A^j+1\widehat{A}_{j+1}. All the eigenvalues in \mathcal{I} are assumed to be computed when the lowest converged eigenvalue is outside the interval \mathcal{I}. This combination of TRLan is referred to as TRLED.

TRLED is compiled using the icc compiler (version 2021.1) with the optimization flag -O2, and linked to BLAS and LAPACK available in Intel Math Kernel Library (version 2021.1.1). The experiments are conducted on a MacBook with 1.61.6 GHz Intel Core i5 CPU and 8GB of RAM.

For numerical experiments, we set the maximal number of Lanczos vectors m=150m=150. When starting TRLED for A^j+1\widehat{A}_{j+1}, the maximal number of the starting vectors is m0=75m_{0}=75. The convergence tolerance for the residual norm was set to tol=108tol=10^{-8} as a common practice for solving large scale eigenvalue problems with double precision; see for example [11].

Numerical results of TRLED are summarized in Table 4, where the 2nd column is the number n^e\widehat{n}_{e} of the computed eigenpairs (λ^i,x^i)(\widehat{\lambda}_{i},\widehat{x}_{i}) in the interval \mathcal{I}, the 3rd column is the number jmaxj_{\max} of steps of EED performed, the 4th column is the loss of orthogonality ωn^e\omega_{\widehat{n}_{e}}, and the 5th column is the relative residual norm Rn^eF/𝙰𝚗𝚘𝚛𝚖\|R_{\widehat{n}_{e}}\|_{\rm F}/{\tt Anorm} of the computed eigenpairs (Λ^ne,V^ne)(\widehat{\Lambda}_{n_{e}},\widehat{V}_{n_{e}}). From the quantities nen_{e} in Table 3 and n^e\widehat{n}_{e} in Table 4, we see that for all test matrices the eigenvalues in the prescribed intervals \mathcal{I} are successfully computed with the desired backward stability.

Table 4: Numerical results of TRLED.
matrix n^e\widehat{n}_{e} jmaxj_{\max} ωn^e\omega_{\widehat{n}_{e}} Rn^eF/𝙰𝚗𝚘𝚛𝚖\|R_{\widehat{n}_{e}}\|_{\rm F}/{\tt Anorm} CPU time (sec.)
TRLED TRLan
Laplacian 205205 6060 1.931081.93\cdot 10^{-8} 6.331086.33\cdot 10^{-8} 66.566.5 86.086.0
worms20 289289 8686 2.631082.63\cdot 10^{-8} 7.241087.24\cdot 10^{-8} 57.357.3 74.874.8
SiO 182182 4141 2.331082.33\cdot 10^{-8} 4.711084.71\cdot 10^{-8} 42.442.4 47.147.1
Si34H36 310310 7272 3.411083.41\cdot 10^{-8} 7.501087.50\cdot 10^{-8} 309.9309.9 310.4310.4
Ge87H76 318318 6666 4.081084.08\cdot 10^{-8} 8.501088.50\cdot 10^{-8} 388.7388.7 421.0421.0
Ge99H100 372372 7474 3.651083.65\cdot 10^{-8} 7.631087.63\cdot 10^{-8} 501.1501.1 533.4533.4

The left plot of Figure 4 is a profile of the number of converged eigenvalues at each external deflation of a total of 74 EEDs for the matrix Ge99H100. The right plot of Figure 4 shows the relative residual norms of all 372 computed eigenpairs in the interval. We observe that a large number of converged eigenvalues are deflated and shifted away at some EED steps.

Refer to caption
Refer to caption
Figure 4: The number of deflated eigenpairs at each EED for the matrix Ge99H100 (left). The relative residual norms of 372 computed eigenpairs (right).

To examine whether the multiple explicit external deflations lead to a significant increase in execution time, in the 6th and 7th columns of Table 4, we record the CPU time of TRLED and TRLan for computing all eigenvalues in the same intervals. For TRLan, we set the maximal number of Lanczos vectors to ne+150n_{e}+150. The restart scheme with restart=1 is used. TRLan is compiled and executed under the same setting as TRLED. We observe comparable execution time of TRLED and TRLan. In fact, TRLED is slightly faster. We think this might be due to the facts that TRLED uses fewer Lanczos vectors, which compensates the cost of the matrix-vector products y:=A^jxy:=\widehat{A}_{j}x and the warm start with those unconverged eigenvectors from the previous deflated matrix A^j1\widehat{A}_{j-1}. These are subjects of future study.

6 Concluding remarks

Based on the governing equations of the EED procedure in finite precision arithmetic, we derived the upper bounds on the loss of the orthogonality of the computed eigenvectors and the symmetric backward error norm of the computed eigenpairs in terms of two key quantities, namely the spectral gaps and the shift-gap ratios. Consequently, we revealed the required conditions of these two key quantities for the backward stability of the EED procedure. We present a practical strategy on the dynamically selection of the shifts such that the spectral gaps and the shift-gap ratios satisfy the required conditions.

There are a number of topics that can be explored further. One of them is the acceleration effect of the warm starting the EIGSOL with the unconverged eigenvectors of A^j1\widehat{A}_{j-1} for computing of the lowest eigenpairs of A^j\widehat{A}_{j}. Another one is how to efficiently perform the matrix-vector products y:=A^jxy:=\widehat{A}_{j}x in EIGSOL by exploiting the sparse plus low rank structure of the matrix A^j\widehat{A}_{j}. Communication-avoid algorithms for sparse-plus-low-rank matrix-vector products can be found in [10, 7]. A preliminary numerical study for high performance computing is reported in [21].

References

  • [1] Z. Bai, J. Demmel, J. Dongarra, A. Ruhe, and H. van der Vorst. Templates for the Solution of Algebraic Eigenvalue Problems: A Practical Guide. SIAM, Philadelphia, PA, 2000.
  • [2] Z. Bai, R.-C. Li, and W. Lin. Linear response eigenvalue problem solved by extended locally optimal preconditioned conjugate gradient methods. Sci. China Math., 259:1443–1460, 2016.
  • [3] P.-Y. Chen, B. Zhang, and M. Al Hasan. Incremental eigenpair computation for graph laplacian matrices: theory and applications. Social Network Analysis and Mining, 8(1):4, 2018.
  • [4] T. Davis and Y. Hu. The University of Florida sparse matrix collection. ACM Trans. Math. Software (TOMS), 38(1):1–25, 2011.
  • [5] D. Higham and N. Higham. Structured backward error and condition of generalized eigenvalue problems. SIAM J. Mat. Anal. Appl., 20(2):493–512, 1998.
  • [6] H. Hotelling. Analysis of a complex of statistical variables into principal components. J. of Educational Psychology, 24(6):417, 1933.
  • [7] N. Knight, E. Carson, and J. Demmel. Exploiting data sparsity in parallel matrix powers computations. In the proceedings of the international conference on parallel processing and applied mathematics (PPAM), 2014.
  • [8] R. Lehoucq and D. Sorensen. Deflation techniques for an implicitly restarted Arnoldi iteration. SIAM J. Matrix Anal. Appl., 17(4):789–821, 1996.
  • [9] R. Lehoucq, D. Sorensen, and C. Yang. ARPACK Users’ Guide: Solution of Large-Scale Eigenvalue Problems with Implicitly Restarted Arnoldi Methods. SIAM, Philadelphia, PA, 1998.
  • [10] C. Leiserson, S. Rao, and S. Toledo. Efficient out-of-core algorithms for linear relaxation using blocking covers. J. Comput. Syst. Sci. Int., 54:332––344, 1997.
  • [11] R. Li, Y. Xi, E. Vecharynskio, C. Yang, and Y. Saad. A thick-restart Lanczos algorithm with polynomial filtering for Hermitian eigenvalue problems. SIAM J. Sci. Comput., 38:A2512–A2534, 2016.
  • [12] J. Money and Q. Ye. Algorithm 845: EIGIFP: a MATLAB program for solving large symmetric generalized eigenvalue problems. ACM Trans. Math. Software (TOMS), 31(2):270–279, 2005.
  • [13] B. N. Parlett. The Symmetric Eigenvalue Problem. SIAM, Philadelphia, PA, 1998.
  • [14] Y. Saad. Numerical solution of large nonsymmetric eigenvalue problems. Comput. Phys. Commun., 53(1-3):71–90, 1989.
  • [15] Y. Saad. Numerical Methods for Large Eigenvalue Problems: revised edition, volume 66. SIAM, Philadelphia, PA, 2011.
  • [16] B. Smith and A. Knyazev. Laplacian in 1D, 2D or 3D, 2015 (accessed March 24, 2020). https://www.mathworks.com/matlabcentral/fileexchange/27279-laplacian-in-1d-2d-or-3d.
  • [17] Danny C Sorensen. Implicit application of polynomial filters in a kk-step Arnoldi method. SIAM J. Matrix. Anal. Appl., 13(1):357–385, 1992.
  • [18] J.-G. Sun. A note on backward perturbations for the Hermitian eigenvalue problem. BIT Numer. Math., 35(3):385–393, 1995.
  • [19] J. H. Wilkinson. The algebraic eigenvalue problem. Clarendon Press, Oxford, 1965.
  • [20] K. Wu and H. Simon. Thick-restart Lanczos method for large symmetric eigenvalue problems. SIAM J. Matrix Anal. Appl., 22(2):602–616, 2000.
  • [21] I. Yamazaki, Z. Bai, D. Lu, and J. Dongarra. Matrix powers kernels for thick-restart Lanczos with explicit external deflation. In 2019 IEEE International Parallel Distributed Processing Symposium (IPDPS), pages 472–481. IEEE, 2019.
  • [22] I. Yamazaki, Z. Bai, H. Simon, L.-W. Wang, and K. Wu. Adaptive projection subspace dimension for the thick-restart Lanczos method. ACM Trans. Math. Software (TOMS), 37(3):27, 2010.
  • [23] I. Yamazaki, K. Wu, and H. Simon. nu-TRLan User Guide. Technical report, LBNL-1288E, 2008. Software https://codeforge.lbl.gov/projects/trlan/.