On a formula of Thompson and McEnteggert for the adjugate matrix
Abstract
For an eigenvalue of a Hermitian matrix , the formula of Thompson and McEnteggert gives an explicit expression of the adjugate of , , in terms of eigenvectors of for and all its eigenvalues. In this paper Thompson-McEnteggert’s formula is generalized to include any matrix with entries in an arbitrary field. In addition, for any nonsingular matrix , a formula for the elementary divisors of is provided in terms of those of . Finally, a generalization of the eigenvalue-eigenvector identity and two applications of the Thompson-McEnteggert’s formula are presented.
keywords:
Adjugate, eigenvalues, eigenvectors, elementary divisors, rank-one matrices.MSC:
15A18 , 15A151 Introduction
Let be a commutative ring with identity. Following [16, Ch. 30], for a polynomial its derivative is . Recall that if is a square matrix of order with entries in and is the minor obtained from by deleting the th row and th column then the adjugate of , , is the matrix whose entry is ; that is, .
Formula (1) below, from now on TM formula, was proved, with and the normalization , for a Hermitian matrix by Thompson and McEnteggert (see [33, pp. 212-213]). Inspection of the proof shows that the formula also holds for normal matrices over (see [28]). With the same arguments we can go further. Recently, Denton, Parke, Tao, and Zhang pointed out that the TM formula has an extension to a non-normal matrix , so long as it is diagonalizable (see [12, Rem. 4]). Even more, as shown in Remark 5 of [12] it holds for matrices over commutative rings (see [17] for an informal proof). A more detailed proof of this result will be given in Section 2. However, for matrices over fields (or over integral domains) with repeated eigenvalues, (1) does not provide meaningful information (see Remark 2.4). We will exhibit in Section 2 a generalization of the TM formula which holds for matrices over arbitrary fields with repeated eigenvalues. This new TM formula will be used to generalize the so-called eigenvector-eigenvalue identity (see (20)) for non-diagonalizable matrices over arbitrary fields. In addition we will provide a complete characterization of the similarity invariants of in terms of those of , generalizing a result about the eigenvalues and the minimal polynomial in [18]. Then in Section 3 two additional consequences of the TM formula will be analysed.
2 The TM formula and its generalization
Let be a square matrix of order with entries in . An element is said to be an eigenvalue of if for some nonzero vector ([7, Def. 17.1]). This vector is said to be a right eigenvector of for (or associated with) . The left eigenvectors of for are the right eigenvectors for of , the transpose of , or, if is the field of complex numbers, of , the conjugate transpose of . That is to say, is a left eigenvector of for if (or if ). The characteristic polynomial of is and is an eigenvalue of if and only if , where is the set of zero divisors of ([7, Lem. 17.2]).
The following result, in a slightly different form, was proved by D. Grinberg in [17].
Theorem 2.1.
Let and let be an eigenvalue of . Let be a right and a left eigenvector, respectively, of for . Then
(1) |
The proof in [17] is based on the Lemma 2.2 below which is interesting in its own right. According to McCoy’s theorem ([7, Th. 5.3]) there is a non-zero vector such that if and only if , where is the (McCoy) rank of ([7, Def. 4.10]). In other words, is an eigenvalue of if and only if . Note that .
Lemma 2.2.
Let be a matrix such that and let be a left eigenvector of for the eigenvalue . For , let be the th column of . Then, for all ,
(2) |
where .
This is Lemma 3 of [17]. The author himself considers the proof to be informal. So a detailed proof of Lemma 2.2, following Grinberg’s ideas333Grinberg’s permission was granted to include the proofs of this Lemma and Theorem 2.1, is given next for reader’s convenience.
Proof of Lemma 2.2.
Let us take and assume that ; otherwise, there is nothing to prove. We assume also, without lost of generality, that . Let and, for , let be the th row of . Define to be the matrix whose th row, , is equal to if and if . A simple computation shows that and . We claim that =. This would prove the lemma.
It follows from that and so
(3) |
Let
This matrix is invertible in (its determinant is ) and by (3),
Then, and, since is invertible, . Hence and for
But in the th column of the only nonzero entry is in position . Therefore, . Now, taking into account that is the matrix with rows th and th interchanged and recalling that is the minor of obtained by deleting the th row and th column of , we get
as claimed. ∎
There is a “row version” of Lemma 2.2 which can be proved along the same lines.
Lemma 2.3.
Let be a matrix such that and let be a right eigenvector of for the eigenvalue . For let be the th row of . Then, for all ,
(4) |
where .
The proof of Theorem 2.1 which follows is very much that of Grinberg in [17]. It is included for completion and reader’s convenience.
Proof of Theorem 2.1.
Let and its characteristic polynomial. Then where, for , , and is the principal submatrix of formed by the rows and columns , …, . In particular, where is the principal minor of obtained by deleting the th row and column. Thus .
On the other hand, . It follows from the definition of derivative of a polynomial that
Hence, proving (1) is equivalent to proving
(5) |
where . It follows from and that and , respectively. So we can apply to properties (2) and (4). It follows from (2) that for all . Then and from (4), . Hence,
Adding on and taking into account that , we get
This is equivalent to (5) and the theorem follows. ∎
Remark 2.4.
Assume that is an integral domain and note that in this case ; i.e., the McCoy rank and the usual rank coincide. It is an interesting consequence of (1) that implies . The converse is not true in general. For example, if then satisfies both and , but and . However, if and then, necessarily, because is not the zero matrix. In particular, if is a field of characteristic zero (see [16, Ch. 30]) then it follows from (1) that if then is an eigenvalue of algebraic multiplicity at least . On the other hand, it is easily checked that if is an eigenvalue of algebraic multiplicity bigger that and geometric multiplicity then for any right and left eigenvectors, and respectively, of for . This is the case, for example, of . For this matrix, the TM formula (1) does not provide any substantial information about because, in this case, and . Thus, the TM formula (1) is relevant for matrices with simple eigenvalues.
Our next goal is to provide a generalization of the TM formula (1) which is meaningful for nondiagonalizable matrices over fields. We will use the following notation: will denote an arbitrary field. If then ,…, will be its (possibly repeated) elementary divisors in ([15, Ch. VI, Sec. 3]). These are powers of monic irreducible polynomials of (the ring of polynomials with coefficients in ). We will assume that for ,
Let denote the determinant of and the set of eigenvalues (the spectrum) of in, perhaps, an extension field, , of . Thus if and only if it is a root in of for some . In particular, is the characteristic polynomial of .
Item (ii) of the following theorem is an elementary result that is included for completion.
Theorem 2.5.
With the above notation:
-
(i)
If then the elementary divisors of are ,…, where for ,
(6) -
(ii)
If and there are two indices , , such that then .
-
(iii)
If , for only one value and are arbitrary right and left eigenvectors of , respectively, for the eigenvalue , then and
(7) where is the -th derivative of .
Proof.
For , let the companion matrix of be
(8) |
Then (see [15, Ch. VI, Sec. 6]) there is an invertible matrix such that
(9) |
An explicit computation shows that
Bearing in mind that , we obtain where, for ,
(10) |
Therefore, from (9) we get
(11) |
-
(i)
Assume that . This means that for all and we can write
Taking into account the definition of of (6),
Let us see that is a power of an irreducible polynomial in . In fact, put
This polynomial is sometimes called the reversal polynomial of (see, for example, [22]). Since is an elementary divisor of in , it is a power of an irreducible polynomial of . By [1, Lemma 4.4], is also a power of an irreducible polynomial. Now, it is not difficult to see that is a power of an irreducible polynomial too. As a consequence, are the elementary divisors of . Since this and are similar matrices (cf. (11)), , are the elementary divisors of . This proves (i).
-
(ii)
If for , then . Hence all minors of of order are equal to zero and so .
-
(iii)
Assume now that there is only one index such that . Then because it is a power of an irreducible polynomial. Thus for and by (8) and (10), and
(12) respectively. Also, it follows from that for , .
Recall now that and split and accordingly:
with and , . Then
(13) For let and be the -th column and row of and , respectively:
Bearing in mind that (cf. (11)), the representation of as a rank-one matrix of (12) and that for , we get
(14) Now, it follows from (13) that
Henceforth, and are right and left eigenvectors of for the eigenvalue . Also, is a cyclic -invariant subspace with as generating vector. Similarly, is a cyclic -invariant subspace with as generating vector. Thus (14) is an explicit rank-one representation of in terms of a right and a left eigenvectors of for the eigenvalue zero. Actually this representation depends on a particular normalization of the vectors which span the cyclic subspaces and . Specifically, . However, we are looking for a more general representation in terms of arbitrary right and left eigenvectors for which such a normalization may fail to hold.
Let us assume that are arbitrary right and left eigenvectors of for the eigenvalue . Then and and since , there are nonzero scalars such that and . Put , and for define
with arbitrary scalars. Using these scalars we define the following triangular matrices
It is plain that and also . Since and are nonsingular matrices for any choice of (because and ), we conclude that and . In addition, for
and
In other words, and are generating vectors of and and and are the given right and left eigenvectors of for the eigenvalue . Now, it follows from , and (14) that
(15) Since ,
But is a lower triangular matrix whose diagonal elements are all equal to . Thus, for , . Since and , as claimed. Now, from (15)
(16)
∎
As a first consequence of Theorem 2.5 we present a generalization of the formula for the eigenvalues of the adjugate matrix (see [18]).
Corollary 2.6.
Let be a nonsingular matrix. Let and let be its partial multiplicities (i.e., the sizes of the Jordan blocks associated to in any Jordan form of in, perhaps, a extension field . Then is an eigenvalue of with as partial multiplicities.
Proof.
Corollary 2.7.
Let , and let be its partial multiplicities. Let be arbitrary right and left eigenvectors of for . Then
(17) |
where is the algebraic multiplicity of and is the Kronecker delta.
Proof.
The following result is an immediate consequence of Corollary 2.7.
Corollary 2.8.
Let and let be its spectrum. Assume that and let and be the algebraic and geometric multiplicities of for the eigenvalue , . Fix and let and be right and left eigenvectors of for . Then
(18) |
The TM formula (1) can be used to provide an easy proof of the so-called eigenvector-eigenvalue identity (see [12, Sec. 2.1]). In fact, under the hypothesis of Theorem 2.1, it follows from (1) that , (see [12, Rem 5]). Hence, recalling that for , is the principal minor of obtained by removing its th row and column,
(19) |
In particular, if is Hermitian, are its eigenvalues and, for , is a unitary right and left eigenvector of for (that is , and ; (recall that we must change transpose by conjugate transpose in the complex case) then
Equivalently, if are the eigenvalues of ,
(20) |
This is the classical eigenvector-eigenvalue identity (see [12, Thm. 1]).
As mentioned in Remark 2.4, if is a field of characteristic zero and then (19) is meaningful if and only if is a simple eigenvalue. If is defective and its geometric multiplicity is bigger than then (19) becomes a trivial identity because, in this case, (item (ii) of Theorem 2.5) and so . However, if is defective and its geometric multiplicity is , then (17) can be used to obtain a generalization of the eigenvector-eigenvalue identity. In fact, one readily gets from (17):
(21) |
where and are the algebraic and geometric multiplicities of , respectively. Moreover, if both and split in then, with the notation of Corollary 2.8, the following identity follows from (18) for the non-repeated eigenvalues of and for :
(22) |
where , , and is the algebraic multiplicity of , and .
In the following section two additional applications will be presented.
3 Two additional consequences of the TM formula
The well-known formula (23) below gives the derivative of a simple eigenvalue of a matrix depending on a (real or complex) parameter. The investigation about the eigenvalue sensitivity of matrices depending on one or several parameters can be traced back to the work of Jacobi ([19]). However a systematic study of the perturbation theory of the eigenvalue problem starts with the books of Rellich (1953), Wilkinson (1965) and Kato (1966), as well as the papers by Lancaster [20], Osborne and Michaelson [27], Fox and Kapoor [14], Crossley and Porter [9] (see also [31] and the references therein). Since then this topic has become classical as evidenced by an extensive literature including books and papers addressed to mathematicians and a broad spectrum of scientist and engineers. In addition to the above early references, a short, and by no means exhaustive, list of books could include [4, p. 463], [24, Ch. 8, Sec. 9], [10, Sec.4.2] or [21, pp. 134-135].
In proving (23), one first must prove, of course, that the eigenvalues smoothly depend on the parameter. It is also a common practice to prove or assume (see [23],[13, Ch. 11, Th. 2] and the referred books), the existence of eigenvectors which depend smoothly on the parameter. It is worth-remarking that in the proof by Lancaster in [20] only the existence of eigenvectors continuously depending on the parameter is required. We propose a simple and alternative proof of (23) where no assumption is made on the right and left eigenvector functions.
Let be the open disc of radius with center . For the following result will be either the field of real numbers or of the complex numbers . Recall that is a left eigenvector of for an eigenvalue if where is the transpose conjugate of . Hence, we will change T by ∗ to include complex vectors in our discussion.
Proposition 3.1.
Let be a square matrix-valued function whose entries are analytic at . Let be a simple eigenvalue of . Then there exist and so that is the unique eigenvalue of with for each . Moreover, is analytic on and
(23) |
where, for , and are arbitrary right and left eigenvector, respectively, of for .
Proof.
Since is a simple root of , by the analytic implicit function theorem, we have, in addition to the first part of the result, that
By the Jacobi formula for the derivative of the determinant and TM formula (1), we have (note that since is a simple eigenvalue, for any right and left eigenvectors and )
and the result follows. ∎
Remark 3.2.
-
(a)
The same conclusion can be drawn in Proposition 3.1 if is a complex or real matrix-valued differentiable function of a real variable. In the first case, we would need a non-standard version of the implicit function theorem like the one in [3, Theorem 2.4]. In the second case the standard implicit function theorem is enough.
-
(b)
It is shown in [2] that the existence of eigenvectors smoothly depending on the parameter can be easily obtained from the properties of the adjugate matrix. In fact, since is a simple eigenvalue of for each , and so by the TM formula, (see Remark 2.4). Now, is a differentiable matrix function of and . Henceforth, all nonzero columns of , which are all proportional, are (right and left) eigenvectors of for .
The second application is related to the problem of characterizing the admissible eigenstructures and, more generally, the similarity orbits of the rank-one updated matrices. There is a vast literature on this problem. A non-exhaustive list of publications is [32, 29, 34, 26, 6, 25, 8, 5] and the references therein. It is a consequence of Theorem 2 in [32] that if is an eigenvalue of with geometric multiplicity and then may or may not be an eigenvalue of . It is then proved in [25, Th. 2.3] that in the complex case, generically, is not an eigenvalue of . That is to say, there is a Zariski open set such that for all , is not an eigenvalue of . With the help of the TM formula we can be a little more precise about the set . Form now on, will be again an arbitrary field.
Proposition 3.3.
Let and let be an eigenvalue of in, perhaps, an extension field . Assume that the geometric multiplicity of is and its algebraic multiplicity is . Let be right and left eigenvectors of for . If then is an eigenvalue of if and only if or .
Proof.
Let . Then . Taking into account that is invertible in , where the field of rational functions, and using the formula of the determinant of updated rank-one matrices, we get
In particular,
(24) |
It follows from (17) that (recall that )
Since , the Proposition follows. ∎
Remark 3.4.
Note that, by (24) and item (ii) of Theorem 2.5, if the geometric multiplicity of as eigenvalue of is then and so, is necessarily an eigenvalue of . This is an easy consequence of the interlacing inequalities of [32, Th. 2]. However, proving that those interlacing inequalities are necessary conditions that the invariant polynomials of and must satisfy is by no means a trivial matter.
The eigenvalues of rank-one updated matrices are at the core of the divide and conquer algorithm to compute the eigenvalues of real symmetric or complex hermitian matrices (see, for example, [11, Sec. 5.3.3], [30, Sec. 2.1]). At each step of the algorithm a diagonal matrix and a vector are given such that the eigenvalues and eigenvectors of are to be computed. In order the algorithm to run smoothly, it is required, among other things, that the diagonal elements of are all distinct. Thus, a so-called deflation process must be carried out. This amounts to check at each step the presence of repeated eigenvalues and, if so, remove and save them. The result that follows is related to the problem of detecting repeated eigenvalues but for much more general matrices over arbitrary fields.
Proposition 3.5.
Let with , . Let and split into blocks such that , . Assume also that the eigenvalues of and have geometric multiplicity equal to and . Then
Proof.
.- If then , as eigenvalue of , has geometric multiplicity . By Remark 3.4, . Assume that but . Let us see that this assumption leads to a contradiction. Let be a right and a left eigenvectors of , respectively. Then and are right and left eigenvectors of , respectively, for . Since , the geometric multiplicity of as eigenvalue of is . Then, by Proposition 3.3, or because . Let us assume that , on the contrary we would proceed similarly with . If we put and , with , then and . It follows from Proposition 3.3 that , contradicting the hypothesis . That is proved similarly. ∎
Remark 3.6.
- (i)
- (ii)
References
- [1] A. Amparan, S. Marcaida, and I. Zaballa. On the structure invariants of proper rational matrices with prescribed finite poles. Linear and Multilinear Algebra, 61(11):1464–1486, 2013.
- [2] A. L. Andrew, K.-W. E. Chu, and P. Lancaster. Derivatives of eigenvalues and eigenvectors of matrix functions. SIAM J. Matrix Anal. Appl., 14(4):903–926, 1993.
- [3] M. S. Ashbaugh and E. M. Harrell II. Perturbation theory for shape resonances and large barrier potentials. Comm. Math. Phys., 83(2):151–170, 1982.
- [4] F. V. Atkinson. Discrete and continuous boundary problems, volume 8 of Mathematics in Science and Engineering. Academic Press, New York-London, 1964.
- [5] I. Baragaña. The number of distinct eigenvalues of a regular pencil and of a square matrix after rank perturbation. Linear Algebra Appl., 588:101–121, 2020.
- [6] M. A. Beitia, I. de Hoyos, and I. Zaballa. The change of the Jordan structure under one row perturbations. Linear Algebra Appl., 401:119 – 134, 2005.
- [7] W. C. Brown. Matrices over Commutative Rings. Marcel Dekker Inc., New York, 1993.
- [8] R. Bru, R. Cantó, and A. M. Urbano. Eigenstructure of rank one updated matrices. Linear Algebra Appl., 485:372–391, 2015.
- [9] T. R. Crossley and B. Porter. Eigenvalue and eigenvector sensitivities in linear system theory. Int. J. Control, 10:163–170, 1969.
- [10] Hinrichsen D. and Pritchard A. J. Mathematical System Theory I. Modelling, State Space Analysis, Stability and Robustness. Springer, Berlin, 2005.
- [11] J. W. Demmel. Applied Numerical Linear Algebra. SIAM, Philadelphia, 1997.
- [12] P. B. Denton, S. J. Parke, T. Tao, and X. Zhang. Eigenvectors from eigenvalues: a survey of a basic identity in linear algebra. arXiv:1908.03795, 2020.
- [13] L. C. Evans. Partial differential equations, volume 19 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, second edition, 2010.
- [14] R. L. Fox and M. P. Kapoor. Rate of change of eigenvalues and eigenvectors. AIAA J., 6:2426–2429, 1968.
- [15] F. R. Gantmacher. The Theory of Matrices. AMS Chelsea Publishing, Providence, Rhode Island, 1988.
- [16] R. Godement. Cours d’algèbre. Hermann Éditeurs, Paris, 2005.
- [17] D. Grinberg. Eigenvectors from eigenvalues: a survey of a basic identity in linear algebra — what’s new. https://terrytao.wordpress.com/2019/12/03/eigenvectors-from-eigenvalues-a-survey-of-a-basic-identity-in-linear-algebra/#comment-531597, 2019.
- [18] R. D. Hill and E. E. Underwood. On the matrix adjoint (adjugate). SIAM J. Algebraic Discrete Methods, 6(4):731–737, 1985.
- [19] C. G. J. Jacobi. Über ein leichtes verfahren die in der theorie der säcularstörungen vorkommenden gleichungen numerisch aufzulösen. J. für die Reine und Angew. Math., 1846(30):51–94, 1846.
- [20] P. Lancaster. On eigenvalues of matrices dependent on a parameter. Numer. Math., 6:377–387, 1964.
- [21] P. D. Lax. Linear Algebra and its Applications. Pure and Applied Mathematics (Hoboken). Wiley-Interscience [John Wiley & Sons], Hoboken, NJ, second edition, 2007.
- [22] D. S. Mackey, N. Mackey, C. Mehl, and V. Mehrmann. Vector spaces of linearizations for matrix polynomials. SIAM J. Matrix Anal. Appl., 28(4):971–1004, 2006.
- [23] J. R. Magnus. On differentiating eigenvalues and eigenvectors. Econometric Theory, 1:179–191, 1985.
- [24] J. R. Magnus and H. Neudecker. Matrix Differential Calculus with Applications in Statistics and Econometrics. John Wiley & Sons, Chichester, 1988.
- [25] C. Mehl, V. Mehrmann, A. C. M. Ran, and L. Rodman. Eigenvalue perturbation theory of classes of structured matrices under generic structured rank one perturbations. Linear Algebra Appl., 435(3):687–716, 2011.
- [26] J. Moro and F. M. Dopico. Low rank perturbation of Jordan structure. SIAM J. Matrix Anal. Appl., 25(2):495–506, 2003.
- [27] M. R. Osborne and S. Michaelson. The numerical solution of eigenvalue problems in which the eigenvalue appears nonlinearly, with an application to differential equations. Computer J., 7:66–71, 1964.
- [28] D. S. Scott. How to make the Lanczos algorithm converge slowly. Math. Comp., 33:239–247, 1979.
- [29] F. C. Silva. The rank of the difference of matrices with prescribed similarity classes. Linear and Multilinear Algebra, 24(1):51–58, 1988.
- [30] G. W. Stewart. Matrix Algorithms, Volume II: Eigensystems. SIAM, Philadelphia, 2001.
- [31] J. G. Su. Multiple eigenvalue sensitivity analysis. Linear Algebra Appl., 137(4):183–211, 1990.
- [32] R. C. Thompson. Invariant factors under rank one perturbations. Canad. J. Math., 32(1):240–245, 1980.
- [33] R. C. Thompson and P. McEnteggert. Principal submatrices. II: The upper and lower quadratic inequalities. Linear Algebra Appl., 1:211–243, 1968.
- [34] I. Zaballa. Pole assignment and additive perturbations of fixed rank. SIAM J. Matrix Anal. Appl., 12(1):16–23, 1991.