Eigenvalues and Eigenvectors

Let A be an n´ n matrix over a field F . We recall that a scalar l Î F is said to be an eigenvalue (characteristic value, or a latent root) of A, if there exists a nonzero vector x such that Ax = l x, and that such an x is called an eigen-vector (characteristic vector, or a latent vector) of A corresponding to the eigenvalue l and that the pair (l , x) is called an eigen-pair of A. If l is an eigenvalue of A, the equation: (l I-A)x = 0, has a non-trivial (non-zero) solution and conversely. Thus, this being a homogeneous equation, it follows that l is an eigenvalue of A iff |l I-A| = 0. The expression

fA(x) = |xI-A|

is a monic (the coefficient of the highest power of x in it is 1) polynomial in x of degree n. It is known as the characteristic polynomial of A. Thus l is an eigen value of A iff l is a zero (or root) of the characteristic polynomial fA(x) in F. The equation

fA(x) = 0,

is called the characteristic equation of A. Note that the coefficients of the characteristic polynomial are elements of field F under consideration. If F is an algebraically closed field (for instance, ) we know that a polynomial of precise degree n has precisely n roots counted after multiplicities. Thus, for instance, a complex n´ n matrix has n eigen values counted after multiplicities.

Note that l = 0 is an eigenvalue of A iff |A| = 0, i.e., A is a singular matrix. If an eigenvalue l of A is known, the corresponding eigenvector(s) may be obtained by elementary row operations (ero's) performed on the matrix part of the augmented matrix [A-l I|0].

Example: Computation of eigen-values and eigen-vectors of a 2´ 2 complex matrix:

The algebraic multiplicity of an eigenvalue l of A is the highest k such that (x-l )k is a factor of fA(x). The geometric multiplicity of an eigenvalue l of a matrix A is the maximum number of linearly independent eigen vectors x of A associated with the eigenvalue l , which is the same as the dimension of the eigenspace of A associated with the eigenvalue l consisting of all x such that Ax = l x. This eigenspace is the same as N(A-l I).

The geometric multiplicity of the eigenvalue 0 of is 2, while that of the eigenvalue 0 of is 1. The algebraic multiplicities of the eigenvalue 0 of both these matrices equal 2.

A collection of eigenvalues of A in which an eigenvalue is repeated according to is algebraic multiplicity is called the spectrum of A, and is denoted by s (A). Clearly, if s (A) = {l , l , ... , l }, fA(x) = P (x-l ). Thus, e.g., for an identity matrix: s (I) = {1, 1, ... , 1}, for a null or zero matrix s (0) = {0, 0, …, 0},

s ( ) = {1, 3}, s ( ) = {1, 1}, s ( ) = {2+Ö 5, 2-Ö 5}, and, s ( ) = {0, 4}.

A matrix B is said to be similar to a matrix A (written: B ~ A) if there exists a non-singular matrix C such that C-1AC = B. This relation is interpreted as follows : if instead of the standard basis of the unit vectors e1 = (1, 0, 0, ... , 0)¢ , e2 = (0, 1, 0, ... , 0)¢ , ... , en = (0, 0, 0, ... , 1)¢ , we choose the basis of Fn to consist of the column vectors of the matrix C, then the matix of the transformation A turns out to be B. It follows that A ~ A (reflexivity); B ~ A implies A ~ B (symmetry); and A ~ B and B ~ C implies A ~ C (transitivity). The similarity relation therefore is an equivalence relation. Consequently the set of all n´ n matrices over a field F can be decomposed into (written as a union of) mutually exhaustive equivalence classes of all matrices which are similar to each other.

 

Companion Matrices and Characteristic Polynomial

Each of the following four matrices Q1, Q2, Q3, Q4 is referred to as a companion matrix of the monic polynomial q(x) = xn + an-1xn-1 + ... + a1x + a0. It is easy to verify that the characteristic polynomial of a companion matrix of a monic polynomial q(x) is the monic poynomial q(x) itself (e.g., for Q4, multiply i-th row of xI-Q4 by x and add to row i-1, i = n, n-1, … , 2. The (1, n)-th element is now q(x), the rest of the elements in the first row being zero and expanding the determinant by the first row gives | xI-Q4| = q(x).

,

,

,

.

Method of Danilevsky for Characteristic Polynomial

A matrix transformation of type: A Î F n´ n ® C-1AC, (C non-singular) is called a similarity transform. The matrix C-1AC is said to be similar to A. The characteristic polynomial remains invariant under a similarity transform, i.e., similar matrices have the same characteristic polynomial:

.

The characteristic polynomial of each of the above companion matrices being the polynomial q(x), to find the characteristic polynomial of A, Danilevsky's method applies a succession of similarity transforms to A to reduce it to a companion matrix form Q4. Let

.

(1) X ¬ A (i.e., copy A in X);

(2) k ¬ 2;

(3) If xj k-1 = 0, for all j ³ k, X has the reducible form ,

in which Q is already in a companion matrix form and the characteristic polynomial of X equals the product of those of Q and Z. For the latter, the method could be separately applied to Z. #

(4) If xj k-1 ¹ 0, for some j ³ k, interchange the j-th row with the k-th row and the j-th column with the k-th column (a similarity transform) after which xk k-1 ¹ 0. Next:

(i) divide the k-th row of X by xk k-1 and multiply the k-th column of X by xk k-1 (it constitutes a similarity transform). X looks like:

(ii) for all i ¹ k, from the i-th row subtract xi1 times the k-th row and to the k-th column add xi1 times the i-th column (it constitutes a similarity transform). X looks like:

(5) If k < n, k ¬ k+1. Go to step (3).

(6) If k = n, X is in a companion form Q4. #

 

PROBLEMS

1. Find the characteristic polynomials (both by expanding the determinant |xI-A| and by the Danilevsky's method), eigenvalues, the associated algebraic and geometric multiplicities and the eigenspaces of the following matrices:

2. Construct 2´ 2 and 33 matrices that, if possible, have (a) no eigenvalues, (b) all distinct eigenvalues, (c) all coincident eigenvalues, (d) some but not all distinct eigenvalues, (e) a full set of eigenvectors, (f) only one linearly independent eigenvector, and (g) every non-zero vector as an eigenvector.

3. Characterize 2´ 2 matrices to each of which there corresponds another matrix differing from it in only one element but having the same spectrum.

4. If v1, v2, …, vn are the eigenvectors associated with the respective eigenvalues l 1, l 2, … , l n of an n´ n matrix A and none of the l 's is zero and b = S 1£ k£ n ckvk, show that the solution of the system Ax = b is given by

x = S 1£ k£ n (ck/l k)vk.

5. When is it possible to write every vector in Fn as a linear combination of eigenvectors of a matrix A? Give an example of a matrix A for which it is possible. Give another example for which it is not.

6. If the geometric multiplicity of an eigenvalue l of an n´ n matrix A is n-k+1, prove that every k´ k principal submatrix (i.e., a submatrix whose diagonal lies along the diagonal of the matrix) of A has l as one of its eigenvalues.

7. Let E be n´ k, W n´ (n-k) and A an n´ n matrix over a field F. If the n´ n partitioned matrix C = (E!W) is non-singular and the columns of E are eigen-vectors of A show that

C-1AC = ,

where D is a diagonal matrix, 0 is the (n-k)´ k zero matrix and X and Y are matrix blocks satisfying

AW = (E|W) = EX+WY.

8. Show that similar matrices have the same eigenvalues, each of which has the same algebraic and the same geometric multiplicities.

9. Verify that similarity is an equivalence relation, i.e., it is reflexive (A ~ A), symmetric (B ~ A Þ A ~ B) and transitive (A ~ B, B ~ C Þ A ~ C).

10. Prove that the geometric multiplicity of an eigenvalue cannot exceed its algebraic multiplicity.

11. Verify that the transformations mentioned in the Danilevsky's method indeed constitute similarity transforms.

12. Verify that the matrices Q1 - Q4 are similar and that their characteristic polynomials are equal to q(x).

 

Matrices with a Full-Set of Eigen-Vectors

An n´ n matrix is said to have a full set of eigenvectors if there exist n linearly independent vectors v1, v2, ... , vn, which at the same time are also eigenvectors of A. In other words an n´ n matrix A has a full set of eigenvectors iff there exists a basis of Fn, consisting entirely of the eigenvectors of A. If the associated eigenvalues are l 1, l 2, … , l n and C denotes the matrix whose j-th column vector coincides with the eigenvector vj, we have AC = C diag (l 1, l 2, … , l n). Further, since the vectors vj 's are linearly independent, the matrix C is non-singular and we may write the above equation as C-1AC = D, where D = diag (l 1, l 2, … , l n). Conversely if C-1AC = D, the n column vectors of the matrix C constitute n linearly independent eigenvectors of the matrix A corresponding to the eigenvalues l 1, l 2, … , l n that are the diagonal entries of the diagonal matrix D. Thus we have proved the following:

Theorem (Diagonalization). There exists a non-singular C and a diagonal D such that C-1AC = D iff A has a full set of eigenvectors.

A precise characterization of diagonalizable matrices involves the notion of the minimal polynomial of a matrix and will appear later. For the present, to determine if A is diagonalizable one has to first determine the eigenvalues of A and then if some of them are multiple, to check if there exist a corresponding number of linearly independent eigenvectors.

 

PROBLEMS

1. If A is diagonalizable and its distinct eigenvalues are c1, c2, ... , ck, show that (c1I-A)(c2I-A) ... (ckI-A) = 0.

2. Find a matrix A with distinct eigenvalues c1, c2, ... , ck, for which (c1I-A)(c2I-A) ... (ckI-A) ¹ 0.

3. Prove that eigenvectors corresponding to distinct eigenvalues of a matrix are linearly independent.

4. Relate fA(x) and fB(x) if B = AT.

5. Relate fA(x) and fB(x) if B = A = (A¢ )* , a complex matrix.

6. If A is a complex matrix relate fA(x) and fB(x) if B = A* .

7. If P(x) is polynomial function and s (A) = {l 1, l 2, ..., l n}, when does there hold s (P(A)) = {P(l 1), P(l 2), ..., P(l n)}?

8. Prove that a hermitian matrix has only real eigenvalues?

9. Prove that the eigenvalues of a unitary matrix are complex numbers of magnitude 1.

10. Prove that eigenvalues of a triangular matrix are its diagonal elements.

11. Which of the following matrices is diagonalizable over R:

.

12. There are 24 distinct 2´ 2 matrices with distinct entries taking the values 1, 2, 3 and 4, only. Show that: (a) all of these are diagonalizable over C ; (b) all of these are diagonalizable over R ; and (c) only those of these are diagonalizable over Q that have 5 as an eigenvalue.

13. Can you show that given any n´ n complex (real) matrix A and a positive number e , there exists another complex (real) matrix A having a full set of eigenvectors such that no entry of A differs from that of the corresponding entry of A by more than e ?

14. If an invertible complex matrix A equals its inverse A-1, prove that it has a full set of eigenvectors. Is the result true if F ¹ C .

 

The Cayley-Hamilton Theorem

If A is an n´ n matrix over a field F and k = n2, the matrices I, A, A2, ... , Ak (being n2+1 in number) are linearly dependent in Fn´ n . Hence there exist scalars c0, c1, c2, ... , ck, not all zero (termed non-trivial scalars) such that c0I + c1A + c2A2 + ... + ckAk = 0. Thus with P(x) = c0 + c1x + c2x2 + ... + ckxk, we have P(A) = 0, i.e., any n´ n matrix A satisfies a non-trivial polynomial equation of degree £ n2. The Cayley-Hamilton theorem asserts that the familiar characteristic polynomial fA(x) itself kills (annihilates) A:

Theorem (Cayley-Hamilton). fA(A) = 0, i.e., A satisfies its own characteristic equation.

Proof: Consider the linear polynomial matrix A(x) = xI-A (its entries are linear polynomials in x). Let Cij(x) denote its (i, j)-th cofactor, which for each pair (i, j) is a matrix polynomial of degree n-1, at most. Using cofactor expansion identities we have d ki fA(x) = S 1£ j£ n Ckj(x)(d ijx-aij), where the elements are from the i-th row whereas the cofactors are corresponding to the k-th row. Evaluating the two polynomial sides at the matrix A, we have d ki fA(A) = S 1£ j£ n Ckj(A)(d ijA-aijI). With ej 's as the standard unit vectors, we therefore have

fA (A)ek = S 1£ i£ n d ki fA(A)ei = S 1£ i£ n S 1£ j£ n Ckj(A)(d ijA-aijI)ei

= S 1£ j£ n Ckj(A) S 1£ i£ n (d ijA-aijI)ei = S 1£ j£ n Ckj(A)(Aj -Aj) = 0.

Hence, for any vector x, fA(A)x = fA(A) S 1£ k£ n xkek = S 1£ k£ n xkfA(A)ek = 0, and consequently f (A) = 0. #

Another simple proof: of the Cayley-Hamilton theorem is as follows: Let B(x) denote the classical adjoint (transpose of the matrix of cofactors) of A(x) = xI-A. If A is n´ n, the entries in B(x) are polynomials in x of a degree not exceeding n-1. Hence we can write

B(x) = B0 + xB1 + x2B2 + ... + xn-1Bn-1,

for certain matrices Bi 's not involving x. Let det (xI-A) = a0 + xa1 + x2a2 + ... + xnan, (an = 1). Comparing the coefficients of powers of x in (xI-A)B(x) = det (xI-A) I, we get

- AB0 = a0 I,

B0 - AB1 = a1 I,

B1 - AB2 = a2 I,

Bn-2 - ABn-1 = an-1 I,

Bn-1 = an I

Premultiplying both sides of these equations by I, A, A2, ... , An, respectively, and adding we have

0 = a0I + a1A + a2A2 + ... + anAn = fA (A),

completing the proof. #

Corollary. Any polynomial in an n´ n matrix A can be written as a polynomial in A of degree n-1 at most.

Proof: If P is a polynomial, write P = fAQ + R, where R, the residual (remainder) polynomial is of degree < deg fA = n. Hence P(A) = R(A) and deg R £ n-1. #

The collection of all polynomials P such that P(A) = 0, is non-empty. Hence, as the degree of a non-zero polynomial is a non-negative integer valued funciton, it follows that there exists a monic polynomial pA(x) of least degree such that pA(A) = 0. The polynomial pA(x) is known as the minimal polynomial of A.The minimal polynomial pA of a matrix A is unique and divides every polynomial P such that P(A) = 0. For, writing P = p Q + R, and evaluating both sides at x = A we get R(A) = 0, but if R ¹ 0, dividing it by its coefficient of highest degree, we have a monic polynomial of a smaller degree than p , a contradiction. In particular, the minimal polynomial divides the characteristic polynomial. Indeed, pA(x) | fA(x) is a rephrasing of the Cayley-Hamilton theorem.

 

PROBLEMS

1. If AB and BA make sense then I-AB and I-BA are singular or regular together.

2. AB and BA have the same distinct non-zero eigenvalues and have the same eigenvalues if A and B are square, in which case also fAB = fBA.

3. |A| = l 1 l 2l n and tr (A) = l 1 + l 2 + … + l n, where l i 's are eigenvalues of A counted after multiplicities.

4. fA and pA have the same distinct roots.

5. If l is an eigenvalue of A, P(l ) is an eigenvalue of P(A). Conversely, if m is an eigenvalue of P(A), m = P(l ) for some eigenvalue l of A.

6. The polynomial whose roots are squares of the roots of P(x) = ax2 + bx + c is Q(x) = a2x2 + (2ac-b2)x + c2.

7. If 1 is not an eigenvalue of n´ n matrices A and B, (A+B)x = x iff (I-B)-1A(I-A)-1Bx =x. [K.C. Deomurari]

8. If A is an n´ n matrix with distinct eigenvalues c1, c2, … , ck, show that [(c1I-A)(c2I-A) ... (ckI-A)]n-1 = 0.

 

Triangulization and Unitary Diagonalization of a Matrix

The interest in the existence and construction of possible similar matrices of a particular type is due to several problems for similar matrices being related: e.g., if A ~ B = C-1AC, Bx = l x Þ ACx = l Cx; By = C-1b Þ A(Cy) = b. Thus, if B has a relatively simpler structure, it may enable an easier solution of problems involving A. The case when B could be managed to be a triangular matrix is discussed below.

Theorem. If A is a square n´ n matrix over a field F, there exists a non-singular matrix C such that C-1AC = T, a triangular (upper) matrix iff the characteristic polynomial fA(x) = det(xI-A) of A is a product of linear factors.

Proof: Let l be any of the eigenvalues of A and let x be a corresponding eigen-vector. Choose {x, x2, x3, … , xn} to be an ordered basis of F n of which the first member is the vector x. Arrange these basis members as columns of a matrix S = (x | X), where X is an n´ (n-1) block. Then S is a non-singular matrix and

S-1AS = S-1 [l x | AX] = ,

say, where the eigenvalues of A other than l are precisely those of B counting multiplicities (algebraic). Note that the characteristic polynomial of B is a product of linear factors, and so using induction we may assume that B is triangulable, i.e., for some matrix R, R-1BR = S, an upper triangular matrix. Then

,

an upper triangular matrix. Thus A has been triangulaized by the matrix

.

Since for 1´ 1 matrices the result is trivial, we have proved it for matrices of all sizes. The converse statement is trivial, since if A is triangulable, its characteristic polynomial is the same as that of the similar matrix T, which being upper triangular has its characteristic polynomial as P (x-tii), a product of linear factors. #

It follows from the above proof that along the diagonal of T, any permutation of the eigenvalues of A could be managed.

The triangulization C-1AC = T could be equivalently written as AC = CT, which says that A maps a k-th column of C into a linear combination of the first k columns of C, k = 1, 2, … , n. Conversely, if C is non-singular and A does the above, then A is triagulized by C. In particular, if C triangulizes A and E is any matrix whose k-th column is a linear combination of the first k columns of C, with a non-zero coefficient of the k-th column, then E also triangulizes A.

 

Schur's Lemma and the Spectral Theorem

Consider the case when the field F is the complex field C. Since C is algebraically closed, any polynomial over C is a product of linear factors. Hence any square matrix A over C is triangulabe by some C. If we apply the Gram-Schmidt orthonormalization process to the columns of C and arrange the resulting orthonormalized vectors as the respective columns of E, then E is of the above type and so it also triangulizes A. But E, in addition is unitary, i.e., E* E = I. Thus we have proved the following:

Schur's Lemma: Given any square complex matrix matrix A, there exists a unitary matrix U such that U* AU = T, a triangular matrix, i.e., any complex matrix can be unitarily triangulized.

As before, the ordering of the eigen-values of A along the diagonal of T could be chosen at will.

As a corollary, we may deduce the following important result known as unitary diagonalization, or, spectral theorem for normal matrices. Recall that a complex matrix A is called normal if A* A = AA* .

Theorem (Spectral). If A is a square complex matrix, there exists a unitary matrix U such that U* AU = D, a diagonal matrix, iff A is normal.

Proof: If A is unitarily diagonalizable, A = UDU* , and so AA* = UDD* U = UD* DU = A* A, i.e., A is normal. Conversely, let A be normal. By Schur's lemma, A = UTU* and so the normalcy implies that UTT* U* = UT* TU* . Then TT* = T* T. Since T is upper triangular, its elements below the main diagonal are zero, i.e., tij = 0, if i > j. Equating the first diagonal elements on both sides of the normalcy equation for T we get |t11|2 + |t12|2 + |t13|2 + ... + |t1n|2 = |t11|2, so that all strictly upper-triangular terms in the first row are zero. Next equating the second diagonal elements |t22|2 + |t23|2 + ... + |t2n|2 = |t22|2, so that all elements in the strictly upper-triangular part in the second row are zero. Comparing the third elements then gives |t33|2 + ... + |t3n|2 = |t33|2, so that all strictly upper-triangular terms in the third row are zero, and so on till |tn-1 n-1|2 + |tn-1 n |2 = |tn-1 n-1|2, which says that tn-1 n = 0. Thus we find that tij = 0 whenever i ¹ j, i.e., T is a diagonal matrix. This completes the proof of the unitary diagonalization theorem. #

The above proof implies that if a matrix is normal its unitary triangular form is automatically a diagonal form, i.e., if one tries to unitarily triangulize it, one actually ends up in its unitary diagonalization..

A matrix A which is hermitian (i.e., A* = A) is automatically normal and so it is unitarily diagonalizable. Since the eigenvalues of a hermitian matrix are real, the corresponding diagonal form is real. Since a real symmetric matrix A (i.e., A¢ = A, and A is with real entries) is hermitian, regarding it as a matrix over C we find that its characteristic polynomial is a product of real linear factors. But its characteristic polynomial over R is the same as over C. Hence regarded as a matrix over R its characteristic polynomial is a product of linear factors. Since to a real eigenvalue of a real matrix there always corresponds a real eigen-vector, repeating everything in the above for this special case of a real symmetric A with F as R , we obtain the following:

Spectral Theorem for Real Symmetric Matrices. If A is a real matrix, there exists a real orthogonal matrix U (i.e., a real matrix U such that U¢ U = I) such that U¢ AU = D, a diagonal matrix, iff A is symmetric.

Note a real hermitian matrix is just a real symmetric matrix, and, a real orthogonal matrix a real unitary matrix.

 

PROBLEMS

1. For an n´ n A, (A+A* )/2 and (A-A* )/2 are respectively known as the hermitian and skew-hermitian parts of A; for a real matrix these respectively become the symmetric and skew-symmetric parts. Prove that A is normal iff its hermitian and skew-hermitian parts commute. Deduce that a hermitian, or a skew-hermitian matrix is normal.

2. When is a triangular matrix unitarily diagonalizable?

3. If a matrix equals its skew-hermitian part show that the matrix can have only purely imaginary eigenvalues. Does the converse hold?

4. Inter-relate the hermitian and skew-hermitian parts of A and iA. [i = Ö (-1)].

5. Prove that a normal matrix is hermitian iff its eigenvalues are real.

6. If a real matrix A is unitarily diagonalizable and its eigenvalues are real then it is orthogonally diagonalizable, i.e., there exists an orthogonal matrix C (i.e., CTC = I, C real) such that C-1AC is diagonal.

7. If a matrix A is orthogonally diagonalizable then its eigenvalues are real iff A is real

8. A real matrix A is orthogonally diagonalizable iff its eigenvalues are real and it is normal.

9. Let A be a real matrix with real eigenvalues. Let C be a non-singular complex matrix such that C-1AC is diagonal. Then there is a real matrix G such that G-1AG = C-1AC.

10. If l 1, … , l n are the eigenvalues of a complex n´ n matrix A, prove that |l 1|2 … + |l n|2 £ S 1£ i£ n S 1£ j£ n |aij|2, and that the equality holds iff A is normal.

 

The Equation: AX - XB = C

We use the result on triangulization to infer about a unique solvability of the matrix system AX - XB = C of equations, in which A is m´ m, X m´ n, B n´ n and C an m´ n matrix. The size constraints make the equation meaningful. The set of eigenvalues of a matrix A is denoted by s (A) and is called the spectrum of A.

Theorem. A complex system AX - XB = C has a unique solution iff s (A)Ç s (B) = f , the null set.

Proof. If n = 1, the system reduces to (A - b11I)X = C. Here b11 is also the eigenvalue of B. If b11 Î s (A), the system either has no solution or an infinity of solutions, verifying the assertion. Similar is the case when m = 1, when the system reduces to (B¢ - a11I)X¢ = C¢ . Hence it is enough to assume the result for the sizes m-1 and n-1 and prove it for m, n. Since the field of complex numbers is algebraically closed, all complex matrices are triangulable. Let Y triangulize A and Z, B. Then we can write

and putting

,

the equation AX - XB = C can be rewritten as

,

which is

.

In the above since the order of the elements along the diagonal of a triangular form is at our wish, we may assume that if s (A)Ç s (B) ¹ f , then a1 Î s (B) and b1 Î s (A).

If s (A)Ç s (B) = f , this system has a unique solution when looked at its rearrangement in N (­ ] ­ ) fashion:

TA z - b1 z = d, (solvable if b1 Ï s (TA) Ì s (A)),

(a1 - b1)x1 + a¢ z = e1, (solvable if a1 ¹ b1 ),

TA X1 - X1 TB - zb¢ = C1, (solvable if a1 Ï s (TB) Ì s (B)),

a1 x¢ + a¢ X1 - x1 b¢ - x¢ TB = c¢ , (solvable if s (TA)Ç s (A) = f ).

However, if f ¹ s (A)Ç s (B) э a1, and the system has a solution, then any nonzero solution of x¢ (a1I - TB) = 0, i.e., of (a1I - TB¢ )x = 0, can be added to the "x" block to give rise to more than one solution. Hence either the system has no solution or more than one solution, completing the proof. #

Remarks

1. Since only the triangulation of A and B has been used in the above proof, the result is valid for an arbitrary field over which the characteristic polynomials of A and B completely split (i.e., into products of linear factors). In particular the result is valid for all such systems over an algebraically closed field. Moreover, in the above, for infinite fields more than one solution means an infinity of solutions.

2. An alternate proof of the above result may be based on an an easily verifiable fact that a polynomial in a triangulable matrix is singular iff the polynomial vanishes at least for one eigenvalue of the matrix.

Consider AX - XB = 0. If consistent Þ A2X = A(AX) = (AX)B = XBB = XB2, A3X = XB3, and so on …, so that for any polynomial P, P(A)X = XP(B). In the non-uniqueness case X ¹ 0, and so we have fB(A)X = 0, contradicting s (A)Ç s (B) = f . Conversely, if s (A)Ç s (B) ¹ f , let (l , x) and (l , y) be eigenpairs of A and B¢ respectively. Then X = xy¢ is a non-trivial solution of AX - XB = 0. #

3. If A and B are not necessarily triangulable, a generalization of the above result is as follows (a formal proof is left as an exercise for the reader):

Theorem. The matrix system AX - XB = C over a filed F is uniquely solvable iff (fA, fB), the HCF of the characteristic polynomials of A and B, equals unity.

 

PROBLEMS

1. Give an example of a real system AX-XB = C, which has a complex solution matrix X.

2. Solve, if possible, the real system

.

Is the solution unique?

3. Characterize a, b, c and d for which the real system

,

possesses a real solution? Is the solution unique?

4. If the real parts of the eigenvalues of a complex matrix A are all positive, show that the system AX-XA* = C, has a unique solution for every complex C.

 

Circulants and the Discrete Fourier Transform (DFT)

A complex matrix of the form

is called a circulant. Note that the subsequent rows of this matrix are obtained by once circularly shifting the entries of the previous rows to the right.

The discrete Fourier transform (DFT) matrix of order n is defined as F = (fjk)1£ j, , k£ n = ( n-1/2 w -(j-1) ( k-1) ), where w is the n-th root of unity, w = exp(2p i/n) = exp(2p Ö (-1)/n). Note that F is symmetric, and for p, q = 1, 2, ... , n,

,

so that the DFT matrix F is a unitary matrix. Here we have used that for p ¹ q,

.

 

PROBLEMS

1. Show that the column vectors of the DFT matrix form a basis of C n that diagonalizes every circulant.

2. Prove that a circulant is a normal matrix that commutes with every other circulant.

3. Show that the set of all circulants is closed under the operations of scalar multiplication, addition and matrix multiplication and consequently forms an algebra over C .

4. If a matrix A commutes with every circulant C, show that A is itself a circulant.

5. Prove that inverse of a circulant is a circulant.

 

DFT of a Vector Signal and Solution of a Circulant System

Observe that with w j = w j , j = 0, 1, ... , n-1,

,

and CF* = F* L , where L = diag (l 0, l 1, … , l n-1), and

l j = S 0£ k£ n-1 ck w jk, j = 0, 1, ... , n-1.

Hence C = UL U*, (the spectral decomposition of a circulant) where U = F* is unitary. If L is non-singular, i.e. if C is non-singular then the solution of the circulant system

Cx = b

is given by

x = F* L -1 Fb = IDFT[DFT(b)/DFT(c)],

where the division is in the pointwise sense,

DFT º Ö n F,

IDFT º (1/Ö n) F* ,

and c denotes the first column vector

(c0 , cn-1 , … , c2 , c1)¢

of C, as l j = c0 w j0 + c1 w j1 + … + cn-2 w jn-2 + cn-1 w jn-1 = c0 w j0 + c1 w j-1 + … + cn-2 w j-(n-2) + cn-1 w j-(n-1). #

 

PROBLEMS

1. Determine the eigenvalues, range and the null space of the n´ n matrix

2. For a given circulant C such that |C| = 0, use DFT to characterize the vectors b for which Cx = b has a solution.

 

l -Matrices and Similarity

A square matrix A(l ) = (aij(l )) where the entries aij(l ) are polynomials in l over a certain field F (i.e., aij(l ) Î F[l ]) is called a l -matrix. We can write A(l ) = Ap l p + Ap-1 l p-1 + ... + A1 l + A0 where A0, A1, ... , Ap are square matrices with scalar entries from F and Bl k is to be interpreted as l kB (i.e., Bl k = (bijl k) = (l kbij), B = (bij)). A l -matrix is also called a matric polynomial in the indeterminate scalar l .

Thus A(l ) could be regarded as a polynomial in l with coefficients as matrices A0, A1, ... , Ap Î Fn´ n. If Ap ¹ 0, A(l ) is called a l -matrix of degree p (i.e., p = deg A(l )). A(l ) is called proper if the highest degree coefficient matrix Ap is non-singular. A(l) is called non-singular if F[l ] э det A(l ) ¹ 0. As a convention, the degree of the zero l -matrix is taken to be -¥ , i.e., deg (0) = -¥ . The rank of a l -matrix A(l ) is simply the rank of A(l ) regarded as a matrix over the field QF[l ].

If det A(l ) is a non-zero scalar, i.e., an element of F, there exists a l -matrix B(l ) such that A(l )B(l ) = B(l )A(l ) = I, the identity matrix. Here bij(l ) = (det A(l ))-1 Cji(l ), i, j = 1, 2, ... , n where Cji(l ) is the (j, i)-th cofactor in A(l ). Thus if 0 ¹ det A(l ) is a scalar, (A(l ))-1 is itself a l -matrix; whereas in general the entries in (A(l ))-1 if det A(l ) ¹ 0, are elements of QF[l ], the field of fractions (rational functions) of F in the indeterminate l . Note that deg A(l ) = maxi, j {deg aij(l )}.

Lemma. If A(l ) is a proper l -matrix and B(l ) ¹ 0 is a l -matrix, deg (A(l )B(l )) = deg (B(l )A(l )) = deg A(l ) + deg B(l ).

Proof: We can write A(l ) = Ap l p + Ap-1 l p-1 + ... + A1 l + A0, 0 £ p B(l ) = Bq l q + Bq-1 l q-1 + ... + B1 l + B0 , 0 £ q where |Ap| ¹ 0 and Bq ¹ 0. Consequently ApBq and BqAp are non-zero matrices. As A(l )B(l ) = ApBq l p+q + ... + (S i+j=k AiBj)l k + ... + A0 B0 and B(l )A(l ) = BqAp l p+q + ... + ( S i+j=k BjAi)l k + ... + B0A0, the result follows. #

Corollary (degree reduction). If A(l ) is a proper l -matrix, C(l ) is a l -matrix, and B(l ) = A(l )C(l ), then deg B(l ) < deg A(l ) iff C(l ) = B(l ) = 0.

Proof: Else, by the Lemma deg B(l ) = deg A(l ) + deg C(l ) ³ deg A(l ) if C(l ) ¹ 0, a contradiction. #

Remainder Theorem (for l -Matrices) . If A(l ), B(l ) are l -matrices and B(l ) is proper, there exist unique l -matrices Ql(l ), Rl(l ), Qr(l), Rr(l ) such that Ql(l )B(l ) + Rl(l ) = A(l ) = B(l )Qr(l ) + Rr(l ), where deg Rl(l ), deg Rr(l ) < deg B(l ).

Proof: The uniqueness follows from Corollary 2, since, e.g., Ql(l )B(l ) + Rl(l ) = Q(l )B(l ) + R(l ) implies that (Ql(l )-Q(l ))B(l ) = R(l )-Rl(l ), and deg B(l ) > deg (R(l )-Rl(l )). The existence follows from the long-hand left and right division process. #

Theorem (Similarity via l -Matrices). Two n´ n matrices A and B are similar iff there exist l -matrices L(l ) and R(l ) such that |L(l )| and |R(l )| are non-zero scalars, and, L(l )(l I-A) = (l I-B)R(l ).

Proof: As l I-B and l I-A are first degree proper l -matrices, let L(l ) = (l I-B)R1(l ) + C and R(l ) = L1(l)(l I-A) + E, where C and E are scalar matrices. Then (l I - B)(R1(l ) - L1(l ))(l I - A) = l (E - C) + CA - BE. If R1(l ) - L1(l ) ¹ 0, the degree of left hand l -matrix is ³ 2 > degree of right hand l-matrix, a contradiction. Hence R1(l ) = L1(l ). It follows that E = C and CA = BE, so that CA = BC. It remains, therefore, to show that E = C is non-singular.

Since det (R(l )) is a non-zero scalar, (R(l ))-1 is a l -matrix. Hence we can write (R(l )) = S1(l )(l I-B) + S2 (S2 a scalar matrix). Then (R(l ))-1R(l ) = I = (R(l ))-1 (L1(l )(l I-A) + E) = (R(l ))-1L1(l )(l I-A) + S1(l )(l I-B)E + S2E = (R(l ))-1L1(l )(l I-A) + S1(l )C(l I-A)+ S2E = ((R(l ))-1L1(l )+S1(l )C)(l I-A) + S2E, and unless (R(l ))-1 L1(l ) + S1(l )C is zero, the degree of the right hand l -matrix would be ³ 1 ¹ deg (I) = 0. Hence S2E = I and so E is non-singular. Note that the converse part is trivial, since if A = C-1BC, we could choose L(l ) = R(l ) = C. #

 

PROBLEMS

1. Examine the proof of the previous theorem carefully so that if you are given two l-matrices L(l ) and R(l ) such that |L(l )| and |R(l )| are non-zero scalars and L(l )(l I-A) = (l I-B)R(l ), you could actually determine a C such that A = C-1BC.

2. Construct appropriate l -matrices to check if the matrices

are similar and if so find a similarity relation for the same.

3. If A(l ), B(l ) are proper, show that A(l )B(l ) is proper.

4. Give an example of l -matrices A(l ) and B(l ) for which no two of deg (A(l )B(l )), deg (B(l )A(l )), and, deg A(l ) + deg B(l ) are equal.

 

Elementary Transformations on l -Matrices

The following three types of row (column) operations are called l -elementary row (column) operations on l -matrices, or just elementary operations, or transformations on l -matrices:

(i) Multiplying a row (column) by a non-zero scalar.

(ii) Interchanging two rows (columns).

(iii) Adding to a row (column) a scalar polynomial in l times another row (column).

The operations (i)-(ii) are the same as the elementary row (column) operations discussed earlier. The difference lies in only the operation of type (iii) where here instead of just a scalar multiple we allow a more general l -polynomial multiple of a row (column) to be added to another.

A matrix obtained from the identity matrix by performing a l -elementary row (column) operation would be called an elementary l -matrix. It is interesting to note that every elementary l -matrix can be obtained by both a l -elementary row operation and a l -elementary column operation on an identity matrix. For, the three types of elementary l-matrices are:

, ,

Type (i) l -elementary matrix Type (ii) l -elementary matrix Type (iii) l -elementary matrix

and each type (i)-(ii) of a l -elementary row operation has the same effect on the identity matrix as the corresponding type of a l -elementary column operation. While adding a (l ) times the j-th row to the i-th row of I results in the same matrix as obtained by adding a (l ) times the i-th column to the j-th column.

As has been the case with the elementary row and column operations on a matrix, it follows that the rank of a l -matrix does not change by performing a succession of l -elementary row or column operations (as they constitute elementary row and column operations with respect to the field QF[l ]).

Two l -matrices are called equivalent if one of them can be obtained from the other by a succession of l -elementary row or column operations. Since the l -elementary row or column operations are reversible, it follows that this equivalence is actually an equivalence relation on the set of all l -matrices.

Theorem (Smith Normal Form). Every l -matrix A(l ) is equivalent to a l -matrix of the form diag (p1(l ), p2(l ), ... , pr(l ), 0, ... , 0), where r = rank A(l ), and pi(l ) are monic polynomials in l such that pi | pi+1 , i = 1, 2, ... , r-1.

Proof: Consider the lowest degree (non-zero) term in A(l ). If it does not divide every other term in A(l ), we show that A(l ) is equivalent to a matrix B(l ) in which the lowest degree term is of a lower degree than that in A(l ): For this, if a term which occurs in the same row or column as the lowest degree term aij(l ), say, is not divisible by aij(l ) we can subtract appropriate polynomial multiple of j-th column (or i-th row) from the column (row) containing the element to replace the element by a remainder of degree < deg aij(l ). Hence for the remaining cases we can assume that the i-th row and the j-th column contain only elements divisible by aij(l ). Now suppose apq(l ) is not divisible by aij(l ). First subtract appropriate polynomial multiple of i-th row from the p-th so as to make the (p, j)-th element zero and then add the i-th row so as to turn this element equal to aij(l ) itself. These operations only add a polynomial multiple of aij(l ) to apq(l ). Hence subtracting appropriate polynomial multiple of j-th column from the q-th column we can turn the (p, q)-th element to a remainder of degree less than deg (aij(l )), as required.

It follows from the above that using a succession of l -elementary row and column operations A(l ) is seen to be equivalent to a matrix in which the lowest degree term divides every other term. By an interchange of rows and columns this term could be brought to (1, 1)-position. Then subtracting appropritate polynomial multiples of the first row and column from the remaining rows and columns we can turn all other elements in the first column and row to zero. The remaining elements in the matrix continue to remain divisible by this (1, 1)-element. The matrix now is of the block-partitioned form

with p1(l ) dividing every entry in B(l ). The polynomial p1(l ) could be made monic by dividing the first row or column by the scalar coefficient of the highest degree term in it. Hence the result follows by an induction on the size of A(l ) and the fact that the l -elementary row and column operations do not change the rank of a l -matrix. #

Corollary. If det A(l ) is a non-zero scalar l -matrix, the Smith normal form of A(l ) is I, the identity matrix.

Proof: Since the l -elementary row and column operations change the determinant only by a non-zero scalar factor, it follows that in the present case no pk(l ) can be of a degree > 1. Since the rank is full it follows that pk(l ) = 1, for all k. #

The polynomials p1(l ), p2(l ), ... , pr(l ) in the Smith canonical form of A(l ) are called invariant factors of A(l ). The following is a characterization of the invariant factors:

Theorem (Characterization Of Invariant Factors). If hk(l ) denotes the HCF of all k´ k minors of a l -matrix A(l ) of rank r, the invariant factors pk(l ), 1 £ k £ r, of A(l ) are given by p1(l ) = h1(l ), and pk(l ) = hk(l )/hk-1(l ), 2 £ k £ r.

Proof: Consider a set (ordered in some form) of polynomials in l . It is clear that the HCF of this set is not affected by elementary operations of (i) interchanging the position of two elements in the set, (ii) multiplying an element by a scalar, and (iii) adding a polynomial multiple of one element to another. Next we have to note that when a l -elementary operation is performed on a l -matrix the set of the k´ k minors is affected only in one of the above fashions (i)-(iii) and consequently the HCF hk(l ) remains invariant. Hence, as for the Smith normal form of A(l ) there hold hk(l ) = P 1£ i£ k pi(l ), 1 £ k £ r, the result follows. #

Corollary. The Smith normal form of a l -matrix A(l ) is unique.

Proof: Since the hk(l ) of the Smith normal form of A(l ) are the same as those of A(l ) itself, i.e., they are independent of the l -elementary row and column operations that brings A(l ) to this form, the result follows from the previous theorem which determine pk(l ) uniquely in terms of the hk(l ). #

To compute the invariant factors of a l -matrix, we could use the constructive procedure in the proof of the Smith normal form theorem, or else resort to the theorem on the characterization of the invariant factors via the HCF of minors of various size. Since, in this fashion, the invariant factors pk(l ) of a l -matrix can be computed in a finite number of steps, the following result may be used to determine if two given matrices are similar:

Theorem. For two n´ n matrices A and B the following statements are equivalent:

(a) A and B are similar;

(b) l I-A and l I-B have the same Smith normal form;

(c) l I-A and l I-B have the same invariant factors pk(l ); and

(d) l I-A and l I-B are equivalent l -matrices.

Proof: It is clear that (a) Þ (b) Þ (c) Þ (d). Finally, if (d) holds we have L(l )(l I-A)C(l ) = (l I-B), where L(l ) is a product of elementary l -matrices corresponding l -row operations and C(l ) corresponds to the l -column operations that take l I-A to l I-B. Since both det L(l ) and det C(l ) are non-zero scalars, R(l ) = (C(l ))-1 is a l -matrix and det R(l ) is also a scalar, (a) follows from a previous lemma. #

 

PROBLEMS

1. The invariant factors of l I-A, where A is the second of the following matrices, are l , l and l 2 -4l :

.

Find the same if A is the first matrix to determine if the two matrices are similar.

2. Show that the determinant of a l -matrix A(l ) is a non-zero scalar iff A(l ) is a product of elementary l -matrices.

3. Find the inverses of the the three types of l -elementary matrices.

4. If D(l ) = det A(l ) ¹ 0, regarding A(l) as a matrix over the field QF[l ], show that (A(l ))-1 is a l -matrix

iff D(l ) does not depend on l , i.e., it is a scalar.

5. Find the Smith normal form of l I-A for the following A's:

and thereby determine whichever of these matrices are similar.