The objective of this section is to find invariant subspaces of a linear operator. For a given vector space V
over the field of complex numbers \( \mathbb{C} \) (or real numbers \( \mathbb{R} \) ),
let \( T:\,V\,\to\,V \) be a linear transformation, we want to find subspaces M of
V such that \( T(M) \subseteq M . \) The operator T can be a matrix transformation, a linear integral operator,
or an unbounded linear differential operator. Obviously, the zero subspace {0} and the whole space V are
invariant for every linear transformation. We refer to these as trivial invariant subspaces.
Let us begin by looking for one-dimensional invariant subspaces. If M is spanned by a nonzero vector x,
that is, M = span {x}, then \( T(M) \subseteq M \) if and only if there
exists a scalar \( \lambda \in \mathbb{C} \) such that Tx = λx.
If A is a square \( n\times n \) matrix (or linear transformation),
then a nonzero vector \( {\bf x} \in \mathbb{R}^n \) is called eigenvector of
A (or a linear operator T) if Ax is a scalar multiple of x; that is,
\[
{\bf A}\,{\bf x} = \lambda \,{\bf x}
\]
for some scalar λ. The scalar λ is called an eigenvalue of A (or of linear operator T),
and x is said to be an eigenvector corresponding to λ. An orderred pair
\( \left( \lambda , {\bf x} \right) \) of eigenvalue and corresponding eigenvector is
called the eigenpair.
The set of all eigenvalues of matrix A or linear transformation T is called spectrum of matrix
A or linear transformation and denoted σ(A) or σ(T), respectively.
In general, the image of a vector x under multiplication by a square matrix A (or corresponding linear transformation
T) differs from x in
both magnitude and direction. However, in the special case where x is an eigenvector of A, multiplication
by A leaves the direction unchanged. Let M be a subspace spanned by the eigenvector x.
If y is another nonzero vector from M, then y = cx for some scalar c. Hence,
\( T\,{\bf y} = T \left( c\,{\bf x} \right) = c\,T \,{\bf x} = \lambda\,c\,{\bf x} = \lambda \,{\bf y} . \)
Thus, we see that λ depends on linear transformation T and subspace of eigenvectors M but not on
a particular choice of a vector spanning M. So any nonzero vector \( {\bf z} \in V \)
satisfying \( T\,{\bf z} = \lambda\,{\bf z} \) is an eigenvector of T corresponding to,
or associated with, the eigenvalue λ.
The set of all eigenvectors corresponding to an eigenvalue together with the zero vector form the vector space, called
the eigenspace corresponding to the eigenvalue, and denoted by \( E_{\lambda} \ \mbox{or}\ E(\lambda ) . \)
The dimension of this eigenspace is called the geometric multiplicity
of the eigenvalue.
Let I denote the identity operator (or the identity matrix) on vector space V (or on \( \mathbb{R}^n \) ), then
\( \lambda \in \mathbb{C} \) is an eigenvalue of the linear operator T
(corresponding matrix operator TA) if and only if \( \lambda {\bf I} - T \) is not
injective (one-to-one). The corresponding vector equation \( {\bf A} \,{\bf x} = \lambda\,{\bf x} \)
or \( \left( \lambda\,{\bf I} - {\bf A} \right) {\bf x} = {\bf 0} \) has a nontrivial solution only when the matrix
\( \lambda\,{\bf I} - {\bf A} \) is singular, that is, its determinant is zero:
\( \det \left( \lambda\,{\bf I} - {\bf A} \right) =0 . \) Since this algebraic equation
has degree n, we conclude that an n×n matrix has at most n posssible eigenvalues.
For a square n-by-n matrix A, the polynomial of degree n\( \chi_A (\lambda ) = \lambda^n + c_{n-1} \lambda^{n-1} + \cdots + c_0 = \det \left( \lambda\,{\bf I} - {\bf A} \right) \)
is called the characteristic polynomial of matrix A.
We call the set Eλ of all eigenvectors with the zero vector an eigenspace because
Eλ is the kernel (null space) of the matrix
\( \lambda\,{\bf I} - {\bf A} , \) which is known to be a vector space.
Now suppose that p(λ) is an arbitrary polynomial. Then for any square matrix A we can define
a matrix polynomial according to the formula:
Theorem: Every Linear operator (or matrix) on a finite-dimensional
complex vector space has an eigenvalue.
To show that a linear operator T on n-dimensional vector space V has an eigenvalue, fix any nonzero vector
\( {\bf x} \in V . \) The vectors
\( {\bf x}, T\,{\bf x} , T^2 {\bf x} , \ldots , T^n {\bf x} \) cannot be linearly
independent because V has dimension N and we have n+1 vectors. Thus, there exist complex numbers
\( a_0 , a_1 , \ldots , a_n , \) not all 0, such that
where c is a non-zero complex number, each λj is complex, and the equation holds for all complex λ. Then we have
\[
{\bf 0} = \left( a_0 I + a_1 T + \cdots + a_n T^n \right) {\bf x} = c \left( T - \lambda_1 {\bf I} \right) \cdots \left( T - \lambda_s {\bf I}\right) ,
\]
which means that T - λI is not injective for at least one j. In other words, T
has an eigenvalue. ■
Theorem: Let \( \left( \lambda , {\bf x} \right) \)
be an eigenpair of a square matrix A, and let p(λ) be a polynomial. Then
\[
p \left( {\bf A} \right) {\bf x} = p \left( \lambda \right) {\bf x} ,
\]
that is, \( \left( p \left( \lambda \right) , {\bf x} \right) \) is an eigenpair of
p(A). ■
It is sufficient to show that Ak has the same eigenvectors as A associated with the eigenvalue
λk. We proceed by induction to show that
\( {\bf A}^k {\bf x} = \lambda^k {\bf x} \) for all \( k \ge 0 . \)
For the base case k=0, observe that A0 = I, the identity matrix and hence
λ0 = 1. For the inductive step, suppose that
\( {\bf A}^k {\bf x} = \lambda^k {\bf x} \) for some \( k \ge 0 . \)
Then
Theorem: For every n-by-n matrix
A, eigenvectors associated with distinct eigenvalues are linearly independent. ■
According to the Lagrange interpolation theorem, there are polynomials
\( p_1 (\lambda ), p_2 (\lambda ) , \ldots , p_s (\lambda_s ) \) such that
\[
p_i \left( \lambda_k \right) = \begin{cases} 1 , & \ \mbox{ if } i=k , \\
0 , & \ \mbox{ if } i \ne k , \end{cases}
\]
for each \( i = 1, 2, \ldots , s . \) Suppose that
\( c_1 , c_2 , \ldots , c_s \) are scalars so that for s eigenvectors associated with
s distinct eigenvalues we have
We must show that each ci = 0. The previous theorem ensures us that
\( p_i \left( {\bf A} \right) {\bf x}_k = p \left( \lambda_k \right) {\bf x}_k , \) so for each
\( i = 1, 2, \ldots , s \)
Theorem: For every square matrix
A, if λ is its eigenvalue then its complex conjugate \( \overline{\lambda} \)
is an eigenvalue of the adjoint matrix \( {\bf A}^{\ast} = \overline{{\bf A}^{\mathrm T}} =
\left( \overline{\bf A} \right)^{\mathrm T} , \) and \( \mbox{dim}\,E_{\lambda} \left( {\bf A} \right) =
\mbox{dim} \,E_{\overline{\lambda}} \left( {\bf A}^{\ast} \right) . \) ■
If u is an eigenvector of the adjoint operator T*, we take an inner product
where x2 an arbitrary nonzero scalar. Since the corresponding eigenspace E0 is
spanned on the vector \( \left[ -1, 1, 0 \right]^{\mathrm T} , \) its geometric multiplicity
is 1. Solving the system (b), we find that x1 = -3x2 and x3 =
x2. Therefore, the associated eigenvector becomes
So the eigenspace E-1 is spanned on the vector
\( \left[ -3, 1, 1 \right]^{\mathrm T} . \) Finally, solving system (c), we obtain
x1 = 2x2 and x3 = x2. Hence, the eigenvectors
associated with λ = 4 are expressed via the formula
Here j is the unit vector in the positive vertical direction on the complex plane, so
\( {\bf j}^2 = -1 . \)
Therefore, the given 2×2 matrix has two simple complex conjugate eigenvalues both having magnitude 1.
To find eigenvectors corresponding to λ+, we need to compute nonzero solutions of the homogeneous equation
because the second equation \( x_1 \sin \theta + x_2 \cos\theta = x_2 e^{{\bf j} \theta} \)
is a multiple of the above one. Solving it, we get \( x_1 = {\bf j{ \,x_2 . \) Hence,
\( \left( e^{{\bf j} \theta} , \left[ 1 , -{\bf j} \right]^{\mathrm T} \right) \) is an
eigenpair of A. Taking complex conjugate, we conclude that
\( \left( e^{-{\bf j} \theta} , \left[ 1 , {\bf j} \right]^{\mathrm T} \right) \) is another
eigenpair. The pair of eigenvectors \( \left[ 1 , -{\bf j} \right]^{\mathrm T} , \left[ 1 , {\bf j} \right]^{\mathrm T} \)
form a basis for \( \mathbb{C} . \) It is noteworthy that the eigenvalues of
Aθ on \( \theta \in \left( 0, \frac{\pi}{2} \right) , \) but
associated eigenvectors do not. Each of the vectors \( \left[ 1 , \pm {\bf j} \right]^{\mathrm T} \)
is an eigenvector of all the matrices Aθ independently of θ.
Example: If an integer \( n \ge 2 , \) then the identity matrix
In of dimensions \( n \times n \) has n identical eigenvalues,
namely λ = 1. Any basis of \( \mathbb{R}^n \) comprises eigenvectors of
In because any nonzero vector is an eigenvector of In associated with
λ = 1.
■
Example: The infinite set of monomials \( \left\{ 1, x, x^2 , \ldots , x^n , \ldots \right\} \)
form a basis in the set of all polynomials.
■
A defective matrix is a square matrix that does not have a complete basis of eigenvectors.
In other words, a square matrix A is called defective if A has an eigenvalue λ of
algebraic multiplicity m greater than 1, but for which the associated eigenspace has a basis of fewer than
m vectors; that is, the dimension of the eigenspace associated with λ is less than m.
that has only two distinct eigenvalues λ = 1 and λ = 2. Its characteristic polynomial
\( \chi (\lambda ) = \det \left( \lambda {\bf I} - {\bf A} \right) = \left( \lambda -2 \right) \left( \lambda -1 \right)^2 \)
has two roots: λ2 =2, which is simple one, and λ1 =1 that has algebraic multiplicity 2.
To these two eigenvalues correspond only two eigenvectors:
Since we have only two eigenvectors in three dimensional space
\( \mathbb{R}^n , \) the given matrix is defective.
■
Theorem: For every n-by-n matrix
A, the geometric multiplicity of each eigenvalue is less than or equal to its algebraic multiplicity of that
eigenvalue. ■
Let L0 be the eigenspace of A corresponding to an eigenvalue λ0 and let
dim L0 = m. We denote by \( {\bf e}_1 , {\bf e}_2 , \ldots , {\bf e}_m \)
a basis of the eigenspace L0. We extend this basis to a basis of \( \mathbb{R}^n \)
by some additional vectors \( {\bf e}_{m+1} , {\bf e}_{m+2} , \ldots , {\bf e}_n . \) Since
\( {\bf A} \,{\bf e}_i = \lambda_0 {\bf e}_i , \ i=1,2,\ldots , m , \) it follows that matrix
A is similar to a block diagonal matrix
where Λ0 is a diagonal m-by-m matrix all of whose diagonal entries are equal to
λ0. Hence, the characteristic polynomial of transformation corresponding to matrix A can be written as
All other entries are assumed to be zero.
Because of the zeroes in columns 2 through (n-1) of an
N-matrix, computation of its chracteristic polynomial
det(λI - A) is very easy. Expanding by cofactors
of any column, we obtain
By proper selection of the elements in an N-matrix one can derive a
great variety of examples of matrices with desiered eigenvalues. For
example, a symmetric 3 × 3 matrix has the followng eigenvalues:
\[
\det \left( \lambda {\bf I} -{\bf A} \right) = \left( \lambda - d
\right) \left(\lambda - a -b \right) \left( \lambda - a + b \right) ,
\qquad \mbox{for} \quad {\bf A} = \begin{bmatrix} a&0&b \\ c&d& c \\ b
& 0 & a \end{bmatrix} .
\]
By choosing appropriate values of entries, one can obtain a
diagonalizable matrix with either all
distinct eigenvalues or one double eigenvalue when d = a
- b . More variety are seen for 4 × 4 matrix
\[
{\bf A} = \begin{bmatrix} a&0&0& b \\
c_1 & d & 0 & c_1 \\
c_2 & 0 & e & c_2 \\
b&0&0&a \end{bmatrix} .
\]
Councilman, S., Eigenvalues and eigenvectors of "N-matrices", The
American Mathematical Monthly, 1986, Vol. 93, No. 5,
pp. 392--395. doi: 10.2307/2323607