In
linear algebra
Linear algebra is the branch of mathematics concerning linear equations such as
:a_1x_1+\cdots +a_nx_n=b,
linear maps such as
:(x_1, \ldots, x_n) \mapsto a_1x_1+\cdots +a_nx_n,
and their representations in vector spaces and through matrix (mathemat ...
and
functional analysis
Functional analysis is a branch of mathematical analysis, the core of which is formed by the study of vector spaces endowed with some kind of limit-related structure (for example, Inner product space#Definition, inner product, Norm (mathematics ...
, a spectral theorem is a result about when a
linear operator
In mathematics, and more specifically in linear algebra, a linear map (also called a linear mapping, linear transformation, vector space homomorphism, or in some contexts linear function) is a mapping V \to W between two vector spaces that pr ...
or
matrix can be
diagonalized (that is, represented as a
diagonal matrix
In linear algebra, a diagonal matrix is a matrix in which the entries outside the main diagonal are all zero; the term usually refers to square matrices. Elements of the main diagonal can either be zero or nonzero. An example of a 2×2 diagon ...
in some basis). This is extremely useful because computations involving a diagonalizable matrix can often be reduced to much simpler computations involving the corresponding diagonal matrix. The concept of diagonalization is relatively straightforward for operators on
finite-dimensional vector spaces but requires some modification for operators on infinite-dimensional spaces. In general, the spectral theorem identifies a class of
linear operator
In mathematics, and more specifically in linear algebra, a linear map (also called a linear mapping, linear transformation, vector space homomorphism, or in some contexts linear function) is a mapping V \to W between two vector spaces that pr ...
s that can be modeled by
multiplication operators, which are as simple as one can hope to find. In more abstract language, the spectral theorem is a statement about commutative
C*-algebras. See also
spectral theory
In mathematics, spectral theory is an inclusive term for theories extending the eigenvector and eigenvalue theory of a single square matrix to a much broader theory of the structure of operator (mathematics), operators in a variety of mathematical ...
for a historical perspective.
Examples of operators to which the spectral theorem applies are
self-adjoint operators or more generally
normal operators on
Hilbert space
In mathematics, a Hilbert space is a real number, real or complex number, complex inner product space that is also a complete metric space with respect to the metric induced by the inner product. It generalizes the notion of Euclidean space. The ...
s.
The spectral theorem also provides a
canonical decomposition, called the
spectral decomposition, of the underlying vector space on which the operator acts.
Augustin-Louis Cauchy proved the spectral theorem for
symmetric matrices, i.e., that every real, symmetric matrix is diagonalizable. In addition, Cauchy was the first to be systematic about
determinant
In mathematics, the determinant is a Scalar (mathematics), scalar-valued function (mathematics), function of the entries of a square matrix. The determinant of a matrix is commonly denoted , , or . Its value characterizes some properties of the ...
s. The spectral theorem as generalized by
John von Neumann
John von Neumann ( ; ; December 28, 1903 – February 8, 1957) was a Hungarian and American mathematician, physicist, computer scientist and engineer. Von Neumann had perhaps the widest coverage of any mathematician of his time, in ...
is today perhaps the most important result of
operator theory.
This article mainly focuses on the simplest kind of spectral theorem, that for a
self-adjoint operator on a Hilbert space. However, as noted above, the spectral theorem also holds for normal operators on a Hilbert space.
Finite-dimensional case
Hermitian maps and Hermitian matrices
We begin by considering a
Hermitian matrix
In mathematics, a Hermitian matrix (or self-adjoint matrix) is a complex square matrix that is equal to its own conjugate transpose—that is, the element in the -th row and -th column is equal to the complex conjugate of the element in the ...
on
(but the following discussion will be adaptable to the more restrictive case of
symmetric matrices on We consider a
Hermitian map on a finite-dimensional
complex inner product space endowed with a
positive definite sesquilinear inner product
In mathematics, an inner product space (or, rarely, a Hausdorff pre-Hilbert space) is a real vector space or a complex vector space with an operation called an inner product. The inner product of two vectors in the space is a scalar, ofte ...
. The Hermitian condition on
means that for all ,
An equivalent condition is that , where is the
Hermitian conjugate of . In the case that is identified with a Hermitian matrix, the matrix of is equal to its
conjugate transpose. (If is a
real matrix, then this is equivalent to , that is, is a
symmetric matrix
In linear algebra, a symmetric matrix is a square matrix that is equal to its transpose. Formally,
Because equal matrices have equal dimensions, only square matrices can be symmetric.
The entries of a symmetric matrix are symmetric with ...
.)
This condition implies that all eigenvalues of a Hermitian map are real: To see this, it is enough to apply it to the case when is an eigenvector. (Recall that an
eigenvector of a linear map is a non-zero vector such that for some scalar . The value is the corresponding
eigenvalue. Moreover, the
eigenvalues are roots of the
characteristic polynomial.)
We provide a sketch of a proof for the case where the underlying field of scalars is the
complex number
In mathematics, a complex number is an element of a number system that extends the real numbers with a specific element denoted , called the imaginary unit and satisfying the equation i^= -1; every complex number can be expressed in the for ...
s.
By the
fundamental theorem of algebra, applied to the
characteristic polynomial of , there is at least one complex eigenvalue and corresponding eigenvector , which must by definition be non-zero. Then since
we find that is real. Now consider the space
, the
orthogonal complement of . By Hermiticity,
is an
invariant subspace of . To see that, consider any
so that
by definition of
. To satisfy invariance, we need to check if
. This is true because,
. Applying the same argument to
shows that has at least one real eigenvalue
and corresponding eigenvector
. This can be used to build another invariant subspace
. Finite induction then finishes the proof.
The matrix representation of in a basis of eigenvectors is diagonal, and by the construction the proof gives a basis of mutually orthogonal eigenvectors; by choosing them to be unit vectors one obtains an orthonormal basis of eigenvectors. can be written as a linear combination of pairwise orthogonal projections, called its spectral decomposition. Let
be the eigenspace corresponding to an eigenvalue
. Note that the definition does not depend on any choice of specific eigenvectors. In general, is the orthogonal direct sum of the spaces
where the
ranges over the
spectrum
A spectrum (: spectra or spectrums) is a set of related ideas, objects, or properties whose features overlap such that they blend to form a continuum. The word ''spectrum'' was first used scientifically in optics to describe the rainbow of co ...
of
.
When the matrix being decomposed is Hermitian, the spectral decomposition is a special case of the
Schur decomposition (see the proof in case of
normal matrices below).
Spectral decomposition and the singular value decomposition
The spectral decomposition is a special case of the
singular value decomposition
In linear algebra, the singular value decomposition (SVD) is a Matrix decomposition, factorization of a real number, real or complex number, complex matrix (mathematics), matrix into a rotation, followed by a rescaling followed by another rota ...
, which states that any matrix
can be expressed as
, where
and
are
unitary matrices and
is a diagonal matrix. The diagonal entries of
are uniquely determined by
and are known as the
singular values of
. If
is Hermitian, then
and
which implies
.
Normal matrices
The spectral theorem extends to a more general class of matrices. Let be an operator on a finite-dimensional inner product space. is said to be
normal if .
One can show that is normal if and only if it is unitarily diagonalizable using the
Schur decomposition. That is, any matrix can be written as , where is unitary and is
upper triangular.
If is normal, then one sees that . Therefore, must be diagonal since a normal upper triangular matrix is diagonal (see
normal matrix
In mathematics, a complex square matrix is normal if it commutes with its conjugate transpose :
:A \text \iff A^*A = AA^* .
The concept of normal matrices can be extended to normal operators on infinite-dimensional normed spaces and to nor ...
). The converse is obvious.
In other words, is normal if and only if there exists a
unitary matrix such that
where is a
diagonal matrix
In linear algebra, a diagonal matrix is a matrix in which the entries outside the main diagonal are all zero; the term usually refers to square matrices. Elements of the main diagonal can either be zero or nonzero. An example of a 2×2 diagon ...
. Then, the entries of the diagonal of are the
eigenvalues of . The column vectors of are the eigenvectors of and they are orthonormal. Unlike the Hermitian case, the entries of need not be real.
Compact self-adjoint operators
In the more general setting of Hilbert spaces, which may have an infinite dimension, the statement of the spectral theorem for
compact self-adjoint operators is virtually the same as in the finite-dimensional case.
As for Hermitian matrices, the key point is to prove the existence of at least one nonzero eigenvector. One cannot rely on determinants to show existence of eigenvalues, but one can use a maximization argument analogous to the variational characterization of eigenvalues.
If the compactness assumption is removed, then it is ''not'' true that every self-adjoint operator has eigenvectors. For example, the multiplication operator
on
which takes each
to
is bounded and self-adjoint, but has no eigenvectors. However, its spectrum, suitably defined, is still equal to