Sylvester's law of inertia

Last updated

Sylvester's law of inertia is a theorem in matrix algebra about certain properties of the coefficient matrix of a real quadratic form that remain invariant under a change of basis. Namely, if is the symmetric matrix that defines the quadratic form, and is any invertible matrix such that is diagonal, then the number of negative elements in the diagonal of is always the same, for all such ; and the same goes for the number of positive elements.

Contents

This property is named after James Joseph Sylvester who published its proof in 1852. [1] [2]

Statement

Let be a symmetric square matrix of order with real entries. Any non-singular matrix of the same size is said to transform into another symmetric matrix , also of order , where is the transpose of . It is also said that matrices and are congruent. If is the coefficient matrix of some quadratic form of , then is the matrix for the same form after the change of basis defined by .

A symmetric matrix can always be transformed in this way into a diagonal matrix which has only entries , , along the diagonal. Sylvester's law of inertia states that the number of diagonal entries of each kind is an invariant of , i.e. it does not depend on the matrix used.

The number of s, denoted , is called the positive index of inertia of , and the number of s, denoted , is called the negative index of inertia. The number of s, denoted , is the dimension of the null space of , known as the nullity of . These numbers satisfy an obvious relation

The difference, , is usually called the signature of . (However, some authors use that term for the triple consisting of the nullity and the positive and negative indices of inertia of ; for a non-degenerate form of a given dimension these are equivalent data, but in general the triple yields more data.)

If the matrix has the property that every principal upper left minor is non-zero then the negative index of inertia is equal to the number of sign changes in the sequence

Statement in terms of eigenvalues

The law can also be stated as follows: two symmetric square matrices of the same size have the same number of positive, negative and zero eigenvalues if and only if they are congruent [3] (, for some non-singular ).

The positive and negative indices of a symmetric matrix are also the number of positive and negative eigenvalues of . Any symmetric real matrix has an eigendecomposition of the form where is a diagonal matrix containing the eigenvalues of , and is an orthonormal square matrix containing the eigenvectors. The matrix can be written where is diagonal with entries , and is diagonal with . The matrix transforms to .

Law of inertia for quadratic forms

In the context of quadratic forms, a real quadratic form in variables (or on an -dimensional real vector space) can by a suitable change of basis (by non-singular linear transformation from to ) be brought to the diagonal form

with each . Sylvester's law of inertia states that the number of coefficients of a given sign is an invariant of , i.e., does not depend on a particular choice of diagonalizing basis. Expressed geometrically, the law of inertia says that all maximal subspaces on which the restriction of the quadratic form is positive definite (respectively, negative definite) have the same dimension. These dimensions are the positive and negative indices of inertia.

Generalizations

Sylvester's law of inertia is also valid if and have complex entries. In this case, it is said that and are -congruent if and only if there exists a non-singular complex matrix such that , where denotes the conjugate transpose. In the complex scenario, a way to state Sylvester's law of inertia is that if and are Hermitian matrices, then and are -congruent if and only if they have the same inertia, the definition of which is still valid as the eigenvalues of Hermitian matrices are always real numbers.

Ostrowski proved a quantitative generalization of Sylvester's law of inertia: [4] [5] if and are -congruent with , then their eigenvalues are related by

where are such that .

A theorem due to Ikramov generalizes the law of inertia to any normal matrices and : [6] If and are normal matrices, then and are congruent if and only if they have the same number of eigenvalues on each open ray from the origin in the complex plane.

See also

Related Research Articles

In mathematics, a symmetric matrix with real entries is positive-definite if the real number is positive for every nonzero real column vector where is the transpose of . More generally, a Hermitian matrix is positive-definite if the real number is positive for every nonzero complex column vector where denotes the conjugate transpose of

<span class="mw-page-title-main">Symmetric matrix</span> Matrix equal to its transpose

In linear algebra, a symmetric matrix is a square matrix that is equal to its transpose. Formally,

<span class="mw-page-title-main">Square matrix</span> Matrix with the same number of rows and columns

In mathematics, a square matrix is a matrix with the same number of rows and columns. An n-by-n matrix is known as a square matrix of order . Any two square matrices of the same order can be added and multiplied.

<span class="mw-page-title-main">Orthogonal group</span> Type of group in mathematics

In mathematics, the orthogonal group in dimension n, denoted O(n), is the group of distance-preserving transformations of a Euclidean space of dimension n that preserve a fixed point, where the group operation is given by composing transformations. The orthogonal group is sometimes called the general orthogonal group, by analogy with the general linear group. Equivalently, it is the group of n × n orthogonal matrices, where the group operation is given by matrix multiplication (an orthogonal matrix is a real matrix whose inverse equals its transpose). The orthogonal group is an algebraic group and a Lie group. It is compact.

In mathematics, particularly in linear algebra, a skew-symmetricmatrix is a square matrix whose transpose equals its negative. That is, it satisfies the condition

In mathematics, a Hermitian matrix is a complex square matrix that is equal to its own conjugate transpose—that is, the element in the i-th row and j-th column is equal to the complex conjugate of the element in the j-th row and i-th column, for all indices i and j:

In linear algebra, a square matrix  is called diagonalizable or non-defective if it is similar to a diagonal matrix. That is, if there exists an invertible matrix  and a diagonal matrix such that . This is equivalent to . This property exists for any linear map: for a finite-dimensional vector space , a linear map  is called diagonalizable if there exists an ordered basis of  consisting of eigenvectors of . These definitions are equivalent: if  has a matrix representation as above, then the column vectors of  form a basis consisting of eigenvectors of , and the diagonal entries of  are the corresponding eigenvalues of ; with respect to this eigenvector basis,  is represented by .

In mathematics, a quadratic form is a polynomial with terms all of degree two. For example,

In the mathematical discipline of linear algebra, a matrix decomposition or matrix factorization is a factorization of a matrix into a product of matrices. There are many different matrix decompositions; each finds use among a particular class of problems.

In mathematics, the signature(v, p, r) of a metric tensor g (or equivalently, a real quadratic form thought of as a real symmetric bilinear form on a finite-dimensional vector space) is the number (counted with multiplicity) of positive, negative and zero eigenvalues of the real symmetric matrix gab of the metric tensor with respect to a basis. In relativistic physics, v conventionally represents the number of time or virtual dimensions, and p the number of space or physical dimensions. Alternatively, it can be defined as the dimensions of a maximal positive and null subspace. By Sylvester's law of inertia these numbers do not depend on the choice of basis and thus can be used to classify the metric. The signature is often denoted by a pair of integers (v, p) implying r = 0, or as an explicit list of signs of eigenvalues such as (+, −, −, −) or (−, +, +, +) for the signatures (1, 3, 0) and (3, 1, 0), respectively.

In the mathematical field of graph theory, the Laplacian matrix, also called the graph Laplacian, admittance matrix, Kirchhoff matrix or discrete Laplacian, is a matrix representation of a graph. Named after Pierre-Simon Laplace, the graph Laplacian matrix can be viewed as a matrix form of the negative discrete Laplace operator on a graph approximating the negative continuous Laplacian obtained by the finite difference method.

In probability theory and mathematical physics, a random matrix is a matrix-valued random variable—that is, a matrix in which some or all elements are random variables. Many important properties of physical systems can be represented mathematically as matrix problems. For example, the thermal conductivity of a lattice can be computed from the dynamical matrix of the particle-particle interactions within the lattice.

In linear algebra, it is often important to know which vectors have their directions unchanged by a given linear transformation. An eigenvector or characteristic vector is such a vector. Thus an eigenvector of a linear transformation is scaled by a constant factor when the linear transformation is applied to it: . The corresponding eigenvalue, characteristic value, or characteristic root is the multiplying factor .

In mathematics, two square matrices A and B over a field are called congruent if there exists an invertible matrix P over the same field such that

In mathematics, the square root of a matrix extends the notion of square root from numbers to matrices. A matrix B is said to be a square root of A if the matrix product BB is equal to A.

In linear algebra, an idempotent matrix is a matrix which, when multiplied by itself, yields itself. That is, the matrix is idempotent if and only if . For this product to be defined, must necessarily be a square matrix. Viewed this way, idempotent matrices are idempotent elements of matrix rings.

In geometry and linear algebra, a principal axis is a certain line in a Euclidean space associated with an ellipsoid or hyperboloid, generalizing the major and minor axes of an ellipse or hyperbola. The principal axis theorem states that the principal axes are perpendicular, and gives a constructive procedure for finding them.

In linear algebra, eigendecomposition is the factorization of a matrix into a canonical form, whereby the matrix is represented in terms of its eigenvalues and eigenvectors. Only diagonalizable matrices can be factorized in this way. When the matrix being factorized is a normal or real symmetric matrix, the decomposition is called "spectral decomposition", derived from the spectral theorem.

In the mathematical field of linear algebra, an arrowhead matrix is a square matrix containing zeros in all entries except for the first row, first column, and main diagonal, these entries can be any number. In other words, the matrix has the form

In mathematics, symmetric cones, sometimes called domains of positivity, are open convex self-dual cones in Euclidean space which have a transitive group of symmetries, i.e. invertible operators that take the cone onto itself. By the Koecher–Vinberg theorem these correspond to the cone of squares in finite-dimensional real Euclidean Jordan algebras, originally studied and classified by Jordan, von Neumann & Wigner (1934). The tube domain associated with a symmetric cone is a noncompact Hermitian symmetric space of tube type. All the algebraic and geometric structures associated with the symmetric space can be expressed naturally in terms of the Jordan algebra. The other irreducible Hermitian symmetric spaces of noncompact type correspond to Siegel domains of the second kind. These can be described in terms of more complicated structures called Jordan triple systems, which generalize Jordan algebras without identity.

References

  1. Sylvester, James Joseph (1852). "A demonstration of the theorem that every homogeneous quadratic polynomial is reducible by real orthogonal substitutions to the form of a sum of positive and negative squares" (PDF). Philosophical Magazine. 4th Series. 4 (23): 138–142. doi:10.1080/14786445208647087 . Retrieved 2008-06-27.
  2. Norman, C.W. (1986). Undergraduate algebra. Oxford University Press. pp. 360–361. ISBN   978-0-19-853248-4.
  3. Carrell, James B. (2017). Groups, Matrices, and Vector Spaces: A Group Theoretic Approach to Linear Algebra. Springer. p. 313. ISBN   978-0-387-79428-0.
  4. Ostrowski, Alexander M. (1959). "A quantitative formulation of Sylvester's law of inertia" (PDF). Proceedings of the National Academy of Sciences. A quantitative formulation of Sylvester's law of inertia (5): 740–744. Bibcode:1959PNAS...45..740O. doi: 10.1073/pnas.45.5.740 . PMC   222627 . PMID   16590437.
  5. Higham, Nicholas J.; Cheng, Sheung Hun (1998). "Modifying the inertia of matrices arising in optimization". Linear Algebra and Its Applications. 275–276: 261–279. doi: 10.1016/S0024-3795(97)10015-5 .
  6. Ikramov, Kh. D. (2001). "On the inertia law for normal matrices". Doklady Mathematics. 64: 141–142.