Fact-checked by Grok 2 weeks ago

Skew-symmetric matrix

In linear algebra, a skew-symmetric matrix (also called an antisymmetric matrix) is a A satisfying A^T = -A, which implies that its elements satisfy a_{ij} = -a_{ji} for all indices i and j. This condition ensures that the matrix is equal to the negative of its , distinguishing it from where A^T = A. Skew-symmetric matrices form a vector subspace of the of all , closed under addition and , and any real can be uniquely decomposed as the sum of a and a skew-symmetric matrix. Key properties of skew-symmetric matrices include zero diagonal elements, as a_{ii} = -a_{ii} forces a_{ii} = 0 for all i, making the matrix . For real skew-symmetric matrices, all eigenvalues are either zero or purely imaginary, and they occur in pairs; moreover, the matrix is singular if its dimension is odd. Over the real numbers, the of an n \times n skew-symmetric matrix is always even. These matrices are also (commuting with their ) when considered over the complexes, and their exponentials yield orthogonal matrices with 1, connecting them to the special SO(n). Skew-symmetric matrices play a fundamental role in geometry and physics, particularly in representing infinitesimal rotations and the of the rotation group SO(3), where elements correspond to angular velocities or infinitesimal motions. In and , they model operators and cross-product structures, facilitating the study of conservation laws and symmetries. Their invariant further links them to determinants and combinatorial problems in even dimensions.

Definition and Basic Concepts

Definition

In linear algebra, a skew-symmetric matrix is a square matrix A over the real numbers satisfying A^T = -A, where A^T denotes the transpose of A. This condition implies that the entries of the matrix obey a_{ij} = -a_{ji} for all indices i and j, with the diagonal entries necessarily zero since a_{ii} = -a_{ii} forces a_{ii} = 0. The definition extends to square matrices over any field of characteristic not equal to 2, where the relation A^T = -A remains meaningful and distinct from the symmetric case A^T = A. In fields of characteristic 2, however, the condition A^T = -A simplifies to A^T = A because -1 = 1, causing skew-symmetric matrices to coincide with symmetric ones and rendering the notion less useful without additional structure like alternating forms. Skew-symmetric matrices serve as the antipodal counterpart to symmetric matrices in the decomposition of arbitrary square matrices.

Elementary Properties

A skew-symmetric matrix A satisfies A^T = -A, which immediately implies that all diagonal entries are zero. For the (i,i)-th entry, a_{ii} = -a_{ii}, so $2a_{ii} = 0 and thus a_{ii} = 0. The trace of A, defined as the sum of its diagonal entries \operatorname{tr}(A) = \sum_{i} a_{ii}, is therefore zero. The set of skew-symmetric matrices is closed under and . If A and B are skew-symmetric, then (A + B)^T = A^T + B^T = -A - B = -(A + B), so A + B is skew-symmetric. Similarly, for any scalar c, (cA)^T = c A^T = c (-A) = - (cA), confirming that cA is skew-symmetric. The product of two skew-symmetric matrices need not be skew-symmetric. If A and B are skew-symmetric, then (AB)^T = B^T A^T = (-B)(-A) = BA, which equals -AB only if AB = -BA. However, the square A^2 of a skew-symmetric matrix is symmetric: (A^2)^T = (A^T)^2 = (-A)^2 = A^2. For real matrices, the skew-symmetric condition A^T = -A coincides with the skew-Hermitian condition A^* = -A, where A^* denotes the , since conjugation has no effect on real entries.

Examples

Low-Dimensional Cases

In the one-dimensional case, the only skew-symmetric matrix is the {{grok:render&&&type=render_inline_citation&&&citation_id=0&&&citation_type=wikipedia}}, since the single diagonal entry a must satisfy a = -a, which implies a = 0. For two dimensions, a general real skew-symmetric $2 \times 2 takes the form \begin{pmatrix} 0 & a \\ -a & 0 \end{pmatrix}, where a \in \mathbb{R} is the single independent parameter above the diagonal. An example is \begin{pmatrix} 0 & 1 \\ -1 & 0 \end{pmatrix}, which generates a scaled 90-degree when exponentiated. In three dimensions, a general real skew-symmetric $3 \times 3 has three independent parameters and takes the form \begin{pmatrix} 0 & a & b \\ -a & 0 & c \\ -b & -c & 0 \end{pmatrix}, where a, b, c \in \mathbb{R}. A concrete example is the associated with the operation for the \mathbf{v} = (x, y, z), given by [\mathbf{v}]_\times = \begin{pmatrix} 0 & -z & y \\ z & 0 & -x \\ -y & x & 0 \end{pmatrix}, which satisfies the skew-symmetric condition and maps vectors via matrix-vector multiplication to mimic the . Skew-symmetric matrices in low dimensions can be interpreted as representing antisymmetric bilinear forms on vectors: for a matrix A, the form B(\mathbf{u}, \mathbf{v}) = \mathbf{u}^T A \mathbf{v} satisfies B(\mathbf{u}, \mathbf{v}) = -B(\mathbf{v}, \mathbf{u}). All such matrices have zero , as their diagonal entries are necessarily zero.

General Constructions

A skew-symmetric matrix of size n \times n has exactly \frac{n(n-1)}{2} independent entries, as the diagonal elements must be zero and the entries below the diagonal are determined by those above it via the relation a_{ji} = -a_{ij}. One general method to construct larger skew-symmetric matrices is through the (or block-diagonal form) of smaller skew-symmetric matrices. If A and B are skew-symmetric matrices of sizes p \times p and q \times q, respectively, their A \oplus B is the (p+q) \times (p+q) block-diagonal matrix with A and B on the diagonal blocks and zeros elsewhere; this is skew-symmetric because (A \oplus B)^T = A^T \oplus B^T = (-A) \oplus (-B) = -(A \oplus B). Low-dimensional skew-symmetric matrices can serve as building blocks in such constructions. In three dimensions, a skew-symmetric $3 \times 3 can be constructed from a \mathbf{v} = (v_1, v_2, v_3) as [\mathbf{v}]_\times = \begin{pmatrix} 0 & -v_3 & v_2 \\ v_3 & 0 & -v_1 \\ -v_2 & v_1 & 0 \end{pmatrix}, which satisfies [\mathbf{v}]_\times^T = -[\mathbf{v}]_\times. The provides another construction route: for a skew-symmetric matrix A, the matrix e^{tA} is orthogonal for any real scalar t, since (e^{tA})^T = e^{tA^T} = e^{-tA} = (e^{tA})^{-1}.

Algebraic Properties

Vector Space Structure

The set of all n \times n skew-symmetric matrices over the real numbers \mathbb{R} forms a vector space under matrix addition and scalar multiplication. This space is closed under these operations: if A^T = -A and B^T = -B, then (aA + bB)^T = aA^T + bB^T = -aA - bB = -(aA + bB) for any scalars a, b \in \mathbb{R}. The dimension of this vector space is n(n-1)/2, corresponding to the number of independent entries above the main diagonal, as the diagonal elements must be zero and the lower triangle is determined by the upper triangle via the skew-symmetry condition. A for this consists of the matrices E_{ij} - E_{ji} for $1 \leq i < j \leq n, where E_{kl} denotes the n \times n matrix with a 1 in position (k,l) and zeros elsewhere. These basis elements are linearly independent and span the space, as any skew-symmetric matrix A = [a_{kl}] can be uniquely expressed as A = \sum_{1 \leq i < j \leq n} a_{ij} (E_{ij} - E_{ji}). The space of skew-symmetric matrices admits the Frobenius inner product, defined for any two n \times n matrices A, B by \langle A, B \rangle = \operatorname{tr}(A^T B) = \sum_{i,j=1}^n a_{ij} b_{ij}, where \operatorname{tr} denotes the trace. For skew-symmetric matrices A and B, this simplifies to \langle A, B \rangle = \operatorname{tr}(A^T B) = \operatorname{tr}((-A) B) = -\operatorname{tr}(A B), reflecting the anti-symmetric nature of the pairing. This inner product establishes orthogonality between the space of skew-symmetric matrices and the space of symmetric matrices: if A is skew-symmetric and S is symmetric (S^T = S), then \langle A, S \rangle = \operatorname{tr}(A^T S) = \operatorname{tr}((-A) S) = -\operatorname{tr}(A S). But \operatorname{tr}(A S) = \operatorname{tr}(S A) = \operatorname{tr}((A S)^T) = \operatorname{tr}(S^T A^T) = \operatorname{tr}(S (-A)) = -\operatorname{tr}(S A) = -\operatorname{tr}(A S), implying \operatorname{tr}(A S) = 0 and thus \langle A, S \rangle = 0. This decomposition highlights how the full matrix space \mathbb{R}^{n \times n} splits into orthogonal subspaces of symmetric and skew-symmetric matrices under the .

Determinant and Pfaffian

For a skew-symmetric matrix A of odd order n, the determinant is always zero. This follows from the property that \det(A) = \det(A^T) = \det(-A) = (-1)^n \det(A); since n is odd, \det(A) = -\det(A), implying \det(A) = 0. Thus, such matrices are singular. For a real skew-symmetric matrix A of even order n = 2m, the determinant is non-negative and equals the square of the Pfaffian: \det(A) = \Pf(A)^2 \geq 0. The Pfaffian \Pf(A) is defined for such matrices as the sum over all perfect matchings \sigma of \{1, 2, \dots, 2m\}, where each matching pairs the indices into m disjoint transpositions (i_1 j_1), \dots, (i_m j_m) with i_k < j_k and i_1 < i_2 < \dots < i_m: \Pf(A) = \sum_{\sigma} \operatorname{sgn}(\sigma) \prod_{k=1}^m a_{i_k j_k}, with \operatorname{sgn}(\sigma) the sign of the corresponding permutation. This polynomial in the entries of A satisfies the square relation, providing a square root of the determinant. The Pfaffian was introduced by Arthur Cayley in 1852 and named after the German mathematician Johann Friedrich Pfaff, who studied related systems of differential equations in 1814–1815. It has applications in statistical mechanics, such as computing partition functions for the Ising model on planar graphs via the Pfaffian of an associated .

Spectral Theory

For a real A \in \mathbb{R}^{n \times n}, the eigenvalues are either zero or purely imaginary, occurring in conjugate pairs \pm i \lambda where \lambda > 0. This property arises because iA is symmetric (hence Hermitian over \mathbb{C}), and the for Hermitian matrices guarantees real eigenvalues for iA, implying pure imaginary eigenvalues for A. The p(\lambda) = \det(\lambda I - A) is an even function, satisfying p(\lambda) = p(-\lambda). This follows from the relation \det(\lambda I - A) = \det((\lambda I - A)^T) = \det(\lambda I + A), reflecting the skew-symmetry. Consequently, non-zero eigenvalues appear in \pm pairs, ensuring the algebraic multiplicity of each pair is even. In odd dimensions (n odd), the odd degree of p(\lambda) requires zero to be an eigenvalue with odd multiplicity, implying A is singular. Over the complex numbers, A is diagonalizable, as iA is Hermitian and thus unitarily diagonalizable, yielding eigenvalues of the form i \theta and -i \theta for real \theta. The contains no non-trivial Jordan blocks for non-zero eigenvalues, due to this diagonalizability. The geometric multiplicity matches the algebraic multiplicity for all eigenvalues. In the real canonical form, there exists an Q such that Q^T A Q is block diagonal, consisting of 2×2 rotation-scaled blocks \begin{pmatrix} 0 & \lambda_k \\ -\lambda_k & 0 \end{pmatrix}, \quad \lambda_k > 0, along with 1×1 zero blocks corresponding to the kernel. The dimension of the kernel (nullity, or corank, of A) has the same parity as n; for a non-singular A (which requires n even), this dimension is zero. This form underscores the pairing of non-zero eigenvalues.

Geometric Applications

Cross Product Representation

In three-dimensional Euclidean space, skew-symmetric matrices provide a matrix representation of the cross product operation. For a vector \mathbf{v} = (v_1, v_2, v_3)^\top \in \mathbb{R}^3, the associated skew-symmetric matrix K_{\mathbf{v}} is defined as K_{\mathbf{v}} = \begin{pmatrix} 0 & -v_3 & v_2 \\ v_3 & 0 & -v_1 \\ -v_2 & v_1 & 0 \end{pmatrix}. This matrix satisfies K_{\mathbf{v}} \mathbf{w} = \mathbf{v} \times \mathbf{w} for any \mathbf{w} \in \mathbb{R}^3, where \times denotes the standard cross product. The mapping \mathbf{v} \mapsto K_{\mathbf{v}} is linear over \mathbb{R}, preserving addition and : K_{\mathbf{u} + \mathbf{v}} = K_{\mathbf{u}} + K_{\mathbf{v}} and K_{c \mathbf{v}} = c K_{\mathbf{v}} for \mathbf{u}, \mathbf{v} \in \mathbb{R}^3 and c \in \mathbb{R}. Moreover, this mapping extends to the Lie bracket structure, establishing an of Lie algebras between \mathfrak{so}(3), the space of $3 \times 3 skew-symmetric matrices under the bracket, and \mathbb{R}^3 under the . A useful norm relation for this representation is the Frobenius norm \|K_{\mathbf{v}}\|_F = \sqrt{2} \|\mathbf{v}\|_2, where \|\cdot\|_2 is the Euclidean norm on \mathbb{R}^3; this follows from squaring the off-diagonal entries of K_{\mathbf{v}} and summing under the Frobenius norm definition \|A\|_F = \sqrt{\sum_{i,j} a_{ij}^2}. While analogous representations arise in higher dimensions through the adjoint action of \mathfrak{so}(n) on \mathbb{R}^n, the direct equivalence to a cross product is unique to three dimensions.

Infinitesimal Rotations

Skew-symmetric matrices form the \mathfrak{so}(n) of the special \mathrm{SO}(n), consisting of all n \times n real skew-symmetric matrices under the Lie bracket [A, B] = AB - BA. This bracket captures the infinitesimal composition of rotations, reflecting the non-commutative nature of the . The \exp: \mathfrak{so}(n) \to \mathrm{SO}(n), defined by the matrix exponential \exp(A) = \sum_{k=0}^\infty \frac{A^k}{k!}, associates each skew-symmetric matrix A with a , providing a local parametrization of the group near the identity. The basis elements of \mathfrak{so}(n) serve as infinitesimal generators of rotations in specific planes. For each pair of distinct indices i < j, the matrix E_{ij} with a 1 in position (i,j), a -1 in position (j,i), and zeros elsewhere generates an rotation in the i-j . These n(n-1)/2 basis matrices \mathfrak{so}(n), and any general is a \sum_{i<j} \theta_{ij} E_{ij}, corresponding to simultaneous rotations by angles \theta_{ij} in each . The n(n-1)/2 matches the number of independent rotation parameters in \mathrm{SO}(n). In physics, particularly , operators play the role of generators for but are Hermitian to ensure real eigenvalues for observables. The generators are then skew-Hermitian, related to these operators by a factor of i (up to constants like \hbar), linking the real skew-symmetric structure of \mathfrak{so}(3) to the complex skew-Hermitian \mathfrak{su}(2), which is isomorphic. This connection underlies the representation of in , where the commutation relations [J_x, J_y] = i \hbar J_z (and cyclic permutations) mirror the structure. Noether's theorem establishes that rotational invariance of the in physical systems implies the conservation of , with skew-symmetric matrices encoding the associated symmetries. For a system invariant under continuous rotations, the theorem yields a conserved current corresponding to the angular momentum tensor, whose components are generated by elements of \mathfrak{so}(n). This principle explains the persistence of in isolated systems, from to .

Coordinate-Free Perspective

In the coordinate-free framework, a skew-symmetric matrix arises as the matrix representation of an alternating on a finite-dimensional V over a F of not equal to 2. An alternating \omega: V \times V \to F is a satisfying \omega(v, v) = 0 for all v \in V, which equivalently implies \omega(u, v) = -\omega(v, u) for all u, v \in V. The space of all such forms, denoted \mathrm{Alt}^2(V; F), is naturally isomorphic to the second exterior power \Lambda^2 V^* of the V^*, where elements of \Lambda^2 V^* are skew-symmetric contravariant 2-tensors, or bivectors in the dual sense. Given a basis \{e_1, \dots, e_n\} for V, the associated dual basis \{e^1, \dots, e^n\} for V^* yields the coordinate matrix A = (a_{ij}) of \omega, defined by a_{ij} = \omega(e_i, e_j). This matrix satisfies A^T = -A, confirming its skew-symmetry, and the form can be expressed in the exterior algebra as \omega = \sum_{i < j} a_{ij} \, e^i \wedge e^j. This identification highlights that skew-symmetric matrices parametrize the components of bivectors in \Lambda^2 V^*, emphasizing the intrinsic tensorial nature independent of any specific inner product. Under a specified by an P \in \mathrm{GL}(n, F), where the new basis vectors are the columns of P applied to the old basis, the matrix representation transforms via the A' = P^T A P. This law preserves skew-symmetry, as (P^T A P)^T = P^T A^T P = -P^T A P, ensuring the abstract form \omega remains unchanged while its coordinates adapt to the new . Certain quantities are basis-independent: the trace of A is always zero, reflecting \omega(e_i, e_i) = 0 for all basis vectors, and the transforms as \det(A') = \det(P)^2 \det(A), rendering it up to positive scaling in contexts where basis changes preserve up to sign, such as orthogonal transformations. This perspective connects directly to the study of alternating forms, where \omega defines pairings on V.

Connections to Forms

Alternating Bilinear Forms

A skew-symmetric matrix A \in M_n(\mathbb{R}) (or more generally over a of not equal to 2) defines an alternating \omega: \mathbb{R}^n \times \mathbb{R}^n \to \mathbb{R} by \omega(\mathbf{u}, \mathbf{v}) = \mathbf{u}^T A \mathbf{v}. This form is bilinear by the properties of and alternating because \omega(\mathbf{v}, \mathbf{v}) = \mathbf{v}^T A \mathbf{v} = 0 (since A^T = -A implies the vanishes) and \omega(\mathbf{u}, \mathbf{v}) = -\omega(\mathbf{v}, \mathbf{u}). The map sending a skew-symmetric matrix A to the alternating bilinear form \omega(\mathbf{u}, \mathbf{v}) = \mathbf{u}^T A \mathbf{v} is an isomorphism between the vector space of n \times n skew-symmetric matrices and the space of alternating bilinear forms on \mathbb{R}^n. Under this identification, the matrix A is the coordinate matrix of \omega with respect to the standard basis. The form \omega is non-degenerate—meaning \omega(\mathbf{u}, \mathbf{v}) = 0 for all \mathbf{v} implies \mathbf{u} = \mathbf{0}—if and only if the matrix A is invertible. Equivalently, the radical of \omega (the set of \mathbf{u} such that \omega(\mathbf{u}, \mathbf{v}) = 0 for all \mathbf{v}) is trivial precisely when \ker A = \{\mathbf{0}\}. In even dimensions n = 2m, a non-degenerate alternating bilinear form is known as a symplectic form, and linear transformations preserving such a form maintain the volume up to sign, with the determinant being 1 for the standard symplectic group. The Pfaffian of A, satisfying \operatorname{pf}(A)^2 = \det A, relates to the oriented volume induced by the form. Any alternating bilinear form admits a canonical form under basis change: if the rank is $2k (always even), there is a basis in which the matrix is block diagonal with k copies of the $2 \times 2 block \begin{pmatrix} 0 & 1 \\ -1 & 0 \end{pmatrix} along the diagonal and zeros elsewhere. For non-degenerate cases in even dimensions, the Darboux theorem guarantees a basis (called a symplectic or Darboux basis) where the matrix takes the standard block form consisting of m such $2 \times 2 blocks.

Relation to Orthogonal and Symplectic Groups

Skew-symmetric matrices constitute the \mathfrak{so}(n, \mathbb{R}) of the O(n, \mathbb{R}), which comprises all n \times n real matrices g satisfying g^\top g = I_n, thereby preserving the standard \langle u, v \rangle = u^\top v. This preservation condition implies that for a one-parameter g(t) with g(0) = I_n and g'(0) = A, yields A^\top + A = 0, confirming that elements of the are precisely the skew-symmetric matrices. The \exp: \mathfrak{so}(n, \mathbb{R}) \to SO(n, \mathbb{R}) associates each skew-symmetric matrix A with a \exp(A), which preserves the associated q(x) = x^\top x. In even dimensions $2n, skew-symmetric matrices also play a central role in defining alternating bilinear forms preserved by the Sp(2n, \mathbb{R}), the subgroup of GL(2n, \mathbb{R}) consisting of matrices g such that g^\top J g = J, where J is the standard J = \begin{pmatrix} 0 & I_n \\ -I_n & 0 \end{pmatrix}, a block-diagonal skew-symmetric matrix representing the non-degenerate alternating form \omega(u, v) = u^\top J v. The \mathfrak{sp}(2n, \mathbb{R}) consists of all $2n \times 2n matrices X satisfying X^\top J + J X = 0, derived from the infinitesimal preservation condition. Although elements of \mathfrak{sp}(2n, \mathbb{R}) are not skew-symmetric in general, the fixed matrix J underscores how skew-symmetric matrices parameterize such alternating forms. The orthogonal and symplectic groups thus arise as preservers of distinct classes of bilinear forms: symmetric forms for the former (with Lie algebra skew-symmetric relative to the identity) and alternating forms for the latter (with Lie algebra satisfying the twisted skew-symmetry X^\top J + J X = 0). For a general invertible skew-symmetric matrix A, the group \{ g \in GL(2n, \mathbb{R}) \mid g^\top A g = A \} is isomorphic to Sp(2n, \mathbb{R}) via change of basis, preserving the associated non-degenerate alternating form \omega(u, v) = u^\top A v; if A is singular, this group enlarges accordingly, corresponding to a degenerate case.

Generalizations

Skew-Symmetrizable Matrices

A skew-symmetrizable matrix is a square real A for which there exists a D with positive diagonal entries such that D A D^{-1} is skew-symmetric. This condition implies that A is similar to a skew-symmetric matrix via a , generalizing the class of skew-symmetric matrices (which correspond to the special case D = I). The eigenvalues of a skew-symmetrizable matrix are either zero or purely imaginary, occurring in pairs, mirroring the spectral properties of skew-symmetric matrices due to the similarity. These matrices are also linked to specific patterns in combinatorial matrix theory, where the existence of such a D requires compatible signs in off-diagonal entries to allow balancing via positive scalings. In applications, skew-symmetrizable matrices arise in stability analysis of dynamical systems, such as epidemic models, where the Jacobian matrix satisfying a weighted ensures global asymptotic under certain positivity assumptions. They also connect to generalized inverses, where the structure facilitates computations analogous to those for skew-symmetric cases, and to totally nonnegative matrices through combinatorial frameworks like , where skew-symmetrizable exchange matrices underpin classifications of finite-type structures.

Matrices over Complex Numbers

Over the complex numbers, a skew-symmetric matrix is defined as a square matrix A \in M_n(\mathbb{C}) satisfying A^T = -A, where ^T denotes the transpose without conjugation. This definition parallels the real case but lacks the additional structure imposed by reality of entries, leading to distinct spectral behavior. Unlike real skew-symmetric matrices, whose eigenvalues are purely imaginary or zero, complex skew-symmetric matrices can have eigenvalues with nonzero real parts. The characteristic polynomial p_A(\lambda) = \det(\lambda I - A) satisfies p_A(\lambda) = p_A(-\lambda), implying that nonzero eigenvalues occur in pairs \lambda and -\lambda, where \lambda \in \mathbb{C} need not lie on the imaginary axis. For instance, consider the $2 \times 2 A = \begin{pmatrix} 0 & 1+i \\ -(1+i) & 0 \end{pmatrix}, which is skew-symmetric since A^T = -A. Its is \lambda^2 + (1+i)^2 = \lambda^2 + (1 + 2i - 1) = \lambda^2 + 2i, with \lambda = \pm \sqrt{-2i} = \pm (1 - i), eigenvalues having real part 1 and -1, respectively. Complex skew-symmetric matrices are not necessarily diagonalizable over \mathbb{C}; blocks can appear, particularly for the eigenvalue 0, though blocks for paired eigenvalues \lambda and -\lambda must have matching structures. The possible canonical forms are constrained by the skew-symmetry, often involving blocks of even size for nonzero eigenvalues. A key distinction arises in comparison to skew-Hermitian matrices, defined by A^* = -A where ^* is the . Skew-Hermitian matrices serve as the of real skew-symmetric matrices, inheriting properties like purely imaginary eigenvalues and unitarily diagonalizable forms (up to the factor of i). Specifically, if K is skew-Hermitian, then iK is Hermitian, with real eigenvalues. In contrast, complex skew-symmetric matrices do not generally satisfy A^* = -A unless their entries are purely imaginary, and thus lack these Hermitian-like guarantees. In applications, complex skew-symmetric matrices play a role in the structure of complex , particularly as the defining elements of \mathfrak{so}(n, \mathbb{C}), the Lie algebra of the complex O(n, \mathbb{C}), consisting of all n \times n skew-symmetric matrices over \mathbb{C}. This algebra is semisimple for n \geq 3 and intersects with \mathfrak{sl}(n, \mathbb{C}) (trace-zero matrices) in a of 1 when n is odd. However, due to the prevalence of Hermitian structures in and other contexts, skew-symmetric matrices over \mathbb{C} receive less attention in the literature compared to their real or skew-Hermitian counterparts.

References

  1. [1]
    [PDF] Properties of Matrix Arithmetic
    Definition. (a) A matrix X is symmetric if XT = X. (b) A matrix X is skew symmetric if XT = −X. Both definitions imply that X is a square matrix. Using the ...
  2. [2]
    [PDF] Chapter 3. Basic Properties of Matrices
    Jun 15, 2020 · Note. A skew symmetric matrix must satisfy aii = −aii for 1 ≤ i ≤ n and so is. necessarily hollow. Page 3.
  3. [3]
    [PDF] Notes Matrix and Linear Algebra - University of Washington
    Any real square matrix A may be written as the sum of a symmetric matrix R and a skew- symmetric matrix S, where. R = 1. 2. A + AT , S = 1. 2. A − AT. If A ...
  4. [4]
    [PDF] FIRST-ORDER LINEAR SYSTEMS WHOSE COEFFICIENT ...
    If the dimension n of the space is odd, a (real) skew-symmetric n × n matrix always has zero as a real. eigenvalue (with a real eigenvector). The reason is ...
  5. [5]
    [PDF] COMPUTING EXPONENTIALS OF SKEW-SYMMETRIC MATRICES ...
    Observe that the eigenvalues of B are ±iθj, or 0, recon- firming the well-known fact that the eigenvalues of a skew-symmetric matrix are purely imaginary, or ...
  6. [6]
    [PDF] CHAPTER 7 - Purdue Engineering
    We now show that infinitesimal rotations commute!! [ ] [ ][ ] [ ][ ]. 2. 1. 1. 2. Φ = Φ Φ ... Note: is a skew symmetric matrix consider the matrix: r r ω.
  7. [7]
    [PDF] May 20, 1995
    May 20, 1995 · 50(3) = TI SO(3). Consider a one-parameter family of infinitesimal rotations & ε @ & so (3) ... given skew-symmetric matrix € so (3). Additionally ...
  8. [8]
    [PDF] 1. Properties of antisymmetric matrices
    Finally, if M is an odd-dimensional complex antisymmetric matrix, the corresponding pfaffian is defined to be zero. The pfaffian and determinant of an ...
  9. [9]
    Antisymmetric Matrix -- from Wolfram MathWorld
    An antisymmetric matrix, also known as a skew-symmetric or antimetric matrix, is a square matrix that satisfies the identity A=-A^(T) where A^(T) is the matrix ...
  10. [10]
    Skew-symmetric matrix - Encyclopedia of Mathematics
    Sep 12, 2016 · A square matrix A over a field of characteristic ≠2 such that AT=−A. The rank of a skew-symmetric matrix is an even number.Missing: primary | Show results with:primary
  11. [11]
    [PDF] bilinear forms - keith conrad
    ... not in characteristic. 2) skew-symmetric. The only bilinear form that is both symmetric and skew-symmetric outside of characteristic 2 is identically zero.
  12. [12]
    Skew-Symmetric Matrix - an overview | ScienceDirect Topics
    A matrix is skew-symmetric if and only if it is the opposite of its transpose. All main diagonal entries of a skew-symmetric matrix are zero. □. Every square ...
  13. [13]
    Symmetric Matrix Properties and Applications: A Guide | Built In
    In a skew-symmetric matrix, the condition Aᵀ = -A is met, plus all main diagonal entries are zero and the matrix's trace equals zero. The sum of two skew- ...
  14. [14]
    All About Skew Symmetric Matrix - Shiksha Online
    Apr 12, 2024 · A skew-symmetric matrix is a square matrix whose transpose is equal to its negative. In other words, it satisfies the condition AT = -A.
  15. [15]
    Products of symmetric and skew-symmetric Matrices
    Apr 1, 2008 · The class of matrices which can be represented as products of two matrices, each of which is either symmetric or skew-symmetric, is identified.
  16. [16]
    [PDF] The Matrix Cookbook
    Nov 15, 2012 · Symmetric matrices have many important properties, e.g. that their eigenvalues are real and eigenvectors orthogonal. 9.8.2 Skew-symmetric/ ...<|control11|><|separator|>
  17. [17]
    [PDF] Some definitions from linear algebra A matrix m-by-n is a table of ...
    In particular, the diagonal elements of a skew-symmetric matrix are equal to zero. A singular matrix is a square matrix with zero determinant.
  18. [18]
    [PDF] Quadratic Polynomials - Penn Math
    A matrix C is called skew-symmetric if CT = −C . (a) Give an example of a skew-symmetric 2 × 2 matrix (other than the 0 matrix). (b) Show that C is skew- ...
  19. [19]
    Math21b, Fall 2003, Linear Algebra and Differential Equations
    For (1x1) matrices, this is the usual exponential function. Verify that if A is a skew symmetric matrix, then eA is an orthogonal matrix. Hint. Check first ...
  20. [20]
    [PDF] Cross-Products and Rotations
    Nov 23, 2003 · k/k! . Now, every 3-by-3 skew-symmetric matrix S determines a vector v such that S = v. ¢. ; then. Sv = o , S. 2. = vv. T. – ||v||. 2. ·I , S. 3.
  21. [21]
    [PDF] BILINEAR FORMS The geometry of Rn is controlled algebraically by ...
    Among the three types of bilinear forms we have defined (symmetric, skew-symmetric, alternating), the first and third types are more basic than the second.
  22. [22]
    [PDF] Exponentials and Rotations - UMD MATH
    Jul 21, 2021 · If A is a skew-symmetric n × n matrix, then eA is orthog- onal. Additionally, det(eA) = 1, so eA is a rotation matrix. So what does a skew- ...
  23. [23]
    [PDF] Vector spaces - Purdue Math
    Aug 23, 2024 · Dimension of the space of skew-symmetric matrices of size n × n is n(n −. 1)/2. The main use of a basis is the possibility to represent ...
  24. [24]
    [PDF] Chapter 10 Euclidean Spaces - UPenn CIS
    define the inner product of two matrices A, B 2 Mn(R) as. hA, Bi = n. X i,j=1 ... The (real) vector space of n⇥n skew symmetric matrices is denoted by so(n).
  25. [25]
    [PDF] Midterm, October 18, 2011, with solutions Linear Algebra I Cross out ...
    The dimension of the space equals the number of elements above the diagonal, i.e., equals n(n − 1)/2. A basis elements can be chosen to vanish except for ...
  26. [26]
    CS 6210: Matrix Computations
    The two spaces are orthogonal to each other in the Frobenius inner product; and for any matrix A ∈ R n × n , there is a unique decomposition into a symmetric ...
  27. [27]
    A note on skew-symmetric determinants - Cambridge University Press
    There are two classical results about a skew-symmetric matrix A: (I) When n is odd, then deM = 0. (II) When n is even, then det A = (pn(A))2, where pn(A) is a ...
  28. [28]
    Pfaffian | Wolfram Function Repository
    The Pfaffian of a skew-symmetric matrix m is an integer-coefficient polynomial in the entries of m whose square is the determinant of m.
  29. [29]
  30. [30]
    Graph theory and Pfaffian representations of Ising partition function
    Dec 27, 2013 · The partition function of an Ising model on an arbitrary planar graph can be represented as the Pfaffian of a skew-symmetric matrix associated to the graph.
  31. [31]
    [PDF] Chapter 6 Eigenvalues and Eigenvectors
    A symmetric matrix (ST = S) can be compared to a real number. A skew-symmetric matrix (AT = −A) can be compared to an imaginary number. An orthogonal matrix. ( ...
  32. [32]
    Eigenvalues of Real Skew-Symmetric Matrix are Zero or Purely ...
    We prove that eigenvalues of a real skew-symmetric matrix are zero or purely imaginary and the rank of the matrix is even. We use the diagonalization of ...
  33. [33]
    [PDF] Spectral Theorems in Euclidean and Hermitian Spaces - CIS UPenn
    Spectral Theorems. Theorem 11.4.4 For every skew symmetric matrix A there is an orthog- onal matrix P and a block diagonal matrix D such that A = PDP. , where.Missing: theory | Show results with:theory
  34. [34]
    [PDF] A QUICK NOTE ON ORTHOGONAL LIE ALGEBRAS
    (R3, ×) is a Lie algebra, which is isomorphic to so(3,R). Proof: We have already mentioned that × is skew-symmetric. Moreover, the cyclic sum over x, y and ...
  35. [35]
    [PDF] Lie Groups and Mechanics: an introduction - arXiv
    The Lie algebra so(n) of the rotation group SO(n) consists of skew-symmetric n × n matrices. 0. 2.2. Adjoint representation of G. The composition Ig = Rg−1 ...
  36. [36]
    [PDF] Lecture Notes in Lie Groups - arXiv
    Apr 7, 2011 · so(n, R) skew– symmetric square real matrices, with. Lie bracket the ... unitary matrices; the Lie algebra will then map to skew–Hermitian ...
  37. [37]
    [PDF] arXiv:2010.05020v1 [math.OC] 10 Oct 2020
    Oct 10, 2020 · Any matrix Q ∈ SO(N) can be represented as a matrix ex- ponential function of a skew-symmetric matrix. The converse is also true: the.
  38. [38]
    [PDF] Introduction to Lie Groups and Lie Algebras Alexander Kirillov, Jr.
    ... Lie algebra with the commutator given by the cross-product, is isomorphic (as a Lie algebra) to so(3, R). Page 37. Exercises. 37. (2) Let ϕ: so(3, R) → R3 be ...
  39. [39]
    [PDF] Ensemble Control on Lie Groups - arXiv
    Aug 7, 2020 · Basics of the Lie algebra so(n). The Lie algebra so(n) is the vector space containing all n × n real skew-symmetric matrices, which has ...
  40. [40]
    [PDF] Lectures 6 and 7 SO(3) and Conserved Quantities
    The generators of the group are the angular momentum operators. • Note that these operators obey the same commutation relation as the three generators did for.
  41. [41]
    [PDF] Lecture 14 Angular momentum operator algebra
    Consider the three Hermitian angular momentum operators. ˆ. Jx,. ˆ. Jy and ˆJz ... commutation relations (14.1) represent the Lie algebra of the SO(3) group.Missing: skew- | Show results with:skew-
  42. [42]
    [PDF] Noether's Theorem - Physics Courses
    This is the conserved angular momentum about the z axis. If instead the particle moved in a potential U(y), independent of x, then writing. L = 1.
  43. [43]
    [PDF] Conservation Laws
    ◇ translation invariance leads to linear momentum conservation. ◇ rotational invariance leads to angular momentum conservation. □ Familiar conserved ...<|control11|><|separator|>
  44. [44]
    [PDF] 4 Exterior algebra - People
    We can add alternating forms and multiply by scalars so they form a vector space, isomorphic to the space of skew-symmetric n × n matrices. ... 2-vector of ...
  45. [45]
    [PDF] Bilinear forms and their matrices
    Mar 11, 2011 · A bilinear form H is called symmetric if H(v, w) = H(w, v) for all v, w ∈ V . A bilinear form H is called skew-symmetric if H(v, w) = −H(w, v) ...
  46. [46]
    [PDF] Skew-symmetric bilinear forms - Fachbereich Mathematik und Statistik
    Dec 9, 2020 · Let f be a skew-symmetric bilinear form f is alternating if and only if char(K) 6= 2. Proof: Let f be any skew-symmetric bilinear form over a ...Missing: coordinate- free
  47. [47]
    [PDF] bilinear forms
    If <-, -> is alternating, then we know. 0 = <v + w,v + w> = <v,v> + <v, w> + <w,v> + <w, w> = <v, w> + <w,v>. Thus <-, -> is skew-symmetric. If the ...
  48. [48]
    [PDF] 1. Linear algebra preliminaries 1.1. Some facts about bilinear forms ...
    Skew-symmetric bilinear forms. Proposition. Let ψ be a skew-symmetric bilinear form. Then, there exists a basis with respect to which ψ is given by the matrix.
  49. [49]
    Appendix D The Orthogonal Group and Its Lie Algebra
    identifies the skew-symmetric linear maps as the Lie algebra o(d) of O(d) ... The Lie bracket for o(d) is the usual bracket of matrices [ξ1,ξ2] = ξ1ξ2 − ξ2ξ1.<|control11|><|separator|>
  50. [50]
    [PDF] The Exponential Map, Lie Groups, and Lie Algebras - UPenn CIS
    The Lie algebra so(n,R) consisting of real skew symmet- ric n×n matrices is the corresponding set of infinitesimal rotations. The geometric link between a Lie ...
  51. [51]
    [PDF] 18.745 F20 Lecture 06: Classical Lie Groups - MIT OpenCourseWare
    An important role in the theory of Lie groups is played by the algebra of quaternions, which is the only noncommu- tative finite dimensional division algebra ...<|control11|><|separator|>
  52. [52]
    Least Squares Symmetrizable Solutions for a Class of Matrix ...
    Definition 1. A real matrix A is called a symmetrizable (skew-symmetrizable) matrix if A is similar to a symmetric (skew-symmetric) matrix A.
  53. [53]
    Mutation classes of skew-symmetrizable 3×3 matrices
    Sep 27, 2012 · Let us recall that an integer matrix B is skew-symmetrizable if DB is skew-symmetric for some diagonal matrix. D with positive diagonal entries.
  54. [54]
    [PDF] cluster algebras of finite type and positive symmetrizable matrices
    The mutation-equivalence class of a skew-symmetrizable matrix B satisfies the equivalent conditions in Theorem 1.1 if and only if B satisfies: (4) Every ...
  55. [55]
    [PDF] Canonical forms for complex matrix congruence and ... - arXiv
    Sep 16, 2007 · Any square complex matrix A can be represented uniquely as A = S +C, in which S is symmetric and C is skew-symmetric; it can also be represented.
  56. [56]
    The Jordan Canonical Forms of complex orthogonal and skew ...
    We study the Jordan Canonical Forms of complex orthogonal and skew-symmetric matrices, and consider some related results.
  57. [57]
    Canonical form of complex skew symmetric matrix - MathOverflow
    Jul 9, 2021 · ... Horn and Johnson's matrix analysis. So do there exists canonical form of unitary similarity transformation for complex skew symmetric matrix?
  58. [58]
    Hermitian and Skew-Hermitian Matrix - GeeksforGeeks
    Aug 19, 2025 · A complex square matrix is said to be a skew-Hermitian matrix if the conjugate transpose matrix is equal to the negative of the original matrix.
  59. [59]
    [PDF] Classification of Semisimple Lie Algebras
    For completeness, I will list the remaining classical algebras. The. Lie algebra so(n) consists of skew-symmetric matrices. Cartan's classification of so(2n ...