Fact-checked by Grok 2 weeks ago

Bilinear form

In , particularly in linear algebra, a bilinear form on a V over a F is a B: V \times V \to F that is linear in each of its two arguments separately, meaning B(av + bw, u) = aB(v, u) + bB(w, u) and B(u, av + bw) = aB(u, v) + bB(u, w) for all scalars a, b \in F and vectors u, v, w \in V. Bilinear forms generalize the concept of inner products and play a central role in encoding geometric and algebraic structures on vector spaces. They can be represented by matrices with respect to a basis of V: if \{e_1, \dots, e_n\} is a basis, then B(v, w) = ^T M , where M = (B(e_i, e_j)) is an n \times n matrix over F, and , are the coordinate vectors of v and w. Key properties include symmetry, where B(v, w) = B(w, v) for all v, w \in V (corresponding to a symmetric matrix M^T = M), and skew-symmetry or alternation, where B(v, w) = -B(w, v) or B(v, v) = 0 for all v, w \in V (corresponding to a skew-symmetric matrix M^T = -M with zero diagonal). A bilinear form is non-degenerate if its matrix is invertible, ensuring no nonzero vector annihilates the form, which is crucial for preserving dimensions in associated dual spaces. Symmetric bilinear forms are intimately linked to quadratic forms: given a quadratic form Q: V \to F satisfying Q(cv) = c^2 Q(v) and the parallelogram law, there is a unique associated symmetric bilinear form via the polarization identity B(v, w) = \frac{1}{4} [Q(v + w) - Q(v - w)], and conversely Q(v) = B(v, v) when the characteristic of F is not 2. Positive definite symmetric bilinear forms, such as the standard dot product on \mathbb{R}^n, define norms \|v\| = \sqrt{B(v, v)} and angles, enabling the spectral theorem for symmetric matrices, which guarantees orthonormal bases of eigenvectors. Beyond linear algebra, bilinear forms appear in diverse applications, including (e.g., metrics on spaces), representation theory (e.g., invariant forms on Lie algebras like the Killing form), and (e.g., in estimates involving Kloosterman sums and modular forms). Their study extends to classification over specific fields, such as real or finite fields, influencing topics from quadratic space decompositions to arithmetic progressions of primes.

Basic Concepts

Definition

A bilinear form on a vector space V over a field F is a mapping B: V \times V \to F that is linear in each argument when the other is held fixed. This bilinearity is expressed by the following axioms: for all scalars a, b \in F and all vectors u, v, w \in V, \begin{align} B(au + bv, w) &= a B(u, w) + b B(v, w), \\ B(u, av + bw) &= a B(u, v) + b B(u, w). \end{align} These conditions ensure additivity and homogeneity in each variable separately. The set of all such bilinear forms on V is denoted \mathrm{Bil}(V), which itself forms a vector space under pointwise addition and scalar multiplication. More generally, one may consider bilinear maps B: V \times W \to F between two vector spaces, but the standard case focuses on V = W. Typically, V is taken over the real numbers \mathbb{R} or complex numbers \mathbb{C}, with no additional structure such as a imposed at the outset. The concept assumes only the algebraic framework of a , allowing bilinear forms to generalize structures like the without requiring metric properties. The notion of bilinear forms traces back to the late in Joseph-Louis Lagrange's implicit use of matrices to study quadratic forms, and the term was introduced in the context of during the early 20th century as vector spaces were formalized.

Examples

A fundamental example of a symmetric bilinear form is the standard on the \mathbb{R}^n, defined by B(\mathbf{u}, \mathbf{v}) = \sum_{i=1}^n u_i v_i for \mathbf{u} = (u_1, \dots, u_n) and \mathbf{v} = (v_1, \dots, v_n). This form arises naturally in the study of inner product spaces and captures the notion of when B(\mathbf{u}, \mathbf{v}) = 0. Another illustrative example is the determinant form on the \mathbb{R}^2, given by B(\mathbf{x}, \mathbf{y}) = \det\begin{pmatrix} x_1 & y_1 \\ x_2 & y_2 \end{pmatrix} = x_1 y_2 - x_2 y_1, where \mathbf{x} = (x_1, x_2) and \mathbf{y} = (y_1, y_2). This alternating bilinear form, which changes sign upon swapping arguments, is used to define oriented areas in the plane and exemplifies non-symmetric behavior. In the context of , the Minkowski metric provides an example of an indefinite bilinear form on \mathbb{R}^{1,3}, defined by B(\mathbf{u}, \mathbf{v}) = u_1 v_1 + u_2 v_2 + u_3 v_3 - u_4 v_4, where the indices correspond to spatial and temporal components. This form, with signature (3,1), measures and length in four-dimensional while allowing both positive and negative values. The zero form, B(\mathbf{u}, \mathbf{v}) = 0 for all \mathbf{u}, \mathbf{v} in a , represents the trivial bilinear form, which satisfies bilinearity but lacks non-trivial structure. Scalar multiples of non-trivial forms, such as B(\mathbf{u}, \mathbf{v}) = c \sum_{i=1}^n u_i v_i for a constant c \in \mathbb{R}, preserve bilinearity and scale the original form's values without altering its qualitative properties. Over finite fields, bilinear forms exhibit field-dependent behavior; for instance, on the vector space \mathbb{F}_{q^n} over \mathbb{F}_q, the trace form B(a, b) = \mathrm{Tr}_{\mathbb{F}_{q^n}/\mathbb{F}_q}(a b) uses the field trace to pair extension elements bilinearly. This form is symmetric and plays a key role in coding theory and the classification of forms in characteristic p.

Representation

Coordinate representation

In finite-dimensional vector spaces over a F, bilinear forms admit a coordinate representation with respect to a chosen basis. Let V be an n-dimensional with basis \{ e_1, \dots, e_n \}. For a bilinear form B: V \times V \to F, the associated A = (a_{ij})_{1 \leq i,j \leq n} is defined by a_{ij} = B(e_i, e_j). This matrix fully determines B, as the bilinearity ensures that for any u = \sum_{i=1}^n u_i e_i and v = \sum_{j=1}^n v_j e_j, the value is B(u, v) = \sum_{i=1}^n \sum_{j=1}^n u_i a_{ij} v_j = \mathbf{u}^T A \mathbf{v}, where \mathbf{u} = (u_1, \dots, u_n)^T and \mathbf{v} = (v_1, \dots, v_n)^T are the coordinate column vectors. To compute the matrix A, evaluate B directly on the basis vectors: the (i,j)-entry is simply B(e_i, e_j). This yields a straightforward procedure, as only n^2 evaluations are needed, after which the form on arbitrary vectors follows from expanding in coordinates and applying bilinearity. For instance, in the of F^n, the matrix entries correspond to the coefficients in the expression B(\mathbf{u}, \mathbf{v}) = \sum_{i,j} a_{ij} u_i v_j. The representation changes under a basis transformation. Suppose \{ f_1, \dots, f_n \} is a new basis, and let P be the invertible change-of-basis whose k-th column consists of the coordinates of f_k with respect to the old basis \{ e_i \}. Then, the coordinate vectors transform as \mathbf{u} = P \mathbf{u}' and \mathbf{v} = P \mathbf{v}', where primes denote new coordinates. Substituting gives B(u, v) = (P \mathbf{u}')^T A (P \mathbf{v}') = {\mathbf{u}'}^T (P^T A P) \mathbf{v}', so the new is A' = P^T A P. This transformation preserves the bilinear form up to the choice of basis. Such matrix representations rely on the finite dimensionality of V, which ensures a finite basis and thus a finite . In infinite-dimensional spaces, a similar coordinate approach is possible using a Hamel basis, leading to an infinite with entries B(e_i, e_j), but these are typically unwieldy due to the lack of finite support in expansions, and practical treatments often defer to operator-theoretic or kernel representations instead.

Matrix properties

The rank of a bilinear form B on a finite-dimensional V over a F is the of its A with respect to any basis of V. This quantity, denoted \operatorname{rank}(B), is independent of the choice of basis and equals the of the image of the associated \phi_B: V \to V^*, where V^* is the of V and \phi_B(v)(w) = B(v, w) for all v, w \in V. In the symmetric case with \operatorname{char} F \neq 2, the also corresponds to the number of nonzero entries in a obtained by congruence transformation. Two matrices A and B' represent the same bilinear form on V with respect to different bases if and only if B' = P^T A P for some P; matrices related in this manner are said to be . preserves essential properties of the bilinear form, such as its , and provides a framework for classifying forms up to . For symmetric bilinear forms over the real numbers, Sylvester's law of further refines this classification: the of the form, expressed as the triple (p, q, r) where p is the number of positive eigenvalues of A, q the number of negative eigenvalues, and r the multiplicity of the zero eigenvalue (with p + q + r = \dim V), remains invariant under . The of the form is the pair (p, q), which captures the "indefiniteness" structure. Symmetric bilinear forms admit diagonalization by congruence over fields of characteristic not equal to 2, meaning there exists a basis of V in which the matrix representation is diagonal. This holds in particular over algebraically closed fields, such as the complex numbers, where the diagonal entries can often be normalized to 1 for nondegenerate forms. For nondegenerate bilinear forms, the associated matrix A has full , implying \det A \neq 0. In the real symmetric nondegenerate case, the sign of the is (-1)^q, where q is the number of negative eigenvalues (positive if q even, negative if q odd), but the provides a more complete invariant than the determinant alone.

Core Properties

Non-degeneracy

A bilinear form B: V \times V \to F on a vector space V over a field F is defined to be non-degenerate, or regular, if the associated linear map \phi_B: V \to V^*, given by \phi_B(u)(v) = B(u, v) for all u, v \in V, is injective; this is termed left non-degeneracy. Equivalently, the left kernel of B, denoted \{u \in V \mid B(u, v) = 0 \ \forall v \in V\}, is the zero subspace \{0\}. Similarly, right non-degeneracy holds if the map \psi_B: V \to V^*, defined by \psi_B(v)(u) = B(u, v), is injective, which is equivalent to the right kernel \{v \in V \mid B(u, v) = 0 \ \forall u \in V\} being \{0\}. When both left and right non-degeneracy conditions are satisfied, B is said to be non-degenerate in the proper sense. The of B, denoted \mathrm{rad}(B), is the of the left and the right of B. Thus, B is non-degenerate \mathrm{rad}(B) = \{0\}. This unifies the left and right conditions, emphasizing the form's ability to pair vectors non-trivially in both directions. In the finite-dimensional case, where \dim V = n < \infty, the spaces V and V^* have equal dimension n, so left non-degeneracy (injectivity of \phi_B) implies that \phi_B is an isomorphism, and similarly for right non-degeneracy. Consequently, B is non-degenerate \phi_B (or equivalently \psi_B) is an isomorphism from V to V^*. With respect to a basis of V, this holds the matrix A representing B has full rank n, i.e., \mathrm{rank}(A) = n. Representative examples illustrate these concepts. The standard dot product on \mathbb{R}^n, defined by B(\mathbf{x}, \mathbf{y}) = \mathbf{x} \cdot \mathbf{y} = \sum_{i=1}^n x_i y_i, is non-degenerate, as its left kernel (and right kernel, since it is symmetric) is \{[0](/page/0)\}—if \mathbf{x} \cdot \mathbf{y} = 0 for all \mathbf{y} \in \mathbb{R}^n, then \mathbf{x} = \mathbf{0}. In contrast, the zero form B(\mathbf{u}, \mathbf{v}) = 0 for all \mathbf{u}, \mathbf{v} \in V is degenerate, with both kernels equal to the entire space V and radical V \neq \{0\} (unless \dim V = 0).

Symmetry types

Bilinear forms are classified into symmetry types based on their behavior under interchange of arguments. A bilinear form B: V \times V \to F on a vector space V over a field F is symmetric if B(u, v) = B(v, u) for all u, v \in V. In terms of matrix representation with respect to a basis, a symmetric bilinear form corresponds to a symmetric matrix A, satisfying A^T = A. A bilinear form is skew-symmetric (or antisymmetric) if B(u, v) = -B(v, u) for all u, v \in V. The associated matrix representation is then skew-symmetric, with A^T = -A. Skew-symmetric forms satisfy B(u, u) = 0 for all u \in V, but the converse holds only over fields of characteristic not equal to 2. An alternating bilinear form is defined by the condition B(u, u) = 0 for all u \in V. Over fields of characteristic not equal to 2, every alternating form is skew-symmetric, but the converse is false in characteristic 2, where skew-symmetric forms need not vanish on the diagonal. For non-degenerate alternating forms over the real numbers, non-trivial examples exist only when the dimension of V is even. For a non-degenerate skew-symmetric bilinear form on a finite-dimensional even-dimensional vector space, the Pfaffian provides a polynomial invariant that serves as a square root of the determinant of the associated skew-symmetric matrix. Specifically, if A is the skew-symmetric matrix representing the form, then \operatorname{Pf}(A)^2 = \det(A).

Reflexivity and orthogonality

A bilinear form B on a vector space V over a field K is called reflexive if B(u, v) = 0 for all v \in V implies u = 0, and symmetrically, if B(u, v) = 0 for all u \in V implies v = 0. This condition ensures that the form captures the full structure of the space without collapsing any nonzero vectors to zero under orthogonalization. In finite-dimensional spaces, reflexivity is equivalent to the standard notion of non-degeneracy for bilinear forms. Bilateral non-degeneracy, or reflexivity, requires that both the left kernel \{ u \in V \mid B(u, v) = 0 \ \forall v \in V \} and the right kernel \{ v \in V \mid B(u, v) = 0 \ \forall u \in V \} are trivial, i.e., equal to \{0\}. This bilateral condition distinguishes reflexive forms from those that may be degenerate on one side only, ensuring a balanced duality between the two arguments. In coordinate terms, if V is finite-dimensional with basis \{e_i\}, the matrix representation A of B satisfies A_{ij} = B(e_i, e_j), and reflexivity holds if and only if \det A \neq 0. Given a reflexive bilinear form B, two vectors u, v \in V are orthogonal, denoted u \perp v, if B(u, v) = 0. For a subspace S \subseteq V, the orthogonal complement is the set S^\perp = \{ w \in V \mid B(u, w) = 0 \ \forall u \in S \}. Reflexivity implies that \dim S + \dim S^\perp = \dim V in finite dimensions, and S \cap S^\perp = \{0\} if S is non-degenerate as a restriction of B. For a non-degenerate symmetric bilinear form over the real numbers, the spectral theorem guarantees an orthogonal basis of eigenvectors, allowing V to decompose as an orthogonal direct sum of eigenspaces: V = \bigoplus_{\lambda \in \sigma(B)} E_\lambda, where E_\lambda = \{ u \in V \mid B(u, v) = \lambda \langle u, v \rangle for some inner product, but adapted to the form's geometry, with E_\lambda \perp E_{\mu} for \lambda \neq \mu. This decomposition highlights the form's role in classifying quadratic structures via eigenvalues. In the context of duality induced by a reflexive bilinear form B: V \times V \to K, which identifies V with its dual V^* via v \mapsto B(v, \cdot), the orthogonal complement S^\perp of a subspace S \subseteq V is precisely the annihilator of S under this pairing. Thus, reflexivity ensures the duality map is an isomorphism, preserving annihilator properties such as (S^\perp)^\perp = S for finite-dimensional V.

Advanced Properties

Boundedness

In normed vector spaces, a bilinear form B: V \times V \to \mathbb{F} (where \mathbb{F} is \mathbb{R} or \mathbb{C}) on a space (V, \|\cdot\|) is said to be bounded if there exists a constant C > 0 such that |B(u, v)| \leq C \|u\| \|v\| for all u, v \in V. This condition ensures that the form does not grow faster than the product of the norms of its arguments. The of a bounded bilinear form B is defined as \|B\| = \sup \{ |B(u, v)| : \|u\| \leq 1, \|v\| \leq 1 \}, which is finite and equivalent to the infimum of all such constants C. This norm measures the maximum "amplification" of the form on the unit ball. Boundedness is precisely equivalent to of B as a map from V \times V (equipped with the product ) to \mathbb{F}. In finite-dimensional normed spaces, every bilinear form is bounded. This follows from the fact that every linear operator between finite-dimensional normed spaces is continuous (and hence bounded), and a bilinear form induces linear operators by fixing one argument—for instance, the map v \mapsto B(u, v) is linear in v for fixed u, with boundedness uniform across unit vectors due to norm equivalence in finite dimensions. More generally, a bounded bilinear form B induces bounded linear operators, such as the left multiplication map V \to L(V, \mathbb{F}) given by u \mapsto (v \mapsto B(u, v)), which has equal to \|B\|. A prominent example of a bounded bilinear form is an inner product on a , which satisfies |\langle u, v \rangle| \leq \|u\| \|v\| by the Cauchy-Schwarz inequality, yielding operator norm at most 1.

Ellipticity

In the context of , a B: H \times H \to \mathbb{R} is said to be elliptic, or coercive, if there exists a constant \alpha > 0 such that B(u, u) \geq \alpha \|u\|^2 for all u \in H. This condition ensures that the form provides a lower bound relative to the norm induced by the inner product on H, often used to establish stability in variational problems. Over the real numbers, for a finite-dimensional inner product space with matrix representation A of B, ellipticity is equivalent to A being positive definite, meaning all eigenvalues of A are positive. This spectral characterization follows from the spectral theorem for symmetric matrices, where the quadratic form B(u, u) = u^T A u is bounded below by the smallest eigenvalue times \|u\|^2. In more general settings, such as Sobolev spaces H^1_0(\Omega) for bounded domains \Omega \subset \mathbb{R}^n, full may not hold due to lower-order terms, leading to Gårding's inequality. For a bilinear form B[u, v] = \int_\Omega \sum_{i,j=1}^n a_{ij} \partial_i u \partial_j v + \sum_{i=1}^n b_i (\partial_i u) v + c u v \, dx associated with a second-order in divergence form, there exist constants \alpha, \beta > 0 and \gamma \geq 0 such that |B[u, v]| \leq \alpha \|u\|_{H^1_0(\Omega)} \|v\|_{H^1_0(\Omega)} and B[u, u] \geq \beta \|u\|_{H^1_0(\Omega)}^2 - \gamma \|u\|_{L^2(\Omega)}^2 for all u, v \in H^1_0(\Omega), assuming uniform ellipticity of the principal coefficients a_{ij}. This weakened , combined with compactness from the Rellich-Kondrachov , enables solvability via the or modified Lax-Milgram arguments. Elliptic bilinear forms play a central role in the analysis of partial differential equations (PDEs), particularly in ensuring well-posedness of weak solutions. By the Lax-Milgram theorem, if B is continuous (bounded) and elliptic on a H, then for any bounded linear functional f \in H^*, the variational problem B(u, v) = \langle f, v \rangle for all v \in H admits a unique solution u \in H. For second-order uniformly elliptic PDEs like -\sum_{i,j} \partial_i (a_{ij} \partial_j u) + \sum_i b_i \partial_i u + c u = f with suitable boundary conditions, the associated bilinear form satisfies these hypotheses in H^1_0(\Omega), yielding unique weak solutions. In , particularly for hyperelastic materials, ellipticity conditions on the tangent bilinear form distinguish between strong and weak variants. Strong ellipticity requires the bilinear form to be positive definite on the entire space of admissible fields, ensuring the acoustic tensor has positive eigenvalues for all wave directions and guaranteeing material under arbitrary infinitesimal deformations. Weak ellipticity, in contrast, demands positivity only along rank-one directions (corresponding to plane waves), which is a necessary but insufficient for full stability and aligns with the Legendre-Hadamard condition but permits loss of ellipticity in higher-rank perturbations.

Associated Structures

Quadratic forms

A quadratic form associated with a bilinear form B: V \times V \to F on a vector space V over a field F is defined by Q(u) = B(u, u) for all u \in V. This association holds for any bilinear form, but when B is symmetric (i.e., B(u, v) = B(v, u) for all u, v \in V), the quadratic form Q uniquely determines B via the polarization identity, provided the characteristic of F is not 2: B(u, v) = \frac{1}{4} \left[ Q(u + v) - Q(u - v) \right]. This identity follows from expanding Q(u + v) and Q(u - v) using bilinearity and collecting terms. More generally, over fields of characteristic not 2, the can be expressed as B(u, v) = \frac{1}{4} \sum_{\varepsilon = \pm 1} \varepsilon Q(u + \varepsilon v), where the sum accounts for the contributions from the expansions. Over the complex numbers, where bilinear forms are often replaced by (linear in the first argument and conjugate-linear in the second), the is adjusted to recover the from the associated \Phi(u) = \langle u, u \rangle, for example: $4 \langle u, v \rangle = \Phi(u + v) - \Phi(u - v) + i \Phi(u + i v) - i \Phi(u - i v). This ensures compatibility with the Hermitian structure. The quadratic form Q satisfies quadratic homogeneity: Q(a u) = a^2 Q(u) for all scalars a \in F and u \in V, which follows directly from the bilinearity of B. Over the real numbers, a quadratic form Q is positive semi-definite if Q(u) \geq 0 for all u \in V. In coordinates with respect to a basis, if B has matrix A = (B(e_i, e_j)), then Q(u) = ^T A for the column vector $$ of coordinates of u. For symmetric B, A is symmetric. For real symmetric bilinear forms, the associated quadratic form Q admits a classification via its , defined as the pair (p, q) where p is the number of positive eigenvalues, q the number of negative eigenvalues (with the rest zero), of the diagonalized form under Sylvester's law of inertia. This signature is invariant under and determines the inertia of Q.

Tensor product relations

Bilinear forms on a V over a F can be understood through their connection to in . Specifically, for finite-dimensional V, the space of bilinear forms \mathrm{Bil}(V, V; F) is naturally isomorphic to the of the with itself, V^* \otimes V^*. Under this isomorphism, a bilinear form B corresponds to an element \sum_i \phi_i \otimes \psi_i \in V^* \otimes V^*, where B(v, w) = \sum_i \phi_i(v) \psi_i(w) for all v, w \in V. In a basis \{e_j\} for V with dual basis \{e_i^*\} for V^*, the bilinear form expands as B = \sum_{i,j} B(e_i, e_j) e_i^* \otimes e_j^*. This tensor product perspective aligns with the universal property of the tensor product. A bilinear form B: V \times V \to F is equivalently a linear map from the tensor product V \otimes V to F, and thus an element of the dual space (V \otimes V)^*. The universal property ensures that \mathrm{Hom}(V \otimes V, F) \cong \mathrm{Bil}(V, V; F), providing a basis-free characterization that extends the algebraic structure of bilinear forms. For bilinear forms B: V \times W \to F on distinct finite-dimensional spaces V and W, the tensor isomorphism becomes \mathrm{Bil}(V, W; F) \cong V^* \otimes W^*, with B(v, w) = \sum_i \phi_i(v) \psi_i(w). This induces adjoint maps: B defines a V \to W^* by v \mapsto (w \mapsto B(v, w)), and dually W \to V^* by w \mapsto (v \mapsto B(v, w)). These adjoints embed bilinear forms into the space of linear maps involving duals, facilitating connections in and duality. Bilinear forms represent the case k=2 of , which are linear maps on k-fold products V^k \to F, or elements of (V^{\otimes k})^*. The universal property generalizes: the V^{\otimes k} linearizes k-linear maps, so \mathrm{Bil}(V, V; F) is the special case where the multilinear form is separately linear in two arguments. This embedding highlights bilinear forms as building blocks in higher . In infinite-dimensional settings, such as Banach spaces, care is required with tensor products. The algebraic tensor product V \otimes W exists but is generally not complete with respect to natural norms, so continuous bilinear forms may require completed tensor products (e.g., projective or injective) to ensure the isomorphisms hold in the topological category. For Hilbert spaces, no single completion satisfies the universal property for all continuous bilinear maps to Hilbert spaces.

Generalizations

Distinct vector spaces

A bilinear form on distinct vector spaces V and W over a F is a function B: V \times W \to F that is linear in each argument separately. Specifically, for all v_1, v_2 \in V, w \in W, and \lambda \in F, B(v_1 + v_2, w) = B(v_1, w) + B(v_2, w) and B(\lambda v_1, w) = \lambda B(v_1, w), and similarly for linearity in the second argument. The set \mathrm{Bil}(V, W) of all such bilinear forms forms a over F, which is naturally isomorphic to the V^* \otimes W^*, where V^* and W^* denote the dual spaces. Unlike the case where V = W, bilinear forms on distinct spaces exhibit asymmetries in their non-degeneracy properties. A bilinear form B is left non-degenerate if the induced map V \to W^*, defined by v \mapsto (w \mapsto B(v, w)), is injective, meaning that if B(v, w) = 0 for all w \in W, then v = 0. Similarly, B is right non-degenerate if the map W \to V^*, w \mapsto (v \mapsto B(v, w)), is injective. In the non-square case, left non-degeneracy does not imply right non-degeneracy, and vice versa, particularly when \dim V \neq \dim W or in infinite . However, if both conditions hold and V, W are finite-dimensional with equal dimension, then B induces an V \cong W^*. In linear algebra, bilinear forms between distinct spaces arise in applications such as pairing linear maps. For instance, consider V = \mathrm{Hom}(U, X) and W = \mathrm{Hom}(X, U) for vector spaces U and X; the trace form B(f, g) = \mathrm{Tr}(f \circ g) defines a bilinear map V \times W \to F, which is non-degenerate under suitable conditions on dimensions. A fundamental example is the canonical pairing between a vector space and its dual, given by the evaluation map \langle \phi, v \rangle = \phi(v) for \phi \in V^* and v \in V. This bilinear form B: V^* \times V \to F is non-degenerate in both directions when V is finite-dimensional, establishing a natural duality.

Module extensions

In the more general setting of modules over a commutative ring R, a bilinear form on R-modules M and N is an R-bilinear map B: M \times N \to R, meaning B is additive in each argument and satisfies B(rm, n) = r B(m, n) = B(m, rn) for all r \in R, m \in M, and n \in N. This extends the vector space case by incorporating the ring structure, where linearity holds with respect to the module actions rather than field scalar multiplication. Unlike over fields, where the dual module M^\vee = \Hom_R(M, R) often allows straightforward isomorphisms via non-degenerate forms, no such general duality isomorphism exists for arbitrary s over rings, complicating the identification of bilinear forms with elements of M^\otimes N. Non-degeneracy, defined as the associated maps M \to N^\vee and N \to M^\vee (given by m \mapsto (n \mapsto B(m,n)) and similarly) being s, becomes harder to characterize due to torsion elements; for instance, in modules with 2-torsion, skew-symmetric forms may fail to be alternating, affecting properties like the dimension of isotropic subspaces. For free modules such as M = N = R^n, bilinear forms reduce to those represented by n \times n matrices over R, with the yielding B(e_i, e_j) = a_{ij} for the matrix (a_{ij}). On abelian groups, viewed as \mathbb{Z}-modules, examples include the form B: \mathbb{Q} \times \mathbb{Q} \to \mathbb{Q} given by B(r, s) = rs. For torsion groups, the \mathbb{Z}/n\mathbb{Z} \otimes_\mathbb{Z} \mathbb{Z}/m\mathbb{Z} = 0 if \gcd(n,m)=1, illustrating issues with torsion in module , which bilinear forms help study. Sesquilinear forms extend this framework over rings A with involution \sigma: A \to A, defined as biadditive maps s: V \times V \to A on a right A-module V satisfying s(xa, yb) = \sigma(a) s(x, y) b, making s linear in the second argument and anti-linear (via \sigma) in the first; this generalizes Hermitian forms over complexes, where \sigma is complex conjugation. In , bilinear forms appear on coherent sheaves over varieties, such as non-degenerate symmetric forms on sheaves on a X \subset \mathbb{P}^d, which correspond to real symmetric determinantal representations of X and link to properties like hyperbolicity.

References

  1. [1]
    [PDF] BILINEAR FORMS The geometry of Rn is controlled algebraically by ...
    Definition 1.1. Let F be a field and V be a vector space over F. A bilinear form on V is. a function B: V × V → F that is linear in each variable when the ...
  2. [2]
    [PDF] Bilinear and Quadratic Forms This handout should be read just ...
    Bilinear Forms. A bilinear form on V is a function B: V × V → R that is linear in each variable separately; this means that for all v, w, x ∈ V and all a, b ∈ ...
  3. [3]
    [PDF] Further linear algebra. Chapter V. Bilinear and quadratic forms.
    Definition 1.1 Let V be a vector space over k. A bilinear form on V is a function f : V × V → k such that. • f(u + λv, w) = f(u, w) + λf(v, w);.
  4. [4]
    [PDF] Symmetric bilinear forms
    The bilinear form B is called non- degenerate if ker(B) = 0, i.e. if and only if B♭ is an isomorphism.Missing: mathematics | Show results with:mathematics
  5. [5]
    Bilinear forms with Kloosterman sums and applications
    We then derive applications to the second moment of cusp forms twisted by characters modulo primes, and to the distribution in arithmetic progressions to large ...
  6. [6]
    [PDF] Bilinear forms with Kloosterman fractions
    There are also applications to sums of the type occurring in the works [BFI] on the distribution of divisor functions and of primes in arithmetic progressions.
  7. [7]
    [PDF] Structure of the group preserving a bilinear form
    We shall often write v, w instead of ϕ(v, w). The space of all bilinear forms on V will be denoted by Bil(V ). ... The transpose of φ ∈ Bil(V ) is the bilinear ...
  8. [8]
    Bilinear Form -- from Wolfram MathWorld
    A bilinear form on a real vector space is a function b:V×V->R that satisfies the following axioms for any scalar alpha and any choice of vectors.
  9. [9]
    [PDF] A Brief History of Linear Algebra and Matrix Theory
    In contrast, the first implicit use of matrices occurred in Lagrange's work on bilinear forms in the late. 1700s. Lagrange desired to characterize the maxima ...
  10. [10]
    [PDF] LADR4e.pdf - Linear Algebra Done Right - Sheldon Axler
    ... linear algebra textbook, the proof given here uses linear algebra techniques and makes nice use of a basis of 𝒫𝑛(𝐅), which is the (𝑛 + 1)-dimensional ...
  11. [11]
    [PDF] Lecture 4.7. Bilinear and quadratic forms - Purdue Math
    Apr 9, 2020 · A function which assigns to. every pair of vectors x,y a real number B(x,y) is called a bilinear form if it.Missing: mathematics | Show results with:mathematics
  12. [12]
    [PDF] Bilinear and quadratic forms. Special relativity - Purdue Math
    This is where the Minkowski quadratic form comes from. So we conclude that the correct transformation is not (12) but the Lorenz boost (8).
  13. [13]
    [PDF] Minkowski space
    Mar 21, 2013 · In the case of significant gravitation, spacetime becomes curved and one must abandon special relativity in favor of the full theory of general ...
  14. [14]
    [PDF] 18.704: Classification of Bilinear Forms over Finite Fields
    Mar 7, 2005 · If B is a nonzero symmetric bilinear form on V with quadratic form Q, then Q(v) 6= 0 for some v ∈ V . Proof: If Q(u) = 0 for all u, then 0 = Q(u ...
  15. [15]
    None
    ### Summary of Matrix Representation of Bilinear Forms
  16. [16]
    [PDF] THE STRUCTURE OF SPIN SYSTEMS - Indian Statistical Institute ...
    infinite matrix (cij ) with 0, 1 entries satisfying ... This bilinear form is skew-symmetric in that it ... Thus a symplectic basis for V is a countable Hamel basis ...
  17. [17]
    [PDF] Bilinear Forms over a field F Let V be a vector space.
    The most famous alternating bilinear form is the cross product B(x, y) = x × y = x1y2 − x2y1 on F2. Every n × n matrix A = (aij) gives rise to a bilinear form ...
  18. [18]
    MA22020: Advanced linear algebra
    A symmetric bilinear form B on V is positive definite if and only if it is an inner product on V . The signature is easy to compute: Theorem 5.8 (Sylvester's ...
  19. [19]
    [PDF] Contents 5 Bilinear and Quadratic Forms - Evan Dummit
    This is precisely the condition we require in order to diagonalize a symmetric bilinear form. ◦ Hence: we may diagonalize a symmetric bilinear form over R by ...
  20. [20]
    [PDF] Real representations - MIT Mathematics
    Oct 21, 2016 · A bilinear form (or sesquilinear form) is called nondegenerate if its left kernel is 0, or equivalently its right kernel is 0, or equivalently.
  21. [21]
    [PDF] The kernel of a function from a module to a set - IJS
    Jul 31, 2018 · Let U and V be vector spaces over a field F, and let β: U × V → F be a bilinear form with a left kernel U0 and a right kernel V0. Since ...
  22. [22]
    Symmetric Bilinear Form -- from Wolfram MathWorld
    A symmetric bilinear form on a vector space V is a bilinear function Q:V×V->R (1) which satisfies Q(v,w)=Q(w,v). For example, if A is a n×n symmetric matrix ...
  23. [23]
    matrix representation of a bilinear form - PlanetMath.org
    Mar 22, 2013 · Thus we see that a symmetric bilinear form is represented by a symmetric matrix , and similarly ...
  24. [24]
    skew-symmetric bilinear form - PlanetMath
    Mar 22, 2013 · A bilinear form is skew-symmetric iff its representing matrix is skew-symmetric. (The fact that the representing matrix is skew-symmetric is independent of the ...
  25. [25]
    [PDF] Notes on antisymmetric matrices and the pfaffian
    We now specialize to skew-symmetric bilinear forms (which are called antisymmetric by physicists). A bilinear form is called skew-symmetric if f(v,w) = −f ...
  26. [26]
    [PDF] Unit 17: Spectral theorem
    The spectral theorem states that every symmetric matrix has an orthonormal eigenbasis. Symmetric matrices have only real eigenvalues.
  27. [27]
    Bounded Bilinear Maps
    **Summary of Bounded Bilinear Maps (Chapter 6)**
  28. [28]
    None
    Below is a merged and comprehensive summary of the mentions of "bilinear form" and "bounded bilinear" in Erwin Kreyszig's *Introductory Functional Analysis with Applications*, based on the provided segments. To retain all information in a dense and organized manner, I will use a table in CSV format for key details, followed by a narrative summary that integrates additional context and notes. This approach ensures all relevant information is preserved while maintaining clarity and accessibility.
  29. [29]
    [PDF] MAA 4212—Spring 2014—Some facts about normed vector spaces
    Then every linear transformation V → W is continuous. Proof: If dim(V ) = 0 then V has only one element, so any function V → W is continuous.Missing: operator | Show results with:operator
  30. [30]
    [PDF] Second Order Elliptic PDE: The Lax-Milgram Theorem
    Theorem 6.2.1. (Lax-Milgram Theorem) Suppose B : H × H → R is a bilinear form for which there exist constants α,β > 0 so that. |B[u,v]| ≤αkukkvk, ∀u,v ...
  31. [31]
    [PDF] Chapter 4: Elliptic PDEs - UC Davis Math
    We begin by stating the Lax-Milgram theorem for a bilinear form on a Hilbert space. Afterwards, we verify its hypotheses for the bilinear form associated with a ...
  32. [32]
    [PDF] Sobolev Spaces and Garding Inequality
    Apr 27, 2015 · Sobolev space Wk,p(Ω) contains functions with weak derivatives of various orders in Lp spaces. Wk,p(Ω) is a Banach space.Missing: Gårding's | Show results with:Gårding's
  33. [33]
    On Gårding's inequality - SpringerLink
    In this paper we prove Gårding's inequality for linear differential operators in generalized divergence form which satisfy a generalized Ehrling inequality ...Missing: bilinear | Show results with:bilinear
  34. [34]
    [PDF] Cosserat micropolar elasticity: classical Eringen vs. dislocation form
    Jun 6, 2022 · The strong ellipticity conditions (6.16) are weaker than the conditions (6.11) in the sense that they are implied by the necessary and ...
  35. [35]
    [PDF] notes on the algebraic theory of quadratic forms
    Intersection forms on even-dimensional manifolds. A closed oriented 4k-manifold M has an intersec- tion (symmetric bilinear) form. H2k(M;k) × H2k(M;k) → H2k ...
  36. [36]
    [PDF] Chapter 8 Basics of Hermitian Geometry - CIS UPenn
    These are called polarization identities. Proposition 8.1 shows that a sesquilinear form is com- pletely determined by the quadratic form Φ(u) = '(u, u), even ...
  37. [37]
    [PDF] Hilbert Spaces I: Basic Properties - KSU Math
    for all x, y ∈ X. Lemma 2 (The Polarization Identity). If X is a vector space and φ : X ×X → C is a sesquilinear form, then φ(x, y) = 1. 4. 3. X k=0 i−kφ(x ...
  38. [38]
    [PDF] Lecture 4.9. Positive definite and semidefinite forms - Purdue Math
    Apr 10, 2020 · Definitions. Q and A are called positive semidefinite if Q(x) ≥ 0 for all x. They are called positive definite if Q(x) > 0 for all x 6= 0.
  39. [39]
    [PDF] Math 416, Spring 2010 Congruence; Sylvester's Law of Inertia
    Apr 22, 2010 · We'll spend the balance of class proving Sylvester's Law of Inertia. 3. Proof of Sylvester's Law of Inertia. Proof. Suppose that B and C are ...<|separator|>
  40. [40]
    [PDF] SOME NECESSARY LINEAR ALGEBRA. 1. Bilinear forms Let V be ...
    Bilinear forms. Let V be a vector space over a field k. A bilinear form h( , ) is a function V ×V → k that is linear in both arguments.Missing: mathematics | Show results with:mathematics
  41. [41]
    [PDF] Lecture 16: Tensor Products
    We begin with a generalization of a bilinear form. Definition. A function f ... The tensor product is the unique (up to isomorphism) vector space such that for.
  42. [42]
    [PDF] Duality, Bilinearity
    Using the symmetry and bilinearity of Q , one obtains the formulas. Q(ξu) = ξ2Q(u) for all ξ ∈ R, u ∈ V,. (27.16). Q(u + v) = Q(u)+2Q (u,v) + Q(v) for all u,v ∈ ...Missing: via | Show results with:via
  43. [43]
    [PDF] Non-existence of tensor products of Hilbert spaces
    Jul 22, 2010 · Tensor products of infinite-dimensional Hilbert spaces do not exist. That is, for infinite-dimensional Hilbert spaces V,W, there is no ...
  44. [44]
    [PDF] A note on bilinear forms
    “Bilinear form” is just an alternative terminology for a bilinear map whose target is k: Definition 2.9. Let V and W be two vector spaces. A bilinear form on V ...<|control11|><|separator|>
  45. [45]
    [PDF] bilinear forms - keith conrad
    A bilinear space is a vector space equipped with a specific choice of bilinear form. We call a bilinear space symmetric, skew-symmetric, or alternating when ...
  46. [46]
    [PDF] An Advanced Course in Linear Algebra - Jim Brown
    Jul 20, 2015 · ... vector space that ϕ is left non-degenerate if and only if it is right non-degenerate. However, if V is infinite dimensional these may not be ...
  47. [47]
    [PDF] ALGEBRA - LECTURE V 1. Bilinear forms Let R be a commutative ...
    4. [Q(v + w) − Q(v − w)] defines a bilinear (necessarily symmetric) form on V . A quadratic space is a pair (V,Q) consisting of a vector space V and a ...Missing: mathematics | Show results with:mathematics<|control11|><|separator|>
  48. [48]
  49. [49]