Fact-checked by Grok 2 weeks ago

Pairing

In mathematics, a pairing is an R-bilinear map from the Cartesian product of two R-modules M and N to the commutative ring R, often denoted ⟨⋅,⋅⟩: M × N → R. This map is linear in each argument separately and plays a central role in establishing dualities between modules. A pairing is non-degenerate if the induced homomorphisms M → Hom_R(N, R) and N → Hom_R(M, R) are injective, and perfect if these maps are isomorphisms. Pairings appear in various contexts, including algebraic structures like vector spaces and abelian groups, where they generalize inner products. In geometry, they relate to polarizations on varieties. Advanced types include alternating pairings, which are skew-symmetric and underlie , and Hermitian pairings over complex numbers. Applications of pairings are prominent in , particularly bilinear pairings on elliptic curves such as the Weil or pairings, enabling protocols like and attribute-based encryption. In , pairings facilitate the study of module homomorphisms and character tables.

Fundamentals

Definition

In , an R- over a R with identity is an M equipped with a map R \times M \to M that distributes over addition in R and M, and satisfies $1 \cdot m = m for all m \in M. Similarly, the M \otimes_R N of two R-s M and N is an R- equipped with a M \times N \to M \otimes_R N satisfying a : any from M \times N to another R- P factors uniquely through a M \otimes_R N \to P. A pairing on R-modules M, N, and L is an R-bilinear map e: M \times N \to L, meaning it is additive in each argument separately—e(m_1 + m_2, n) = e(m_1, n) + e(m_2, n) and e(m, n_1 + n_2) = e(m, n_1) + e(m, n_2)—and homogeneous over R—e(rm, n) = r e(m, n) = e(m, rn) for all r \in R, m \in M, n \in N. This bilinearity ensures the map respects the module structures of M and N. Equivalently, every such pairing corresponds to an R-linear map from the tensor product M \otimes_R N to L, via the induced map sending m \otimes n to e(m, n), establishing a natural between the space of bilinear maps and \hom_R(M \otimes_R N, L). Basic types of pairings include non-degenerate and s. A pairing e is left non-degenerate if the left \{m \in M \mid e(m, n) = 0 \ \forall n \in N\} is trivial (i.e., equals \{0\}), and right non-degenerate if the right \{n \in N \mid e(m, n) = 0 \ \forall m \in M\} is trivial; it is non-degenerate if both hold. A is one that induces an M \cong \hom_R(N, L) via the map m \mapsto (n \mapsto e(m, n)), or dually N \cong \hom_R(M, L); every is non-degenerate, though the converse requires additional conditions such as finite-dimensionality over a .

Properties

A bilinear pairing e: M \times N \to L on R-modules satisfies bilinearity, meaning it is R-linear in each argument separately. This implies preservation under addition and scalar multiplication: for all m_1, m_2 \in M, n \in N, and r \in R, e(m_1 + m_2, n) = e(m_1, n) + e(m_2, n), \quad e(r m_1, n) = r e(m_1, n), with analogous properties holding when fixing m \in M and varying n \in N. These consequences follow directly from the linearity in each slot, enabling the pairing to extend naturally to multilinear maps on tensor products. Non-degeneracy is a key structural property of pairings, ensuring they do not collapse non-trivial elements to zero. The pairing is left non-degenerate if the left annihilator \{ m \in M \mid e(m, n) = 0 \ \forall n \in N \} = \{0\}, and right non-degenerate if the right annihilator \{ n \in N \mid e(m, n) = 0 \ \forall m \in M \} = \{0\}. Equivalently, left non-degeneracy means the induced map M \to \hom_R(N, L), given by m \mapsto (n \mapsto e(m, n)), is injective, with a similar injectivity condition for the right induced map N \to \hom_R(M, L). This injectivity prevents the pairing from being "degenerate" in either direction, providing a faithful encoding of module elements via linear functionals. A pairing is perfect if the induced map M \to \hom_R(N, L) is an (hence also the dual map N \to \hom_R(M, L) is, under suitable finiteness conditions). This bijectivity establishes a between M and N, identifying M with the dual module N^\vee = \hom_R(N, L) and vice versa, which underpins reflexive structures in . Perfect pairings are necessarily non-degenerate, as isomorphisms imply injectivity, but the converse requires surjectivity as well. Pairings interact compatibly with R-linear maps between modules, preserving structure under homomorphisms. Specifically, if f: M' \to M and g: N \to N' are R-module homomorphisms, then the composition e \circ (f \times g): M' \times N' \to L defines a new bilinear pairing, inheriting bilinearity from e. Moreover, perfect pairings induce natural isomorphisms between spaces of homomorphisms, such as \hom_R(M, N^\vee) \cong \hom_R(N, M^\vee), facilitating the study of module categories via duals.

Examples

Algebraic Examples

One prominent algebraic example of a pairing is the scalar product on a finite-dimensional V over \mathbb{R} or \mathbb{C}. This is a \langle \cdot, \cdot \rangle: V \times V \to \mathbb{R} (or \mathbb{C}) that is symmetric for real spaces and Hermitian for spaces, satisfying \langle u, v \rangle = \overline{\langle v, u \rangle}. It is positive definite, meaning \langle v, v \rangle > 0 for v \neq 0, which implies non-degeneracy: if \langle u, v \rangle = 0 for all u \in V, then v = 0. In an , the scalar product takes the form \langle \sum x_i e_i, \sum y_j e_j \rangle = \sum x_i y_i. This pairing underpins concepts like and norms in linear algebra. A canonical alternating pairing arises on the two-dimensional k^2 over a k of characteristic not equal to 2, defined by e((a, b), (c, d)) = ad - bc. This is skew-symmetric, satisfying e(u, v) = -e(v, u), and alternating since e(v, v) = 0 for all v. It is perfect, meaning both the left and right kernels are trivial, as the matrix representation \begin{pmatrix} 0 & 1 \\ -1 & 0 \end{pmatrix} has full rank. This example illustrates the structure of forms in low dimensions and extends to higher even dimensions via block-diagonal constructions. In matrix algebras, the trace pairing provides another symmetric example on the space of n \times n matrices over \mathbb{R} or \mathbb{C}, given by \langle A, B \rangle = \tr(A^T B). This is bilinear and symmetric, with \langle A, A \rangle = \|A\|_F^2 \geq 0, where \|\cdot\|_F is the Frobenius norm, and equality holds if and only if A = 0, ensuring non-degeneracy. The trace operation sums the diagonal entries of A^T B, making this pairing invariant under simultaneous orthogonal similarity transformations. It is widely used to induce norms and study matrix decompositions. In the context of group representation theory, the character pairing acts on the space of class functions of a finite group G, defined by the inner product \langle \chi, \psi \rangle = \frac{1}{|G|} \sum_{g \in G} \chi(g) \overline{\psi(g)}. This Hermitian form is positive definite on the subspace spanned by irreducible characters, yielding orthonormality: \langle \chi_i, \chi_j \rangle = \delta_{ij}. Non-degeneracy follows from the completeness of irreducible characters as a basis for class functions. This pairing quantifies the multiplicity of irreducibles in representations and facilitates decomposition theorems.

Geometric Examples

In geometric settings, pairings often arise from structures on manifolds and varieties, extending the bilinear forms from algebraic contexts to incorporate topological and sheaf-theoretic features. One prominent example is the , which can be interpreted through quaternionic to induce a non-degenerate bilinear structure. Identifying S^3 with the quaternions and S^2 with the unit pure imaginary quaternions, the map (q, v) \mapsto q v \overline{q} (where \overline{q} is the conjugate) defines an \mathbb{R}-bilinear preserving the spheres, reflecting the geometry of the S^3 \to S^2 via conjugation rotations. This structure is non-degenerate as the of the quaternions on the pure imaginaries is faithful and irreducible. Another key geometric pairing emerges from Serre duality on projective varieties. For a smooth projective variety X of dimension n over an algebraically closed field k, and a coherent sheaf \mathcal{F}, Serre duality provides a perfect pairing \Hom(\mathcal{F}, \omega_X) \times \Ext^n(\mathcal{F}, \mathcal{O}_X) \to k, where \omega_X is the canonical sheaf. This bilinear form is induced by the trace map on the top cohomology and is non-degenerate under the assumptions of smoothness and projectivity, pairing global sections of the dualizing sheaf with Ext groups. The duality extends to a perfect pairing on cohomology groups H^i(X, \mathcal{F}^\vee \otimes \omega_X) \times H^{n-i}(X, \mathcal{F}) \to k, capturing geometric invariants like Euler characteristics. Poincaré duality furnishes a fundamental pairing in manifold . For a closed orientable n-manifold M, the cap product with the fundamental class induces a non-degenerate bilinear pairing H_p(M; \mathbb{R}) \times H^{n-p}(M; \mathbb{R}) \to \mathbb{R}, given by [\alpha] \cap [M] = \langle \alpha \cup \beta, [M] \rangle for dual cycles. This pairing is symmetric or skew-symmetric depending on dimensions and identifies with via the , with non-degeneracy ensured by the and of M. It underpins on manifolds by equating algebraic intersections with topological pairings. In , the intersection pairing on Chow groups provides a on cycles. For a smooth projective variety X, the Chow group CH_p(X) consists of p-dimensional cycles modulo , and the intersection product equips it with a structure where the pairing CH_p(X) \times CH_q(X) \to CH_{p+q}(X) composed with the degree map yields a to \mathbb{Z} for complete . This form is non-degenerate on the numerical Chow groups and reflects geometric transversality conditions, with compatibility to the cycle class map into .

Advanced Concepts

Duality and Perfect Pairings

In the context of over a R, a perfect pairing e: M \times N \to R is a that induces a duality between the modules M and N. Specifically, it defines an R- \phi: M \to \Hom_R(N, R) by sending m \in M to the functional n \mapsto e(m, n), and this map is an . Dually, the map \psi: N \to \Hom_R(M, R) given by n \mapsto (m \mapsto e(m, n)) is also an . This mutual identification with the dual modules establishes a strong form of duality, stronger than mere non-degeneracy, as it requires the induced maps to be bijective rather than merely injective. When M = N, a perfect pairing e: M \times M \to R renders M self-dual, meaning M \cong \Hom_R(M, R). This self-duality implies reflexivity of M, where the natural evaluation map M \to \Hom_R(\Hom_R(M, R), R) is an . For finite modules, such as R^n equipped with the standard pairing, self-duality holds via the choice of a basis, providing a realization of this concept. Self-dual modules play a key role in structures like quadratic forms and orthogonal groups in algebra. In , perfect pairings are closely tied to finite projective modules, where they facilitate local-global principles. A pairing between finitely generated projective modules is perfect globally if and only if it induces isomorphisms locally after localization at every , owing to the exactness of localization and the preservation of projectivity for such modules. This localization property underscores the utility of perfect pairings in descent theory and . Beyond modules, in general abelian categories, perfect pairings connect to through the Yoneda extensions defining the Ext functors. The Yoneda product equips the graded groups \Ext^p(A, B) \times \Ext^q(B, C) \to \Ext^{p+q}(A, C) with a bilinear pairing, interpreting compositions of extensions as higher-dimensional structures. This pairing via Yoneda extensions bridges bilinear forms on extension classes to duality phenomena in derived categories, generalizing module dualities to abstract settings.

Alternating and Hermitian Pairings

In the context of pairings between modules over a , an alternating pairing occurs when the and modules coincide, denoted as V \times V \to k, and the pairing e satisfies e(v, v) = 0 for all v \in V. This condition implies that e is skew-symmetric, meaning e(v, w) = -e(w, v) for all v, w \in V. Such forms are bilinear by assumption and play a fundamental role in linear algebra over fields of characteristic not equal to 2. A prime example of alternating pairings is found in symplectic vector spaces, where e is non-degenerate—meaning if e(v, w) = 0 for all w \in V, then v = 0—and the dimension of V is even. In this setting, the pairing defines a structure, enabling the study of transformations that preserve e. For instance, over numbers, the standard symplectic form on \mathbb{R}^{2n} is given by e((x_1, y_1), (x_2, y_2)) = x_1 \cdot y_2 - y_1 \cdot x_2, which is alternating and non-degenerate. Hermitian pairings extend this notion to complex vector spaces, where the pairing e: V \times V \to \mathbb{C} is sesquilinear—linear in the first argument and conjugate-linear in the second—and satisfies e(v, w) = \overline{e(w, v)} for all v, w \in V. This conjugate symmetry distinguishes Hermitian pairings from purely bilinear forms. When e is additionally positive definite, meaning e(v, v) > 0 for all nonzero v, it defines a , which equips V with a norm \|v\| = \sqrt{e(v, v)} and facilitates orthogonal decompositions. A canonical example is the standard inner product on \mathbb{C}^n, e(z, w) = \sum_{i=1}^n z_i \overline{w_i}. Alternating pairings relate to quadratic forms through the polarization identity, particularly in fields of characteristic 2, where the polar bilinear form associated to a quadratic form Q: V \to k is alternating, given by e(v, w) = Q(v + w) - Q(v) - Q(w). In this case, the alternating pairing e arises as the derivative of Q, though Q is not uniquely recoverable from e without additional structure, such as the Arf invariant. Over fields of odd characteristic, non-degenerate alternating pairings instead yield the trivial quadratic form Q(v) = 0, highlighting their skew-symmetry. The classification of alternating pairings over a k of not 2 relies on the existence of a basis for non-degenerate forms on even-dimensional spaces. Specifically, if \dim V = 2n, there exists a basis \{e_1, f_1, \dots, e_n, f_n\} such that e(e_i, f_i) = 1 and e(e_i, e_j) = e(f_i, f_j) = 0 for all i, j, with the matrix of e being block-diagonal with n copies of the $2 \times 2 \begin{pmatrix} 0 & 1 \\ -1 & 0 \end{pmatrix}. This is unique up to the dimension n, confirming that all such pairings are equivalent under . Degenerate cases reduce to restrictions on symplectic quotients.

Applications

In Cryptography

In , bilinear pairings provide a powerful for constructing secure protocols, particularly on over finite fields. A cryptographic pairing is defined as an admissible e: G_1 \times G_2 \to G_T, where G_1, G_2, and G_T are cyclic groups of prime order r, typically with G_1 and G_2 as additive of points on an elliptic curve and G_T a multiplicative subgroup. The map satisfies bilinearity, meaning e(aP, bQ) = e(P, Q)^{ab} for points P \in G_1, Q \in G_2, and scalars a, b \in \mathbb{Z}_r; non-degeneracy, ensuring e(P, Q) generates G_T when P and Q are generators; and efficient computability via algorithms like Miller's. Pairings are classified as symmetric when G_1 = G_2 (often on supersingular curves for simplicity) or asymmetric when G_1 \neq G_2 (allowing more flexibility and efficiency on curves). Common constructions include the Weil pairing, originally from the of elliptic curves, and the Tate-Lichtenbaum pairing, which modifies the Tate pairing for practical use by raising to a power (q^k - 1)/r to yield a non-degenerate , where q is the base field size and k the embedding degree. These pairings enable novel cryptographic functionalities by allowing computations in G_T that are infeasible directly in G_1 or G_2. Key applications leverage these properties for advanced protocols. The first practical identity-based encryption (IBE) scheme, proposed by Boneh and Franklin in 2001, uses pairings to encrypt messages directly to a user's string as the public key, eliminating management and achieving chosen-ciphertext under the bilinear Diffie-Hellman (BDH) assumption. Short signatures, such as the Boneh-Lynn-Shacham (BLS) scheme, employ pairings for compact signatures verifiable via a single pairing evaluation, enabling efficient aggregation of multiple signatures into one (e.g., 154-bit signatures with 1024-bit ). Attribute-based encryption (ABE) extends this to fine-grained , where ciphertexts embed policies over user attributes, and decryption succeeds only if attributes satisfy the policy, as in the ciphertext-policy ABE construction using bilinear maps for policy evaluation. Security in pairing-based systems relies on the hardness of the BDH problem, where given generators P, aP, bP \in G_1 and Q \in G_2, computing e(P, Q)^{abc} \in G_T is intractable. Efficiency depends on the embedding degree [k](/page/K), the smallest integer such that [r](/page/R) divides [q](/page/Q)^[k](/page/K) - 1; low [k](/page/K) (e.g., 6–12) balances against discrete logarithm attacks in G_T with computational feasibility, as higher [k](/page/K) inflates sizes and pairing times.

In Representation Theory

In representation theory of finite groups, pairings arise prominently through the inner product defined on the space of class functions. For characters \chi and \psi of representations of a finite group G, the inner product is given by \langle \chi, \psi \rangle = \frac{1}{|G|} \sum_{g \in G} \chi(g) \overline{\psi(g)}, where \overline{\psi(g)} denotes the complex conjugate. This bilinear form is Hermitian and positive definite on the space of class functions, and it is non-degenerate when restricted to the irreducible characters, meaning that if \langle \chi, \psi \rangle = 0 for all irreducible \chi, then \psi = 0. The non-degeneracy ensures that irreducible characters uniquely determine the isomorphism class of representations and facilitate key structural results. The relations, established by Schur, assert that the irreducible characters of G form an for the space of s under this pairing: \langle \chi_i, \chi_j \rangle = \delta_{ij}, where \delta_{ij} is the , for distinct irreducible characters \chi_i and \chi_j. This implies that the number of irreducible representations equals the number of conjugacy classes in G, as the dimension of the class function space matches the number of classes, and the irreducibles span it. Column further states that for conjugacy classes C_k and C_l, \sum_{\chi \in \operatorname{Irr}(G)} \chi(C_k) \overline{\chi(C_l)} = |G| \delta_{kl} / |C_k|, reinforcing the basis property. The Frobenius-Schur indicator leverages this pairing framework to classify irreducible representations over the reals. For an irreducible character \chi, the indicator is \nu(\chi) = \frac{1}{|G|} \sum_{g \in G} \chi(g^2), which evaluates to 1 if the representation is realizable over \mathbb{R} with a symmetric invariant bilinear form, -1 if it admits a skew-symmetric form but not symmetric, and 0 otherwise, distinguishing real, quaternionic, and complex types. This value connects directly to the existence of non-degenerate G-invariant pairings on the representation space, as \nu(\chi) = \langle \chi, \chi \circ \sigma \rangle where \sigma(g) = g^2, detecting the reality of the representation via the inner product structure. These pairings enable the decomposition of any finite-dimensional representation \rho of G into irreducibles: the multiplicity of an irreducible \chi_i in \chi_\rho is n_i = \langle \chi_i, \chi_\rho \rangle, yielding \chi_\rho = \sum_i n_i \chi_i. This projection formula, rooted in , ensures complete reducibility over \mathbb{C} and provides an algorithmic tool for computing decomposition matrices. Connections to emerge here, as the implies |G| = \sum_{\chi \in \operatorname{Irr}(G)} \chi(1)^2, counting the order of G via representation dimensions and underscoring the pairing's role in global group structure.

Pairing Functions

A pairing function is a bijection \pi: \mathbb{N} \times \mathbb{N} \to \mathbb{N} that uniquely and reversibly encodes a pair of natural numbers into a single natural number, enabling the enumeration of the Cartesian product \mathbb{N} \times \mathbb{N} despite its infinite size. This construction demonstrates the countability of \mathbb{N} \times \mathbb{N}, a foundational result in set theory. The canonical example is the , introduced by in to support proofs of the uncountability of the real numbers by showing the countability of rational pairs. It is defined as \pi(k, l) = \frac{(k + l)(k + l + 1)}{2} + l, where the formula arises from summing the first k + l natural numbers and adjusting for the position within the antidiagonal. This function is bijective, with explicit inverse functions for decoding: the sum s = k + l is recovered as the smallest where \frac{s(s+1)}{2} \geq \pi(k, l), followed by l = \pi(k, l) - \frac{s(s+1)}{2} and k = s - l. Pairing functions possess key properties that make them computationally tractable: they are primitive recursive, meaning both the encoding and decoding can be implemented using only basic recursive operations like and primitive recursion, without unbounded minimization. This computability underpins their role in schemes, where they encode finite sequences of symbols (such as logical formulas) into unique natural numbers, enabling arithmetization in proofs of incompleteness theorems. A notable variant is the Szudzik pairing function, proposed in 2006 for its simplicity and efficiency in software applications. Defined as \pi(x, y) = \begin{cases} y^2 + x & \text{if } x < y, \\ x^2 + x + y & \text{otherwise}, \end{cases} it fills squares in a grid-like manner, avoiding the triangular overhead of Cantor's approach and simplifying inversion through square roots and modulo operations. In practice, it is valued for generating compact, unique identifiers in programming tasks, such as combining row and column indices for database storage or spatial hashing without collisions.

Combinatorial Pairings

In combinatorics, pairings refer to structured matchings or arrangements of elements in discrete settings, such as sequences, graphs, or formal languages, where elements are coupled under specific constraints to form balanced or optimal configurations. These differ from algebraic pairings by emphasizing and enumeration rather than bilinear forms. Key examples include Langford pairings, perfect matchings in graphs, and bracket pairings modeled as Dyck words. Langford pairings, also known as Langford sequences, are permutations of the {1, 1, 2, 2, \dots, n, n} where the two occurrences of each k are separated by exactly k other numbers, ensuring no two identical numbers are adjacent. For instance, for n=3, the sequence 2 3 1 2 1 3 satisfies the condition: the 1's have one number between them, the 2's have two, and the 3's have three. Such pairings exist n \equiv 0 \pmod{4} or n \equiv 3 \pmod{4}, as proven by analyzing positional constraints 4. In the , E. J. Groth applied Langford pairings to generate sequences with controllable complexity for efficient circuits, leveraging their structured separations to minimize computational overhead. Perfect matchings in graphs represent pairings as edge sets where every vertex is incident to exactly one edge, often termed 1-factors. In bipartite graphs G = (U \cup V, E) with |U| = |V|, a perfect matching exists if and only if Hall's condition holds: for every subset S \subseteq U, the neighborhood N(S) satisfies |N(S)| \geq |S|. This theorem, originally formulated for combinatorial designs, guarantees the pairing of elements across partitions without leftovers. For example, in a complete bipartite graph K_{n,n}, perfect matchings correspond to permutations of n elements, enabling systematic pairings. Bracket pairings model nested or non-crossing matchings, commonly represented as Dyck words—binary strings of balanced opening and closing parentheses, such as (())() for n=3 pairs, where no prefix has more closings than openings and the total counts match. These structures enumerate via the C_n = \frac{1}{n+1} \binom{2n}{n}, capturing valid sequences in formal languages. In bioinformatics, bracket pairings describe secondary structures, where pair non-crossingly to form stems and loops, as in dot-bracket notation for folding predictions. Combinatorial pairings find applications in scheduling and coding. In tournament scheduling, perfect matchings decompose the complete graph K_{2n} into $2n-1 rounds of n disjoint games, ensuring fair round-robin play without repeats, as surveyed in algorithmic constructions for sports leagues. In error-correcting codes, minimum-weight perfect matchings decode syndromes in quantum surface codes by pairing error excitations with minimal paths, optimizing correction thresholds in topological quantum memory, as implemented in efficient decoders like PyMatching.

References

  1. [1]
    Cooper Pairs and the BCS Theory of Superconductivity - HyperPhysics
    Cooper pairs are coupled electron pairs that can act as bosons, and their condensation is the basis for the BCS theory of superconductivity.
  2. [2]
    Pairing in nuclei - Physical Review Link Manager
    It is well known that nucleons can form paired states, analogous to the way electrons pair in superconducting metals.
  3. [3]
    [1405.1652] Overview of Neutron-Proton Pairing - arXiv
    May 7, 2014 · Abstract:The role of neutron-proton pairing correlations on the structure of nuclei along the N=Z line is reviewed.
  4. [4]
    Base Pair - National Human Genome Research Institute (NHGRI)
    A base pair consists of two complementary DNA nucleotide bases that pair together to form a “rung of the DNA ladder.”
  5. [5]
    5.4: Base Pairing in DNA and RNA - Biology LibreTexts
    Mar 17, 2025 · This page explains the rules of base pairing in DNA, where adenine pairs with thymine and cytosine pairs with guanine, enabling the double ...
  6. [6]
    [PDF] TENSOR PRODUCTS 1. Introduction Let R be a commutative ring ...
    The correct tensor product M ⊗R N for noncommutative R uses a right R-module M, a left R- module N, and a “middle-linear” map B where B(mr, n) = B(m, rn). In ...
  7. [7]
    [PDF] ALGEBRA - LECTURE V 1. Bilinear forms Let R be a commutative ...
    Bilinear forms. Let R be a commutative ring with 1, and M and N two R-modules. A map T : M → N is a homomorphism of R-modules if (i) T(v + u) = T(v) + T(u) for ...<|control11|><|separator|>
  8. [8]
    [PDF] Notes and questions for perfect pairings - Mathematics and Statistics
    A R-bilinear map h·,·i: M × N → L is called a pairing. Exercise 2. A good example to keep in mind is the pairing between M∗ and M. Specifically, define h·,· ...
  9. [9]
    [PDF] DUAL MODULES 1. Introduction Let R be a commutative ring. For ...
    So only reflexive modules could be candidates for being part of a perfect pairing. In any event, the point of pairings is to systematize the idea that we can ...
  10. [10]
    [PDF] Modules - OSU Math
    Feb 20, 2024 · perfect pairing of M1 and M2 identifies M1 with M∗. 2 and M2 with M∗. 1 . In the case M1 and M2 are free of finite rank, the bilinear forms ...
  11. [11]
    [PDF] Linear Algebra
    A scalar product on a vector space V is also called a symmetric bilinear ... be a linear map, symmetric with respect to the scalar product. Then V has an ...
  12. [12]
    [PDF] BILINEAR FORMS The geometry of Rn is controlled algebraically by ...
    In quantum mechanics, whose underlying mathematics involves heavy doses of linear algebra, the terms “degenerate” and “non-degenerate” are used with meanings ...
  13. [13]
    [PDF] The Matrix Cookbook
    Nov 15, 2012 · 3.2.6 Rank-1 update of inverse of inner product. Denote A = (XT X)−1 and that X is extended to include a new column vector in the end ˜X ...
  14. [14]
    [PDF] Introduction to representation theory - MIT Mathematics
    Jan 10, 2011 · Furthermore, the standard inner product. (x, y) = Xxiyi. 88. Page 89. on ZN restricts to the inner product B given by Γ on L, since it takes the ...
  15. [15]
    [PDF] An Elementary Introduction to the Hopf Fibration - Niles Johnson
    The Hopf fibration, named after Heinz Hopf, is an important object in mathematics and physics, defined as a mapping h:S3 → S2.
  16. [16]
    [PDF] Faisceaux Algebriques Coherents Jean-Pierre Serre The Annals of ...
    Sep 23, 2007 · Un faisceau algébrique cohérent sur une variété algébrique. V est simple~nent un faisceau cohérent de 8v-modules, 8~désignant le faisceau des ...
  17. [17]
    [PDF] Poincaré duality
    On the middle- dimensional homology Hm(M,R) the intersection form is a bilinear form which is symmetric if m is even and skew if m is odd. Since the pairing is ...
  18. [18]
    [PDF] Intersection Theory
    by corollary 4.2 to the Chow group of cycles modulo rational equivalence. The refined Gysin pullback exists not only for regular embeddings, but for any.
  19. [19]
    Ext in nLab
    Aug 23, 2023 · exhibiting Ext 1 ( X , A ) Ext^1(X,A) as the cokernel of Hom ( i , A ) Hom(i,A) . Yoneda product. The Yoneda product? is a pairing. Ext n ( ...
  20. [20]
    alternating form - PlanetMath
    Mar 22, 2013 · A two-dimensional vector space which carries a non-singular alternating form is sometimes called an alternating or symplectic hyperbolic plane.
  21. [21]
    Symplectic spaces
    Recall that ``alternating'' means that (v,v)=0 for all v in V, from which it follows that (v,w) = -(w,v) for all v,w in V; and ``nondegenerate'' means that the ...
  22. [22]
    [PDF] 1. Linear algebra preliminaries 1.1. Some facts about bilinear forms ...
    A bilinear form ψ : V ×V → k on a vector space V is called symplectic if it is skew-symmetric and non-degenerate. Furthermore, the pair (V,ψ) is called a ...
  23. [23]
    Hermitian form - PlanetMath
    Mar 22, 2013 · A sesquilinear form B:V×V→C B : V × V → ℂ over a single vector space V is called a Hermitian form if it is complex conjugate symmetric.
  24. [24]
    [PDF] Chapter 12 Hermitian Spaces - UPenn CIS
    Definition 12.2. Given a complex vector space E, a function ': E ⇥ E ! C is a sesquilinear form iff it is linear in its first argument and semilinear in its ...
  25. [25]
    [PDF] Study/read: Bilinear/quadratic forms - Penn Math
    Dec 11, 2017 · 2) Prove the polarization identity: f(x,y) = 1. 4 h qf (x + y) − qf ... For every bilinear alternating f ∈ L2 alt(V ,F) there exists a ...
  26. [26]
    [PDF] 18.704: Classification of Bilinear Forms over Finite Fields
    Mar 7, 2005 · Today's goal will be to classify all of the bilinear forms over finite fields of odd order. In order to classify them, we need to know when ...
  27. [27]
    Programming ECC - Bilinear Pairings
    Bilinear pairings are central to pairing-based cryptosystems, using a bilinear nondegenerate map. The Weil pairing is an example, but the Tate pairing is often ...
  28. [28]
    [PDF] Improved Weil and Tate pairings for elliptic and hyperelliptic curves
    The Weil and Tate pairings have been proposed for use in cryptography, includ- ing one-round 3-way key establishment, identity-based encryption, and short.
  29. [29]
    [PDF] Identity-Based Encryption from the Weil Pairing
    An extended abstract of this paper appears in the Proceedings of Crypto 2001, volume 2139 of Lecture Notes in Computer Science, pages. 213–229, Springer-Verlag, ...
  30. [30]
    Boneh, Lynn, Shacham – Short signatures from the Weil pairing
    No information is available for this page. · Learn why
  31. [31]
    [PDF] Ciphertext-Policy Attribute-Based Encryption - UT Computer Science
    In its simplest form, the decryption al- gorithm could require two pairings for every leaf of the access tree that is matched by a private key attribute and (at ...
  32. [32]
    A Note on the Bilinear Diffie-Hellman Assumption
    The Bi-linear Diffie-Hellman (BDH) intractability assumption is required to establish the security of new Weil-pairing based cryptosystems.
  33. [33]
    [PDF] Constructing Pairing-Friendly Elliptic Curves with Embedding ...
    Abstract. We present a general framework for constructing families of elliptic curves of prime order with prescribed embedding degree. We.
  34. [34]
    [PDF] Notes on Representations of Finite Groups
    the inner product between characters and the regular representation. Theorem 10.1 (Sum of Squares Formula). For G any finite group,. ∑ ρ∈Irrep(G). (dim ρ). 2.Missing: citation | Show results with:citation
  35. [35]
    [PDF] Finite Groups and Character Theory - Columbia Math Department
    The characters give us an explicit formula for the inner product on the representation ring. <V,W >= 1. |G|. X g∈G. χV χW. The map [V ] → χV extends to an ...
  36. [36]
    [PDF] representation theory of finite groups and burnside's theorem
    Aug 17, 2015 · In this paper we develop the basic theory of representations of finite groups, especially the theory of characters. With the help of the ...
  37. [37]
    [PDF] Math 210B. Frobenius-Schur indicator 1. Introduction Let G be a ...
    For example, the representation theory of symmetric groups is an extremely well-understood subject (relevant to many topics, ranging from cohomology rings.
  38. [38]
    Pairing Function -- from Wolfram MathWorld
    A pairing function is a function that reversibly maps Z^*×Z^* onto Z^*, where Z^*={0,1,2,...} denotes nonnegative integers. Pairing functions arise ...
  39. [39]
    [PDF] The Rosenberg-Strong Pairing Function - arXiv
    Jan 28, 2019 · Cantor's pairing function serves as an important example in elementary set theory (Enderton, 1977). It is also used as a fundamental tool in ...
  40. [40]
    Recursive Functions - Stanford Encyclopedia of Philosophy
    Apr 23, 2020 · ... Gödel number of an axiom of T is primitive recursive. This is so precisely because membership in the schemes in question is determined by a ...
  41. [41]
    [PDF] An Elegant Pairing Function - Matthew Szudzik
    It is possible to describe the points on a surface with only one coordinate, and a method for doing this was first described by Georg Cantor in 1878. Definition.