Fact-checked by Grok 2 weeks ago

Bialgebra

In mathematics, a bialgebra over a field K is a vector space over K that carries the structure of both a unital associative algebra and a counital coassociative coalgebra, with the two structures compatible via the requirement that the comultiplication map is an algebra homomorphism (or equivalently, that the multiplication map is a coalgebra homomorphism). The algebra structure consists of a multiplication map \mu: A \otimes A \to A and a unit map \eta: K \to A satisfying associativity \mu \circ (\mu \otimes \mathrm{id}) = \mu \circ (\mathrm{id} \otimes \mu) and unit axioms, while the coalgebra structure includes a comultiplication \Delta: A \to A \otimes A and a counit \epsilon: A \to K obeying coassociativity ( \Delta \otimes \mathrm{id} ) \circ \Delta = ( \mathrm{id} \otimes \Delta ) \circ \Delta and counit properties (\epsilon \otimes \mathrm{id}) \circ \Delta = \mathrm{id} = (\mathrm{id} \otimes \epsilon) \circ \Delta. The compatibility condition ensures that \Delta \circ \mu = (\mu \otimes \mu) \circ (\mathrm{id} \otimes \gamma \otimes \mathrm{id}) \circ (\Delta \otimes \Delta), where \gamma is the twist map, allowing the bialgebra to act simultaneously as a ring-like and coring-like object. This dual nature distinguishes bialgebras from ordinary or , enabling applications where both multiplicative and comultiplicative operations are needed. Prominent examples include the group algebra K[G] of a group G, where basis elements multiply as group elements and \Delta(g) = g \otimes g for g \in G, making it a bialgebra that captures the group's symmetries in a linear algebraic setting. Another key example is the universal enveloping algebra U(\mathfrak{g}) of a \mathfrak{g}, equipped with \Delta(x) = x \otimes 1 + 1 \otimes x for x \in \mathfrak{g}, which linearizes the Lie bracket while preserving a coalgebra structure derived from the primitive elements. Polynomial rings like K also form bialgebras under the standard multiplication and a coproduct such as \Delta(x) = x \otimes 1 + 1 \otimes x. Bialgebras serve as the foundational framework for Hopf algebras, which extend them by adding an antipode map S: A \to A that acts as a , satisfying \mu \circ (\mathrm{id} \otimes S) \circ \Delta = \eta \circ \epsilon = \mu \circ (S \otimes \mathrm{id}) \circ \Delta. This structure arises prominently in , where it models cohomology rings of spaces, and in , linking algebras to their enveloping algebras via functors between categories of restricted algebras and Hopf algebras. More broadly, bialgebras underpin quantum groups and , providing tools to study symmetries in deformed or quantized settings, with historical roots tracing back to topological applications by Milnor and in the mid-20th century.

Introduction

Overview

A bialgebra is a equipped with both an algebra structure, consisting of a and a , and a structure, consisting of a comultiplication and a counit, such that these structures are compatible with one another. This compatibility ensures that the operations interact seamlessly, allowing the vector space to function dually as both an algebraic and ic object. The motivation for bialgebras stems from the need to unify algebraic and coalgebraic frameworks, which traditionally operate in parallel but complementary ways. By combining (from the algebra side) with comultiplication (from the coalgebra side), bialgebras provide a powerful tool for exploring dualities and extensions of classical structures. Bialgebras play a central role in modern algebra, particularly in , where they enable the construction of tensor products and dual representations, extending insights from group and representations. They also arise naturally in quantum groups, facilitating the study of deformed symmetries, and in duality theory, where they highlight connections between algebras and their coalgebraic counterparts. This unification is essential for investigating symmetries and deformations across algebraic contexts.

Historical development

The origins of bialgebras trace back to the in , where and others explored group representations and the of H-spaces. Hopf's 1941 paper introduced the topological notion of H-spaces, whose cohomology rings possessed a compatible algebra and structure that later formalized as bialgebras, leading to the development of Hopf algebras as a special case with an antipode. This work, extended by Hans Samelson in 1941 through homology products, marked the early algebraic recognition of these dual structures in representing group-like objects. In the 1950s and early , the concepts were refined in the context of groups and hyperalgebras. coined the term "algèbre de Hopf" in 1953 to describe the cohomology algebras of H-spaces, honoring Hopf's contributions. Pierre Cartier formalized hyperalgebras as bialgebras in 1956, defining comultiplication compatible with the algebra structure for groups over fields of positive characteristic, thus establishing bialgebras as a foundational tool in . and John C. Moore provided a modern axiomatic definition of Hopf algebras in 1965, emphasizing the bialgebra core without initially requiring bijectivity of the antipode. The term "bialgebra" was formally introduced in 1969 by Moss E. Sweedler in his book Hopf Algebras, as a structure generalizing Hopf algebras by omitting the antipode to focus on the essential compatibility between algebra and coalgebra operations over commutative rings. In the 1970s, Mitsuhiro Takeuchi and others extended these ideas, with Takeuchi's work providing key structural insights, including generalizations to ×_R-bialgebras and extensions of to these settings. This broadened the applicability beyond topological origins, enabling studies of arbitrary rings. The 1970s and 1980s saw bialgebras evolve through links to quantum groups, revitalizing the field. and Michio Jimbo independently defined quantum groups in 1985 as deformed bialgebras (Hopf algebras) of universal enveloping algebras and function algebras on Lie groups, motivated by quantum integrable systems and Yang-Baxter equations. Drinfeld's formulation emphasized quantization of Lie bialgebras, while Jimbo's focused on q-deformations, establishing bialgebras as central to and physics.

Foundational Structures

Associative algebras

An over a k is a A over k equipped with a bilinear \mu: A \otimes_k A \to A that is associative, satisfying \mu \circ (\mu \otimes_k \mathrm{id}_A) = \mu \circ (\mathrm{id}_A \otimes_k \mu), and a distinguished element e \in A such that \mu(e \otimes_k a) = \mu(a \otimes_k e) = a for all a \in A. This structure makes (A, \mu, e) a unital associative algebra, where bilinearity ensures the multiplication respects the scalar operations. Prominent examples include the of n \times n matrices M_n(k) over k, which carries the standard as its bilinear operation and the as ; this is a non-commutative of n^2. Another example is the k[x_1, \dots, x_n] in n indeterminates, forming a commutative under the usual multiplication, with the constant 1 as the unit. An algebra homomorphism between associative algebras A and B over the same field k is a k- f: A \to B that preserves the multiplicative structure, satisfying f(\mu_A(a \otimes_k b)) = \mu_B(f(a) \otimes_k f(b)) for all a, b \in A, and maps the unit of A to the unit of B. Such morphisms form the arrows in the of associative algebras over k.

Coalgebras

A coalgebra over a k is a C equipped with two k-linear maps: a comultiplication \Delta: C \to C \otimes C and a counit \varepsilon: C \to k. These maps satisfy specific axioms that ensure a consistent structure on C. The coassociativity axiom states that the comultiplication allows unambiguous iterated decomposition, expressed as (\Delta \otimes \mathrm{id}_C) \circ \Delta = (\mathrm{id}_C \otimes \Delta) \circ \Delta. This equality means that applying \Delta twice and then projecting via one of the two possible tensor associations yields the same result in C \otimes C \otimes C. Geometrically, it corresponds to a :
       Δ
C ──────────→ C ⊗ C
  ↘              ↙
   ↖───── Δ ─────↙
    C ⊗ C ⊗ C
where the two paths from C to C \otimes C \otimes C coincide, ensuring that ternary decompositions are well-defined regardless of . The counit axioms provide a projection back to the original space, stating that (\varepsilon \otimes \mathrm{id}_C) \circ \Delta = \mathrm{id}_C, \quad (\mathrm{id}_C \otimes \varepsilon) \circ \Delta = \mathrm{id}_C. These ensure that the counit acts as a retraction for the comultiplication, allowing recovery of elements from their decompositions along either tensor factor. Examples of coalgebras include the dual of a finite-dimensional algebra, where for an algebra A with multiplication m: A \otimes A \to A and unit \eta: k \to A, the dual maps define \Delta(f)(a \otimes b) = f(m(a \otimes b)) and \varepsilon(f) = f(\eta(1)) for f \in A^*. Another example is the n-dimensional matrix coalgebra over k, with basis \{e_{ij} \mid 1 \leq i,j \leq n\}, comultiplication \Delta(e_{ij}) = \sum_{k=1}^n e_{ik} \otimes e_{kj}, and counit \varepsilon(e_{ij}) = \delta_{ij}, which models the decomposition of matrix units into row-column factors. A morphism of coalgebras between (C, \Delta_C, \varepsilon_C) and (D, \Delta_D, \varepsilon_D) is a k- f: C \to D such that \Delta_D \circ f = (f \otimes f) \circ \Delta_C and \varepsilon_D \circ f = \varepsilon_C, preserving the and structures. Coalgebras are to algebras in the sense that finite-dimensional examples arise as linear duals, reversing the arrows of and maps.

Bialgebra Definition

Core axioms

A bialgebra over a k is a A equipped with the structure of both an and a , where the algebraic and coalgebraic operations satisfy specific compatibility conditions ensuring that the and counit are algebra homomorphisms. Specifically, let (A, \mu, e) denote the associative algebra structure, with multiplication \mu: A \otimes A \to A and unit e: k \to A, and let (A, \Delta, \varepsilon) denote the coalgebra structure, with \Delta: A \to A \otimes A and counit \varepsilon: A \to k. The core axioms require that \Delta and \varepsilon preserve the algebra operations. The \Delta is an algebra morphism if it satisfies the multiplicativity condition \Delta \circ \mu = (\mu \otimes \mu) \circ (\mathrm{id}_A \otimes \tau \otimes \mathrm{id}_A) \circ (\Delta \otimes \Delta), where \tau: A \otimes A \to A \otimes A is the twist map defined by \tau(a \otimes b) = b \otimes a, for all a, b \in A, and the unitality condition \Delta(e(\lambda)) = e(\lambda) \otimes e(1) = e(1) \otimes e(\lambda) for all \lambda \in k, where the multiplication on A \otimes A is given by (a \otimes b)(c \otimes d) = ac \otimes bd. In Sweedler notation, this multiplicativity is expressed as \Delta(ab) = \Delta(a)\Delta(b), and unitality as \Delta(1) = 1 \otimes 1. Similarly, the counit \varepsilon is an algebra morphism if \varepsilon(\mu(a \otimes b)) = \varepsilon(a) \varepsilon(b) and \varepsilon(e(1)) = 1. These axioms extend naturally to the setting where k is replaced by a R, with A a over R that is flat as an R- to ensure tensor products behave well. In this more general context, the definitions of the and structures, along with the morphism properties, remain formally identical, allowing bialgebras to be studied over rings beyond fields.

Compatibility conditions

In a bialgebra A, the core compatibility condition requires that the comultiplication \Delta: A \to A \otimes A is an homomorphism with respect to the algebra structure on A and the induced algebra structure on A \otimes A, where the multiplication in the tensor product is defined by (a \otimes b)(c \otimes d) = (ac) \otimes (bd) for all a, b, c, d \in A. This ensures \Delta(ab) = \Delta(a) \Delta(b) for all a, b \in A, guaranteeing that the structure respects the multiplication in A. The counit \varepsilon: A \to k (where k is the base field) must similarly be an algebra homomorphism, implying \varepsilon(ab) = \varepsilon(a) \varepsilon(b) for all a, b \in A, along with \varepsilon(1) = 1 if A is unital. This multiplicative property of the counit follows directly from the homomorphism condition and ensures the coalgebra structure aligns with the algebra's unit. Equivalently, a bialgebra can be viewed as a in the of comonoids (with respect to the monoidal structure given by the ) or dually as a comonoid in the of monoids. These categorical perspectives highlight the intertwined nature of the structures without altering the axiomatic requirements. While the standard theory assumes associative unital algebras and coassociative coalgebras, non-unital variants such as weak multiplier bialgebras and non-associative extensions (e.g., certain nonassociative Hopf algebras) have been explored, though the foundational framework remains less complete and more specialized in these cases.

Properties and Morphisms

Coassociativity and counit

In a bialgebra A, the comultiplication \Delta: A \to A \otimes A inherits coassociativity from the underlying coalgebra structure, satisfying the axiom (\Delta \otimes \mathrm{id}_A) \circ \Delta = (\mathrm{id}_A \otimes \Delta) \circ \Delta. This ensures that the decomposition of elements into tensor products is unambiguous under different associations, a property fundamental to the coalgebra component. The coassociativity is verified in the bialgebra context where \Delta acts as a of algebras from A to the algebra A \otimes A, equipped with the induced . This property aligns the operation with the , allowing \Delta to respect both the associativity of and the coassociative . The counit \varepsilon: A \to [k](/page/k), where [k](/page/k) is the base , satisfies the counit axioms (\varepsilon \otimes \mathrm{id}_A) \circ \Delta = \mathrm{id}_A = (\mathrm{id}_A \otimes \varepsilon) \circ \Delta, which confirm that \varepsilon serves as a retraction for the comultiplication. In the bialgebra setting, compatibility with the m: A \otimes A \to A requires \varepsilon to be an algebra homomorphism, so \varepsilon(ab) = \varepsilon(a) \varepsilon(b) for all a, b \in A. This ensures the counit behaves multiplicatively, preserving scalar-like properties across products. A key consequence of coassociativity is the well-defined iterative comultiplication \Delta^n: A \to A^{\otimes n} for n \geq 1, constructed recursively by \Delta^1 = \Delta and \Delta^n = (\Delta \otimes \mathrm{id}^{\otimes (n-1)}) \circ \Delta^{n-1} (or symmetrically via the other association). This n-fold tensor extension facilitates the study of representations, where elements decompose into multi-component forms useful for analyzing symmetries and invariants in bialgebra modules. To sketch how the compatibility conditions preserve coassociativity, consider that \Delta being an homomorphism implies \Delta(m) = m_{A \otimes A} \circ (\Delta \otimes \Delta), where m_{A \otimes A} is the multiplication on the . Applying this to the algebra's associativity m \circ (m \otimes \mathrm{id}_A) = m \circ (\mathrm{id}_A \otimes m) and composing with the coassociativity on both sides yields equality of the two possible decompositions, confirming that the algebraic upholds the coalgebra's coassociative integrity without contradiction.

Bialgebra homomorphisms

A bialgebra homomorphism between two bialgebras A and B over a k is a k- f: A \to B that preserves both the and structures. Specifically, f is an if it satisfies f(ab) = f(a)f(b) for all a, b \in A and f(1_A) = 1_B, where $1_A and $1_B denote the respective unit elements. Simultaneously, f is a if it satisfies (f \otimes f) \Delta_A = \Delta_B f and \varepsilon_B f = \varepsilon_A, where \Delta_A, \Delta_B are the coproducts and \varepsilon_A, \varepsilon_B are the counits. The compatibility condition inherent to bialgebras—namely, that the coproduct is an homomorphism (or dually, the product is a homomorphism)—is automatically preserved by such homomorphisms. This follows because f respects each structure separately, ensuring that the intertwined operations remain compatible in the B. A bijective bialgebra homomorphism admits an inverse that is also a bialgebra homomorphism, yielding a bialgebra . Bialgebras over k, denoted \mathbf{Bialg}_k, form a category where the objects are k-bialgebras and the morphisms are bialgebra homomorphisms. This category supports standard constructions such as kernels (defined as bialgebra ideals, subspaces I \subseteq A with \Delta_A(I) \subseteq I \otimes A + A \otimes I) and coproducts.

Examples

Group bialgebra

The group bialgebra provides a fundamental example of a bialgebra arising from the structure of a finite group. Given a finite group G and a field k, the group bialgebra kG is constructed as the vector space over k with basis \{ g \mid g \in G \}. The algebra structure on kG is defined by the multiplication g \cdot h = gh, where gh denotes the group operation in G, and the unit element is the identity e of G, extended linearly to the entire space. The structure is introduced via the comultiplication \Delta(g) = g \otimes g and the counit \varepsilon(g) = 1 for each basis element g \in G, with both maps extended linearly to kG. This endows kG with a bialgebra structure, as the comultiplication and counit are algebra homomorphisms: \Delta(gh) = gh \otimes gh = (g \otimes g)(h \otimes h) = \Delta(g) \Delta(h), and \varepsilon(gh) = 1 = \varepsilon(g) \varepsilon(h), satisfying the compatibility condition. Coassociativity of the comultiplication follows directly from the group-like nature of the basis elements, since (\Delta \otimes \mathrm{id}) \Delta(g) = (g \otimes g) \otimes g = g \otimes (g \otimes g) = (\mathrm{id} \otimes \Delta) \Delta(g). This bialgebra construction captures the multiplicative structure of G in a linear algebraic setting and serves as a model for more general quantum group constructions. Modules over kG correspond precisely to representations of the group G, thereby representing group actions algebraically. Moreover, kG is dual to the coalgebra of representative functions on G, highlighting the interplay between algebraic and functional perspectives on .

Representative functions on groups

In the context of an algebraic group G over a field k, the space \operatorname{Rep}(G) consists of the representative functions on G, which are the k-linear combinations of matrix coefficients arising from finite-dimensional representations of G. These functions are polynomial in the matrix entries when G is embedded in \mathrm{GL}_n(k) for some n, and \operatorname{Rep}(G) spans the coordinate algebra O(G) of G. Specifically, for a representation \rho: G \to \mathrm{GL}(V) on a finite-dimensional vector space V with basis \{v_i\}, the matrix coefficient f(g) = \langle \rho(g) v_i, v_j \rangle for some dual basis lies in \operatorname{Rep}(G). The algebra structure on \operatorname{Rep}(G) is given by pointwise multiplication: for f, h \in \operatorname{Rep}(G), (f \cdot h)(g) = f(g) h(g) for all g \in [G](/page/G), making \operatorname{Rep}(G) a commutative associative k- with unit the constant function $1. The coalgebra structure is defined by the comultiplication \Delta: \operatorname{Rep}(G) \to \operatorname{Rep}(G) \otimes \operatorname{Rep}(G) given by \Delta(f)(g, h) = f(gh) for g, h \in G, and the counit \varepsilon: \operatorname{Rep}(G) \to k by \varepsilon(f) = f(e), where e is the of G. This endows \operatorname{Rep}(G) with a coassociative coalgebra structure, as coassociativity follows directly from the associativity of the group in G: (\Delta \otimes \mathrm{id}) \Delta(f)(g, h, k) = f((gh)k) = f(g(hk)) = (\mathrm{id} \otimes \Delta) \Delta(f)(g, h, k). The compatibility condition for the bialgebra structure holds because \Delta and \varepsilon are algebra homomorphisms: for f, h \in \operatorname{Rep}(G), \Delta(f \cdot h)(g, h) = (f \cdot h)(gh) = f(gh) h(gh) = \Delta(f)(g, h) \Delta(h)(g, h), with a similar verification for the unit. Under suitable conditions, such as when G is finite, \operatorname{Rep}(G) is isomorphic to the dual of the group algebra kG as Hopf algebras, reflecting the duality between the functional perspective on G and its group element basis. For general affine algebraic groups, this duality manifests in the Hopf algebra pairing between O(G) and the space of distributions on G, with \operatorname{Rep}(G) playing a central role in generating the former.

Other bialgebras

One prominent example of a bialgebra is the polynomial algebra k over a k, equipped with the standard and a structure defined by the comultiplication \Delta(x) = x \otimes 1 + 1 \otimes x and counit \varepsilon(x) = 0. This structure makes k a connected graded bialgebra, where the primitive elements generate the algebra, and it serves as the Borel Hopf subalgebra in certain contexts related to quantum groups. Another fundamental example arises from the universal enveloping algebra U(\mathfrak{g}) of a \mathfrak{g} over a k. Here, the algebra multiplication is the standard one induced by the tensor algebra quotient, while the comultiplication is defined on generators by \Delta(X) = X \otimes 1 + 1 \otimes X for X \in \mathfrak{g}, making these elements primitive, and extended multiplicatively to the full algebra. The counit is \varepsilon(X) = 0 for X \in \mathfrak{g}. The compatibility condition for the bialgebra structure holds because the comultiplication on primitive generators satisfies the required distributivity over the multiplication, as verified by the fact that U(\mathfrak{g}) is freely generated by primitives and the Lie bracket is preserved in the tensor product via the Leibniz rule. This construction is cocommutative and plays a central role in and . Non-commutative examples include coordinate algebras like the quantum plane k_q[x, y], generated by x and y with the relation xy = q yx for a scalar q \in k \setminus \{0, 1\}, equipped with a twisted comultiplication such as \Delta(x) = x \otimes x, \Delta(y) = y \otimes 1 + x \otimes y, and counit \varepsilon(x) = 1, \varepsilon(y) = 0. This structure deforms the classical and exemplifies non-cocommutative bialgebras arising in theory. While introductory treatments often emphasize finite-dimensional or cocommutative bialgebras, the diversity includes infinite-dimensional cases like k and non-cocommutative ones like the quantum plane, highlighting broader applications in and physics.

References

  1. [1]
    [PDF] Bialgebras and Hopf algebras - UChicago Math
    A Hopf algebra is a bialgebra with a given antipode. If A and B have antipodes χ, then A ⊗ B has the antipode χ ⊗ χ. Remark 3.2. The original definition of ...Missing: mathematics | Show results with:mathematics
  2. [2]
    [PDF] AN INTRODUCTION TO HOPF ALGEBRAS 1. Tensor products and ...
    First, the definition of a bialgebra is really sym- metric with respect to its multiplicative and comultiplicative structures. While this is not apparent ...<|control11|><|separator|>
  3. [3]
    Coalgebras, bialgebras and Hopf algebras
    We have seen that we like things that are both algebras and coalgebras,. compatibly. These are known as bialgebras. ... (i) (B, µ, u) is an algebra. (ii) (B, ∆, ε) ...
  4. [4]
    [PDF] Introduction to Hopf algebras and representations
    We will soon also motivate this definition with properties of representations. Exercise 15 (Alternative definitions of bialgebra). Let B be a vector space ...
  5. [5]
    [0901.2460] The beginnings of the theory of Hopf algebras - arXiv
    Jan 16, 2009 · Hopf algebras are named from the work of Heinz Hopf in the 1940's. In this note we trace the infancy of the subject back to papers from the ...
  6. [6]
    A characterization of the Galois subbialgebras Hk(KF) - ScienceDirect
    , Pages 315-362. Journal of Algebra. A characterization of the Galois subbialgebras H k ( K F ). Author links open overlay panel. Mitsuhiro Takeuchi.
  7. [7]
    [PDF] A primer of Hopf algebras
    Jun 17, 2021 · ... associative algebra U(g) coincide. One defines a linear map δ : g ... definition of the Hopf algebra Γ. U(g). Theorem 3.8.2. (Cartier ...
  8. [8]
    [PDF] Coalgebras
    Oct 28, 2016 · This is a coalgebra with Δ(eij)=Σeik⊗ekj and ε(eij)=δij. It's called the matrix coalgebra. 3.4) Let G be a finite group and k(G) the vector ...
  9. [9]
    coalgebra in nLab
    Jul 15, 2025 · A coalgebra is just like an associative algebra, but with all the structure maps 'turned around' (a “co-monoid”).
  10. [10]
    [PDF] 13. Hopf algebras - Columbia Math Department
    Reversing all the arrows, we obtain the definition of a coalgebra. Def. A Ik-linear coalgebra C is a lk-vector space C equipped with linear maps satisfying ...<|control11|><|separator|>
  11. [11]
    [PDF] Math 207A: Hopf Algebras (Lecture Notes)
    Definition 3.2 (Hopf algebra). A bialgebra H is a Hopf algebra if, in the convolution algebra Homk(H, H), idH : H → H has an inverse S: H → H under ...
  12. [12]
    [PDF] Monoidal Categories, Bialgebras, and Automata - Cornell Mathematics
    Oct 29, 2009 · Bialgebras, Bimonoids. Definition. A bimonoid B is a monoid in a category of comonoids, or equivalently, a comonoid in a category of monoids.
  13. [13]
    [1306.1466] Weak multiplier bialgebras - arXiv
    Jun 6, 2013 · A non-unital generalization of weak bialgebra is proposed with a multiplier-valued comultiplication. Certain canonical subalgebras of the ...
  14. [14]
    [PDF] Bialgebras and Hopf algebras - MIT OpenCourseWare
    An algebra H equipped with a comultiplication Δ and a counit ε satisfying properties (i),(ii) of Theorem 1.21.1 is called a bialgebra. Thus, Theorem 1.21.1 ...
  15. [15]
    [PDF] Bialgebras 1. Sweedler's sigma notation
    Definition 1: A bialgebra is a quintuple (H, µ, η, ∆,) such that (H, µ, η) is an algebra, (H, ∆,) is a coalgebra and one of the equivalent conditions of Theorem ...
  16. [16]
    Bialgebras - Category Framework
    Indeed, the category of bialgebras defines no additional structure: a morphism of coalgebras and of algebras between two bialgebras is a bialgebra morphism.Missing: homomorphism | Show results with:homomorphism
  17. [17]
    Foundations of Quantum Group Theory - Shahn Majid - Google Books
    Foundations of Quantum Group Theory ... Now in paperback, this is a graduate level text for theoretical physicists and mathematicians which systematically lays ...
  18. [18]
    [PDF] Basic Theory of Affine Group Schemes - James Milne
    Mar 11, 2012 · This book is a modern exposition of the basic theory of affine group schemes, emphasizing those of finite type over a field, and also ...
  19. [19]
    [PDF] Three variations on the linear independence of grouplikes in a ... - HAL
    Sep 22, 2020 · A (x) = x ⊗ 1 + 1 ⊗ x and ǫ (x) = 0. Then, it is easy to check that the k-algebra k [x], equipped with the comultiplication. A and the counit ǫ, ...
  20. [20]
    [PDF] arXiv:2211.15885v3 [math.RA] 5 Sep 2023
    Sep 5, 2023 · This gives a strongly graded twisting map by construction, and the twisted tensor product is isomorphic to the quantum plane kq[x, y] = khx ...