Fact-checked by Grok 2 weeks ago

Frobenius algebra

A Frobenius algebra is a finite-dimensional A over a k, equipped with a nondegenerate \beta: A \times A \to k that is associative, meaning \beta(xy, z) = \beta(x, yz) for all x, y, z \in A, or equivalently, a linear functional \varepsilon: A \to k such that the kernel of \varepsilon contains no nontrivial left ideals. This structure endows A with both an algebra (multiplication \mu: A \otimes A \to A and unit \eta: k \to A) and a coalgebra (comultiplication \delta: A \to A \otimes A and counit \varepsilon: A \to k) in a compatible manner, satisfying the Frobenius condition (\mu \otimes \mathrm{id})(\mathrm{id} \otimes \delta) = \delta \circ \mu = (\mathrm{id} \otimes \mu)(\delta \otimes \mathrm{id}), along with associativity, coassociativity, and unit/counit axioms. The nondegeneracy of the form ensures that \beta induces an isomorphism A \cong A^* (the dual space), highlighting the self-dual nature of Frobenius algebras. These algebras are finite-dimensional by necessity, and the Frobenius structure is unique up to scaling by invertible elements in A. Frobenius algebras admit several equivalent formulations, including as Frobenius objects in a symmetric like \mathrm{Vect}_k, where they serve as bimonoids with the specified . They may be commutative (if multiplication and comultiplication respect the braiding) or symmetric (if the is symmetric), with the of two Frobenius algebras again forming one. Prominent examples include the matrix algebra \mathrm{Mat}_n(k) with the trace form \beta(X, Y) = \mathrm{tr}(XY), group algebras kG for finite groups G with the augmentation map as counit, and cohomology rings of compact oriented manifolds, which often carry a Frobenius structure via . In applications, Frobenius algebras underpin two-dimensional topological quantum field theories (TQFTs), establishing a categorical equivalence between commutative Frobenius algebras over k and 2D TQFTs with values in \mathrm{Vect}_k, as developed in works connecting algebraic and geometric structures. They also arise in , (e.g., quasi-Frobenius rings), and physics, bridging with and .

Algebraic Definition and Structure

Definition

A Frobenius algebra is a finite-dimensional unital A over a k equipped with a \sigma: A \times A \to k that is non-degenerate and satisfies the compatibility condition \sigma(ab, c) = \sigma(a, bc) for all a, b, c \in A. This form, often called the Frobenius form or pairing, endows the algebra with a duality that aligns its multiplicative structure with a trace-like functional. The non-degeneracy of \sigma means that the induced linear maps A \to A^*, defined by a \mapsto (b \mapsto \sigma(a, b)) and b \mapsto (a \mapsto \sigma(a, b)), are isomorphisms, where A^* = \mathrm{Hom}_k(A, k) is the of A. This ensures that \sigma provides a perfect pairing between A and itself, with no nontrivial annihilators. A Frobenius algebra is symmetric if the bilinear form satisfies \sigma(a, b) = \sigma(b, a) for all a, b \in A. The concept is named after , who introduced it around 1900 in the context of group representations through his work on hypercomplex numbers.

Nakayama Automorphism

In a Frobenius algebra (A, \sigma), where \sigma: A \times A \to k is a non-degenerate associative over a field k, the Nakayama automorphism \nu: A \to A is the algebra satisfying \sigma(a, b) = \sigma(\nu(b), a) for all a, b \in A. This condition ensures that the bilinear form is "twisted symmetric" via \nu. The automorphism \nu arises from the bimodule structure induced by \sigma. Specifically, the map \phi: A \to A^* defined by \phi(b)(a) = \sigma(a, b) is an isomorphism of left A-modules, as associativity of \sigma implies \phi(bc) = b \cdot \phi(c), where the left action on A^* is (a \cdot f)(x) = f(x a). The Nakayama automorphism \nu is the unique (up to inner automorphisms) algebra automorphism such that \phi \circ \nu is an isomorphism of A-bimodules, with the right action on A^* given by (f \cdot a)(x) = f(a x). Existence and uniqueness of \nu (up to inner automorphisms) follow from the non-degeneracy of \sigma, which guarantees \phi is bijective, and the rigidity of the bimodule category for finite-dimensional algebras over fields, ensuring a unique twisting automorphism matches the right module structures. If two automorphisms \nu_1, \nu_2 satisfy the condition, then \nu_1 \circ \nu_2^{-1} commutes with left and right multiplications, hence is inner. The Frobenius algebra is symmetric if there exists a symmetric non-degenerate associative bilinear form \tau on A, in which case the corresponding Nakayama automorphism is inner, i.e., \nu(a) = u a u^{-1} for some invertible u \in A. Conversely, if \nu is inner for some choice of \sigma, then \tilde{\sigma}(a, b) = \sigma(\nu(a), b) defines a symmetric form. To compute \nu in coordinates, choose a basis \{e_i\}_{i=1}^n for A. Let S = (s_{pq}) with s_{pq} = \sigma(e_p, e_q). The c_j of \nu(e_j) satisfies S^T c_j = S e_j, or equivalently c_j = (S^T)^{-1} S e_j, where e_j is the . For a symmetric form, one can choose a basis such that S = I, in which case \nu is the identity automorphism.

Properties

Basic Properties

A Frobenius algebra A over a k is self-injective, meaning that A is an injective left (and right) A-module. This property arises from the non-degenerate bilinear form \sigma: A \times A \to k, which induces an isomorphism \Phi: A \to A^* given by \Phi(a)(b) = \sigma(a, b), where A^* = \mathrm{Hom}_k(A, k). Since A is projective as a module over itself and \Phi preserves the module structure via the associativity condition \sigma(ab, c) = \sigma(a, bc), the injectivity follows by constructing explicit injections using the form to split extensions. In particular, the socle \mathrm{soc}_l(A) (the sum of simple left submodules) and the Jacobson radical \mathrm{rad}(A) are related via \sigma, as the form identifies \mathrm{soc}_l(A) with the annihilator of \mathrm{rad}(A) in the dual space, ensuring duality between minimal injective and projective structures. The bilinear form \sigma induces a trace functional \mathrm{tr}: A \to k defined by \mathrm{tr}(a) = \sigma(a, 1). By the associativity of \sigma, we have \mathrm{tr}(ab) = \sigma(ab, 1) = \sigma(a, b) for all a, b \in A. The Nakayama automorphism \nu: A \to A, characterized by \sigma(a, b) = \sigma(b, \nu(a)) for all a, b \in A, enables the cyclicity property \mathrm{tr}(ab) = \mathrm{tr}(ba) by twisting the form to yield a symmetric pairing. Non-degeneracy ensures that \mathrm{tr} is nonzero on the socle and that its kernel contains no nontrivial left ideals, as any such ideal would contradict the isomorphism \Phi. The non-degeneracy of \sigma implies that \dim_k A = \dim_k A^*, with the explicit isomorphism \Phi providing the equality. To see this, the map \Phi is injective because if \Phi(a) = 0, then \sigma(a, b) = 0 for all b \in A, so a = 0 by non-degeneracy; surjectivity follows similarly from the finite-dimensional setting and the dual basis argument using associativity to match module actions. If A = A_1 \oplus A_2 as a direct sum of subalgebras that are orthogonal with respect to \sigma (i.e., \sigma(A_i, A_j) = 0 for i \neq j), then each A_i is a Frobenius subalgebra with the restricted form \sigma|_{A_i \times A_i}. Non-degeneracy on A implies non-degeneracy on each A_i, as the orthogonal complement of A_i is A_j and the form vanishes across summands. Associativity preserves the condition on each component: for a, b, c \in A_i, \sigma(ab, c) = \sigma(a, bc) holds within the restriction, while cross terms vanish by orthogonality.

Structural Properties

Commutative Frobenius algebras over a are precisely the zero-dimensional Gorenstein rings. Specifically, for a finite-dimensional commutative algebra A over a k, A admits a Frobenius structure if and only if it is Gorenstein, meaning its socle is one-dimensional as a k- when A is local. This equivalence arises because the Gorenstein condition ensures the existence of a non-degenerate functional whose contains no nonzero ideals, defining the Frobenius form. A Frobenius algebra A is symmetric if its Nakayama automorphism \nu is inner, meaning there exists a unit u \in A^\times such that \nu(a) = u a u^{-1} for all a \in A. This is equivalent to the existence of a linear functional \lambda: A \to k such that \ker \lambda contains no nonzero left ideals and \lambda(rs) = \lambda(sr) for all r, s \in A, making the induced bilinear form symmetric. For local algebras, A is symmetric if and only if \soc(A_A) \not\subseteq [A, A], the commutator subspace. The of two Frobenius algebras is again Frobenius. If A and B are Frobenius over the same field k with forms \langle -, - \rangle_A and \langle -, - \rangle_B, then A \otimes_k B admits the \langle a_1 \otimes b_1, a_2 \otimes b_2 \rangle = \langle a_1, a_2 \rangle_A \langle b_1, b_2 \rangle_B, which is non-degenerate and associative. The corresponding Nakayama is \nu_A \otimes \nu_B, given explicitly by (\nu_A \otimes \nu_B)(a \otimes b) = \nu_A(a) \otimes \nu_B(b). This construction follows from the bimodule (A \otimes B)^* \cong A^* \otimes B^* twisted appropriately by the forms. In a Frobenius algebra A, the socle \soc(A) is to the dual of the Jacobson as bimodules up to the Nakayama twist, but more precisely, the Frobenius form induces an \soc(A) \cong \Hom_k(A / \rad(A), k) as left A-modules. To see this, note that \soc(A) = \ann_A(\rad(A)), and the form B: A \times A \to k restricts to a perfect pairing between \soc(A) and A / \rad(A), since elements of \rad(A) pair non-trivially only outside the socle due to non-degeneracy, yielding the dual via s \mapsto ( \overline{a} \mapsto B(s, a) ). This duality swaps the socle and the top in the module category. Any two Frobenius forms on the same algebra A differ by multiplication by a nonzero scalar in k^\times. If B_1 and B_2 are two non-degenerate associative bilinear forms, then B_2 = \lambda B_1 for some \lambda \in k^\times, because the induced maps A \to A^* differ by an automorphism of A^*, but since both realize the same bimodule isomorphism A \cong {}_A A^*, the scalar follows from trace considerations on the identity endomorphism.

Examples

Finite-Dimensional Algebras

Finite-dimensional Frobenius algebras over a k provide illustrations of the abstract definition, where the algebra A admits a non-degenerate k- \sigma: A \times A \to k satisfying \sigma(ab, c) = \sigma(a, bc) for all a, b, c \in A. These examples highlight how the Frobenius structure arises in familiar algebraic settings, with explicit forms allowing direct verification of the axioms. A prominent example is the group algebra kG for a finite group G, which is a finite-dimensional associative algebra with basis the elements of G and multiplication extended linearly from the group operation. The Frobenius form is given by \sigma(a, b) = \epsilon(ab), where \epsilon is the augmentation map sending the identity element to 1 and all other group elements to 0; equivalently, this extracts the coefficient of the identity in the product ab. To verify non-degeneracy, note that the form induces an isomorphism between kG and its dual as bimodules, ensuring the kernel is zero. For the associativity condition, \sigma(ab, c) = \epsilon(abc) = \epsilon(a(bc)) = \sigma(a, bc), which holds by associativity of the algebra multiplication. Another example is the full matrix algebra M_n(k), the algebra of n \times n over k with standard matrix multiplication. This is a symmetric Frobenius algebra with the form \sigma(A, B) = \operatorname{tr}(AB), where \operatorname{tr} denotes the matrix trace. Non-degeneracy follows from the fact that if \operatorname{tr}(AB) = 0 for all B, then A = 0, as the trace form is the standard inner product on matrices up to basis choice. The associativity \sigma(AB, C) = \operatorname{tr}(ABC) = \operatorname{tr}(A(BC)) = \sigma(A, BC) holds by the cyclic property of the trace and associativity of matrix multiplication. In the symmetric case, the Nakayama automorphism is the . The algebra of dual numbers k[\varepsilon]/(\varepsilon^2), a two-dimensional algebra with basis \{1, \varepsilon\} and multiplication (a + b\varepsilon)(c + d\varepsilon) = ac + (ad + bc)\varepsilon, admits the Frobenius form \sigma(a + b\varepsilon, c + d\varepsilon) = ad + bc. To check non-degeneracy, represent the form in the basis \{1, \varepsilon\}: the matrix is \begin{pmatrix} 0 & 1 \\ 1 & 0 \end{pmatrix}, with determinant -1 \neq 0. For associativity, compute \sigma((a + b\varepsilon)(c + d\varepsilon), e + f\varepsilon) = \sigma(ac + (ad + bc)\varepsilon, e + f\varepsilon) = (ac)f + (ad + bc)e = acf + ade + bce, and \sigma(a + b\varepsilon, (c + d\varepsilon)(e + f\varepsilon)) = \sigma(a + b\varepsilon, ce + (cf + de)\varepsilon) = a(cf + de) + b(ce) = acf + ade + bce, confirming equality. This structure is one of the classified two-dimensional Frobenius algebras over k. Path algebras of quivers without oriented cycles offer further examples under certain conditions, specifically when the quiver has no arrows, yielding the semisimple commutative algebra k^I for a finite set I of vertices, with basis \{e_i \mid i \in I\} and multiplication e_i e_j = \delta_{ij} e_i. The form \sigma(\sum a_i e_i, \sum b_i e_i) = \sum a_i b_i is symmetric Frobenius. Non-degeneracy is evident as the matrix is the identity, hence invertible. Associativity follows from \sigma((\sum a_i e_i)(\sum b_j e_j), \sum c_k e_k) = \sigma(\sum a_i b_i e_i, \sum c_k e_k) = \sum a_i b_i c_i = \sigma(\sum a_i e_i, (\sum b_j e_j)(\sum c_k e_k)). More generally, such path algebras are hereditary, and the Frobenius property holds precisely when they are semisimple, as in this case.

Ring-Theoretic Examples

In the commutative setting, a local is Frobenius if and only if its socle is . This condition ensures the existence of a non-degenerate associative compatible with the multiplication. For example, over a k, the k[x,y]/(x^2, y^2) has basis \{1, x, y, xy\} and socle generated by the ideal (xy), making it Frobenius with the form defined by \langle ab, cd \rangle = \delta_{b,c} \delta_{d,x} \delta_{a,y} + \cdots (explicitly checked via the socle generator). Similarly, truncated like k/(x^n) are Frobenius, with socle (x^{n-1}). These examples extend to over non- of finite type, such as complete local with of finite length , where the socle condition verifies the Frobenius property through injectivity of the as a over itself. A counterexample illustrates the necessity of the simple socle: the ring k[x,y]/(x,y)^2 = k[x,y]/(x^2, xy, y^2) over a field k is local Artinian with maximal ideal (x,y) and socle spanned by \{x, y\}, which is two-dimensional and thus not simple. Any attempted bilinear form on this ring fails non-degeneracy, as the left annihilator of the socle equals the maximal ideal, whose dimension exceeds the socle's corank, violating the isomorphism between the ring and its dual. This failure highlights that while the ring is self-injective on one side, the lack of simple socle prevents a compatible non-degenerate form. In the non-commutative case, finite-dimensional Hopf algebras over a field provide prominent examples of Frobenius algebras. By the Larson-Sweedler theorem, every finite-dimensional Hopf algebra admits a non-degenerate form via its modular integral trace, where the form \phi(h) = \int h for the left integral \int satisfies \phi(hg) = \phi(gh) under the antipode. For instance, the group algebra kG for finite group G is Frobenius with trace \sum_{g \in G} \epsilon_g, and more generally, any semisimple Hopf algebra inherits this structure. Compatibility is checked by verifying the Frobenius condition \phi(ab) = \phi(ba) using the integral's invariance. The exterior algebra \Lambda(V) on a finite-dimensional V over a also carries a Frobenius structure, defined by the \langle \alpha, \beta \rangle = \int \alpha \wedge \beta, where the integral projects to the one-dimensional top-degree component \Lambda^{\dim V}(V) paired with a fixed . This form is non-degenerate because wedging with elements of positive degree annihilates lower forms, ensuring the dual basis matches the algebra basis in the graded sense. Over commutative rings of finite type (e.g., Artinian), the exterior algebra remains Frobenius if the module is projective of finite rank, with the form extending via the module. Commutative Frobenius rings coincide with Gorenstein Artinian rings in the local case.

Categorical Perspectives

Category-Theoretical Definition

In category theory, particularly in the context of homological algebra, a Frobenius functor provides a framework for defining Frobenius algebras abstractly. Consider categories \mathcal{C} and \mathcal{D} equipped with exact structures. A Frobenius functor consists of an exact functor F: \mathcal{C} \to \mathcal{D} that admits both a left adjoint F_! and a right adjoint F^*, forming ambidextrous adjunctions F_! \dashv F \dashv F^*. Central to the structure is a natural transformation \eta: \mathrm{id}_\mathcal{C} \to F^* F satisfying compatibility axioms with the unit and counit of the adjunctions F \dashv F^*. These axioms ensure that the composite F \eta_{F^*}: F F^* \to F F^* F F^* coincides with the unit \epsilon_{F F^*} of the further adjunction F F^* \dashv F^* F F^*, and dually for the other side, mirroring the Frobenius relations via the following key diagram (where horizontal compositions denote whiskering): \begin{CD} \mathrm{id}_\mathcal{C} @>\eta>> F^* F \\ @V{F}VV @VV{F F^* F}V \\ F @= F \\ @V{F^*}VV @VV{F^* F F^*}V \\ F F^* @>\epsilon_{F F^*}>> \mathrm{id}_\mathcal{D} \end{CD} The vertical maps and the equality follow from the ambidexterity, ensuring the transformation induces a Frobenius monad on \mathcal{C}. This categorical setup links directly to algebraic structures: finite-dimensional Frobenius algebras over a field k are in equivalence with Frobenius functors between categories of finite-dimensional vector spaces and modules. Specifically, for a finite-dimensional algebra A over k, the tensor functor F = A \otimes_k -: {}_k\mathrm{Vect}_\mathrm{fd} \to A\mathrm{-Mod}_\mathrm{fd} is exact, with right adjoint F^* = \mathrm{Hom}_k(A, -), and the Frobenius bilinear form on A induces the natural transformation \eta_V: V \to \mathrm{Hom}_k(A, A \otimes_k V) via v \mapsto (a \mapsto a \cdot v), satisfying the required axioms if and only if A is Frobenius. Conversely, any such Frobenius functor yields a Frobenius algebra as the image of the unit object under F. The structure emerges from the T = F^* F on {}_k\mathrm{Vect}_\mathrm{fd} induced by the adjunction F \dashv F^*, where T carries both monoidal and comonoidal structures compatible via the Frobenius conditions on \eta and the counit, making T a Frobenius object in the of endofunctors. This 's Eilenberg-Moore recovers the over the underlying . This perspective originated in the through works in , notably by Alex Heller, who explored functorial constructions like loop-space functors in categories, laying groundwork for ambidextrous structures and their relation to Frobenius properties. The equivalence to the classical algebraic definition follows from representability: the Frobenius form \beta: A \times A \to k represents the natural \mathrm{Hom}_k(A \otimes_k V, W) \cong \mathrm{Hom}_k(V, \mathrm{Hom}_k(A, W)) via \beta, yielding \eta as the transpose; non-degeneracy of \beta ensures the exactness and axiom satisfaction, while the Nakayama arises from the comparison between left and right adjoints.

Equivalent Formulations

A Frobenius algebra over a k (or typically a for finite-dimensional cases) admits an equivalent formulation as a finite-dimensional associative unital algebra A equipped with a structure consisting of a \Delta: A \to A \otimes_k A and a counit \varepsilon: A \to k, satisfying coassociativity (\mathrm{id} \otimes \Delta) \circ \Delta = (\Delta \otimes \mathrm{id}) \circ \Delta and counitality (\varepsilon \otimes \mathrm{id}) \circ \Delta = (\mathrm{id} \otimes \varepsilon) \circ \Delta = \mathrm{id}_A, along with the Frobenius condition that ensures compatibility between the multiplication m: A \otimes_k A \to A and \Delta: (m \otimes \mathrm{id}) \circ (\mathrm{id} \otimes \Delta) = \delta \circ m = (\mathrm{id} \otimes m)(\Delta \otimes \mathrm{id}). This twist relates the algebra and coalgebra operations, making A a bialgebra-like structure without an antipode. For the symmetric case, an equivalent definition uses an : A is a symmetric Frobenius algebra if it possesses a k-linear functional \mathrm{tr}: A \to k such that \mathrm{tr}(ab) = \mathrm{tr}(ba) for all a, b \in A ( under ) and the induced \langle a, b \rangle = \mathrm{tr}(ab) is non-degenerate, meaning that for any basis \{e_i\} of A, the set \{\mathrm{tr}(a e_i) \mid i\} spans the for each a \in A, or equivalently, the map A \to A^* given by b \mapsto \mathrm{tr}(b \cdot -) is an . In the semisimple case, where A decomposes into a of matrix algebras over division rings, the Frobenius-Perron of A (defined via the dominant eigenvalue of the fusion matrix from its semisimple decomposition) coincides with its ordinary over k, providing a categorical that aligns with the 's . These formulations are equivalent to the standard one via canonical constructions. From the trace (in the symmetric case), the bilinear form recovers directly as \langle a, b \rangle = \mathrm{tr}(ab), and non-degeneracy ensures the form's associativity and compatibility with multiplication follow from the trace's properties; conversely, given a non-degenerate associative bilinear form \beta: A \times A \to k that is symmetric, define \mathrm{tr}(a) = \beta(a, 1_A), which yields an invariant trace since \beta(ab, c) = \beta(a, bc) implies \mathrm{tr}(ab) = \beta(ab, 1) = \beta(a, b) and symmetry gives \beta(a, b) = \beta(b, a) = \mathrm{tr}(ba). For the general (possibly non-symmetric) case, the trace \mathrm{tr}(a) = \beta(a, 1_A) satisfies \mathrm{tr}(ab) = \beta(a, b) but \mathrm{tr}(ab) = \mathrm{tr}(ba) only if \beta is symmetric. For the coalgebra side, the coproduct arises as \Delta(a) = \sum_i e_i \otimes f_i where \{e_i\}, \{f_i\} is the dual basis with respect to \beta, satisfying the Frobenius condition by the form's non-degeneracy, and the counit is \varepsilon(a) = \beta(a, 1_A); the reverse constructs the form from \beta(a, b) = \varepsilon(ab). When the bilinear form is symmetric (\langle a, b \rangle = \langle b, a \rangle), the resulting Frobenius algebra is symmetric, corresponding to the case where the Nakayama automorphism is the identity. These algebraic equivalences underpin the category-theoretical perspective, where Frobenius algebras arise as objects with compatible monad and comonad structures.

Applications

Topological Quantum Field Theories

Frobenius algebras play a central role in the construction of topological quantum field theories (TQFTs), particularly in low-dimensional cases where they provide the algebraic structure underlying the functorial assignments to manifolds and cobordisms. In this framework, a Frobenius algebra over a assigns a finite-dimensional to the circle, serving as the state space for closed 1-manifolds. Linear maps corresponding to 2-dimensional cobordisms are then defined using the algebra's and comultiplication: the multiplication arises from the "pants" cobordism, which connects two incoming circles to one outgoing circle, while the (or counit) is induced by the "" cobordism, pairing an incoming circle with the empty manifold. In two dimensions, there is a full equivalence between commutative Frobenius algebras and 2D TQFTs, establishing a one-to-one correspondence via symmetric monoidal functors from the category of 2D cobordisms to the category of vector spaces. This equivalence implies that every such TQFT is determined by a commutative Frobenius algebra, with the algebra's structure capturing the topological invariants assigned to surfaces. For non-commutative cases, the functoriality is projective, meaning assignments are defined up to natural isomorphism, and the non-associativity in the comonoid structure is resolved by incorporating the Nakayama twist, an automorphism that adjusts the pairing to ensure consistency across equivalent cobordisms. A prominent example arises in the Reshetikhin-Turaev TQFT, constructed from modular tensor categories, where the Frobenius algebra structure on objects ensures the theory's modularity by providing Δ-separable symmetric pairings that classify surface defects and maintain the invertibility required for invariants. This Frobenius condition guarantees the absence of non-trivial transparent objects, a key property for the TQFT's topological invariance. The connection between Frobenius algebras and TQFTs originated with Atiyah's axiomatization of TQFTs in , which emphasized functorial assignments to bordisms, and was rigorously developed in subsequent work, including ' 1996 demonstration of the equivalence for 2D cases and extensions by Lekili exploring bordered structures. These contributions highlight Frobenius algebras as the algebraic core of low-dimensional TQFTs, bridging and linear .

Representation Theory and Beyond

In , the group algebra kG over a k is a Frobenius algebra, which facilitates the study of modules and characters through its self-injective property, where projective and injective modules coincide. This structure enables the decomposition of the into simple modules and supports relations for characters of simple modules, ensuring that distinct irreducible representations are orthogonal with respect to the trace inner product. Frobenius algebras thus classify representations via the Wedderburn decomposition of group rings, particularly in modular settings where Brauer characters, defined on p-regular elements, extend the classical to characteristic p, capturing the composition factors of modules over blocks of the algebra. In , symmetric Frobenius algebras provide non-degenerate bilinear forms that define self-duality for codes over Frobenius , ensuring the code equals its dual under the standard inner product. Self-dual codes constructed this way, such as those over Galois rings, yield extremal codes with optimal minimum distances and connect to codes via lifting constructions, where the Frobenius form preserves duality in the embedding. The self-injectivity of the supports code duality without additional stabilizers. Frobenius structures appear in quantum information through entanglement-assisted quantum error-correcting codes (EAQECCs) over local Frobenius rings, where classical additive codes generate quantum codes using isotropic subspaces and hyperbolic pairs from the ring's trace form. Post-2020 developments quantify the minimum number of entangled qudits as half the rank of the code modulo its trace-dual, enabling constructions like the ((6,4,1)) code over \mathbb{Z}_4 with improved distances via lengthening. These frameworks extend to anyon models in topological quantum codes, where Frobenius algebras model fusion rules and error syndromes in non-Abelian anyon theories. In Hopf algebra applications, modular invariant Frobenius algebras arise from automorphisms of factorizable ribbon Hopf algebras, yielding commutative symmetric structures invariant under the mapping class group of the punctured torus. These algebras compute modular invariants in vertex operator algebras, linking to partition functions in logarithmic conformal field theories via bulk-boundary correspondences.

Generalizations

Frobenius Extensions

A Frobenius extension is a ring extension B \subseteq A in which A is a finitely generated projective B-module, and there exists a B-linear trace map \operatorname{tr}: A \to B that is non-degenerate and satisfies \operatorname{tr}(ab) = \operatorname{tr}(ba) for all a \in A and b \in B. The non-degeneracy means that for every nonzero a \in A, there exists b \in A such that \operatorname{tr}(ab) \neq 0, and for every nonzero b \in A, there exists a \in A such that \operatorname{tr}(ba) \neq 0, ensuring the trace induces a B-bimodule duality A \cong {}_B \Hom_B(A, B)_A. This structure generalizes the notion of a , where B is the base ring (often a field), to arbitrary ring inclusions. In many cases, such as when B is a field or a PID, A is free of finite rank over B. The concept of Frobenius extensions was introduced by Friedrich Kasch in the early 1960s as a generalization of separable extensions, allowing for the study of projective dimensions and structures across the extension. In Kasch's framework, such extensions preserve certain homological properties, such as the equality of projective and injective dimensions for modules when passing from B to A. Examples include the group algebra [kG](/page/KG) over a k, where G is a ; here, [kG](/page/KG) is of |G| over k, and the given by the augmentation (sum of coefficients) satisfies the required since it is k-linear and invariant under multiplication. Another example arises from polynomial rings in quotients, such as k/(x^n) over k, which forms a Frobenius extension of n with \operatorname{tr}(a_0 + a_1 x + \cdots + a_{n-1} x^{n-1}) = a_0, satisfying \operatorname{tr}(ab) = \operatorname{tr}(ba) due to the commutative nature of the . Key properties of Frobenius extensions include that A is projective as a B-. Additionally, the Nakayama automorphism extends from the Frobenius algebra case: there exists a B-automorphism \sigma: A \to A such that \operatorname{tr}(ab) = \operatorname{tr}(b \sigma(a)) for all a, b \in A, which captures the twisting in the bimodule isomorphism {}_B \Hom_B(A, B) \cong {}_B A_A and plays a role in ring theorems. The existence of such a symmetric non-degenerate , together with A being finitely generated projective over B, defines the Frobenius extension. Under mild conditions on B (e.g., when B is a ), projectivity implies freeness. To see the duality, the defines a non-degenerate \beta: A \times A \to B by \beta(a, b) = \operatorname{tr}(ab), which is symmetric over B by the given condition. This form induces an A \cong \operatorname{Hom}_B(A, B) as right A-modules via a \mapsto (x \mapsto \operatorname{tr}(xa)), and similarly for left modules, establishing the Frobenius duality.

Frobenius Adjunctions

In category theory, a Frobenius adjunction refers to an adjunction F \dashv G between categories \mathcal{A} and \mathcal{B} where G serves as both the left and right adjoint to F, or equivalently, the functors form a Frobenius pair with mutual adjointness. This structure arises in the context of abelian categories, where the natural isomorphism \Hom_{\mathcal{B}}(FX, Y) \cong \Hom_{\mathcal{A}}(X, GY) is supplemented by the reverse adjunction G \dashv F, ensuring a balanced reciprocity akin to Frobenius reciprocity in its full axiomatic form. Such adjunctions induce a Frobenius monad on \mathcal{A}, where the endofunctor GF is equipped with compatible monoidal and comonoidal structures derived from the unit and counit of the adjunction. The exactness properties of Frobenius adjunctions are particularly notable in abelian categories with sufficient projectives. Specifically, if \mathcal{A} and \mathcal{B} admit finite projective resolutions and the adjunction involves functors preserving these, then both F and G are , preserving short exact sequences and maintaining the homological structure. This exactness facilitates applications in , where the adjunction preserves kernels and cokernels, ensuring compatibility with derived categories. A example occurs in : for finite groups H \subseteq G over a k, the functor \Ind_H^G : \Rep_k(H) \to \Rep_k(G) forms a Frobenius adjunction with the restriction functor \Res_H^G : \Rep_k(G) \to \Rep_k(H), since coinduction coincides with (\Coind_H^G \cong \Ind_H^G) in finite-dimensional representations, yielding mutual adjointness. This realizes Frobenius reciprocity explicitly, with the projection formula \Ind(\Res(W) \otimes V) \cong W \otimes \Ind(V) holding naturally. The connection to Frobenius functors emphasizes the coalgebra-like conditions on the unit \eta : \id_{\mathcal{A}} \to GF and counit \epsilon : FG \to \id_{\mathcal{B}}, which intertwine to form a Frobenius comonad on \mathcal{B}, mirroring the algebraic duality in Frobenius algebras. Developments in this area emerged post-1960s within , with foundational work on such structures in module categories by Auslander and Reiten, who explored their implications for almost split sequences and stable categories in self-injective settings.

References

  1. [1]
    [PDF] Frobenius algebras and 2D topological quantum field theories (short ...
    Feb 3, 2010 · 1 Definition of Frobenius algebra. A Frobenius algebra is a k-algebra A of finite dimension, equipped with a linear functional ε : A → k ...
  2. [2]
    [PDF] constructing frobenius algebras - UChicago Math
    Definition 1. An algebra A over a field K is a vector space over K together with a K-linear vector multiplication µ : A ⊗ A → A ...
  3. [3]
    [PDF] koch-frobenius-algebras.pdf - Cambridge - MIT Mathematics
    2.2.1 Definition of Frobenius algebra. A Frobenius algebra is a k-algebra. A of finite dimension, equipped with a linear functional ε : A → k whose.<|control11|><|separator|>
  4. [4]
  5. [5]
    [PDF] Nakayama automorphisms of Frobenius algebras - arXiv
    Feb 19, 2014 · So the Nakayama automorphism is determined up to com- position with inner automorphisms; equivalently, it is a well-defined element of the group ...
  6. [6]
    [PDF] Frobenius algebras
    This leads to three equivalent ways to define a Frobenius algebra namely as a k-algebra together with a piece of extra structure. ◮ (A,Φ), where Φ: A.
  7. [7]
    [PDF] Bilinear Forms on Frobenius Algebras - arXiv
    Jan 25, 2014 · For any finite-dimensional algebra R, the dual ˆR is isomorphic as a left R- module to the injective hull of R/radR, so the isomorphism type of ...
  8. [8]
    [PDF] Nearly Frobenius Algebras - arXiv
    Jul 11, 2019 · a) For a given finite group G, the group algebra A := k[G] of a finite group together with the trace θ(Pagg) := a1, is a Frobenius algebra.
  9. [9]
    [PDF] arXiv:2109.12106v2 [math.QA] 25 Nov 2021
    Nov 25, 2021 · A Frobenius algebra A is a unital algebra over a field k equipped ... These calculations were done directly using the group algebra. We ...
  10. [10]
    [PDF] arXiv:2308.05246v2 [math.RA] 19 Dec 2024
    Dec 19, 2024 · – Any matrix algebra defined over a field F is a Frobenius algebra with Frobenius form σ(A, B) = tr(A · B), where tr(X) is the trace of ...
  11. [11]
    [PDF] Self-dual and quasi self-dual algebras - arXiv
    Nov 2, 2011 · Taking r = s = 2 it follows that k[x, y]/(x2,y2) is Frobenius. This is free of rank 4 over k, therefore of dimension 4 when k is a field.
  12. [12]
    [PDF] Non-commutative Algebra, SS 2019 Exercises 2
    Show that K[x]/(xn) is a symmetric algebra. Show that K[x, y]/(x2,y2) is a symmetric algebra. Show that. K[x, y]/(x, y)2 is not a Frobenius algebra. (d) Let ...
  13. [13]
    A Larson-Sweedler theorem for Hopf V-categories - ScienceDirect.com
    Jan 6, 2021 · The aim of this paper is to extend the classical Larson-Sweedler theorem, namely that a k-bialgebra has a non-singular integral (and in particular is Frobenius)
  14. [14]
    [PDF] frobenius algebras - Rice University
    RECENT WORK INVOLVING FROBENIUS ALGEBRAS. Page 4. I. HISTORY & THE VARIOUS DEFIUS OF FROBENIUS ALGEBRAJ/IR. A 1903- STYLE GAME. DUE TO F.G. FROBENIUS (1849-1937).
  15. [15]
    [PDF] frobenius algebras and ambidextrous adjunctions
    Jan 21, 2006 · Every Frobenius algebra in the category Vect arises from an ambidex- trous adjunction in the 2-category whose objects are algebras, morphisms ...
  16. [16]
    The Loop-Space Functor in Homological Algebra - jstor
    In a Frobenius category the projective (injective) dimension of any object or map is either 0 or oo. 4. Loop-space and permutation in 3C8. If 3d is an abelian ...
  17. [17]
    Frobenius Algebras and 2-D Topological Quantum Field Theories
    This 2003 book describes a striking connection between topology and algebra, namely that 2D topological quantum field theories are equivalent to commutative ...
  18. [18]
    [PDF] Frobenius Algebra Structure in Hopf Algebras and Cohomology ...
    Oct 17, 2020 · If f in the dual space H∗ is a non-zero left (or right) integral in. H∗, then the bilinear form it defines β : H ⊗ H → k, β(h ⊗ k) = f ...
  19. [19]
    Frobenius-Perron Dimensions of Integral Z + -rings and Applications
    Nov 7, 2019 · We introduce the notion of the Frobenius-Perron dimension of an integral -ring and give some applications of this notion to classification of finite ...
  20. [20]
    [PDF] Two-dimensional topological quantum field theories and Frobenius ...
    The relationship between two-dimensional topological quantum eld theories and Frobenius algebras is then formulated as an equivalence of categories.<|control11|><|separator|>
  21. [21]
    Invariants of 3-manifolds via link polynomials and quantum groups
    Invariants of 3-manifolds via link polynomials and quantum groups · Article PDF · References · Author information · Additional information · Rights and permissions.
  22. [22]
    [PDF] Line and surface defects in Reshetikhin-Turaev TQFT - arXiv
    The surface defects are labelled by. ∆-separable symmetric Frobenius algebras and the line defects by. “multi-modules” which are equivariant with respect to a ...
  23. [23]
    [PDF] Topological quantum field theory - Numdam
    MICHAEL ATIYAH b) Holomorphic Conformal Field Theories. These are not strictly topological QFTs in our sense since they depend on a complex structure, and ...
  24. [24]
    [PDF] A Course in Finite Group Representation Theory
    (e) For a Frobenius algebra (as in (d)), show that the left A-module AA is injective. (f) Show that A is a symmetric algebra (that is, there is a non ...
  25. [25]
    CHARACTERS AND ORTHOGONALITY IN FROBENIUS ALGEBRAS
    There we derive the orthogonality relations and de- monstrate a relation between characters and certain homogeneous modules over Frobenius algebras. In § 4 we ...
  26. [26]
    Entanglement-Assisted Quantum Error-Correcting Codes over Local ...
    Feb 1, 2022 · In this paper, we provide a framework for constructing entanglement-assisted quantum error-correcting codes (EAQECCs) from classical additive codes over a ...Missing: measures anyon post- 2020
  27. [27]
    Quantum error-correcting codes from Abelian anyon theories - PIRSA
    Dec 6, 2023 · In this talk, I will discuss three works in which we extend the stabilizer formalism to topological orders beyond that of the toric code. This ...Missing: Frobenius entanglement measures post-<|separator|>
  28. [28]
    [1106.0210] Modular invariant Frobenius algebras from ribbon Hopf ...
    Jun 1, 2011 · For any finite-dimensional factorizable ribbon Hopf algebra H and any ribbon automorphism omega of H, we establish the existence of the ...
  29. [29]
    [PDF] From Subfactors to Categories and Topology I. - arXiv
    (iii) The Frobenius algebra satisfies the equivalent conditions (ii) of Proposition 5.1. Then E is spherical iff A is spherical and Q is normalized. If ...
  30. [30]
    [PDF] Frobenius functors and transfer
    The functors F : A→B and G : B→A are said to form a Frobenius pair if G is both a left and a right adjoint of F. This concept was introduced.Missing: coalgebra | Show results with:coalgebra
  31. [31]
    Frobenius reciprocity in nLab
    ### Summary of Frobenius Reciprocity (Categorical Definition)
  32. [32]
    [math/0502550] Frobenius algebras and ambidextrous adjunctions
    Feb 25, 2005 · In this paper we explain the relationship between Frobenius objects in monoidal categories and adjunctions in 2-categories.
  33. [33]
    [PDF] relative stable categories and birationality - UCLA Mathematics
    May 7, 2021 · This situation is often referred to as a Frobenius adjunction. V a F a V , see [Mor65]. This occurs when one considers induction, restriction ...
  34. [34]
    [1703.10636] Hilsum-Skandalis maps as Frobenius adjunctions with ...
    Mar 30, 2017 · The charaterisation is that the adjunction is covered by a stable Frobenius adjunction that is a slice and whose right adjoint is monadic.
  35. [35]
    ambidextrous adjunction in nLab
    ### Summary of Ambidextrous Adjunction
  36. [36]
    [PDF] An Introduction to Auslander-Reiten Theory
    Oct 13, 2006 · Aim of these lectures is to provide a brief introduction to the notion of an almost split sequence and its use in the representation theory ...Missing: adjunctions | Show results with:adjunctions