This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Handle element on nearly Frobenius algebras

Dalia Artenstein, Ana González and Gustavo Mata
(July 28, 2025)
Abstract

In this article the concept of handle element of Frobenius algebras, as in [16], will be extended to nearly Frobenius algebras. The main properties of this element will be analyzed in this case and many examples will be constructed. Also the Casimir and Schur elements of symmetric algebras will be considered ([8]) and generalized for Frobenius and nearly Frobenius algebras showing which results still hold in this framework and which ones fail.

1 Introduction

The notion of (commutative) nearly Frobenius algebra was introduced in the thesis of the second author of this article. It was motivated by the result proved in [10], which states that: the homology of the free loop space HL(LM)HL^{\ast}(LM) has the structure of a Frobenius algebra without counit. Nearly Frobenius algebras were studied in [12], and more recently from an algebraic point of view in [4], [5], [6], [7] and [14].

The handle element (also called characteristic element or distinguished element) of a Frobenius algebra was introduced by L. Abrams in his thesis (see [1]). There, he proved the following two results, that show it is an useful concept:

Theorem 2.3.3 of [1]. A Frobenius algebra is semisimple if and only if the handle element is a unit.

Proposition 2.3.4 of [1]. In a Frobenius algebra, the ideal generated by the handle element coincides with the socle of the algebra.

The notion of Schur element of an absolutely irreducible representation of a symmetric algebra was first introduced by M. Geck in [11]. In [8] a slight generalization of that notion was presented for an epimorphism between two symmetric algebras. In that article the following result was proved:

Theorem 4.6 of [8] If λ:AB\lambda:A\rightarrow B is an epimorphism between symmetric algebras, the following properties are equivalent.

  1. 1.

    The Schur element sϕs_{\phi} is invertible in BB.

  2. 2.

    ϕ:AB\phi:A\twoheadrightarrow B is split as a morphism of AA-AA-bimodules.

  3. 3.

    BB is a projective right AA-module.

  4. 4.

    Any projective BB-module is a projective right AA-module.

In this article we study the previous concepts, and it is organized as follows.

In section 3 we introduce the notion of handle element for nearly Frobenius algebras. We prove an analogous to Theorem 2.3.3 from [1]: an algebra AA admits a nearly Frobenius structure where the handle element is a unit of AA if and only if AA is separable (Theorem 3.8). We prove that, under certain hypothesis, the handle element of a nearly Frobenius algebra AA belongs to its Jacobson radical (Proposition 3.10), but it does not belong to its socle generally (Example 3.9). It shows that Proposition 2.3.4 from [1] cannot be generalized to nearly Frobenius algebras.

In section 4 we introduce the notion of Schur element for nearly Frobenius algebras. We prove a generalization of Theorem 4.6 from [8] where λ:AB\lambda:A\rightarrow B is an epimorphism of algebras between Frobenius algebras (Proposition 12), and we also prove that (1)(2)(1)\Rightarrow(2) if AA is a symmetric nearly Frobenius algebra and BB is a symmetric algebra. However, the converse is not true (Example 4.2).

2 Preliminaries

2.1 Path algebras

If Q=(Q0,Q1,s,t)Q=(Q_{0},Q_{1},\mathrm{s},\mathrm{t}) is a finite connected quiver, 𝕜Q\Bbbk Q denotes its associated path algebra. Given ρ\rho a path in 𝕜Q\Bbbk Q, we denote by l(ρ)l(\rho), s(ρ)\mathrm{s}(\rho) and t(ρ)\mathrm{t}(\rho) the length, start and target of ρ\rho, respectively. We denote by JJ the ideal generated by all the arrows from Q1Q_{1}. Let xQ0x\in Q_{0}, we say that xx is a sink (source) if there are no arrows ρ\rho such that s(ρ)=x\mathrm{s}(\rho)=x (t(ρ)=x\mathrm{t}(\rho)=x).
We say that an ideal I𝕜QI\subset\Bbbk Q is admissible if there exist k2k\geq 2 such that JkIJ2J^{k}\subset I\subset J^{2}.

In the following definitions consider II an admisible ideal and A=𝕜QIA=\frac{\Bbbk Q}{I}.

Definition 2.1.

AA is a radical square zero algebra if I=J2I=J^{2}.

Definition 2.2.

The algebra AA is a toupie algebra if QQ has one sink (ω)(\omega) and one source (0)(0), and for every vertex xx different to ω\omega and 0 there is only one arrow α\alpha such that s(α)=x\mathrm{s}(\alpha)=x and only one arrow β\beta such that t(β)=x\mathrm{t}(\beta)=x.

2.2 Frobenius and nearly Frobenius algebras

In what follows RR is an associative commutative ring with unit.

Definition 2.3.

An associative RR-algebra AA is Frobenius if AA is finitely generated and projective as a RR-module, and there exists a left AA-module isomorphism

λl:AA.\lambda_{l}:A\rightarrow A^{*}.
Remark 2.4.

The isomorphism λl\lambda_{l} induces a morphism εA:AR\varepsilon_{A}:A\to R as εA=λl(1A),\varepsilon_{A}=\lambda_{l}\bigl{(}1_{A}\bigr{)}, which is called trace.

Definition 2.5.

An associative RR-algebra AA is symmetric if AA is a Frobenius algebra with trace εA:AR\varepsilon_{A}:A\to R such that εA(ab)=εA(ba)\varepsilon_{A}(ab)=\varepsilon_{A}(ba) for all a,bAa,b\in A.

Another characterization of Frobenius algebra, given by Abrams for example in [2], is that an algebra AA is Frobenius if it has a coassociative counital comultiplication Δ:AARA\Delta:A\rightarrow A\otimes_{R}A which is a map of AA-bimodules.

From this characterization, the following definition arises naturally:

Proposition 2.6.

If AA is a symmetric algebra, then its coproduct verifies the following property

Δ(1A)=τΔ(1A),\Delta\bigl{(}1_{A}\bigr{)}=\tau\circ\Delta\bigl{(}1_{A}\bigr{)},

where τ\tau is the transposition.

Proof.

The coproduct is characterized by the relation

ε(x(ab))=ε(x1a)ε(x2b),whereΔ(x)=x1x2\varepsilon\bigl{(}x(ab)\bigr{)}=\sum\varepsilon\bigl{(}x_{1}a\bigr{)}\varepsilon\bigl{(}x_{2}b\bigr{)},\;\mbox{where}\;\Delta(x)=\sum x_{1}\otimes x_{2}

then, using that ε(ab)=ε(ba)\varepsilon(ab)=\varepsilon(ba) for all a,bAa,b\in A, we have that Δ(1A)=τΔ(1A).\Delta\bigl{(}1_{A}\bigr{)}=\tau\circ\Delta\bigl{(}1_{A}\bigr{)}.

Definition 2.7.

An associative RR-algebra AA is a nearly Frobenius algebra if it admits an homomorphism Δ:AARA\Delta:A\rightarrow A\otimes_{R}A of AA-bimodules.

Let (A,ΔA)(A,\Delta_{A}) and (B,ΔB)(B,\Delta_{B}) be two nearly Frobenius algebras. An homomorphism f:ABf:A\rightarrow B is a nearly Frobenius homomorphism if it is a morphism of RR-algebras and the following diagram commutes

AfΔABΔBARAffBRB.\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 18.80473pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\crcr}}}\ignorespaces{\hbox{\kern-6.75pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 25.75609pt\raise 6.1111pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{f}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 55.15286pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern-18.2501pt\raise-18.79166pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.88889pt\hbox{$\scriptstyle{\Delta_{A}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 0.0pt\raise-30.25pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 55.15286pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 62.19626pt\raise-18.79166pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.88889pt\hbox{$\scriptstyle{\Delta_{B}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 62.19626pt\raise-30.25pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-18.80473pt\raise-37.58331pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{A\otimes_{R}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 20.28905pt\raise-43.69441pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{f\otimes f}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 42.80473pt\raise-37.58331pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 42.80473pt\raise-37.58331pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{B\otimes_{R}B}$}}}}}}}\ignorespaces}}}}\ignorespaces.

Based on Proposition 2.6 we give the following definition.

Definition 2.8.

A nearly Frobenius algebra AA is symmetric if the coproduct satisfies that

Δ(1A)=τΔ(1A).\Delta\bigl{(}1_{A}\bigr{)}=\tau\circ\Delta\bigl{(}1_{A}\bigr{)}.
Example 2.1.

We consider (A,Δ)(A,\Delta) the following nearly Frobenius algebra, AA is the truncated polynomial algebra in one variable 𝕜[x]/xn+1\Bbbk[x]/x^{n+1} and

Δ(1)=i+j=n+1xixj.\Delta(1)=\sum_{i+j=n+1}x^{i}\otimes x^{j}.

It is immediate to see that (A,Δ)(A,\Delta) is a symmetric nearly Frobenius algebra.

Remark 2.9.

The Frobenius space A\mathcal{E}_{A} (see [4] for definition) associated to the algebra AA admits structure of AA-module, where the action is given as follows:

AAAuΔ[uΔ:AAAx(xu)Δ(1)]\begin{array}[]{ccc}A\otimes\mathcal{E}_{A}&\to&\mathcal{E}_{A}\\ u\otimes\Delta&\mapsto&\left[\begin{array}[]{cccc}u\ast\Delta:&A&\to&A\otimes A\\ &x&\mapsto&\bigl{(}x\otimes u\bigr{)}\Delta(1)\end{array}\right]\end{array}

Observe that

x(uΔ)(1)=(x1)(1u)Δ(1)=(1u)(x1)Δ(1)=(1u)Δ(1)(1x)=(uΔ)(1)xx\cdot(u\ast\Delta)(1)=(x\otimes 1)(1\otimes u)\Delta(1)=(1\otimes u)(x\otimes 1)\Delta(1)=(1\otimes u)\Delta(1)(1\otimes x)=(u\ast\Delta)(1)\cdot x

3 Handle element

In the context of Frobenius algebras, in [16], is defined the handle element as follows.
If AA is a Frobenius algebra, introduce the handle operator as the RR-linear map ω:AA\omega:A\to A defined as the composite mΔm\circ\Delta, where mm is the multiplication of AA and Δ\Delta is the coproduct. This operator is a right (and left) AA-module homomorphism, and it is given by multiplication by a central element ωA,Δ\omega_{A,\Delta}. This element is called the handle element and ωA,Δ=mΔ(1)\omega_{A,\Delta}=m\circ\Delta(1).

3.1 Constructing handle elements

Definition 3.1.

If (A,Δ)(A,\Delta) is a nearly Frobenius RR-algebra we define its handle element as:

ωA,Δ:=mΔ(1)A.\omega_{A,\Delta}:=m\circ\Delta(1)\in A.
Remark 3.2.

The handle element is central: since Δ\Delta is an AA-bimodule morphism, if Δ(1)=xiyi\Delta(1)=\sum x_{i}\otimes y_{i} then xxiyi=xiyix\sum xx_{i}\otimes y_{i}=\sum x_{i}\otimes y_{i}x, for all xAx\in A. Therefore, applying the multiplication, xxiyi=xiyix\sum xx_{i}y_{i}=\sum x_{i}y_{i}x and x(ixiyi)=(ixiyi)xx\left(\sum_{i}x_{i}y_{i}\right)=\left(\sum_{i}x_{i}y_{i}\right)x which means xωA=ωAxx\omega_{A}=\omega_{A}x, for all xAx\in A and ωAZ(A)\omega_{A}\in Z(A).

Proposition 3.3.

The handle element respects product and tensor product structure, i.e.

ωA1×A2=(ωA1,ωA2),\omega_{A_{1}\times A_{2}}=(\omega_{A_{1}},\omega_{A_{2}}),
ωA1A2=ωA1ωA2.\omega_{A_{1}\otimes A_{2}}=\omega_{A_{1}}\otimes\omega_{A_{2}}.
Proof.

If we consider A=A1×A2A=A_{1}\times A_{2}, by Remark 6 of [6], we have that Δ(1)=(a1,0)(a2,0)+(0,b1)(0,b2)AA\Delta(1)=\sum(a_{1},0)\otimes(a_{2},0)+\sum(0,b_{1})\otimes(0,b_{2})\in A\otimes A. Then

ωA1×A2=(a1a2,0)+(0,b1b2)\omega_{A_{1}\times A_{2}}=\sum(a_{1}a_{2},0)+\sum(0,b_{1}b_{2})

By the other hand, ωA1=a1a2\omega_{A_{1}}=\sum a_{1}a_{2} and ωA2=b1b2\omega_{A_{2}}=\sum b_{1}b_{2}. Therefore

ωA1×A2=(a1a2,0)+(0,b1b2)=(ωA1,ωA2).\omega_{A_{1}\times A_{2}}=\left(\sum a_{1}a_{2},0\right)+\left(0,\sum b_{1}b_{2}\right)=\bigl{(}\omega_{A_{1}},\omega_{A_{2}}\bigr{)}.

Now, if we consider A=A1A2A=A_{1}\otimes A_{2}, again by Remark 7 of [6], we have that

ΔA:=τ(ΔA1ΔA2):A1A2(A1A2)(A1A2)\Delta_{A}:=\tau\circ\bigl{(}\Delta_{A_{1}}\otimes\Delta_{A_{2}}\bigr{)}:A_{1}\otimes A_{2}\rightarrow(A_{1}\otimes A_{2})\otimes(A_{1}\otimes A_{2})

where τ:(A1A1)(A2A2)(A1A2)(A1A2)\tau:(A_{1}\otimes A_{1})\otimes(A_{2}\otimes A_{2})\rightarrow(A_{1}\otimes A_{2})\otimes(A_{1}\otimes A_{2}) it the transposition map. Then

ωA=mAΔA=(mA1mA2)τ1τ(ΔA1ΔA2)=(mA1mA2)(ΔA1ΔA2)=(mA1ΔA1)(mA2ΔA2)=ωA1ωA2.\begin{array}[]{rcl}\omega_{A}=m_{A}\circ\Delta_{A}&=&\bigl{(}m_{A_{1}}\otimes m_{A_{2}}\bigr{)}\circ\tau^{-1}\circ\tau\circ\bigl{(}\Delta_{A_{1}}\otimes\Delta_{A_{2}}\bigr{)}\\ &=&\bigl{(}m_{A_{1}}\otimes m_{A_{2}}\bigr{)}\circ\bigl{(}\Delta_{A_{1}}\otimes\Delta_{A_{2}}\bigr{)}\\ &=&\bigl{(}m_{A_{1}}\circ\Delta_{A_{1}}\bigr{)}\otimes\bigl{(}m_{A_{2}}\circ\Delta_{A_{2}}\bigr{)}\\ &=&\omega_{A_{1}}\otimes\omega_{A_{2}}.\end{array}

Proposition 3.4.

If (A,Δ)(A,\Delta) is a nearly Frobenius RR-algebra and ω\omega is its handle element, then for all uZ(A)u\in Z(A), uωu\omega is the handle element of (A,uΔ)(A,u\ast\Delta).

Proof.

If Δ(1)=x1x2\Delta(1)=\sum x_{1}\otimes x_{2}, then uΔ(1)=x1ux2u\ast\Delta(1)=\sum x_{1}\otimes ux_{2}. Therefore

ωuΔ=x1ux2=ux1x2=ux1x2=uω.\omega_{u\ast\Delta}=\sum x_{1}ux_{2}=\sum ux_{1}x_{2}=u\sum x_{1}x_{2}=u\omega.

Corollary 3.5.

If (A,Δ)(A,\Delta) is a nearly Frobenius commutative RR-algebra and ω\omega is its handle element, then for all uAu\in A, uωu\omega is the handle element of the nearly Frobenius algebra (A,uΔ)(A,u\ast\Delta).

The next result appears in [9] in the context of Frobenius algebras.

Theorem 3.6.

Let (A,Δ)(A,\Delta) be a nearly Frobenius RR-algebra. If its handle element is a unit then AA is separable.

Proof.

If Δ(1)=xiyi\Delta(1)=\sum x_{i}\otimes y_{i}, then ωA=xiyi\omega_{A}=\sum x_{i}y_{i}. As ωAZ(A)\omega_{A}\in Z(A) we have that ωA1Z(A)\omega_{A}^{-1}\in Z(A). Then, defining e=ωA1ixiyie=\omega_{A}^{-1}\sum_{i}x_{i}\otimes y_{i} we obtain that:

  1. 1.

    xe=xωA1ixiyi=ωA1xixiyi=ωA1ixiyix=ex.x\cdot e=x\omega_{A}^{-1}\sum_{i}x_{i}\otimes y_{i}=\omega_{A}^{-1}x\sum_{i}x_{i}\otimes y_{i}=\omega_{A}^{-1}\sum_{i}x_{i}\otimes y_{i}x=e\cdot x.

  2. 2.

    m(e)=m(ωA1ixiyi)=ωA1xiyi=ωA1ωA=1m(e)=m\left(\omega_{A}^{-1}\sum_{i}x_{i}\otimes y_{i}\right)=\omega_{A}^{-1}\sum x_{i}y_{i}=\omega_{A}^{-1}\omega_{A}=1.

Then, by Proposition 19 of [6], we conclude that AA is separable. ∎

Some examples of applications of Theorem 3.6 are given below.

Example 3.1.

Let be AA the matrix algebra Mn×n(𝕜)M_{n\times n}(\Bbbk). We consider the canonical basis of AA, ={Eij:i,j=1,,n}\mathcal{B}=\bigl{\{}E_{ij}:\;i,j=1,\dots,n\bigr{\}}. Then 𝒜={Δkj:k,l=1,,n}\mathcal{A}=\bigl{\{}\Delta_{kj}:\;k,l=1,\dots,n\bigr{\}} is a basis of A\mathcal{E}_{A} where

Δkl(Eij)=EikElj,k,l=1,,n.\Delta_{kl}(E_{ij})=E_{ik}\otimes E_{lj},\;\forall k,l=1,\dots,n.

For this family of coproducts their corresponding handle elements are

ωkl=mΔkl(I)=mΔkl(iEii)=imΔkl(Eii)=im(EikEli)={0 if lk,I if k=l.\omega_{kl}=m\circ\Delta_{kl}(I)=m\circ\Delta_{kl}\Bigl{(}\sum_{i}E_{ii}\Bigr{)}=\sum_{i}m\circ\Delta_{kl}\bigl{(}E_{ii}\bigr{)}=\sum_{i}m\bigl{(}E_{ik}\otimes E_{li}\bigr{)}=\begin{cases}0&\text{ if }l\neq k,\\ I&\text{ if }k=l.\end{cases}

Then ωkk=I\omega_{kk}=I, k=1,,n\forall k=1,\dots,n and Δkk\Delta_{kk} is a normalized nearly Frobenius coproduct (see Definition 14 of [6]).
The coproduct Δ0=i=1nΔii\displaystyle{\Delta_{0}=\sum_{i=1}^{n}\Delta_{ii}} corresponds to the Frobenius structure of Mn×n(𝕜)M_{n\times n}(\Bbbk) and

ω0=mΔ0(I)=i,k=1nEikEki=i,k=1nEii=k=1nI=nI.\omega_{0}=m\circ\Delta_{0}(I)=\sum_{i,k=1}^{n}E_{ik}E_{ki}=\sum_{i,k=1}^{n}E_{ii}=\sum_{k=1}^{n}I=nI.

In particular, by Theorem 3.6, we conclude that AA is a separable algebra.

Example 3.2.

Let GG be a cyclic finite group of order nn and the group algebra 𝕜G\Bbbk G, with the natural basis {gi:i=1,,n}\bigl{\{}g^{i}:\;i=1,\dots,n\bigr{\}}. A basis of the Frobenius space is

={Δk:𝕜G𝕜G𝕜G:k1,,n},\mathcal{B}=\bigl{\{}\Delta_{k}:\Bbbk G\rightarrow\Bbbk G\otimes\Bbbk G:k\in{1,\dots,n}\bigr{\}},

where Δ1(1)=i=1ngign+1i\displaystyle{\Delta_{1}(1)=\sum_{i=1}^{n}g^{i}\otimes g^{n+1-i}} and Δk(1)=i=1k1gigki+i=kngign+ki\displaystyle{\Delta_{k}\bigl{(}1\bigr{)}=\sum_{i=1}^{k-1}g^{i}\otimes g^{k-i}+\sum_{i=k}^{n}g^{i}\otimes g^{n+k-i}} for k=2,,nk=2,\dots,n. In this case we have that

ωk=ngk,k=1,,n.\omega_{k}=ng^{k},\;\forall k=1,\dots,n.

Again, we can conclude that this algebra is separable.

Remark 3.7.

The converse of Theorem 3.6 is not fulfilled, consider Δkl\Delta_{kl} with klk\neq l in Example 3.1 as a counterexample. Even if we asked ω0\omega\neq 0, the converse of Theorem 3.6 would still fail, just look at the following example to check it.

Example 3.3.

Let RR be a commutative ring and A=RRA=R\oplus R be the ring direct sum of two copies of RR. Let e1=(1,0)e_{1}=(1,0) and e2=(0,1)e_{2}=(0,1) be the orthogonal idempotents in AA. We prove that AA is separable over RR. For this we consider the element e=e1e1+e2e2AAe=e_{1}\otimes e_{1}+e_{2}\otimes e_{2}\in A\otimes A, it is a separability element for AA:

m(e)=m(e1e1+e2e2)=m(e1e1)+m(e2e2)=e1+e2=1A.m(e)=m\bigl{(}e_{1}\otimes e_{1}+e_{2}\otimes e_{2}\bigr{)}=m\bigl{(}e_{1}\otimes e_{1}\bigr{)}+m\bigl{(}e_{2}\otimes e_{2}\bigr{)}=e_{1}+e_{2}=1_{A}.

As RR is a commutative ring and the tensor product is over RR it is clear that re=erre=er, for all rRr\in R. Therefore AA is separable over RR.
On the other hand we study the nearly Frobenius structures of AA. It is not difficult to prove that a general nearly Frobenius coproduct has the expression

Δ(1)=ae1e1+be2e2,a,bR.\Delta(1)=ae_{1}\otimes e_{1}+be_{2}\otimes e_{2},\;\forall a,b\in R.

In particular the Frobenius structure is Δ0(1)=e1e1+e2e2\Delta_{0}(1)=e_{1}\otimes e_{1}+e_{2}\otimes e_{2}, and ε(e1)=ε(e2)=1\varepsilon(e_{1})=\varepsilon(e_{2})=1.
The handle element associated to the Frobenius structure is

ωΔ0=m(Δ0(1))=e1+e2=1A.\omega_{\Delta_{0}}=m\bigl{(}\Delta_{0}(1)\bigr{)}=e_{1}+e_{2}=1_{A}.

The handle element ωΔ0\omega_{\Delta_{0}} is a unit of AA. But, if we fix the nearly Frobenius structure Δ1(1)=e1e1\Delta_{1}(1)=e_{1}\otimes e_{1} on AA, then the handle element associated ωΔ1=e1\omega_{\Delta_{1}}=e_{1} is not a unit of AA.

If we adapt Theorem 3.6 we can get an equivalence as follows:

Theorem 3.8.

An algebra AA admits a nearly Frobenius structure where the handle element is a unit of AA if and only if AA is separable.

Proof.

()(\Rightarrow) It is Theorem 3.6.
()(\Leftarrow) Applying Theorem 22 of [7] we have that AA admits a normalized nearly Frobenius coproduct Δ\Delta, that is mΔ=IdAm\circ\Delta=\operatorname{Id}_{A}. Then ωΔ=m(Δ(1))=1A\omega_{\Delta}=m\bigl{(}\Delta(1)\bigr{)}=1\in A. Therefore ωΔ\omega_{\Delta} is a unit of AA. ∎

Example 3.4.

Let AA be the truncated polynomial algebra in one variable 𝕜[x]/xn+1\Bbbk[x]/x^{n+1}. In this case, a basis of the Frobenius space is ={Δ0,,Δn}\mathcal{B}=\{\Delta_{0},\dots,\Delta_{n}\}, where

Δk(1)=i+j=n+kxixj,k{0,1,,n}.\Delta_{k}(1)=\sum_{i+j=n+k}x^{i}\otimes x^{j},\;\forall k\in\{0,1,\dots,n\}.

We can determine the handle elements associated to the nearly Frobenius coproducts described.

ωk=i+j=n+kxi+j=0,k=1,,n\omega_{k}=\sum_{i+j=n+k}x^{i+j}=0,\;\forall k=1,\dots,n

and

ω0=i+j=nxi+j=(n+1)xn.\omega_{0}=\sum_{i+j=n}x^{i+j}=(n+1)x^{n}.

Note that the only coproduct that admits a non zero handle element is the Frobenius coproduct. Using Theorem 3.8 we can conclude that if n>0n>0 AA is not separable.

3.2 Handle element of algebras over fields

We will now study some properties of the handle element in the particular case of 𝕜\Bbbk-algebras, where 𝕜\Bbbk is a field.

Proposition 3.9.

Let be AA a nearly Frobenius algebra of finite dimension over 𝕜\Bbbk and ω\omega its corresponding handle element. If ω\omega is not a unit, then there exist kk\in\mathbb{N} such that ωksoc(A)\omega^{k}\in\operatorname{soc}(A).

Proof.

If Δ\Delta is a nearly Frobenius coproduct in AA it is an AA-bimodule morphism. Then, if AA has finite dimension Δ(J)JJ\Delta(J)\subset J\otimes J, where JJ is the Jacobson radical of AA.
If we define ψ=mΔ|J:JJ2\psi=m\circ\Delta|_{J}:J\rightarrow J^{2} we obtain that ψ(x)=xω\psi(x)=x\omega. Composing the function ψ\psi with itself mm times we obtain that ψm:JJ2m\psi^{m}:J\rightarrow J^{2m} is ψm(x)=xωm\psi^{m}(x)=x\omega^{m}.
As AA is a finite dimensional algebra, there exists kk such that J2k={0}J^{2k}=\{0\}, in particular the map ψk=0\psi^{k}=0. Then ψk(x)=xωk=0\psi^{k}(x)=x\omega^{k}=0xJ\forall x\in J.
As ω\omega is not a unit we have that

Jωk={0}ωksoc(A).J\omega^{k}=\{0\}\Rightarrow\omega^{k}\in\operatorname{soc}(A).

Proposition 3.10.

Let be AA a connected finite dimensional nearly Frobenius path algebra which is not a single vertex over 𝕜\Bbbk (𝕜\Bbbk algebraically closed) and ω\omega its corresponding handle element. Then ωJ\omega\in J, where JJ is the Jacobson radical of AA.

Proof.

Consider the path of length zero associated to the vertex ii, eie_{i}. Then Δ(ei)=jajbj\Delta(e_{i})=\sum_{j}a_{j}\otimes b_{j} with s(bj)=t(aj)=eis(b_{j})=t(a_{j})=e_{i}. Suppose eieie_{i}\otimes e_{i} is a summand in Δ(ei)\Delta(e_{i}). Since AA is connected there is at least one arrow that starts or ends in ii. Let us suppose that α\alpha starts on ii and ends in kk (the other case is analogous). Then Δ(α)=Δ(ei)α=αΔ(ek)\Delta(\alpha)=\Delta(e_{i})\alpha=\alpha\Delta(e_{k}) so eiαe_{i}\otimes\alpha is a summand in Δ(α)\Delta(\alpha) but cannot be of the form αvu\alpha v\otimes u with u,vu,v some paths and this is absurd. Then the elements of the form eieie_{i}\otimes e_{i} do not appear in Δ(ei)\Delta(e_{i}) and so do not appear in Δ(1)=iΔ(ei)\Delta(1)=\sum_{i}\Delta(e_{i}) which implies that every summand of ω\omega is a path of positive lenght and then ωJ\omega\in J. ∎

Remark 3.11.

In the hypothesis of the above proposition we conclude that ω\omega is nilpotent.

3.3 Families of nearly Frobenius algebras with zero handle element

In this section we describe some families of algebras that admit nontrivial nearly Frobenius structures but whose corresponding handle elements are null.

Example 3.5.

A family of algebras where we can find its handle element was presented in 5.3 of [4].
If A=𝕜CIC\displaystyle{A=\frac{\Bbbk C}{I_{C}}}, where C=C(n1,n2,,nm)C=C\bigl{(}n_{1},n_{2},\dots,n_{m}\bigr{)}, with m>1m>1 and IC=αnmmα11,αniα1i+1,i=1,,m1I_{C}=\bigl{\langle}\alpha_{n_{m}}^{m}\alpha_{1}^{1},\alpha_{n_{i}}\alpha_{1}^{i+1},i=1,\dots,m-1\bigr{\rangle},

[Uncaptioned image]

by the description of nearly coproduct in Theorem 5 of [4] we conclude that ω=0\omega=0.

Proposition 3.12.

Let A=𝕜QIA=\frac{\Bbbk Q}{I} be an finite dimensional algebra. If every cycle of QQ is 0 in AA, then every handle element of AA is 0.

Proof.

Suppose there is a coproduct Δ\Delta on AA, such that ωΔ=mΔ(1)0\omega_{\Delta}=m\circ\Delta(1)\not=0. Then, there is a vertex pQ0p\in Q_{0} such that mΔ(ep)0m\circ\Delta(e_{p})\not=0. Note that for every nearly Frobenius coproduct Δ\Delta and vertex pQ0p\in Q_{0}, Δ(ep)=aiαiβi\Delta(e_{p})=\sum a_{i}\alpha_{i}\otimes\beta_{i} where t(βi)=s(αi)=p,i\mathrm{t}(\beta_{i})=\mathrm{s}(\alpha_{i})=p,\forall i. So we have an index ii such that αiβi0\alpha_{i}\beta_{i}\not=0 in AA, in particular s(βi)=t(αi)\mathrm{s}(\beta_{i})=\mathrm{t}(\alpha_{i}). Then αiβi\alpha_{i}\beta_{i} is a non zero cycle. ∎

The following results are obtained from the previous proposition.

Corollary 3.13.

If AA a toupie algebra over a field 𝕜\Bbbk then every handle element is 0.

Proof.

Since toupie algebras are acyclic the result follows immediately from Proposition 3.12. ∎

Corollary 3.14.

If AA is a radical square zero algebra different to the loop then every handle element is 0.

Proof.

If every cycle has length bigger to or equal to 22 the result follows from Proposition 3.12. By Remark 2 of [5] if the indegree and the outdegree of a vertex pp are both bigger to or equal to 22 we have that mΔ(ep)=0m\circ\Delta(e_{p})=0, so we can suppose that for a vertex pp there is a loop α\alpha at pp and arrows βi\beta_{i} that starts at pp. Again by Remark 2 of [5] Δ(ep)=biβiα\Delta(e_{p})=\sum b_{i}\beta_{i}\otimes\alpha. Thus mΔ(ep)=biβiα=0m\circ\Delta(e_{p})=\sum b_{i}\beta_{i}\alpha=0 since AA is a radical square zero algebra. ∎

3.4 Handle element and socle

In the context of commutative Frobenius algebras, Theorem 3.15 is valid, which allows to determine the socle of the algebra from its handle element. In this section we answer the question, is it possible to generalize this result to non commutative algebras or nearly Frobenius algebras?

Theorem 3.15 (Proposition 3.3 of [3]).

In a commutative Frobenius algebra AA with handle element ω\omega, the ideal ωA\omega A is the socle of AA.

Remark 3.16.

The next example shows that the Theorem 3.15 is not valid in the case of non commutative Frobenius algebras.

Example 3.6.

Let be A=𝕜Q/IA=\Bbbk Q/I, where Q=1α2βQ=\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 7.74307pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-7.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bullet_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 14.14369pt\raise 10.50694pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{\alpha}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 31.74515pt\raise 3.78268pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}{\hbox{\kern 31.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bullet_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 14.47588pt\raise-12.1111pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\beta}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 7.74112pt\raise-3.78268pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces}}}}\ignorespaces and I=αβ,βαI=\langle\alpha\beta,\beta\alpha\rangle. In this example the nearly Frobenius structures that it admits are the following

Δ(e1)=ae1β+bαe1+cαβΔ(e2)=be2α+aβe2+dβα\begin{array}[]{rcl}\Delta(e_{1})&=&ae_{1}\otimes\beta+b\alpha\otimes e_{1}+c\alpha\otimes\beta\\ \Delta(e_{2})&=&be_{2}\otimes\alpha+a\beta\otimes e_{2}+d\beta\otimes\alpha\end{array}

where a,b,c,d𝕜.a,b,c,d\in\Bbbk.

A Frobenius coproduct is

Δ(1)=e1β+αe1+e2α+βe2αββα,\Delta(1)=e_{1}\otimes\beta+\alpha\otimes e_{1}+e_{2}\otimes\alpha+\beta\otimes e_{2}-\alpha\otimes\beta-\beta\otimes\alpha,

where the trace is defined as

ε:A𝕜,ε(e1)=ε(e2)=ε(α)=ε(β)=1.\varepsilon:A\rightarrow\Bbbk,\;\varepsilon(e_{1})=\varepsilon(e_{2})=\varepsilon(\alpha)=\varepsilon(\beta)=1.

In this case, soc(A)=αAβA\operatorname{soc}(A)=\alpha A\oplus\beta A, and ω=mΔ(1)=0.\omega=m\circ\Delta(1)=0.
Then, the handle element does not generate the socle.

Given a finite dimensional algebra AA and its Jacobson radical JJ, for a left ideal MAM\subset A, JM=0JM=0 if and only if Msoc(A)M\subset\operatorname{soc}(A). As a consequence, if soc(A)=S1S2Sm=A\operatorname{soc}(A)=S_{1}\subset S_{2}\subset\cdots\subset S_{m}=A is a socle series, then JSiSi1JS_{i}\subset S_{i-1}.
Recall that for a Frobenius algebra AA we have that soc(AA)=soc(AA)\operatorname{soc}(A_{A})=\operatorname{soc}(_{A}A).

In the context of non commutative Frobenius algebras we have the following result, which is a weakening of Theorem 3.15.

Theorem 3.17.

In a non-commutative Frobenius algebra AA with handle element ω\omega, the ideal ωA\omega A is a submodule of the socle of AA.

Proof.

Let soc(A)=S1S2Sm=A\operatorname{soc}(A)=S_{1}\subset S_{2}\subset\cdots\subset S_{m}=A be a socle series for AA as a left AA-module. Choose a basis for S1S_{1}. Now, starting with i=1i=1, iteratively take the basis for SiS_{i} and extend it to a basis for Si+1S_{i+1}. Denote the elements of the basis for Sm=AS_{m}=A by e1,,ene_{1},\cdots,e_{n}. Suppose that eiSkSk1e_{i}\in S_{k}\setminus S_{k-1} and xJx\in J, then xeiSk1xe_{i}\in S_{k-1} and xei=jinαjej\displaystyle{xe_{i}=\sum_{j\neq i}^{n}\alpha_{j}e_{j}}.
Remember that AA is a Frobenius algebra if it possesses a left AA-module isomorphism λl:AA\lambda_{l}:A\rightarrow A^{*} with its dual vector space, where AA is viewed as the left module over itself, and AA^{*} is made a left AA-module by the action (af)(b)=f(ba)(a\cdot f)(b)=f(ba), for all aAa\in A and fAf\in A^{*}. In this context we can construct the counit ε:A𝕜\varepsilon:A\rightarrow\Bbbk of AA as ε(a)=λl(1)(a)\varepsilon(a)=\lambda_{l}(1)(a), for all aAa\in A. In particular, we have e1#,,en#e_{1}^{\#},\dots,e_{n}^{\#} the corresponding dual basis elements of AA, which verify ε(eiej#)=δij\varepsilon\bigl{(}e_{i}e_{j}^{\#}\bigl{)}=\delta_{ij} for all i,j=1,,ni,j=1,\dots,n. Then

ε(xeiei#)=ε(jiαjejei#)=jiαjε(ejei#)=0.\varepsilon\bigl{(}xe_{i}e_{i}^{\#}\bigr{)}=\varepsilon\left(\sum_{j\neq i}\alpha_{j}e_{j}e_{i}^{\#}\right)=\sum_{j\neq i}\alpha_{j}\varepsilon\bigl{(}e_{j}e_{i}^{\#}\bigr{)}=0.

Therefore Jeiei#Ker(ε)Je_{i}e_{i}^{\#}\subset\operatorname{Ker}(\varepsilon). But, Ker(ε)\operatorname{Ker}(\varepsilon) can contain no nontrivial left ideals, then

Jeiei#={0}.Je_{i}e_{i}^{\#}=\{0\}.

Thus eiei#soc(A)e_{i}e_{i}^{\#}\in\operatorname{soc}(A) for each ii, so ω=i=1neiei#soc(A).\displaystyle{\omega=\sum_{i=1}^{n}e_{i}e_{i}^{\#}\in\operatorname{soc}(A)}.

The following examples show that Theorem 3.17 does not hold in the context of nearly Frobenius algebras.

The first two examples illustrate that there are nearly Frobenius (non Frobenius) algebras where the ideal generated by the handle element is strict contained in the socle of the algebra.

Example 3.7.

We consider a small modification of the Example 3.6, let be A=𝕜Q/IA=\Bbbk Q/I, where Q=1α2βQ=\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 7.74307pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-7.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bullet_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 14.14369pt\raise 10.50694pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{\alpha}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 31.74515pt\raise 3.78268pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}{\hbox{\kern 31.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bullet_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 14.47588pt\raise-12.1111pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\beta}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 7.74112pt\raise-3.78268pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces}}}}\ignorespaces and I=βαI=\langle\beta\alpha\rangle. In this case we can prove that the nearly Frobenius structures are the following

Δ(e1)=a1(e1αβ+αβe1)+a2αβ+a3ααβ+a4αββ+a5αβαβΔ(e2)=a1βα\begin{array}[]{rcl}\Delta(e_{1})&=&a_{1}(e_{1}\otimes\alpha\beta+\alpha\beta\otimes e_{1})+a_{2}\alpha\otimes\beta+a_{3}\alpha\otimes\alpha\beta+a_{4}\alpha\beta\otimes\beta+a_{5}\alpha\beta\otimes\alpha\beta\\ \Delta(e_{2})&=&a_{1}\beta\otimes\alpha\end{array}

for a1,,a5𝕜a_{1},\dots,a_{5}\in\Bbbk.
A first observation we can make is that this algebra does not support Frobenius algebra structure, because if we suppose that there exist ε:A𝕜\varepsilon:A\rightarrow\Bbbk trace of Δ\Delta, then

(ε1)Δ(e2)=a1ε(β)α=e2,(\varepsilon\otimes 1)\Delta(e_{2})=a_{1}\varepsilon(\beta)\alpha=e_{2},

and this is a contradiction.
On the other hand, we can determine the handle element for (A,Δ)(A,\Delta),

ω=mΔ(1)=mΔ(e1+e2)=2a1αβ+a2αβ=(2a1+a2)αβ,\omega=m\circ\Delta(1)=m\circ\Delta(e_{1}+e_{2})=2a_{1}\alpha\beta+a_{2}\alpha\beta=(2a_{1}+a_{2})\alpha\beta,

and the socle of AA is

soc(A)=αβAβA,\operatorname{soc}(A)=\alpha\beta A\oplus\beta A,

then ωAsoc(A).\omega A\subsetneq\operatorname{soc}(A).

Example 3.8 (Local rings which are not Gorenstein, see [16]).

In A=𝕜[x,y]/(x2,xy2,y3)A=\Bbbk[x,y]/\bigl{(}x^{2},xy^{2},y^{3}\bigr{)}, the socle is (xy,y2)\bigl{(}xy,y^{2}\bigr{)}. This ring does not support Frobenius algebra structure but it does admit nearly Frobenius algebra structure. A general nearly Frobenius coproduct is

Δ(1)=a1(xy2+yxy+y2x+xyy)+a2(xxy+xyx)+a3y2y2+a4y2xy\Delta(1)=a_{1}\bigl{(}x\otimes y^{2}+y\otimes xy+y^{2}\otimes x+xy\otimes y\bigr{)}+a_{2}\bigl{(}x\otimes xy+xy\otimes x\bigr{)}+a_{3}y^{2}\otimes y^{2}+a_{4}y^{2}\otimes xy
+a5xyy2+a6xyxy,+a_{5}xy\otimes y^{2}+a_{6}xy\otimes xy,

where ai𝕜,i=1,,6.a_{i}\in\Bbbk,\forall i=1,\dots,6.
Note that there are no scalars that make any coproduct admit a trace, therefore none can be a Frobenius coproduct.
On the other hand, the handle element is zero for all coproduct

ωΔ=0,ai𝕜,i=1,,6ωΔA={0}Soc(A)=(xy,y2).\omega_{\Delta}=0,\;\forall a_{i}\in\Bbbk,i=1,\dots,6\Rightarrow\omega_{\Delta}A=\{0\}\subsetneq\operatorname{Soc}(A)=\bigl{(}xy,y^{2}\bigr{)}.

In the ring A=𝕜[x,y,z]/(x2,y2,z2,xy)A=\Bbbk[x,y,z]/\bigl{(}x^{2},y^{2},z^{2},xy\bigr{)}, the socle is (xz,yz)(xz,yz). As before, this ring does not support Frobenius algebra structure but it does admit nearly Frobenius algebra structure, where the general Frobenius coproduct is generated, as a linear combination, by

Δ1(1)=xxz+xzx,Δ2(1)=xyz+xzy,Δ3(1)=yxz+yzx,\Delta_{1}(1)=x\otimes xz+xz\otimes x,\;\Delta_{2}(1)=x\otimes yz+xz\otimes y,\;\Delta_{3}(1)=y\otimes xz+yz\otimes x,
Δ4(1)=yyz+yzy,Δ5(1)=xzxz,Δ6(1)=xzyz,\Delta_{4}(1)=y\otimes yz+yz\otimes y,\;\Delta_{5}(1)=xz\otimes xz,\;\Delta_{6}(1)=xz\otimes yz,
Δ7(1)=yzxz,Δ8(1)=yzyz.\Delta_{7}(1)=yz\otimes xz,\;\Delta_{8}(1)=yz\otimes yz.

In particular FrobdimA=8\operatorname{Frobdim}A=8, and none of the nearly Frobenius coproducts can be completed to a Frobenius coproduct. Also, all the handle elements are zero (ωΔ=0\omega_{\Delta}=0). Then {0}=ωΔASoc(A)=(xz,yz).\{0\}=\omega_{\Delta}A\subsetneq\operatorname{Soc}(A)=\bigl{(}xz,yz\bigr{)}.

The following example shows that Theorem 3.17 does not generalize to nearly Frobenius algebras.

Example 3.9.

Let be A=𝕜Q/IA=\Bbbk Q/I, where QQ is

2\textstyle{\bullet_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β\scriptstyle{\beta}1\textstyle{\bullet_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}3\textstyle{\bullet_{3}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γ\scriptstyle{\gamma}

and I=αβγI=\langle\alpha\beta\gamma\rangle. In this example we can describe all the nearly Frobenius structures and as consequence determine the handle element

ω=a(βγα+γαβ),wherea𝕜.\omega=a\bigl{(}\beta\gamma\alpha+\gamma\alpha\beta\bigr{)},\quad\mbox{where}\;a\in\Bbbk.

On the other hand, the socle of AA is

soc(AA)=γαβAβγαβA.\operatorname{soc}(A_{A})=\gamma\alpha\beta A\oplus\beta\gamma\alpha\beta A.

Then ωAsoc(AA)\omega A\not\subset\operatorname{soc}(A_{A}).

4 Casimir and Schur elements

Given an algebra AA, there is a natural isomorphism ψ:AAEnd𝕜(A)\psi:A\otimes A^{*}\to\operatorname{End}_{\Bbbk}(A) such that ψ(xβ)(y)=β(x)y\psi(x\otimes\beta)(y)=\beta(x)y. If AA is Frobenius then AAA\cong A^{*} and we can consider an isomorphism ψ~:AAEnd𝕜(A)\tilde{\psi}:A\otimes A\to\operatorname{End}_{\Bbbk}(A). The Casimir element (see [15]) for (A,λ)(A,\lambda) be a symmetric algebra is defined as the element CAAAC_{A}\in A\otimes A that corresponds to IdAEnd𝕜(A)\operatorname{Id_{A}}\in\operatorname{End}_{\Bbbk}(A) under the isomorphism End𝕜(A)AA\operatorname{End}_{\Bbbk}(A)\cong A\otimes A.
If {ei}\bigl{\{}e_{i}\bigr{\}} and {ei}\bigl{\{}e^{i}\bigr{\}} are dual basis with respect to λ\lambda, the Casimir element is:

CA=eiei=eiei.C_{A}=\sum e_{i}\otimes e^{i}=\sum e^{i}\otimes e_{i}.

This element satisfies that

aeiei=eieia,for allaA.\sum ae_{i}\otimes e^{i}=\sum e_{i}\otimes e^{i}a,\quad\mbox{for all}\;a\in A.

The central Casimir element of AA is defined as

ZA=eiei.Z_{A}=\sum e^{i}e_{i}.
Remark 4.1.

Note that the Casimir element of AA coincide with Δ(1)\Delta(1), if Δ\Delta is the coproduct of this symmetric algebra (see [2]), and the central Casimir element coincide with the handle element.

In the context of Frobenius algebras we can define the Casimir element in a similar way.
Given a basis {e1,,en}\{e_{1},\dots,e_{n}\} of the Frobenius algebra AA with isomorphism λl:AA\lambda_{l}:A\to A^{*}, let {e1,,en}\bigl{\{}e^{1},\dots,e^{n}\bigr{\}} be the dual basis of AA relative to λl\lambda_{l}, that means the elements satisfy λl(ej)(ei)=δij\lambda_{l}\bigl{(}e^{j}\bigr{)}(e_{i})=\delta_{ij}, for all i,j=1,,ni,j=1,\dots,n. In particular, if {e1,,en}\{e_{1}^{*},\dots,e_{n}^{*}\} denotes the basis of AA^{*} satisfying ei(ej)=δije_{i}^{*}(e_{j})=\delta_{ij}, using λl:AA\lambda_{l}:A\rightarrow A^{*}, we conclude that ej=λl1(ej)e^{j}=\lambda_{l}^{-1}(e_{j}^{*}) for all j=1,,nj=1,\dots,n.
In this context, if we use the isomorphism λl:AA\lambda_{l}:A\to A^{*} as left AA-modules, the Casimir element, as before, corresponds to IdAEnd𝕜(A)\operatorname{Id_{A}}\in\operatorname{End}_{\Bbbk}(A) under the isomorphism End𝕜(A)AA\operatorname{End}_{\Bbbk}(A)\cong A\otimes A. But, in this case we only have the following expression since AA is not necessarily symmetric:

CA=i=1neiei.C_{A}=\sum_{i=1}^{n}e^{i}\otimes e_{i}.

Let us recall the Proposition 2 proved in [7].

Proposition 4.2.

The coproduct Δ:AAA\Delta:A\rightarrow A\otimes A of the Frobenius algebra AA satisfies

  1. (i)

    Δ(1A)=i=1neiei\displaystyle{\Delta(1_{A})=\sum_{i=1}^{n}e^{i}\otimes e_{i}}.

  2. (ii)

    Δ(x)=i=1nxeiei=i=1neieix\displaystyle{\Delta(x)=\sum_{i=1}^{n}xe^{i}\otimes e_{i}=\sum_{i=1}^{n}e^{i}\otimes e_{i}x}, for all xAx\in A.

Then in this case, as before, the Casimir element coincides with Δ(1)\Delta(1).

Definition 4.3.

If AA is a nearly Frobenius algebra, we define the Casimir element of AA as

CA=Δ(1).C_{A}=\Delta(1).

This allows us to view the properties of the Casimir element or related constructions in terms of the coproduct of the (nearly) Frobenius algebra.

Let (A,εA)(A,\varepsilon_{A}) and (B,εB)(B,\varepsilon_{B}) be Frobenius algebras and ϕ:AB\phi:A\twoheadrightarrow B an epimorphism of algebras. Note that εBϕ\varepsilon_{B}\circ\phi is in AA^{\ast}, so there exists an element u0BAu_{0}^{B}\in A such that u0BεA=εBϕu_{0}^{B}\cdot\varepsilon_{A}=\varepsilon_{B}\circ\phi. Consider sϕ=ϕ(u0B)Bs_{\phi}=\phi\bigl{(}u_{0}^{B}\bigr{)}\in B.

Definition 4.4.

The element sϕs_{\phi} is called the Schur element of the epimorphism ϕ:AB\phi:A\twoheadrightarrow B.

Remark 4.5.

The Schur element can be describe as

sϕ=i=1nεB(ϕ(ei))ϕ(ei):s_{\phi}=\sum_{i=1}^{n}\varepsilon_{B}\bigl{(}\phi\bigl{(}e^{i}\bigr{)}\bigr{)}\phi\bigl{(}e_{i}\bigr{)}:

Let ΔA(1)=i=1neiei\displaystyle{\Delta_{A}(1)=\sum_{i=1}^{n}e^{i}\otimes e_{i}} with ΔA(a)=i=1naeiei=i=1neieia\displaystyle{\Delta_{A}(a)=\sum_{i=1}^{n}ae^{i}\otimes e_{i}=\sum_{i=1}^{n}e^{i}\otimes e_{i}a}\; for all aAa\in A.
As u0BAu_{0}^{B}\in A we have that

u0B=(εAIdA)ΔA(u0B)=i=1nεA(u0Bei)ei=i=1n(u0BεA)(ei)ei=i=1n(εBϕ)(ei)eiu_{0}^{B}=\bigl{(}\varepsilon_{A}\otimes\operatorname{Id}_{A}\bigr{)}\Delta_{A}\bigl{(}u_{0}^{B}\bigr{)}=\sum_{i=1}^{n}\varepsilon_{A}\bigl{(}u_{0}^{B}e^{i}\bigr{)}e_{i}=\sum_{i=1}^{n}\bigl{(}u_{0}^{B}\cdot\varepsilon_{A}\bigr{)}\bigl{(}e^{i}\bigr{)}e_{i}=\sum_{i=1}^{n}\bigl{(}\varepsilon_{B}\circ\phi\bigr{)}\bigl{(}e^{i}\bigr{)}e_{i}

Then

sϕ=ϕ(u0B)=i=1n(εBϕ)(ei)ϕ(ei)=i=1nεB(ϕ(ei))ϕ(ei).s_{\phi}=\phi\bigl{(}u_{0}^{B}\bigr{)}=\sum_{i=1}^{n}\bigl{(}\varepsilon_{B}\circ\phi\bigr{)}\bigl{(}e^{i}\bigr{)}\phi\bigl{(}e_{i}\bigr{)}=\sum_{i=1}^{n}\varepsilon_{B}\bigl{(}\phi\bigl{(}e^{i}\bigr{)}\bigr{)}\phi\bigl{(}e_{i}\bigr{)}.

The previous remark allows us to generalize the definition of Schur element in the context of nearly Frobenius algebras. We will denote Δ(1)=i=1neiei\displaystyle{\Delta(1)=\sum_{i=1}^{n}e^{i}\otimes e_{i}} in the nearly Frobenius case although {e1,e2,en}\bigl{\{}e_{1},e_{2},\cdots e_{n}\bigr{\}} and {e1,e2,en}\bigl{\{}e^{1},e^{2},\cdots e^{n}\bigr{\}} are not dual basis.

Definition 4.6.

Let AA be a nearly Frobenius algebra and BB a Frobenius algebra with ϕ:AB\phi:A\twoheadrightarrow B an epimorphism of algebras. We define the Schur element of ϕ\phi, sϕBs_{\phi}\in B, as the element

sϕ=i=1nεB(ϕ(ei))ϕ(ei).s_{\phi}=\sum_{i=1}^{n}\varepsilon_{B}\bigl{(}\phi\bigl{(}e^{i}\bigr{)}\bigr{)}\phi\bigl{(}e_{i}\bigr{)}.
Proposition 4.7.

Let AA be a nearly Frobenius algebra and BB a Frobenius algebra with ϕ:AB\phi:A\twoheadrightarrow B an epimorphism of algebras. Then:

  1. 1.

    (ϕϕ)(CA)=sϕCB(\phi\otimes\phi)(C_{A})=s_{\phi}C_{B},

  2. 2.

    ϕ(ZA)=sϕZB\phi(Z_{A})=s_{\phi}Z_{B},

with CA=i=1neieiC_{A}=\sum_{i=1}^{n}e^{i}\otimes e_{i} the Casimir element of AA and ZA=i=1neieiZ_{A}=\sum_{i=1}^{n}e^{i}e_{i} the central Casimir element of AA.

Proof.

Consider the map ψ:BRBEndR(B)\psi:B\otimes_{R}B\to End_{R}(B) as in the definition of the Casimir element: given xyBRBx\otimes y\in B\otimes_{R}B, ψ(xy)\psi(x\otimes y) is the map in EndR(B)End_{R}(B) that sends bBb\in B to εB(bx)y\varepsilon_{B}(bx)y. If CB=fifiC_{B}=\sum f^{i}\otimes f_{i} then

ψ(CB)(b)=εB(bfi)fi=b\psi(C_{B})(b)=\sum\varepsilon_{B}(bf^{i})f_{i}=b

and ψ(CB)=IdB.\psi(C_{B})=Id_{B}.
Let us now prove that ψ(ϕϕ)(CA)=sϕIdB\psi(\phi\otimes\phi)(C_{A})=s_{\phi}Id_{B}:

ψ(ϕϕ)(CA)(b)\displaystyle\psi(\phi\otimes\phi)(C_{A})(b) =ψ(ϕ(ei)ϕ(ei))(b)\displaystyle=\psi\Bigl{(}\sum\phi(e^{i})\otimes\phi(e_{i})\Bigr{)}(b)
=εB(bϕ(ei))ϕ(ei)\displaystyle=\sum\varepsilon_{B}(b\phi(e^{i}))\phi(e_{i})
=εB(ϕ(a)ϕ(ei))ϕ(ei)\displaystyle=\sum\varepsilon_{B}(\phi(a)\phi(e^{i}))\phi(e_{i})
=εB(ϕ(aei))ϕ(ei)\displaystyle=\sum\varepsilon_{B}(\phi(ae^{i}))\phi(e_{i})

Since AA is a nearly Frobenius algebra

Δ(a)=aeiei=eieia\Delta(a)=\sum ae^{i}\otimes e_{i}=\sum e^{i}\otimes e_{i}a

and

(ϕϕ)(aeiei)=ϕ(aei)ϕ(ei)=ϕ(ei)ϕ(eia).(\phi\otimes\phi)\Bigl{(}\sum ae^{i}\otimes e_{i}\Bigr{)}=\sum\phi(ae^{i})\otimes\phi(e_{i})=\sum\phi(e^{i})\otimes\phi(e_{i}a).

Then applying εB1\varepsilon_{B}\otimes 1 we obtain

εB(ϕ(ei))ϕ(eia)=εB(ϕ(aei))ϕ(ei)\sum\varepsilon_{B}(\phi(e^{i}))\phi(e_{i}a)=\sum\varepsilon_{B}(\phi(ae^{i}))\phi(e_{i})

and

ψ(ϕϕ)(CA)(b)=sϕb.\psi(\phi\otimes\phi)(C_{A})(b)=s_{\phi}b.

To prove (2) observe that

ϕ(ZA)=ϕ(eiei)=ϕ(ei)ϕ(ei)=m(ϕϕ)(CA)=m(sϕCB)=sϕZB.\phi(Z_{A})=\phi\Bigl{(}\sum e^{i}e_{i}\Bigr{)}=\sum\phi(e^{i})\phi(e_{i})=m\circ(\phi\otimes\phi)(C_{A})=m\circ(s_{\phi}C_{B})=s_{\phi}Z_{B}.

As the central Casimir element of AA is the handle element of AA the previous proposition can be written as follows.

Proposition 4.8.

Let AA be a nearly Frobenius algebra and BB a Frobenius algebra with ϕ:AB\phi:A\twoheadrightarrow B an epimorphism of algebras. Then:

  1. 1.

    (ϕϕ)(ΔA(1A))=sϕΔB(1B)(\phi\otimes\phi)(\Delta_{A}(1_{A}))=s_{\phi}\Delta_{B}(1_{B}).

  2. 2.

    ϕ(ωA)=sϕωB\phi(\omega_{A})=s_{\phi}\omega_{B}.

Lemma 4.9.

Let AA be a nearly Frobenius algebra and BB a Frobenius algebra with ϕ:AB\phi:A\twoheadrightarrow B an epimorphism of algebras. Then the Schur element sϕs_{\phi} is central in BB.

Proof.

Using that ϕ\phi is an epimorphism it is enough to prove that sϕϕ(a)=ϕ(a)sϕs_{\phi}\phi(a)=\phi(a)s_{\phi} for all aAa\in A.

sϕϕ(a)=εB(ϕ(ei))ϕ(ei)ϕ(a)=εB(ϕ(ei))ϕ(eia)as ϕ is a morphism of algebras=εB(ϕ(aei))ϕ(ei)Δ is a morphism of A-bimodules=εB(ϕ(a)ϕ(ei))ϕ(ei)as ϕ is a morphism of algebras=εB(ϕ(a)fj)fjsϕProposition 4.8=ϕ(a)sϕεB trace of B\begin{array}[]{rclc}s_{\phi}\phi(a)&=&\displaystyle{\sum\varepsilon_{B}\bigl{(}\phi\bigl{(}e^{i}\bigr{)}\bigr{)}\phi\bigl{(}e_{i}\bigr{)}\phi(a)}&\\ &=&\displaystyle{\sum\varepsilon_{B}\bigl{(}\phi\bigl{(}e^{i}\bigr{)}\bigr{)}\phi\bigl{(}e_{i}a\bigr{)}}&\mbox{as $\phi$ is a morphism of algebras}\\ &=&\displaystyle{\sum\varepsilon_{B}\bigl{(}\phi\bigl{(}ae^{i}\bigr{)}\bigr{)}\phi\bigl{(}e_{i}\bigr{)}}&\mbox{$\Delta$ is a morphism of $A$-bimodules}\\ &=&\displaystyle{\sum\varepsilon_{B}\bigl{(}\phi(a)\phi\bigl{(}e^{i}\bigr{)}\bigr{)}\phi\bigl{(}e_{i}\bigr{)}}&\mbox{as $\phi$ is a morphism of algebras}\\ &=&\displaystyle{\sum\varepsilon_{B}\bigl{(}\phi(a)f^{j}\bigr{)}f_{j}s_{\phi}}&\mbox{Proposition 4.8}\\ &=&\phi(a)s_{\phi}&\mbox{$\varepsilon_{B}$ trace of $B$}\end{array}

Lemma 4.10.

Let AA be a Frobenius RR-algebra and X,XX,X^{\prime} right AA-modules such that XX is a finitely generated projective RR-module then:
The submodule HomAproj(X,X)Hom^{proj}_{A}(X,X^{\prime}) of HomA(X,X)Hom_{A}(X,X^{\prime}) consisting of maps that factorize through a finitely generated projective right AA-module coincides with Im(TrA)Im(Tr_{A}) where

TrA:HomR(X,X)HomA(X,X)α[xα(xei)ei]\begin{array}[]{cccc}Tr_{A}:&Hom_{R}(X,X^{\prime})&\to&Hom_{A}(X,X^{\prime})\\ &\alpha&\mapsto&\Bigl{[}\displaystyle{x\mapsto\sum\alpha\bigl{(}xe^{i}\bigr{)}e_{i}}\Bigr{]}\end{array}
Proof.

We will first consider βHomAproj(X,X)\beta\in Hom^{proj}_{A}(X,X^{\prime}) and prove that βIm(TrA)\beta\in Im(Tr_{A}). Since βHomAproj(X,X)\beta\in Hom^{proj}_{A}(X,X^{\prime}), there exist γ:XP\gamma:X\to P and δ:PX\delta:P\to X^{\prime} morphisms of right AA-modules such that δγ=β\delta\circ\gamma=\beta. Since PP is finitely generated projective right AA-module, given yPy\in P, y=yigi(y)y=\sum y_{i}g_{i}(y) with gi:PAg_{i}:P\to A of right AA-modules.

We will now consider α(x)=(tgiγ)(x)δ(yi)\alpha(x)=\sum(t\circ g_{i}\circ\gamma)(x)\delta(y_{i}) with tt the trace of AA and prove that TrA(α)=βTr_{A}(\alpha)=\beta:

TrA(α)(x)\displaystyle Tr_{A}(\alpha)(x) =jα(xej)ej=ji(tgiγ)(xej)δ(yi)ej\displaystyle=\sum_{j}\alpha\bigl{(}xe^{j}\bigr{)}e_{j}=\sum_{j}\sum_{i}(t\circ g_{i}\circ\gamma)(xe^{j})\delta(y_{i})e_{j}
=jit(gi(γ(x)ej))δ(yi)ej\displaystyle=\sum_{j}\sum_{i}t(g_{i}(\gamma(x)e^{j}))\delta(y_{i})e_{j}
=jit((gi(γ(x))ej)δ(yi)ej\displaystyle=\sum_{j}\sum_{i}t((g_{i}(\gamma(x))e^{j})\delta(y_{i})e_{j}
=iδ(yi)jt((gi(γ(x))ej)ej\displaystyle=\sum_{i}\delta(y_{i})\sum_{j}t((g_{i}(\gamma(x))e^{j})e_{j}
=iδ(yi)gi(γ(x))\displaystyle=\sum_{i}\delta(y_{i})g_{i}(\gamma(x))
=δ(iyigi(γ(x)))\displaystyle=\delta\left(\sum_{i}y_{i}g_{i}\bigl{(}\gamma(x)\bigr{)}\right)
=δ(γ(x))=β(x).\displaystyle=\delta(\gamma(x))=\beta(x).

Now we will start with βIm(TrA)\beta\in Im(Tr_{A}) and prove that βHomRproj(X,X)\beta\in Hom^{proj}_{R}(X,X^{\prime}).
Let β(x)=TrA(α)(x)=jα(xej)ej\beta(x)=Tr_{A}(\alpha)(x)=\sum_{j}\alpha(xe^{j})e_{j} for some αHomR(X,X)\alpha\in Hom_{R}(X,X^{\prime}). Since XX is a finitely generated projective RR-module then xej=i=1nfi(xej)xi\displaystyle{xe^{j}=\sum_{i=1}^{n}f_{i}\bigl{(}xe^{j}\bigr{)}x_{i}} with fi:XRf_{i}:X\to R of RR-modules and

β(x)=jα(xej)ej=jα(ifi(xej)xi)ej=i,jfi(xej)α(xi)ej.\beta(x)=\sum_{j}\alpha\bigl{(}xe^{j}\bigr{)}e_{j}=\sum_{j}\alpha\Bigl{(}\sum_{i}f_{i}(xe^{j})x_{i}\Bigr{)}e_{j}=\sum_{i,}\sum_{j}f_{i}(xe^{j})\alpha(x_{i})e_{j}.

Let us now define γ:XAn\gamma:X\to A^{n} such that

γ(x)=(jf1(xej)ej,jf2(xej)ej,,jfn(xej)ej).\gamma(x)=\Bigl{(}\sum_{j}f_{1}(xe^{j})e_{j},\sum_{j}f_{2}(xe^{j})e_{j},\cdots,\sum_{j}f_{n}(xe^{j})e_{j}\Bigr{)}.

To see that γ\gamma is a right AA-module morphism we use that ΔA\Delta_{A} is an AA-bimodule morphism, that is

γ(xa)=(jf1(xaej)ej,jf2(xaej)ej,,jfn(xaej)ej)=(jf1(xej)eja,jf2(xej)eja,,jfn(xej)eja)=γ(x)a\begin{array}[]{rcl}\gamma(x\cdot a)&=&\Bigl{(}\sum_{j}f_{1}(xae^{j})e_{j},\sum_{j}f_{2}(xae^{j})e_{j},\cdots,\sum_{j}f_{n}(xae^{j})e_{j}\Bigr{)}\\ &=&\Bigl{(}\sum_{j}f_{1}(xe^{j})e_{j}a,\sum_{j}f_{2}(xe^{j})e_{j}a,\cdots,\sum_{j}f_{n}(xe^{j})e_{j}a\Bigr{)}\\ &=&\gamma(x)\cdot a\end{array}

Let be δ:AnX\delta:A^{n}\to X^{\prime} defined as

δ(a1,a2,,an)=i=1nα(xi)ai.\delta(a_{1},a_{2},\cdots,a_{n})=\sum_{i=1}^{n}\alpha(x_{i})a_{i}.

It is clear that δ\delta is a right AA-module morphism and β=δγ\beta=\delta\circ\gamma.
Finally, as AnA^{n} is finitely generated projective right AA-module, we conclude that βHomAproj(X,X)\beta\in Hom^{proj}_{A}(X,X^{\prime}). ∎

Proposition 4.11.

Let AA be a Frobenius RR-algebra with Casimir element CA=ieieiC_{A}=\sum_{i}e^{i}\otimes e_{i}. Let XX be a finitely generated RR-module. Then XX is a projective right AA-module if and only if there exists a morphism of RR-modules α\alpha such that x=α(xei)eix=\sum\alpha\bigl{(}xe^{i}\bigr{)}e_{i}, xX\forall x\in X.

Proof.

(\Rightarrow) Consider XX a projective right AA-module. Trivially IdXHomRproj(X,X)Id_{X}\in Hom^{proj}_{R}(X,X) and using the above lemma there exists αEndR(X)\alpha\in End_{R}(X) such that IdX=TrA(α)Id_{X}=Tr_{A}(\alpha) and α(xei)ei=x\sum\alpha\bigl{(}xe^{i}\bigr{)}e_{i}=x for all xXx\in X.
(\Leftarrow) We have that TrA(α)=IdXTr_{A}(\alpha)=Id_{X} for some αEndR(X)\alpha\in End_{R}(X) so IdXHomRproj(X,X)Id_{X}\in Hom^{proj}_{R}(X,X). This means that there exist PP a finitely generated projective right AA-module and right AA-module morphisms f:XPf:X\to P, g:PXg:P\to X such that gf=IdXg\circ f=Id_{X}. This information allows us to prove in a simple way that XX is a right projective AA-module.

Next we will prove a weaker version of Proposition 4.4 of [8] with AA and BB Frobenius algebras. Note that if we take AA and BB non-symmetric Frobenius algebras in condition 2 we can only say that the morphism splits as right AA-module morphism.

Proposition 4.12.

Let AA and BB Frobenius algebras and ϕ:AB\phi:A\twoheadrightarrow B an epimorphism of algebras. The following conditions are equivalent:

  1. 1.

    sϕs_{\phi} is invertible in BB.

  2. 2.

    ϕ:AB\phi:A\twoheadrightarrow B splits as a right AA-module morphism.

  3. 3.

    BB is a projective right AA-module.

  4. 4.

    Every projective BB-module is a projective right AA-module.

Proof.

For (1) \Rightarrow (2) consider CA=eieiC_{A}=\sum e^{i}\otimes e_{i} and define the map σ:BA\sigma:B\to A such that

σ(b)=εB(bsϕ1ϕ(ei))ei.\sigma(b)=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(e^{i})\bigr{)}e_{i}.

By definition, ϕσ(b)=εB(bsϕ1ϕ(ei))ϕ(ei)\phi\circ\sigma(b)=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(e^{i})\bigr{)}\phi(e_{i}).

Since (ϕϕ)(CA)=sϕCB(\phi\otimes\phi)(C_{A})=s_{\phi}C_{B} and sϕs_{\phi} is invertible we deduce that sϕ1(ϕϕ)(CA)=CBs_{\phi}^{-1}(\phi\otimes\phi)(C_{A})=C_{B}. This means that CB=fifi=sϕ1ϕ(ei)ϕ(ei)C_{B}=\sum f^{i}\otimes f_{i}=\sum s_{\phi}^{-1}\phi(e^{i})\otimes\phi(e_{i}). Using that εB(bfi)fi=b\sum\varepsilon_{B}(bf^{i})f_{i}=b we conclude that εB(bsϕ1ϕ(ei))ϕ(ei)=b\sum\varepsilon_{B}(bs_{\phi}^{-1}\phi(e^{i}))\phi(e_{i})=b and ϕ\phi splits.

To prove that σ\sigma is a right AA-module morphism we use that sϕZBs_{\phi}\in ZB.

σ(ba)=σ(bϕ(a))=εB(bϕ(a)sϕ1ϕ(ei))ei=εB(bsϕ1ϕ(a)ϕ(ei))ei\sigma\bigl{(}b\cdot a\bigr{)}=\sigma\bigl{(}b\phi(a)\bigr{)}=\sum\varepsilon_{B}\bigl{(}b\phi(a)s_{\phi}^{-1}\phi\bigl{(}e^{i}\bigr{)}\bigr{)}e_{i}=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(a)\phi\bigl{(}e^{i}\bigr{)}\bigr{)}e_{i}
=εB(bsϕ1ϕ(aei))ei=εB(bsϕ1ϕ(ei))eia=σ(b)a.=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi\bigl{(}ae^{i}\bigr{)}\bigr{)}e_{i}=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi\bigl{(}e^{i}\bigr{)}\bigr{)}e_{i}a=\sigma(b)\cdot a.

For (2) \Rightarrow (3) we need to prove that BB is a projective right AA-module considering the action ba=bϕ(a)b\cdot a=b\phi(a). Given an epimorphism of A-modules f:MNf:M\to N and an epimorphism g:BMg:B\to M we need to find h:BMh:B\to M, a morphism of AA-modules such that fh=gf\circ h=g. Consider gϕ:ANg\circ\phi:A\to N. Since AA is a projective AA-module then there exists p:AMp:A\to M such that fp=gϕf\circ p=g\circ\phi. If h=pσh=p\circ\sigma then fh=fpσ=gϕσ=gf\circ h=f\circ p\circ\sigma=g\circ\phi\circ\sigma=g.

(3) \Rightarrow (4) is a known result.

Finally, we will prove (4) \Rightarrow (1).
Since BB is a projective AA-module then there exists αEndR(B)\alpha\in End_{R}(B) such that TrA(α)=IdBTr_{A}(\alpha)=Id_{B}, that is b=α(bei)eib=\sum\alpha(b\cdot e^{i})\cdot e_{i} for all bBb\in B.
On the other hand, since

ϕ(ei)ϕ(ei)=sϕfjfj=fjfjsϕ\sum\phi(e^{i})\otimes\phi(e_{i})=s_{\phi}\sum f^{j}\otimes f_{j}=\sum f^{j}\otimes f_{j}s_{\phi}

we have that

TrA(α)(b)\displaystyle Tr_{A}(\alpha)(b) =α(bei)ei\displaystyle=\sum\alpha(b\cdot e^{i})\cdot e_{i}
=α(bϕ(ei))ϕ(ei)\displaystyle=\sum\alpha(b\phi(e^{i}))\phi(e_{i})
=α(bfj)fjsϕ\displaystyle=\sum\alpha(bf^{j})f_{j}s_{\phi}
=TrB(α)(b)sϕ,\displaystyle=Tr_{B}(\alpha)(b)s_{\phi},

so TrB(α)(1B)sϕ=TrA(α)(1B)=IdB(1B)=1BTr_{B}(\alpha)(1_{B})s_{\phi}=Tr_{A}(\alpha)(1_{B})=Id_{B}(1_{B})=1_{B} and sϕs_{\phi} is invertible.

It is natural to ask if all the equivalences in Proposition 4.12 hold in the case AA symmetric nearly Frobenius or just nearly Frobenius. The following results and examples answer this questions.

The following example shows that Proposition 4.12 is not valid if we require AA to be a non-symmetric nearly Frobenius algebra.

Example 4.1.

Consider A=𝕂Q1I1\displaystyle{A=\frac{\mathbb{K}Q_{1}}{I_{1}}} and B=𝕂Q2I2\displaystyle{B=\frac{\mathbb{K}Q_{2}}{I_{2}}} with Q1Q_{1} and Q2Q_{2}:

Q1=1αβ2andQ2=1αQ_{1}=\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 7.74307pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-7.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bullet_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{}{{}{{}{{}{{}{{}}{}{{}{{}{{}{{}}{}{{}}{}{{}}{}{{}{{}}}}}}}}}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{}{{}{{}}}\ignorespaces\ignorespaces{\hbox{\kern-5.59938pt\raise 23.59885pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{\alpha}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{}{{}{{}}{}{{}}{}{{}}{}{{}{{}}{}{{}}{}{{}}{}{{}{{}{{}{{}}{}{{}}}}}}}{\hbox{\kern 4.25598pt\raise 4.944pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}{}}{}}}}\ignorespaces{}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 14.47588pt\raise-6.1111pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.75pt\hbox{$\scriptstyle{\beta}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 31.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 31.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bullet_{2}}$}}}}}}}\ignorespaces}}}}\ignorespaces\quad\mbox{and}\quad Q_{2}=\lx@xy@svg{\hbox{\raise 2.5pt\hbox{\kern 7.74307pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr\crcr}}}\ignorespaces{\hbox{\kern-7.74307pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise-2.5pt\hbox{$\textstyle{\bullet_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{}{{}{{}{{}{{}{{}}{}{{}{{}{{}{{}}{}{{}}{}{{}}{}{{}{{}}}}}}}}}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{}{{}{{}}}\ignorespaces\ignorespaces{\hbox{\kern-5.59938pt\raise 23.59885pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{\alpha}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{}{{}{{}}{}{{}}{}{{}}{}{{}{{}}{}{{}}{}{{}}{}{{}{{}{{}{{}}{}{{}}}}}}}{\hbox{\kern 4.25598pt\raise 4.944pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}{}}{}}}}\ignorespaces{}\ignorespaces}}}}\ignorespaces

I1=α2,βαI_{1}=\langle\alpha^{2},\beta\alpha\rangle and I2=α2I_{2}=\langle\alpha^{2}\rangle.
The algebra BB is a symmetric algebra with Casimir element CB=1α+α1C_{B}=1\otimes\alpha+\alpha\otimes 1, and counit ε:B𝕂\varepsilon:B\to\mathbb{K} defined as ε(e1)=1\varepsilon(e_{1})=1 and ε(α)=0\varepsilon(\alpha)=0.
The algebra AA is a nearly Frobenius algebra and the coproducts have the form ΔA(1)=aαα+aαβ\Delta_{A}(1)=a\alpha\otimes\alpha+a^{\prime}\alpha\otimes\beta where a,a𝕂a,a^{\prime}\in\mathbb{K}.
Consider the epimorphism of algebras ϕ:AB\phi:A\to B with ϕ(e1)=e1\phi(e_{1})=e_{1}, ϕ(e2)=0\phi(e_{2})=0, ϕ(α)=α\phi(\alpha)=\alpha and ϕ(β)=0\phi(\beta)=0.
Then sϕ=ε(aα)α+ε(aα)0=0s_{\phi}=\varepsilon(a\alpha)\alpha+\varepsilon(a^{\prime}\alpha)0=0 so is not invertible but if we consider the inclusion of Q2Q_{2} in Q1Q_{1} i:BAi:B\to A the morphism ϕ\phi splits.

Proposition 4.13.

Let AA be a symmetric nearly Frobenius algebra and BB symmetric algebra with ϕ:AB\phi:A\twoheadrightarrow B an epimorphism of algebras. Then if sϕs_{\phi} is invertible,ϕ,\phi splits as a morphism of AA-bimodules.

Proof.

We can use the same map as the Proposition 4.12, that is σ:BA\sigma:B\to A such that

σ(b)=ϵB(bsϕ1ϕ(ei))ei.\sigma(b)=\sum\epsilon_{B}(bs_{\phi}^{-1}\phi(e^{i}))e_{i}.

We only need to prove that σ\sigma is a morphism of left AA-modules.

σ(ab)=σ(ϕ(a)b)=εB(ϕ(a)bsϕ1ϕ(ei))ei=εB(bsϕ1ϕ(ei)ϕ(a))ei.\sigma(a\cdot b)=\sigma\bigl{(}\phi(a)b\bigr{)}=\sum\varepsilon_{B}\bigl{(}\phi(a)bs_{\phi}^{-1}\phi(e^{i})\bigr{)}e_{i}=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(e^{i})\phi(a)\bigr{)}e_{i}.

The last identification is a consequence of the symmetry property of BB. On the other hand, using that ϕ\phi is a morphism of algebras we have that:

σ(ab)=εB(bsϕ1ϕ(eia))ei.\sigma(a\cdot b)=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(e^{i}a)\bigr{)}e_{i}.

The symmetry property on AA means that eiei=eiei\sum e^{i}\otimes e_{i}=\sum e_{i}\otimes e^{i}, using that Δ\Delta is a morphism of AA-bimodules we have

aeiei=eieia=aeiei=eieia.\sum ae^{i}\otimes e_{i}=\sum e^{i}\otimes e_{i}a=\sum ae_{i}\otimes e^{i}=\sum e_{i}\otimes e^{i}a.

Then

σ(ab)=εB(bsϕ1ϕ(eia))ei=εB(bsϕ1ϕ(ei))aei=aεB(bsϕ1ϕ(ei))ei=aσ(b).\sigma(a\cdot b)=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(e^{i}a)\bigr{)}e_{i}=\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(e^{i})\bigr{)}ae_{i}=a\sum\varepsilon_{B}\bigl{(}bs_{\phi}^{-1}\phi(e^{i})\bigr{)}e_{i}=a\sigma(b).

The following example shows that a version of Proposition 4.12 in the context of Proposition 4.13 does not hold.

Example 4.2.

Let be A=𝕜[x]x2\displaystyle{A=\frac{\Bbbk[x]}{x^{2}}} and Δ1(1)=xx\Delta_{1}(1)=x\otimes x, Δ2(x)=x1+1x\Delta_{2}(x)=x\otimes 1+1\otimes x. Note that (A,Δ1)\bigl{(}A,\Delta_{1}\bigr{)} is a symmetric nearly Frobenius algebra and (A,Δ2)\bigl{(}A,\Delta_{2}\bigr{)} is a symmetric algebra. In this case we consider ϕ=IdA:AA\phi=\operatorname{Id_{A}}:A\to A. For this map the Schur element is

sIdA=εA(ei)ei=εA(x)x=x.s_{\operatorname{Id_{A}}}=\sum\varepsilon_{A}\bigl{(}e^{i}\bigr{)}e_{i}=\varepsilon_{A}(x)x=x.

The Schur element is not invertible in AA, but the map IdA\operatorname{Id_{A}} split as a morphism of AA-bimodule.

References

  • [1] L. Abrams, PhD thesis, Frobenius Algebra Structures in Topological Quantum Field Theory and Quantum Cohomology (1997).
  • [2] L. Abrams, Modules, Comodules, and Cotensor Products over Frobenius algebras. Journal of Algebra 219, pp. 201-213 (1999).
  • [3] L. Abrams. The Quantum Euler class and the Quantum Cohomology of the Grassmannians. Israel Journal of Mathematics 117, pp. 335-352 (2000).
  • [4] D. Artenstein, A. González, and M. Lanzilotta, Constructing Nearly Frobenius Algebras. Algebras and Representation Theory Volume 18, 339-367 (2015).
  • [5] D. Artenstein, A. González. and G. Mata, Nearly Frobenius structures in some families of algebras. São Paulo J. Math. Sci. 14, 165–184 (2020). DOI: 10.1007/s40863-019-00158-z.
  • [6] D. Artenstein, A. González and G. Mata. Nearly Frobenius theory and semisimplicity of bimodules. Communications in Algebra, 49:7, pp. 3124-3139 (2021). DOI: 10.1080/00927872.2021.1888961.
  • [7] D. Artenstein, A. González and G. Mata. Nearly Frobenius dimension of Frobenius algebras. Communications in Algebra, 50:11, pp. 4906-4916 (2022). DOI: 10.1080/00927872.2022.2077953D.
  • [8] M. Broué, Higman’s criterion revisited. Michigan Math. J. 58 (2009), pp. 125-179. DOI: 10.1307/mmj/1242071686.
  • [9] S. Caenepeel, Bogdan Ion, G. Militaru. The Structure of Frobenius Algebras and Separable Algebras. K-Theory v. 19 (2000), no. 4, 365-402.
  • [10] R. Cohen and V. Godin. A Polarized View of String Topology. Topology, Geometry and Quantum Field Theory, London Math. Soc. Lecture Note Ser. Vol. 308, Cambridge: Cambridge Univ. Press, pp. 127-154 (2004).
  • [11] M. Geck, Beiträge zur Darstellungstheorie von Iwahori-Hecke-Algebren, Habilitationsschrift, RWTH Aachen, 1993.
  • [12] A. González, E. Lupercio, C. Segovia, B. Uribe. Orbifold Topological Quantum Field Theories in Dimension 2. Preprint (2013).
  • [13] Em (John) Grifall-Sabo. Invariants from Group Algebras via Topological Quantum Field Theory. Thesis. University of Nevada, Reno (2020).
  • [14] V. Gubitosi. Open Frobenius cluster tilted algebras. Algebra Colloquium Vol. 29, No. 01, pp. 1-22 (2022).
  • [15] A. Jacoby and M. Lorenz. Frobenius divisibility for Hopf algebras. Contemp. Math. 688, 125-137 (2017).
  • [16] J. Kock. Frobenius Algebras and 2-D Topological Quantum Field Theories. Cambridge University Press (2004).
  • [17] M. Lorenz. Some applications of Frobenius algebras to Hopf Algebras. Groups, algebras and applications. XVIII Latin American algebra colloquium, São Pedro, Brazil, August 3–8, 2009. Proceedings, 269-289, (2011).