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

Flat GL(1|1)GL(1|1)-connections and fatgraphs

Andrea Bourque  and  Anton M. Zeitlin Department of Mathematics, Louisiana State University, Baton Rouge, LA 70803, USA
Abstract.

We study the moduli space of flat GL(1|1)GL(1|1)-connections on a punctured surface from the point of view of graph connections. To each fatgraph, a system of coordinates is assigned, which involves two bosonic and two fermionic variables per edge, subject to certain relations. In the case of trivalent graphs, we provide a closed explicit formula for the Whitehead moves. In addition, we discuss the invariant Poisson bracket.

1. Introduction

Recently, the study of super-analogues of Teichmüller spaces and supermoduli spaces achieved much progress. In particular, Penner-type coordinates were discovered in [penzeit], [ipz], [ipz2] for N=1,N=2N=1,N=2 versions of Teichmüller space and their decorated analogues. In that particular case, these spaces were viewed as a subspace of the character variety Hom(π1(F),G)/G\operatorname{Hom}(\pi_{1}(F),G)/G, where π1(F)\pi_{1}(F) is the fundamental group of a hyperbolic Riemann surface FF with punctures. For the standard Teichmüller space T(F)T(F), G=PSL(2,)G=PSL(2,\mathbb{R}), while the super-analogues STN=1(F)ST_{N=1}(F), STN=2(F)ST_{N=2}(F) cases are related to rank 11 and rank 22 supergroups OSP(1|2)OSP(1|2), OSP(2|2)OSP(2|2) correspondingly.

Penner coordinates [DTT], [penner] are essential not only in the study of hyperbolic geometry, but they also provide a geometric example of a fundamental algebraic object, known as a cluster algebra. Let us briefly characterize the context. One starts from an ideal triangulation of the Riemann surface, or, equivalently, the dual trivalent fatgraph (aka ribbon graph). Penner coordinates assign a parameter for every edge of triangulation/fatgraph related to a suitably renormalized geodesic length. This provides coordinates for a trivial bundle T~(F)\tilde{T}(F) over T(F)T(F), known as the decorated Teichmüller space. There is a simple transformation to coordinates on Teichmüller space so that these new coordinates are subject to linear constraints. One of the benefits of such coordinates is that the action of mapping class group on T~(F)\tilde{T}(F) is described in a combinatorial way by embedding into the Ptolemy groupoid. The Ptolemy groupoid is generated by elementary moves on the fatgraphs, called flips. The action of flips on Penner coordinates gives an example of the so-called cluster transformations. The other benefit of Penner coordinates is that they serve as Darboux-type coordinates for the Weil-Petersson 2-form, which makes them useful for the quantization of T(F)T(F) [Kashaev97], [Chekhov99].

The supergroups OSP(1|2)OSP(1|2), OSP(2|2)OSP(2|2) which give rise to super-Teichmüller spaces both contain G=SL(2,)G=SL(2,\mathbb{R}) as their body subgroup. Penner coordinates have been generalized successfully in both of these cases, leading to the super-analogue of Ptolemy transformation. Currently, there are a lot of attempts to construct a super-analogue of cluster algebras based on these formulas [ovshap],[musikoven], [musikoven2]. The critical ingredient of both constructions were the graph GG-connections: 2\mathbb{Z}_{2}-graph connections describing the spin structures on FF for STN=1(F)ST_{N=1}(F), and two such spin structures accompanied by GL(1)+GL(1)_{+}-graph connection for STN=2(F)ST_{N=2}(F). In particular, part of the decoration for STN=2(F)ST_{N=2}(F) was related to the gauge equivalences of GL(1)+GL(1)_{+}-graph connections. One of the choices of root systems for OSP(2|2)OSP(2|2) is such that simple roots are “grey”, namely each of them gives rise to a GL(1|1)GL(1|1) subgroup. GL(1|1)GL(1|1) is a reductive supergroup of rank 11, which contains two abelian subgroups as its body. Thus, only the odd coordinates are responsible for non-commutativity.

In this note, we study the first nontrivial case of a character variety related to simple supergroups, namely

MGL(1|1)=Hom(π1(F),GL(1|1))/GL(1|1),{\rm M}_{GL(1|1)}=\operatorname{Hom}(\pi_{1}(F),GL(1|1))/GL(1|1),

which is quite interesting on its own. We will consider this space as the space of GL(1|1)GL(1|1)-graph connections on the trivalent fatgraph associated with FF. Then, we will define coordinates on this space through the assignment of specific parameters to edges of the fatgraph. These parameters are related to Gaussian decomposition of GL(1|1)GL(1|1). Using these coordinates, we obtain a characterization of the action of the Ptolemy groupoid. Finally, we discuss a Poisson bracket structure on MGL(1|1){\rm M}_{GL(1|1)}.

We believe that GL(1|1)GL(1|1) character variety is essential in the context of the super-analogue of abelianization [hn]. Another important context is the quantum GL(1|1)GL(1|1) Chern-Simons theory [Rozansky92], which recently attracted some attention [Aghaei18].

The structure of the paper is as follows. Section 2 reviews some of the notions of super mathematics and necessary facts about the GL(1|1)GL(1|1) supergroup. Section 3 discusses fatgraphs, GG-graph connections, and Ptolemy groupoid actions on trivalent fatgraph GG-connections. Section 4 is devoted to the construction of a coordinate system on MGL(1|1){\rm M}_{GL(1|1)} and its decorated version using its fatgraph GL(1|1)GL(1|1)-connection realization through Gaussian decomposition of GL(1|1)GL(1|1). Section 5 is primarily of a computation nature, where we describe the action of flip transformation using the minimal amount of changing variables in the decorated space. Finally, in Section 6, we discuss the Poisson bracket structure on MGL(1|1){\rm M}_{GL(1|1)}.

Acknowledgements

We thank R.C. Penner for useful discussions and his comments on the manuscript. A.M.Z. is partially supported by Simons Collaboration Grant 578501 and NSF grant DMS-2203823.

2. GL(1|1)GL(1|1) supergroup

2.1. Conventions on superspaces and Grassmann algebras

In this section we follow the conventions from [Supermanifolds].

Let S[N]\mathbb{R}^{S[N]} the real Grassmann algebra with generators 1,β[i]1,\beta_{[i]}, for i=1,2,,Ni=1,2,...,N. The generators have relations 1β[i]=β[i]=β[i]11\beta_{[i]}=\beta_{[i]}=\beta_{[i]}1 and β[i]β[j]=β[j]β[i]\beta_{[i]}\beta_{[j]}=-\beta_{[j]}\beta_{[i]}. In particular, (β[i])2=0(\beta_{[i]})^{2}=0. We also use the notation β[λ]=β[λ1]β[λk]\beta_{[\lambda]}=\beta_{[\lambda_{1}]}\cdot\cdot\cdot\beta_{[\lambda_{k}]} for an ordered multi-index λ=λ1,,λk\lambda=\lambda_{1},...,\lambda_{k}. Note that if λ\lambda has a repeated index, then β[λ]=0\beta_{[\lambda]}=0. If the multi-index is empty, the corresponding element is the empty product, 1. By our commutation relations, any β[λ]\beta_{[\lambda]} can have its terms rearranged so that the indices of the terms are increasing. Thus, an element in S[N]\mathbb{R}^{S[N]} can be written in the form x=λxλβ[λ]x=\sum_{\lambda}x_{\lambda}\beta_{[\lambda]} for xλx_{\lambda}\in\mathbb{R} as λ\lambda runs over all strictly increasing multi-indices.

Definition 2.1.

The degree of a term xλβ[λ]S[N]x_{\lambda}\beta_{[\lambda]}\in\mathbb{R}^{S[N]} is defined as the size of the multi-index λ\lambda.

Thus S[N]\mathbb{R}^{S[N]} has a superalgebra structure given by the decomposition S[N]=0S[N]1S[N]\mathbb{R}^{S[N]}=\mathbb{R}^{S[N]}_{0}\oplus\mathbb{R}^{S[N]}_{1} into elements which are sums of terms of even (respectively, odd) degree. Since the β\beta generators anti-commute, S[N]\mathbb{R}^{S[N]} is supercommutative.

Definition 2.2.

The body map ϵ:S[N]\epsilon:\mathbb{R}^{S[N]}\longrightarrow\mathbb{R} is the projection of an element onto its coefficient of 11. The soul map s:S[N]S[N]s:\mathbb{R}^{S[N]}\longrightarrow\mathbb{R}^{S[N]} is defined by s(x)=xϵ(x)1s(x)=x-\epsilon(x)\cdot 1.

Since there are NN anti-commuting generators, we have that s(x)N+1=0s(x)^{N+1}=0. Thus ϵ(x)0\epsilon(x)\neq 0 if and only if xx is invertible. Explicitly,

x1=1ϵ(x)(1s(x)ϵ(x)+(s(x)ϵ(x))2+(1)N(s(x)ϵ(x))N).x^{-1}=\dfrac{1}{\epsilon(x)}\left(1-\dfrac{s(x)}{\epsilon(x)}+\left(\dfrac{s(x)}{\epsilon(x)}\right)^{2}-...+(-1)^{N}\left(\dfrac{s(x)}{\epsilon(x)}\right)^{N}\right).

In a similar vein, we can use series expansions to define, say, x\sqrt{x} for xx with positive body, as the series will terminate.

Remark.

Inequalities such as x>0,x<0,x0x>0,x<0,x\neq 0 for xS[N]x\in\mathbb{R}^{S[N]} will be taken to be inequalities on the body ϵ(x)\epsilon(x).

Definition 2.3.

Given S[N]=0S[N]1S[N]\mathbb{R}^{S[N]}=\mathbb{R}^{S[N]}_{0}\oplus\mathbb{R}^{S[N]}_{1}, the superspace 1|1\mathbb{R}^{1|1} is defined as 0S[N]×1S[N]\mathbb{R}^{S[N]}_{0}\times\mathbb{R}^{S[N]}_{1}.

In other words, we have a two-dimensional space with one even and one odd coordinate. We can define more generally p|q=(0S[N])p×(1S[N])q\mathbb{R}^{p|q}=\left(\mathbb{R}^{S[N]}_{0}\right)^{p}\times\left(\mathbb{R}^{S[N]}_{1}\right)^{q}.

Remark.

From now on, we will use the convention that odd elements are denoted by Greek letters, and even elements are denoted by Latin letters.

2.2. GL(1|1)GL(1|1) supergroup and its Lie superalgebra

Definition 2.4.

A Lie superalgebra is a superalgebra whose multiplication, denoted by [X,Y][X,Y], is super-anticommutative, and furthermore satisfies the super Jacobi identity (1)|X||Z|[X,[Y,Z]]+(1)|Z||Y|[Z,[X,Y]]+(1)|Y||X|[Y,[Z,X]]=0(-1)^{|X||Z|}[X,[Y,Z]]+(-1)^{|Z||Y|}[Z,[X,Y]]+(-1)^{|Y||X|}[Y,[Z,X]]=0.

Let us introduce a Lie superalgebra 𝔤𝔩(1|1)\mathfrak{gl}(1|1). It has two even generators E,NE,N and two odd generators Ψ±\Psi^{\pm}, which satisfy the following commutation relations:

(2.1) [N,Ψ±]=±Ψ±,[Ψ+,Ψ]=E,[E,Ψ±]=[E,N]=0.\displaystyle[N,\Psi^{\pm}]=\pm\Psi^{\pm},\quad[\Psi^{+},\Psi^{-}]=E,\quad[E,\Psi^{\pm}]=[E,N]=0.

In the defining representation, as elements of End(1|1){\rm End}(\mathbb{R}^{1|1}) these generators are given by the supermatrices below:

(2.2) E=(1001),N=(120012),Ψ+=(0100),Ψ=(0010).\displaystyle E=\begin{pmatrix}1&0\\ 0&1\end{pmatrix},N=\begin{pmatrix}\frac{1}{2}&0\\ 0&-\frac{1}{2}\end{pmatrix},\Psi^{+}=\begin{pmatrix}0&1\\ 0&0\end{pmatrix},\Psi^{-}=\begin{pmatrix}0&0\\ 1&0\end{pmatrix}.
Definition 2.5.

GL(1|1)GL(1|1) is the group of invertible linear transformations of 1|1\mathbb{R}^{1|1}. Elements of GL(1|1)GL(1|1) can be identified as supermatrices of the form (aαβb)\begin{pmatrix}a&\alpha\\ \beta&b\end{pmatrix} for a,b0S[N],α,β1S[N]a,b\in\mathbb{R}^{S[N]}_{0},\alpha,\beta\in\mathbb{R}^{S[N]}_{1} with ϵ(a),ϵ(b)0\epsilon(a),\epsilon(b)\neq 0.

In what follows, GL(1|1)GL(1|1) will refer to the identity component of this group, which means the even entries a,ba,b will have positive body.

The Lie superalgebra of GL(1|1)GL(1|1) is given by 𝔤𝔩(1|1)\mathfrak{gl}(1|1), so that any element of GL(1|1)GL(1|1) can be represented as eRe^{R}, where R=nN+eE+ψ+Ψ++ψΨR=nN+eE+\psi_{+}\Psi^{+}+\psi_{-}\Psi^{-}.

To keep the Lie superalgebra/Lie supergroup correspondence explicit, we choose the following multiplication of two elements in GL(1|1)GL(1|1) as follows:

(2.3) (aαβb)(cγδd)=(acαδaγ+dαcβ+dδbdβγ).\displaystyle\begin{pmatrix}a&\alpha\\ \beta&b\end{pmatrix}\begin{pmatrix}c&\gamma\\ \delta&d\end{pmatrix}=\begin{pmatrix}ac-\alpha\delta&a\gamma+d\alpha\\ c\beta+d\delta&bd-\beta\gamma\end{pmatrix}.

In particular, we note the minus signs in the multiplication formula. Let us elaborate on this choice. We identify eαΨ+=1+αΨ+e^{\alpha\Psi^{+}}=1+\alpha\Psi^{+} =(1001)+α(0100)=\begin{pmatrix}1&0\\ 0&1\end{pmatrix}+\alpha\begin{pmatrix}0&1\\ 0&0\end{pmatrix} with (1α01)\begin{pmatrix}1&\alpha\\ 0&1\end{pmatrix}. Then, when multiplying (1α01)(10β1)\begin{pmatrix}1&\alpha\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ \beta&1\end{pmatrix}, the result should agree with (1+αΨ+)(1+βΨ)(1+\alpha\Psi^{+})(1+\beta\Psi^{-}). Upon expanding the latter product, we have the term αΨ+βΨ\alpha\Psi^{+}\beta\Psi^{-}. Since β\beta and Ψ+\Psi+ are both odd, we have Ψ+β=βΨ+\Psi^{+}\beta=-\beta\Psi^{+}. Thus

(1α01)(10β1)=1+αΨ++βΨαβΨ+Ψ=(1αβαβ1),\begin{pmatrix}1&\alpha\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ \beta&1\end{pmatrix}=1+\alpha\Psi^{+}+\beta\Psi^{-}-\alpha\beta\Psi^{+}\Psi^{-}=\begin{pmatrix}1-\alpha\beta&\alpha\\ \beta&1\end{pmatrix},

which agrees with our choice of multiplication.

Other references may define supermatrix multiplication without these extra signs. There is an isomorphism (aαβb)(aαβb)\begin{pmatrix}a&\alpha\\ \beta&b\end{pmatrix}\mapsto\begin{pmatrix}a&-\alpha\\ \beta&b\end{pmatrix} from our convention to the other convention.

There is also a notion of supertrace, namely str(aαβb)=abstr\begin{pmatrix}a&\alpha\\ \beta&b\end{pmatrix}=a-b. This gives rise to a nondegenerate invariant bilinear form on 𝔤𝔩(1|1)\mathfrak{gl}(1|1). This takes the role of the Killing form, which is degenerate in this case.

2.3. Parametrization of GL(1|1)GL(1|1) and its decomposition

Any element in GL(1|1)GL(1|1) admits the following unique Gaussian factorization:

(2.4) (aαβb)=(1αb01)(a+αβb00b)(10βb1).\displaystyle\begin{pmatrix}a&\alpha\\ \beta&b\end{pmatrix}=\begin{pmatrix}1&\frac{\alpha}{b}\\ 0&1\end{pmatrix}\begin{pmatrix}a+\frac{\alpha\beta}{b}&0\\ 0&b\end{pmatrix}\begin{pmatrix}1&0\\ \frac{\beta}{b}&1\end{pmatrix}.

We choose the following parametrization of gGL(1|1)g\in GL(1|1):

(2.5) g(a,b;α,β)=a(1+12b2αβ)(1α01)(b100b)(10β1).\displaystyle g(a,b;\alpha,\beta)=a(1+\frac{1}{2}b^{2}\alpha\beta)\begin{pmatrix}1&\alpha\\ 0&1\end{pmatrix}\begin{pmatrix}b^{-1}&0\\ 0&b\end{pmatrix}\begin{pmatrix}1&0\\ \beta&1\end{pmatrix}.

The scalar factor in front is meant as an element propotional to the unit matrix as an element of GL(1|1)GL(1|1). This presentation is unique and thus provides a parametrization of GL(1|1)GL(1|1). Given an element in the standard form, say (AϕψB)\begin{pmatrix}A&\phi\\ \psi&B\end{pmatrix}, its coordinates in this parametrization are

(2.6) a=AB,b=BA(1ϕψ2AB);α=ϕB,β=ψB.\displaystyle a=\sqrt{AB},b=\sqrt{\dfrac{B}{A}}\left(1-\dfrac{\phi\psi}{2AB}\right);\alpha=\dfrac{\phi}{B},\beta=\dfrac{\psi}{B}.

The reason to introduce the extra factor in front is that it provides compact formula for the inverse element. We write the formulas for multiplication and inversion in these coordinates:

(a,b;α,β)(c,d;γ,δ)=(ac(112βγ)(112b2d2αδ),bd;α+b2γ,d2β+δ),\displaystyle(a,b;\alpha,\beta)(c,d;\gamma,\delta)=(ac(1-\frac{1}{2}\beta\gamma)(1-\frac{1}{2}b^{2}d^{2}\alpha\delta),bd;\alpha+b^{-2}\gamma,d^{-2}\beta+\delta),
(2.7) (a,b;α,β)1=(a1,b1;b2α,b2β).\displaystyle(a,b;\alpha,\beta)^{-1}=(a^{-1},b^{-1};-b^{2}\alpha,-b^{2}\beta).

3. Fatgraphs and graph connections

We consider surfaces FF with genus g0g\geq 0 and s1s\geq 1 punctures, such that 2g2+s>02g-2+s>0.

Definition 3.1.

A fatgraph (also known as ribbon graph) is a graph with a cyclic ordering of the edges at each vertex. An orientation on a fatgraph is an assignment of direction to each edge of the graph.

One can reconstruct FF from a fatgraph τ\tau by fattening the edges and verticies along with cyclic ordering of edges at the vertices (see e.g. [DTT] for details). The condition 2g2+s>02g-2+s>0 is necessary to choose a trivalent fatgraph for a given surface, as the number of vertices of such a graph is equal to 2(2g2+s)2(2g-2+s). There are Whitehead (flip) transformations as in Figure 1, which take the edge ee between vertices u,vu,v, adjacent to c,dc,d and a,ba,b respectively, shrinks it, and then extends an edge ff connecting vertices u,vu^{\prime},v^{\prime} adjacent to b,cb,c and a,da,d respectively. Flips are known to act transitively on the collection of trivalent fatgraphs of FF. Altogether they form a Ptolemy groupoid Pt(F){\rm Pt}(F). Composition of flips are known to give generators for the mapping class group of a surface [DTT].

Given a fatgraph τ\tau and a Lie (super)group GG, we define a GG-graph connection on τ\tau as follows.

Definition 3.2.

A GG-graph connection is the assignment to each edge ee of τ\tau an element geGg_{e}\in G and an orientation, so that ge¯=ge1g_{\bar{e}}=g_{e}^{-1} if e¯\bar{e} is the inverse-oriented edge. We denote the space of GG-graph connections as M~G(τ)\tilde{\rm M}_{G}(\tau). Two graph connections {ge}\{g_{e}\}, {g~e}\{\tilde{g}_{e}\} are gauge equivalent if there is an assignment vhvGv\rightarrow h_{v}\in G, for all vertices vτv\in\tau, such that g~e=hv1gehv\tilde{g}_{e}=h_{v}^{-1}g_{e}h_{v^{\prime}}, where ee starts and ends at vv^{\prime} and vv respectively. We will denote the space of gauge equivalence classes of elements in M~G(τ)\tilde{\rm M}_{G}(\tau) by MG(τ){\rm M}_{G}(\tau).

The space of graph connections modulo gauge equivalences can be identified with a more common differential-geometric object. Namely, there is a natural one-to-one correspondence between MG(τ){\rm M}_{G}(\tau) and the moduli space of flat GG-connections. This correspondence is constructed as follows: one identifies MA(e)M_{A}(e), the monodromy of the flat connection AA along the oriented edge of the fatgraph ee, with the group element geg_{e}. This way, the space M~G(τ)\tilde{\rm M}_{G}(\tau) is identified with the space of flat connections modulo gauge transformations which are equal to identity at the vertices. Altogether, compositions of these group elements along the cycles of the fatgraph contain all information about the gauge classes of AA. However, there is a residual gauge symmetry at the vertices of the fatgraph, which one has to take into account, and that is precisely the equivalence relation for graph connections.

One can formulate this as follows.

Theorem 1.

If FF deformation retracts to τ\tau, then the moduli space of flat GG-connections on FF is isomorphic to the space of gauge equivalent classes of GG-graph connections on τ\tau corresponding to FF, i.e.

(3.1) MG(τ)Hom(π1(F),G)/G{\rm M}_{G}(\tau)\cong\operatorname{Hom}(\pi_{1}(F),G)/G

For any element AHom(π1(F),G)/GA\in\operatorname{Hom}(\pi_{1}(F),G)/G, let us denote A(τ)A(\tau) the corresponding graph connection on a fatgraph τ\tau.

We are interested in understanding the action of Pt(F){\rm Pt}(F) on graph connections. We require that for any ΓPt(F)\Gamma\in{\rm Pt}(F) and any cycle cτc\in\tau, if AA is the flat connection, the transformed connection AΓA^{\Gamma} is such that MAΓ(Γ(c))=MA(c)M_{A^{\Gamma}}(\Gamma(c))=M_{A}(c). This gives a natural action of Pt(F){\rm Pt}(F) on Hom(π1(F),G)/G\operatorname{Hom}(\pi_{1}(F),G)/G.

To write it formally, we look at the elementary flip transformation as in Figure 1, which involves 5 edges.

Refer to caption
Figure 1. Whitehead flip of a fatgraph, at the edge ee.

There are 6 pieces of possible cycles on such a fatgraph, corresponding to a choice of 2 of the 4 boundary edges. The monodromy conservation along these pieces after the flip transformation thus leads to the following equations:

gagegc\displaystyle g_{a}g_{e}g_{c} =gagfgc,\displaystyle=g_{a}^{\prime}g_{f}g_{c}^{\prime},
gbgegd\displaystyle g_{b}g_{e}g_{d} =gbgf1gd,\displaystyle=g_{b}^{\prime}g_{f}^{-1}g_{d}^{\prime},
gagegd\displaystyle g_{a}g_{e}g_{d} =gagd,\displaystyle=g_{a}^{\prime}g_{d}^{\prime},
gd1gc\displaystyle g_{d}^{-1}g_{c} =gd1gfgc,\displaystyle={g_{d}^{\prime}}^{-1}g_{f}g_{c}^{\prime},
gbgegc\displaystyle g_{b}g_{e}g_{c} =gbgc,\displaystyle=g_{b}^{\prime}g_{c}^{\prime},
gagb1\displaystyle g_{a}g_{b}^{-1} =gagfgb1.\displaystyle=g_{a}^{\prime}g_{f}{g_{b}^{\prime}}^{-1}.

The solution to those equations is unique up to equivalence and given by the following theorem.

Theorem 2.

For any flip Γ\Gamma of τ\tau, the transformation AAΓA\longrightarrow A^{\Gamma} is such that the GG-graph connection corresponding to AΓ(Γ(τ))A^{\Gamma}(\Gamma(\tau)) is related to A(τ)A(\tau) as depicted in Figure 2, where

Refer to caption
Figure 2. Flip transformation on a graph connection.
ga=ga\displaystyle g_{a}^{\prime}=g_{a} gb=gbge\displaystyle\hskip 30.00005ptg_{b}^{\prime}=g_{b}g_{e}
gc=gc\displaystyle g_{c}^{\prime}=g_{c} gd=gegd\displaystyle\hskip 30.00005ptg_{d}^{\prime}=g_{e}g_{d}
gf=ge.\displaystyle g_{f}=g_{e}.

One can verify the relations explicitly by the choice of gauge transformation hu=cc1h_{u^{\prime}}=c^{\prime}c^{-1} and hv=(a)1ah_{v^{\prime}}=(a^{\prime})^{-1}a. However, another method is the following trick. One can use gauge transformations to make ge=1g_{e}=1 and then shrink that edge, since it does not participate in monodromy. This results in a 4-valent vertex. Expanding the edge makes two trivalent vertices and the edge ff. After that, one makes a gauge transformation to achieve gf=geg_{f}=g_{e}.

We note that this solution works for any group; however, it is asymmetric. Group elements spread in one direction of the graph but not the other. In the next section we will put the coordinates on MG(τ){\rm M}_{G}(\tau), represented as graph connections in the case of G=GL(1|1)G=GL(1|1). We will find the formulas for the flip transformations in those coordinates which remove this spreading effect.

4. Coordinates on the moduli space of GL(1|1)GL(1|1)-graph connections

We fix a surface FF with genus g0g\geq 0 and s1s\geq 1 punctures such that 2g+s2>02g+s-2>0. We also fix a trivalent fatgraph τF\tau\subset F with orientation oo.

Assigning to every edge a group element of GL(1|1)GL(1|1), or alternatively assigning an ordered tuple (a,b;α,β)(a,b;\alpha,\beta) in the parametrization (2.5), we obtain a vector in the coordinate system C~(F,o,τ)\tilde{C}(F,o,\tau) for the space of GL(1|1)GL(1|1)-graph connections without factoring by the gauge group equivalences. Thus the chart C~(F,o,τ)\tilde{C}(F,o,\tau) gives a diffeomeorphism:

(4.1) M~GL(1|1)(τ)+12g12+6s|12g12+6s,\displaystyle\tilde{\rm M}_{GL(1|1)}(\tau)\cong\mathbb{R}^{12g-12+6s|12g-12+6s}_{+},

However, we need to take into account that one can change the orientation on edges of the graph, thus leading to an equivalent coordinate system on M~G(τ)\tilde{\rm M}_{G}(\tau).

Definition 4.1.

We say that two charts C~(F,o,τ)\tilde{C}(F,o,\tau) and C~(F,o¯,τ)\tilde{C}(F,\bar{o},\tau), corresponding to orientations oo, o¯\bar{o}, are equivalent if the assignments agree on the edges where orientation is the same, but are related by the formula (a¯,b¯;α¯,β¯)=(a1,b1;b2α,b2β)(\bar{a},\bar{b};\bar{\alpha},\bar{\beta})=(a^{-1},b^{-1};-b^{2}\alpha,-b^{2}\beta), where the orientation is reversed.

Our goal is now to reformulate the gauge transformations at the vertices of a fatgraph as relations between coordinates in the charts C~(F,o,τ)\tilde{C}(F,o,\tau).

Definition 4.2.

Let cC~(F,o,τ)\vec{c}\in\tilde{C}(F,o,\tau) be a coordinate vector. Suppose e1,e2,e3e_{1},e_{2},e_{3} are edges oriented towards a vertex uu such that, for each i=1,2,3i=1,2,3, eie_{i} has coordinates (ai,bi;αi,βi)(a_{i},b_{i};\alpha_{i},\beta_{i}). Then a vertex rescaling at uu is one of three coordinate changes for the coordinates of each eie_{i}, where cc is positive and even and γ\gamma is odd:

  • (ai,bi;αi,βi)(aic1,bi;αi,βi)(a_{i},b_{i};\alpha_{i},\beta_{i})\mapsto(a_{i}c^{-1},b_{i};\alpha_{i},\beta_{i});

  • (ai,bi;αi,βi)(ai,bic1;c2αi,βi)(a_{i},b_{i};\alpha_{i},\beta_{i})\mapsto(a_{i},b_{i}c^{-1};c^{2}\alpha_{i},\beta_{i});

  • (ai,bi;αi,βi)(ai(112bi2βiγ),bi;αiγ,βi)(a_{i},b_{i};\alpha_{i},\beta_{i})\mapsto(a_{i}(1-\frac{1}{2}b_{i}^{2}\beta_{i}\gamma),b_{i};\alpha_{i}-\gamma,\beta_{i}).

If e1,e2,e3e_{1},e_{2},e_{3} are oriented away from a vertex uu, then there is also a vertex rescaling at uu for odd γ\gamma:

  • (ai,bi;αi,βi)(ai(112bi2αiγ),bi;αi,βi+γ)(a_{i},b_{i};\alpha_{i},\beta_{i})\mapsto(a_{i}(1-\frac{1}{2}b_{i}^{2}\alpha_{i}\gamma),b_{i};\alpha_{i},\beta_{i}+\gamma).

It turns out that the edge reversal and the vertex rescalings define an equivalence relation on the chart C~(F,o,τ)\tilde{C}(F,o,\tau). The vertex rescalings come from the equivalences on GL(1|1)GL(1|1) graph connections provided by the appropriate 1-parameter subgroups. This leads to the following theorem.

Theorem 3.

Let C(F,τ)=C~(F,o,τ)/C(F,\tau)=\tilde{C}(F,o,\tau)/\sim be the quotient by the equivalences provided by edge reversal and vertex rescalings. Then C(F,τ)C(F,\tau) is in bijection with the moduli space MGL(1|1)(τ){\rm M}_{GL(1|1)}(\tau).

Proof.

Indeed, explicitly, the coordinates (a,b;α,β)(a,b;\alpha,\beta) correspond to

a(1+12b2αβ)(1α01)(b100b)(10β1)\displaystyle a(1+\frac{1}{2}b^{2}\alpha\beta)\begin{pmatrix}1&\alpha\\ 0&1\end{pmatrix}\begin{pmatrix}b^{-1}&0\\ 0&b\end{pmatrix}\begin{pmatrix}1&0\\ \beta&1\end{pmatrix}

in GL(1|1)GL(1|1). The inverse of such an element is

a1(1+12b2αβ)(1b2α01)(b00b1)(10b2β1),\displaystyle a^{-1}(1+\frac{1}{2}b^{2}\alpha\beta)\begin{pmatrix}1&-b^{2}\alpha\\ 0&1\end{pmatrix}\begin{pmatrix}b&0\\ 0&b^{-1}\end{pmatrix}\begin{pmatrix}1&0\\ -b^{2}\beta&1\end{pmatrix},

which corresponds to the coordinates (a1,b1;b2α,b2β)(a^{-1},b^{-1};-b^{2}\alpha,-b^{2}\beta). Therefore, the rule for reversing orientations in C~(F,o,τ)\tilde{C}(F,o,\tau) matches the rule for reversing orientations of a graph connection on τ\tau.

Now consider a graph connection on τ\tau. For an edge ee directed towards a vertex uu in τ\tau, there is a gauge transformation gehu1geg_{e}\mapsto h_{u}^{-1}g_{e}, where geg_{e} is a group element on ee, and huh_{u} is a group element associated to the vertex uu. Similarly, if ee is directed away from uu, the gauge transformation is gegehug_{e}\mapsto g_{e}h_{u}. This gauge transformation at uu acts on each edge adjacent to uu. In the case of a trivalent τ\tau, there are three edges e1,e2,e3e_{1},e_{2},e_{3} at a given vertex uu. Then the vertex rescalings correspond to gauge transformations at uu, where huh_{u} is one of the following:

  • hu=ch_{u}=c,

  • hu=(c100c)h_{u}=\begin{pmatrix}c^{-1}&0\\ 0&c\end{pmatrix},

  • hu=(1γ01)h_{u}=\begin{pmatrix}1&\gamma\\ 0&1\end{pmatrix},

  • hu=(10γ1)h_{u}=\begin{pmatrix}1&0\\ \gamma&1\end{pmatrix}.

Again, cc is positive even and γ\gamma is odd. In the first three gauge transformations, we require that e1,e2,e3e_{1},e_{2},e_{3} all point towards uu, and for the fourth, we require that the edges all point away from uu. The claim that the vertex rescalings correspond to gauge transformations follows from routine multiplication in GL(1|1)GL(1|1). For instance, in the third scenario, we have

g(1,1;γ,0)1g(a,b;α,β)=g(a(112b2βγ),b;α+γ,β).\displaystyle g(1,1;\gamma,0)^{-1}g(a,b;\alpha,\beta)=g(a(1-\frac{1}{2}b^{2}\beta\gamma),b;\alpha+\gamma,\beta).

These four gauge elements generate all possible gauge transformations, since each of the four elements together generate GL(1|1)GL(1|1). ∎

It is not hard to see that one can simply constrain the vertex rescalings. To do that let us introduce the following notation: for a given edge with assignment (a,b;α,β)(a,b;\alpha,\beta), which is adjacent to vertex vv, we denote av=aa^{v}=a, bv=bb^{v}=b, αv=α\alpha^{v}=\alpha, βv=β\beta^{v}=\beta, if edge is oriented towards vv, and av=a1a^{v}=a^{-1} bv=b1b^{v}=b^{-1}, αv=α\alpha^{v}=-\alpha, βv=β\beta^{v}=-\beta otherwise. Then, let us give the following definition.

Definition 4.3.

Let cC~(F,o,τ)\vec{c}\in\tilde{C}(F,o,\tau) so that {(ai,bi;αi,βi)}i=1,2,3\{(a_{i},b_{i};\alpha_{i},\beta_{i})\}_{i=1,2,3} are the assignments for edges adjacent to vertex vv. We refer to the following conditions as gauge constraints at vertex vv:

(4.2) i=13aiv=i=13biv=1,i=13αiv=i=13βiv=0\displaystyle\prod^{3}_{i=1}{a}^{v}_{i}=\prod^{3}_{i=1}{b}^{v}_{i}=1,\quad\sum_{i=1}^{3}{\alpha^{v}_{i}}=\sum_{i=1}^{3}{\beta^{v}_{i}}=0

One can see that, for fixed orientation oo, such constraints pick exactly one element from the equivalence classes provided by vertex rescalings. Moreover, we have the following theorem.

Theorem 4.

The coordinate chart C~(F,o,τ)\tilde{C}(F,o,\tau) modulo gauge constraints at all vertices vv of τ\tau provides a real-analytic isomorphism

(4.3) MGL(1|1)(τ)+4g+2s2|4g+2s2,\displaystyle{\rm M}_{GL(1|1)}(\tau)\cong\mathbb{R}_{+}^{4g+2s-2|4g+2s-2},

so that M~GL(1|1)(τ)\tilde{\rm M}_{GL(1|1)}(\tau) is a trivial bundle over MGL(1|1)(τ)M_{GL(1|1)}(\tau) with the fiber +8g10+4s|8g10+4s\mathbb{R}_{+}^{8g-10+4s|8g-10+4s}.

There are benefits of working with the “decorated space” M~GL(1|1)(τ)\tilde{\rm M}_{GL(1|1)}(\tau) instead of MGL(1|1)(F){\rm M}_{GL(1|1)}(F). As we shall see in the next section, the gauge freedom allows us to write a compact formula for a flip transformation in our coordinates, somewhat in the spirit of Penner coordinates on decorated Teichmüller space.

5. Minimal formula for the flip transformations

First we describe the suitable description of the elements of MGL(1|1)(F){M}_{GL(1|1)}(F) in terms of coordinates in the extended chart C~(F,o,τ)\tilde{C}(F,o,\tau).

Namely, we assign coordinates (a,b;α,β)(a,b;\alpha,\beta), corresponding to a GL(1|1)GL(1|1) group element, to an oriented edge as follows: the odd parameters α,β\alpha,\beta are assigned to the terminal and initial points of the edge, while we put the pair of even parameters (a,b)(a,b) between them, as depicted in Figure 3. This ordering goes along with the Gaussian decomposition from Section 2.

Refer to caption
Figure 3. Coordinates on an edge.

Let us now describe the minimal representation of flip transformation which involves change of the minimal number of parameters. We will fix the odd elements at the ends, as well as eliminate the dependence on fermionic coordinates in the mid-edge after the flip, as depicted in Figure 4.

Refer to caption
Figure 4. Minimal form of the flip transformation
Theorem 5.

Let τ\tau^{\prime} represent the flip of τ\tau at an edge ee, with orientations as depicted in Figure 1. Let cC~(F,o,τ)\vec{c}\in\tilde{C}(F,o,\tau) and let c\vec{c^{\prime}} be the corresponding element in C~(F,o,τ)\tilde{C}(F,o^{\prime},\tau^{\prime}) modulo gauge equivalences. Choose coordinates for c\vec{c} such that {(ai,bi;αi,βi)}i=1,,5\{(a_{i},b_{i};\alpha_{i},\beta_{i})\}_{i=1,\dots,5} represent the edges where the flip occurs. Then there is a choice for c\vec{c^{\prime}}, which is described on the Figure 4, where the primed variables are given by the following formulas:

β1=β1+b52β5,β2=β5+b52β2,α3=α3+b52α5,α4=b52α4+α5\displaystyle\beta_{1}^{\prime}=\beta_{1}+b_{5}^{2}\beta_{5},~{}\beta_{2}^{\prime}=\beta_{5}+b_{5}^{-2}\beta_{2},~{}\alpha_{3}^{\prime}=\alpha_{3}+b_{5}^{2}\alpha_{5},\alpha_{4}^{\prime}=b_{5}^{-2}\alpha_{4}+\alpha_{5}
a1=a1(112b12b52α1β5),a3=a3(112b32b52α5β3),a5=a5(112b52α5β5),\displaystyle a_{1}^{\prime}=a_{1}(1-\frac{1}{2}b_{1}^{2}b_{5}^{2}\alpha_{1}\beta_{5}),~{}a_{3}^{\prime}=a_{3}(1-\frac{1}{2}b_{3}^{2}b_{5}^{2}\alpha_{5}\beta_{3}),~{}a_{5}^{\prime}=a_{5}(1-\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),
(5.1) a2=a2a5(112b22b52α2β5)(112b52α5β5),a4=a4a5(112b42b52α5β4)(112b52α5β5).\displaystyle a_{2}^{\prime}=a_{2}a_{5}(1-\frac{1}{2}b_{2}^{2}b_{5}^{2}\alpha_{2}\beta_{5})(1-\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),~{}a_{4}^{\prime}=a_{4}a_{5}(1-\frac{1}{2}b_{4}^{2}b_{5}^{2}\alpha_{5}\beta_{4})(1-\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}).
Proof.

Recall that the product of elements (a,b;α,β)(a,b;\alpha,\beta) and (c,d;γ,δ)(c,d;\gamma,\delta) is (ac(112βγ)(112b2d2αδ),bd;α+b2γ,d2β+δ)(ac(1-\frac{1}{2}\beta\gamma)(1-\frac{1}{2}b^{2}d^{2}\alpha\delta),bd;\alpha+b^{-2}\gamma,d^{-2}\beta+\delta).

Applying this formula to the general solution of the flip (Theorem 2) means that we can first write c\vec{c^{\prime}} with coordinates

(a1,b1;α1,β1),\displaystyle(a_{1},b_{1};\alpha_{1},\beta_{1}),
(a2a5(1+12α5β2)(112b22b52α2β5),b2b5;α2+b22α5,b52β2+β5),\displaystyle(a_{2}a_{5}(1+\frac{1}{2}\alpha_{5}\beta_{2})(1-\frac{1}{2}b_{2}^{2}b_{5}^{2}\alpha_{2}\beta_{5}),b_{2}b_{5};\alpha_{2}+b_{2}^{-2}\alpha_{5},b_{5}^{-2}\beta_{2}+\beta_{5}),
(a3,b3;α3,β3),\displaystyle(a_{3},b_{3};\alpha_{3},\beta_{3}),
(a4a5(1+12α4β5)(112b42b52α5β4),b4b5;α5+b52α4,b42β5+β4),\displaystyle(a_{4}a_{5}(1+\frac{1}{2}\alpha_{4}\beta_{5})(1-\frac{1}{2}b_{4}^{2}b_{5}^{2}\alpha_{5}\beta_{4}),b_{4}b_{5};\alpha_{5}+b_{5}^{-2}\alpha_{4},b_{4}^{-2}\beta_{5}+\beta_{4}),
(a5,b5;α5,β5).\displaystyle(a_{5},b_{5};\alpha_{5},\beta_{5}).

We see that α1\alpha_{1} and β3\beta_{3} are already in place, so we do not want to change these. We reverse the orientations of the second, fourth, and fifth edges so that we may alter the α\alpha term of the second and the β\beta term of the fourth. This gives

(a21a51(112α5β2)(1+12b22b52α2β5),b21b51;b22b52α2b52α5,b22β2b22b52β5),\displaystyle(a_{2}^{-1}a_{5}^{-1}(1-\frac{1}{2}\alpha_{5}\beta_{2})(1+\frac{1}{2}b_{2}^{2}b_{5}^{2}\alpha_{2}\beta_{5}),b_{2}^{-1}b_{5}^{-1};-b_{2}^{2}b_{5}^{2}\alpha_{2}-b_{5}^{2}\alpha_{5},-b_{2}^{2}\beta_{2}-b_{2}^{2}b_{5}^{2}\beta_{5}),
(a41a51(112α4β5)(1+12b42b52α5β4),b41b51;b42α4b42b52α5,b42b52β4b52β5),\displaystyle(a_{4}^{-1}a_{5}^{-1}(1-\frac{1}{2}\alpha_{4}\beta_{5})(1+\frac{1}{2}b_{4}^{2}b_{5}^{2}\alpha_{5}\beta_{4}),b_{4}^{-1}b_{5}^{-1};-b_{4}^{2}\alpha_{4}-b_{4}^{2}b_{5}^{2}\alpha_{5},-b_{4}^{2}b_{5}^{2}\beta_{4}-b_{5}^{2}\beta_{5}),
(a51,b51;b52α5,b52β5).\displaystyle(a_{5}^{-1},b_{5}^{-1};-b_{5}^{2}\alpha_{5},-b_{5}^{2}\beta_{5}).

Now that the second, third, and fifth edges are pointed towards a vertex, we apply a move which adds b52α5b_{5}^{2}\alpha_{5} to the first fermionic coordinate of these edges.

(a21a51(1+12b22b52α2β5)(1+12b52α5β5)),b21b51;b22b52α2,b22β2b22b52β5),\displaystyle(a_{2}^{-1}a_{5}^{-1}(1+\frac{1}{2}b_{2}^{2}b_{5}^{2}\alpha_{2}\beta_{5})(1+\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5})),b_{2}^{-1}b_{5}^{-1};-b_{2}^{2}b_{5}^{2}\alpha_{2},-b_{2}^{2}\beta_{2}-b_{2}^{2}b_{5}^{2}\beta_{5}),
(a3(112b32b52α5β3),b3;α3+b52α5,β3),\displaystyle(a_{3}(1-\frac{1}{2}b_{3}^{2}b_{5}^{2}\alpha_{5}\beta_{3}),b_{3};\alpha_{3}+b_{5}^{2}\alpha_{5},\beta_{3}),
(a51(1+12b52α5β5),b51;0,b52β5).\displaystyle(a_{5}^{-1}(1+\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),b_{5}^{-1};0,-b_{5}^{2}\beta_{5}).

Now, we can invert the second edge again to get it to its original orientation; its formula is saved for the end of the proof.

The first, fourth, and fifth edges are all pointed away from a common vertex. We can then apply a move that adds b52β5b_{5}^{2}\beta_{5} to the second fermionic coordinate of these edges. This gives

(a1(112b12b52α1β5),b1;α1,β1+b52β5),\displaystyle(a_{1}(1-\frac{1}{2}b_{1}^{2}b_{5}^{2}\alpha_{1}\beta_{5}),b_{1};\alpha_{1},\beta_{1}+b_{5}^{2}\beta_{5}),
(a41a51(1+12b42b52α5β4)(1+12b52α5β5),b41b51;b42α4b42b52α5,b42b52β4),\displaystyle(a_{4}^{-1}a_{5}^{-1}(1+\frac{1}{2}b_{4}^{2}b_{5}^{2}\alpha_{5}\beta_{4})(1+\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),b_{4}^{-1}b_{5}^{-1};-b_{4}^{2}\alpha_{4}-b_{4}^{2}b_{5}^{2}\alpha_{5},-b_{4}^{2}b_{5}^{2}\beta_{4}),
(a51(1+12b52α5β5),b51;0,0).\displaystyle(a_{5}^{-1}(1+\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),b_{5}^{-1};0,0).

We invert the fourth and fifth edges to match the original choice of orientation, giving the final coordinates for c\vec{c^{\prime}}:

(a1(112b12b52α1β5),b1;α1,β1+b52β5),\displaystyle(a_{1}(1-\frac{1}{2}b_{1}^{2}b_{5}^{2}\alpha_{1}\beta_{5}),b_{1};\alpha_{1},\beta_{1}+b_{5}^{2}\beta_{5}),
(a2a5(112b22b52α2β5)(112b52α5β5),b2b5;α2,β5+b52β2),\displaystyle(a_{2}a_{5}(1-\frac{1}{2}b_{2}^{2}b_{5}^{2}\alpha_{2}\beta_{5})(1-\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),b_{2}b_{5};\alpha_{2},\beta_{5}+b_{5}^{-2}\beta_{2}),
(a3(112b32b52α5β3),b3;α3+b52α5,β3),\displaystyle(a_{3}(1-\frac{1}{2}b_{3}^{2}b_{5}^{2}\alpha_{5}\beta_{3}),b_{3};\alpha_{3}+b_{5}^{2}\alpha_{5},\beta_{3}),
(a4a5(112b42b52α5β4)(112b52α5β5),b4b5;b52α4+α5,β4),\displaystyle(a_{4}a_{5}(1-\frac{1}{2}b_{4}^{2}b_{5}^{2}\alpha_{5}\beta_{4})(1-\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),b_{4}b_{5};b_{5}^{-2}\alpha_{4}+\alpha_{5},\beta_{4}),
(a5(112b52α5β5),b5;0,0).\displaystyle(a_{5}(1-\frac{1}{2}b_{5}^{2}\alpha_{5}\beta_{5}),b_{5};0,0).

These coordinates have the desired fixed fermionic ends. ∎

6. Poisson bracket

In this section, we discuss in detail the Poisson structure that can be endowed on the space of graph connections M~G(τ)\tilde{\rm M}_{G}(\tau). It descends onto MG(τ){\rm M}_{G}(\tau), so that the isomorphism with the space of flat connections, modulo the action of the gauge group, is a Poisson manifold isomorphism. We recall this construction from [fock1998poisson].

As before, let us fix a surface FF with genus g0g\geq 0 and s1s\geq 1 punctures, such that 2g+s2>02g+s-2>0, and a trivalent fatgraph τF\tau\subset F. M~G(τ)\tilde{\rm M}_{G}(\tau) is the space of graph connections on τ\tau. Let GτG^{\tau} be the gauge group; that is, a direct product of the group GG with itself, one copy for each vertex in τ\tau.

We have seen that GτG^{\tau} acts on M~G(τ)\tilde{\rm M}_{G}(\tau), which gives the equivalence of graph connections. In order for the action to be a Poisson map, GτG^{\tau} needs a Poisson-Lie structure. Then GG itself needs a Poisson-Lie structure. This can be attained by choosing an rr-matrix, which involves the Lie algebra 𝔤\mathfrak{g} of GG [Chari:1994pz]. This is an object satisfying the so-called classical Yang-Baxter equation (CYBE), namely:

(6.1) [r12,r13]+[r12,r23]+[r13,r23]=0\displaystyle[r^{12},r^{13}]+[r^{12},r^{23}]+[r^{13},r^{23}]=0

We also assume that the symmetric part of the rr-matrix is the Casimir element Ω\Omega, i.e.

(6.2) 12(r12+r21)=Ω=eiei,\displaystyle\frac{1}{2}(r^{12}+r^{21})=\Omega=\sum e_{i}\otimes e_{i},

based on the invariant bilinear form on 𝔤\mathfrak{g} (so that {ei}i=1,dim𝔤\{e_{i}\}_{i=1,\dots\rm{dim}\mathfrak{g}} is an orthonormal basis). For example, consider simple Lie algebras with the standard triangular decomposition 𝔤=𝔫+𝔥𝔫\mathfrak{g}=\mathfrak{n}_{+}\oplus\mathfrak{h}\oplus\mathfrak{n}_{-}. Here, 𝔥\mathfrak{h} is the Cartan part and 𝔫±\mathfrak{n}_{\pm} are spanned by positive and negative root elements. The basic classical rr-matrix of 𝔤\mathfrak{g} is given by r=Ω𝔥+2α>0eαeαr=\Omega_{\mathfrak{h}}+2\sum_{\alpha>0}e_{\alpha}\otimes e_{-\alpha}, where Ω𝔥\Omega_{\mathfrak{h}} is the Casimir element restricted to 𝔥\mathfrak{h}, and (eα,eα)=1(e_{\alpha},e_{-\alpha})=1 for all α\alpha with respect to an invariant form. For 𝔤𝔩(1|1)\mathfrak{gl}(1|1), one can construct a solution of CYBE in a similar fashion: the only difference is that the nondegenerate bilinear form is defined by supertrace in the defining representation, not by the Killing form, which is degenerate. This way, the solution to CYBE with the condition (6.2) is given by (see also [Gizem04]):

r=EN+NE2Ψ+Ψ,\displaystyle r=E\otimes N+N\otimes E-2\Psi^{+}\otimes\Psi^{-},
(6.3) Ω=EN+NEΨ+Ψ+ΨΨ+\displaystyle\Omega=E\otimes N+N\otimes E-\Psi^{+}\otimes\Psi^{-}+\Psi^{-}\otimes\Psi^{+}

This natually provides a Poisson-Lie structure on GL(1|1)GL(1|1).

Let us describe the construction in [fock1998poisson] of a Poisson bracket on M~G(τ)\tilde{\rm M}_{G}(\tau), adapted to our needs (simple supergroups). Let us adopt the following notation for graph connections. To a half-edge adjacent to a given vertex vv, we associate an element gvg_{v}. To the opposite half edge we associate the element gv1g_{v}^{-1}. In this way, we do not need to refer to a specific orientation of the edge. We then have a map

(6.4) GG×G,g(g,g1)\displaystyle G\longrightarrow G\times G,g\mapsto(g,g^{-1})

associated to every edge of τ\tau, giving a map M~G(τ)G2(6g6+3s)\tilde{{\rm M}}_{G}(\tau)\longrightarrow G^{2(6g-6+3s)}.

Given an orthonormal basis {ei}i=1,dim𝔤\{e_{i}\}_{i=1,\dots\rm{dim}\mathfrak{g}}, one can define a tangent vector at Xi(g)=Li(g)Ri(g1)X_{i}(g)=L_{i}(g)-R_{i}(g^{-1}) on the image of map (6.4), where Li,RiL_{i},R_{i} are left- and right-invariant vector fields associated to eie_{i}.

The final ingredient needed to write the Poisson structure is to fix a linear order at each vertex of τ\tau, not just a cyclic one. This is called a ciliation in [fock1998poisson]. Since τ\tau is trivalent, this means each vertex has edges labelled 1,2,31,2,3 in a cyclic manner. For a vertex of τ\tau, we write the labels of the edges as elements of the vertex.

This allows us to write the following Poisson bivector which is a sum of bivectors over the set of vertices V(τ)V(\tau):

(6.5) P=vV(τ)Pv,Pv=α<βi,jrij(v)XiαXjβ+12αi,jrij(v)XiαXjα\displaystyle P=\sum_{v\in V(\tau)}P_{v},\quad P_{v}=\sum_{\alpha<\beta}\sum_{i,j}r^{ij}(v)X_{i}^{\alpha}\wedge X_{j}^{\beta}+\frac{1}{2}\sum_{\alpha}\sum_{i,j}r^{ij}(v)X_{i}^{\alpha}\wedge X_{j}^{\alpha}

where α,β\alpha,\beta are edges adjacent to vv ordered according to ciliation, and r(v)r(v) is an rr-matrix associated to vertex vv.

Then one has the following theorem.

Theorem 6.

[fock1998poisson] The bivector PP defines a Poisson structure on M~G(τ)\tilde{{\rm M}}_{G}(\tau), and the group GτG^{\tau} acts on M~G(τ)\tilde{{\rm M}}_{G}(\tau) in a Poisson way.

If each r(v)r(v) has the same symmetric part, then denoting the antisymmetric part as ra(v)=12(r(v)12r(v)21)r_{a}(v)=\frac{1}{2}(r(v)^{12}-r(v)^{21}), we obtain:

Pv=Pv,h+Pv,t,\displaystyle P_{v}=P_{v,h}+P_{v,t},
(6.6) Pv,h=i,jraij(v)XivXjv,Pv,t=α,β;i(v,α,β;i)XiαXiβ\displaystyle P_{v,h}=\sum_{i,j}r_{a}^{ij}(v)X^{v}_{i}\otimes X^{v}_{j},\quad P_{v,t}=\sum_{\alpha,\beta;i}(v,\alpha,\beta;i)X_{i}^{\alpha}\otimes X_{i}^{\beta}

Here Xiv=αXiαX^{v}_{i}=\sum_{\alpha}X_{i}^{\alpha} is the generator of gauge transformations at vv and is tangent to the orbits of GτG^{\tau}, and

(6.7) (v,α,β;i)={1α>β0α=β(1)p(i)α<β,(v,\alpha,\beta;i)=\begin{cases}1\qquad\qquad\quad\alpha>\beta\\ 0\qquad\qquad\quad\alpha=\beta\\ -(-1)^{p(i)}\qquad\alpha<\beta,\end{cases}

where p(i)p(i) stands for the parity of generator eie_{i}. From here one can see that the choice of antisymmetric part of the rr-matrix is irrelevant when considering the Poisson bracket on the quotient MG(τ){\rm M}_{G}(\tau).

In our case ra=Ψ+ΨΨΨ+r_{a}=-\Psi_{+}\otimes\Psi_{-}-\Psi_{-}\otimes\Psi_{+}. In Pv,tP_{v,t}, we sum over the ii index to get terms corresponding to the symmetric part (Casimir element). Once we sum over ii, there is no need for 6.7, so we use the simplified (v,α,β)=sign(αβ)(v,\alpha,\beta)=\textrm{sign}(\alpha-\beta). Then we have

Pv,hGL(1|1)=XΨ+vXΨvXΨvXΨ+v,\displaystyle P_{v,h}^{GL(1|1)}=-X^{v}_{\Psi_{+}}\otimes X^{v}_{\Psi_{-}}-X^{v}_{\Psi_{-}}\otimes X^{v}_{\Psi_{+}},
(6.8) Pv,tGL(1|1)=α,β;i(v,α,β)(XEαXNβ+XEαXNβXΨ+αXΨβ+XΨαXΨ+β).\displaystyle P^{GL(1|1)}_{v,t}=\sum_{\alpha,\beta;i}(v,\alpha,\beta)(X_{E}^{\alpha}\otimes X_{N}^{\beta}+X_{E}^{\alpha}\otimes X_{N}^{\beta}-X_{\Psi^{+}}^{\alpha}\otimes X_{\Psi^{-}}^{\beta}+X_{\Psi^{-}}^{\alpha}\otimes X_{\Psi^{+}}^{\beta}).

Given our system of coordinates C~(F,o,τ)\tilde{C}(F,o,\tau) for M~G(τ)\tilde{{\rm M}}_{G}(\tau), one can easily write down the presentation of the generators XiαX^{\alpha}_{i}. Let us assume that an edge μ\mu, with coordinates (a,b;α,β)(a,b;\alpha,\beta), is oriented towards a vertex vv. Then

XEμ=aa,XNμ=12bb+αα,\displaystyle X^{\mu}_{E}=a\partial_{a},\quad X^{\mu}_{N}=-\frac{1}{2}b\partial_{b}+\alpha\partial_{\alpha},
(6.9) XΨ+μ=12ab2βa+α,XΨμ=12αaa+b2β.\displaystyle X^{\mu}_{\Psi^{+}}=-\frac{1}{2}ab^{2}\beta\partial_{a}+\partial_{\alpha},\quad X^{\mu}_{\Psi^{-}}=-\frac{1}{2}\alpha a\partial_{a}+b^{-2}\partial_{\beta}.

We introduce a change of coordinates by rearranging the one-parameter subgroups, as follows:

a(1+12b2αβ)(1α01)(b100b)(10β1)=a^(1α^01)(10β^1)(b^100b^),\displaystyle a(1+\frac{1}{2}b^{2}\alpha\beta)\begin{pmatrix}1&\alpha\\ 0&1\end{pmatrix}\begin{pmatrix}b^{-1}&0\\ 0&b\end{pmatrix}\begin{pmatrix}1&0\\ \beta&1\end{pmatrix}=\hat{a}\begin{pmatrix}1&\hat{\alpha}\\ 0&1\end{pmatrix}\begin{pmatrix}1&0\\ \hat{\beta}&1\end{pmatrix}\begin{pmatrix}\hat{b}^{-1}&0\\ 0&\hat{b}\end{pmatrix},
(6.10) g(a,b;α,β)=g(a^(112α^β^),b^;α^,b^2β).\displaystyle g(a,b;\alpha,\beta)=g(\hat{a}(1-\frac{1}{2}\hat{\alpha}\hat{\beta}),\hat{b};\hat{\alpha},\hat{b}^{-2}\beta).

This change of coordinates simplifies the vector fields:

XEμ=a^a^,XNμ=12b^b^+α^α^β^β^,\displaystyle X^{\mu}_{E}=\hat{a}\partial_{\hat{a}},\quad X^{\mu}_{N}=-\frac{1}{2}\hat{b}\partial_{\hat{b}}+\hat{\alpha}\partial_{\hat{\alpha}}-\hat{\beta}\partial_{\hat{\beta}},
(6.11) XΨ+μ=α^,XΨμ=α^a^a^+β^.\displaystyle X^{\mu}_{\Psi^{+}}=\partial_{\hat{\alpha}},\quad X^{\mu}_{\Psi^{-}}=-\hat{\alpha}\hat{a}\partial_{\hat{a}}+\partial_{\hat{\beta}}.

These coordinates, given that all three edges are pointed at the vertex vv, form a Darboux-like chart for the bivector PvP_{v}.

References