1 Introduction

Consider the formal pseudo-differential operator L as

$$\begin{aligned} L=\frac{\partial }{\partial x}+u_1\left( \frac{\partial }{\partial x}\right) ^{-1}+u_2\left( \frac{\partial }{\partial x}\right) ^{-2}+ \dots, \end{aligned}$$

where \(u_1, u_2, \dots,\) are infinite number of variables which are dependent variables of systems of partial differential equations. The classical Kadomtsev–Petviashvili (KP) hierarchy is the infinite system of equations

$$\begin{aligned} \partial _{x_m}L=[(L^m)_+,L], \ \ \ \ m=1,2,3, \dots, \end{aligned}$$

where \((L^m)_+\) denotes the projection of the pseudo-differential operator \(L^m\) into the space of differential operators, and \(x_1, x_2, \dots\) denote independent variables.

The non-commutative KP hierarchy usually has been defined (see e.g. [11]) using non-commutativity of the coordinates

$$\begin{aligned}{}[x_k, x_l]=i\theta ^{kl}, \end{aligned}$$

in Pseudo-differential operators, for the independent variables \(x_k.\) Here \(\theta ^{kl}\) are real constants and called the non-commutative parameters. Hence, the non-commutative KP hierarchy is described by

$$\begin{aligned} \partial _{x_m}L=[(L^m)_+,L]_{\star }, \end{aligned}$$

where the star-product \(\star\) is represented by the Moyal product.

In this note, we are going to show that one can obtain naturally the non-commutative structure on KP equations from some algebraic data. To this end we explain the relation between Lie bialgebroids and quantum groupoids from work of Xu in [23], i.e. “ deformation of a bialgebroid ” (known as a quantum groupoid) induces a Lie bialgebroid as a classical limit. For this, during Sects. 2, 3, and 4 we review several algebraic structures such as bialgebroids, Lie bialgebras and Lie bialgebroids with their examples. Gathering together these concepts helps to understand the relation between them (especially for a non-familiar reader to these subjects) and provides the ingredients which we use later in the last sections. Then, in Sect. 5 we explain deformation of bialgebroids based on [23]. In Sect. 6 we use the result of Xu, to associate a Lie bialgebroid to the algebra of formal Pseudo-differential operators. In Sect. 7 we present non-commutative KP equations based on quantum groupoid associated to the algebra of formal Pseudo-differential operators. Solving KP equations, in particular showing existence and uniqueness of solutions of KP equations has been an interesting subject of study. For instance, in a recent work in [18] (also see references there in), the Cauchy problem of the KP hierarchy has been solved and formulated in several non-standard cases such as non-commutative case described by Moyal product. Now, with the algebraic construction in this paper, one can expect that the non-commutative KP equations can be solved more generally.

1.1 Preliminaries

Consider an associative algebra as a triple \((A,\mu,\eta )\) where A is a vector space over a field k,  the map \(\mu :A\otimes A \rightarrow A\) denotes the multiplication, and \(\eta :k\rightarrow A\) denotes the unit of the algebra. The maps \(\mu\) and \(\eta\) are linear and satisfy the following associative properties.

  • \(\mu \circ (\mu \otimes id)=\mu \circ (id\otimes \mu )\)

  • \(\mu \circ (\eta \otimes id)=\mu \circ (id\otimes \eta )=id.\)

If A is commutative then we also have \(\mu \circ \ \tau _{A,A}=\mu\) where \(\tau _{A,A}\) is the flip switching, that is \(\tau _{A,A}(a\otimes a')=a'\otimes a.\) A morphism between algebras \((A,\mu,\eta )\) and \((A',{\mu },{\eta }')\) is a linear map \(f:A\rightarrow A'\) such that \({\mu }'\circ (f\otimes f)=f\circ \mu\) and \(f\circ \eta ={\eta }'\).

The notion of coalgebra is dual to the notion of algebra in the following sense.

Definition 1

A co-associative coalgebra is a triple \((C,\Delta,\varepsilon )\) where C is a vector space over a field k and the maps \(\Delta :C\rightarrow C\otimes C\) and \(\varepsilon :C\rightarrow k\) are linear satisfying the co-associative properties

  • \((id\otimes \Delta )\circ \Delta =(\Delta \otimes id)\circ \Delta\)

  • \((id\otimes \varepsilon )\circ \Delta =(\varepsilon \otimes id)\circ \Delta =id.\)

If C is co-commutative then we also have \(\tau _{C,C}\circ \Delta =\Delta.\) A morphism between coalgebras \((C,\Delta,\varepsilon )\) and \((C',{\Delta }',{\varepsilon }')\) is a linear map \(f:C\rightarrow C'\) such that \((f\otimes f)\circ \Delta ={\Delta }'\circ f\) and \(\varepsilon ={\varepsilon }'\circ f.\)

Definition 2

Suppose H is a vector space over the field k equipped with an algebra structure \((H,\mu,\eta )\) and a coalgebra structure \((H,\Delta,\varepsilon )\) then \((H,\mu,\eta,\Delta,\varepsilon )\) is a bialgebra verifying the compatibility conditions between these two structures i.e.,

  1. 1.

    The maps \(\mu\) and \(\eta\) are morphisms of coalgebras, or equivalently

  2. 2.

    the maps \(\Delta\) and \(\varepsilon\) are morphisms of algebras.

Given an algebra \((A,\mu,\eta )\) and a coalgebra \((C,\Delta ,\varepsilon )\) one can define a bilinear map called the convolution on the vector space Hom(CA),  denoted by \(\star :Hom(C,A)\rightarrow Hom(C,A),\) of linear maps from C to A. If fg are such linear maps then the convolution \(f\star g\) is the composition of the following maps

$$\begin{aligned} C{\mathop {\rightarrow }\limits ^{\Delta }}C\otimes C{\mathop {\rightarrow }\limits ^{f\otimes g}}A\otimes A{\mathop {\rightarrow }\limits ^{\mu }}A. \end{aligned}$$

Let \((H,\mu,\eta,\Delta,\varepsilon )\) be a bialgebra. An endomorphism S of H is called an antipode for the bialgebra H if \(S\star id_H=id_H\star S =\eta \circ \varepsilon\).

Remark 1

A bialgebra does not necessarily have an antipode, but if it does, it has only one. Because if \(S'\) is another antipode then we have

$$\begin{aligned} S=S\star (\eta \varepsilon )=S\star (id_H\star S')=(S\star id_H)\star S'=(\eta \varepsilon )\star S'=S'. \end{aligned}$$

Definition 3

A Hopf algebra is a bialgebra with antipode, and is denoted by \((H,\mu,\eta, \Delta,\varepsilon,S).\) A morphism of Hopf algebras is a morphism between the underlying bialgebras commuting with the antipode.

Example 1

Any algebra A is the quotient of a free algebra \(K\{X\}\) where K is a field and X is a generating set. Since, it suffices to take any generating set X and \(A=K\{X\}/I\) where I is a two sided ideal of \(K\{X\}\).

Let \(X=\{x_1,...,x_n\}\) and let I be the two sided ideal of \(K\{x_1,...,x_n\}\) generated by elements of the form \(x_ix_j-x_jx_i\) where ij run over all integers from 1 to n. The quotient algebra \(K\{x_1,...,x_n\}/I\) is isomorphic to the polynomial algebra \(K[x_1,...,x_n]\) in n variables with coefficient in the ground field K.

The structure \((K[x],\mu,\eta,\Delta, \varepsilon,S)\) is a Hopf algebra where the algebra K[x] is called affine line and \(\mu\) is the multiplication and \(\eta\) is the unit of K[x]. The co-multiplication, co-unit and antipode are defined by

$$\begin{aligned}&\Delta :K[x]\rightarrow K[x',x'']; \ \ \ \ \Delta (x)=x'+x'' \\&\varepsilon :K[x]\rightarrow K; \ \ \ \ \varepsilon (x)=0 \\&S:K[x]\rightarrow K[x]; \ \ \ \ S(x)=-x. \end{aligned}$$

If \(X=G\) is a finite group with commutative product, then K[G] is a bialgebra with co-multiplication

$$\begin{aligned}&\Delta :k[G]\rightarrow k[G]\otimes k[G]\\&\Delta (f)(x,y)=f(xy) \end{aligned}$$

and co-unit \(\varepsilon (x)=1\). Also, if \(X=G\) is a finite group with convolution product \((f\star g)(x)=\sum _{x_1,x_2}f(x_1)g(x_2)\), where \(x_1x_2=x,\) then K[G] is a bialgebra with \(\Delta (x)=x\otimes x\), and \(\varepsilon (x)=1\).

Example 2

Let I be the two sided ideal of \(K\{x,y\}\) generated by \(xy-yx\), the quotient algebra \(K\{x,y\}/I\) is isomorphic to the polynomial algebra K[xy] with two variables xy with coefficients in the ground field K. The algebra K[xy] is called affine plane. In the affine plane if we define the two sided ideal \(I_q=<xy-qyx>\) then we obtain a new algebra \(K\{x,y\}/I_q=K_q[x,y]\) which is clearly non-commutative and is called quantum plane. In this algebra instead of the relation \(xy=yx\) in commutative case we have \(xy=qyx\). This is a bialgebra equipped with

$$\begin{aligned}&\Delta (x)=x\otimes x,\ \ \Delta (y)=y\otimes 1+x\otimes y \\&\varepsilon (x)=1,\ \ \varepsilon (y)=0. \end{aligned}$$

Example 3

Let V be a vector space. Define \(T^0(V)=k, T^1(V)=V, T^n(V)=V^{\otimes n}\) (The tensor product of n copies of V) if \(n>1\).

The canonical isomorphisms \(T^n(V)\otimes T^m(V)\cong T^{n+m}(V)\) induce an associative product

$$\begin{aligned} (x_1\otimes...\otimes x_n)(x_{n+1}\otimes... \otimes x_{n+m})=x_1\otimes...\otimes x_n\otimes x_{n+1}\otimes...\otimes x_{n+m} \end{aligned}$$

on the vector space \(T(V)=\bigoplus _{n\ge 0}T^n(V)\). The unit for this product is the image of the unit element 1 in \(k=T^0(V)\). The vector space T(V) equipped with this algebra structure, is called the tensor algebra of V. If V is a vector space, the symmetric algebra S(V) is the quotient \(S(V)=T(V)/I(V)\) of the tensor algebra T(V) by the two sided ideal I(V) generated by all elements \(xy-yx\) where xy run over V.

To any Lie algebra L we can assign an (associative) algebra U(L) called envelo** algebra of L,  with a morphism of Lie algebras \(i_L:L\rightarrow Lie(U(L)).\) More precisely, let I(L) be the two sided ideal of the tensor algebra T(L) generated by all elements of the form \(xy-yx-[x,y]\) where \(x,y\in L\) and [., .] is the Lie bracket. We define \(U(L):=T(L)/I(L)\). If L is a Lie algebra then the universal envelo** algebra of L,  denoted by U(L) is a Hopf algebra equipped with co-multiplication

$$\begin{aligned}&\Delta :U(L)\rightarrow U(L)\otimes U(L)\\&\Delta (X)=X\otimes 1+1\otimes X, \end{aligned}$$

and with co-unit and antipode

$$\begin{aligned}&\varepsilon :U(L)\rightarrow k, \ \ \varepsilon (X)=0, \\&S:U(L)\rightarrow U(L), \ \ S(X)=-X, \end{aligned}$$

for \(X\in L.\) As an example, for the Lie algebra \({\mathfrak {sl}}(2)\) of traceless \(2\times 2\) matrices, the envelo** algebra \(U({\mathfrak {sl}}(2))\) is a Hopf algebra, see [12]. Moreover, we can construct a Hopf algebra \(U_q({\mathfrak {sl}}(2))\) which is a one-parameter deformation of the envelo** algebra of the Lie algebra \({\mathfrak {sl}}(2)\).

2 Bialgebroids

Definition 4

A bialgebroid is a pair of algebras A and R together with an algebra homomorphism \(\alpha :R\rightarrow A\), an algebra anti-homomorphism \(\beta :R\rightarrow A\) such that the images of \(\alpha\) and \(\beta\) commute in A i.e. \(\alpha (r_1)\beta (r_2)=\beta (r_2)\alpha (r_1)\), for each \(r_1,r_2\in R\). By these two morphisms we can assign an (RR)-bimodule structure on A in natural way by \(r.a=\alpha (r)a\) and \(a.r=\beta (r)a\), \(r\in R,\ a\in A\). With this bimodule structure on A we can also consider \(A\otimes _R A\) as an (RR)-bimodule which is given as a left module by \(r.(a_1\otimes a_2)=\alpha (r)a_1\otimes a_2\) and as a right module by \((a_1\otimes a_2).r=a_1\otimes \beta (r)a_2\), \(r\in R,\ \ a_1,a_2\in A\).

We define the co-product \(\Delta :A\rightarrow A\otimes _R A\) and the co-unit \(\varepsilon :A\rightarrow R\) as (RR)-bimodule maps satisfying the co-associativity axiom for \(\Delta\) and co-unit as below

  • \((id_A\otimes _R\Delta )\circ \Delta =(\Delta \otimes _R id_A)\circ \Delta\)

  • \(\varepsilon (1_A)=1_R\), \(\varepsilon \beta =\varepsilon \alpha =id_R\), and \((id_A\otimes \varepsilon )\circ \Delta =(\varepsilon \otimes id_A)\circ \Delta =id_A.\)

The co-product \(\Delta\) and the algebra structure on A should be compatible in the sense that the kernel of the following map

$$\begin{aligned} \Phi :A\otimes A\otimes A\rightarrow & {} A\otimes _RA\\ a_1\otimes a_2\otimes a_3\mapsto & {} (\Delta a_1)(a_1\otimes a_2) \end{aligned}$$

is a left ideal of \(A\otimes A^{op}\otimes A^{op}\). Here we are using the fact that \(A\otimes A\) acts on \(A\otimes _RA\) from the right by right multiplications. Also we need the co-unit map is compatible with the algebra structure on A in the sense that the kernel of \(\varepsilon\) is a left ideal of A. We denote a bialgebroid by \((A,R,\alpha,\beta,\Delta,\varepsilon )\) where A is called total algebra, R is called the base algebra, and the maps \(\alpha ,\beta\) are called source map and target map respectively.

Recall that a groupoid G over a set \(G_0\) is a set together with a pair of structure maps \(G{\mathop {\rightarrow }\limits ^{s}}G_0\) and \(G{\mathop {\rightarrow }\limits ^{t}}G_0,\) where s is called source and t is called target. Moreover, G is equipped with a product, identity, and an inversion, (for more details see for example [4]). In fact, we can think of an element \(g\in G\) as an arrow (morphism) from \(x=s(g)\) to \(y=t(g)\) in \(G_0.\)

Example 4

If we consider a group G as a groupoid \(\mathcal G\) over a set \(X=\{x\}\) then the source map and target map are as constant maps \(s,t:G\rightarrow \{x\},\ s(g)=t(g)=x\) for each \(g\in G\). The arrows or elements of G are as \(G=s^{-1}(x)\rightarrow t^{-1}(x)=G.\) We can consider them as \(G{\mathop {\rightarrow }\limits ^{g}}G\) such that \(h\mapsto gh\) for each \(h,g\in G\). Also the unit for our groupoid is \(\varepsilon :X\rightarrow G\) with \(\varepsilon (x)=e\) (e is the identity element of the group G). The inverse map is \(i:G\rightarrow G\) with \(g\mapsto g^{-1}\). The multiplication map

$$\begin{aligned} m:\mathcal G_{(2)}=\mathcal G\times _{\{x\}}\mathcal G\rightarrow \mathcal G \end{aligned}$$

where \(\mathcal G_{(2)}=\{(g,h)\in G\times G \mid s(g)=t(h)\}=G\times G\), is in fact the multiplication of the group G.

Now, we consider the space of smooth functions on G,  i.e. \(C^\infty (G)\) as the total algebra and R the real numbers as the base algebra. We define \(\alpha =s^*\) and \(\beta =t^*\) such that

$$\begin{aligned} s^*=t^*:R=C^\infty (\{x\})\rightarrow C^\infty (G). \end{aligned}$$

It is clear that \((s^*f)(g)=f\circ s(g)=f(x)\in R.\) In other words \(s^*\) assigns to each \(r\in R\) the constant function \(r\in C^\infty (G).\) So \(s^*(r)=r\) similarly \(t^*(r)=r\) for \(r\in R.\)

Moreover,

$$\begin{aligned} \varepsilon ^*:C^\infty (G)\rightarrow C^\infty (\{x\})=R \end{aligned}$$

with \((\varepsilon ^*f)(x)=f\circ \varepsilon (x)=f(e).\) It is clear that \((\varepsilon ^*id)=id(e)=1\in R\) hence \(\varepsilon (1_{C^\infty (G)})=1\). Therefore \(\varepsilon ^*\) can be considered as the co-unit. Finally,

$$\begin{aligned} m^*:C^\infty (G)\rightarrow C^\infty (G)\otimes C^\infty (G) \end{aligned}$$

defines the co-product, with \(\Delta (f)=f\otimes f.\) Clearly \(C^\infty (G)\) is an (RR)-bimodule. It is easy to check that \(\varepsilon ,\Delta\) are (RR)-bimodule maps and they are compatible with the algebra structure on \(C^\infty (G)\). Therefore, we constructed a bialgebroid \((C^\infty (G),R,s^*,t^*,\Delta ,\varepsilon )\) by the groupoid \(\mathcal G\) which is the group G.

Recall that a smooth map \(f:M\rightarrow N\) between smooth manifolds M and N is a local diffeomerphism (or étale map) if \((df)_x\) is an isomorphism for any \(x\in M.\) An étale groupid is a groupoid with the source map s as a local diffeomorphism. For more details and examples see [19].

Example 5

Let \(G_1{\mathop {\rightarrow }\limits ^{{\mathop {\rightarrow }\limits ^{s}}}}G_0\) be an étale groupoid. We define

$$\begin{aligned} \alpha :=s^*,\ \beta :=t^*:C^\infty (G_0)\rightarrow C^\infty (G_1) \end{aligned}$$

and we consider the co-unit as \(\varepsilon ^*:C^\infty (G_1)\rightarrow C^\infty (G_0)\). The multiplication of groupoid is

$$\begin{aligned} m:G_2=G_1\times _{G_0}G_1\rightarrow G_1 \end{aligned}$$

where \(G_2=\{(g,h)\in G_1 | s(g)=t(h)\}\). So, the pull-back of this multiplication, i.e.

$$\begin{aligned} \Delta =m^*:C^\infty (G_1)\rightarrow C^\infty (G_2)=C^\infty (G_1\times _{G_0}G_1)\cong C^\infty (G_1)\otimes _{C^\infty (G_0)}C^\infty (G_1) \end{aligned}$$

defines a co-product. Hence, \(C^\infty (G_1)\) over \(C^\infty (G_0)\) is a bialgebroid.

Example 6

Again consider the étale groupoid \(G{\mathop {\rightarrow }\limits ^{{\mathop {\rightarrow }\limits ^{s}}}}G_0.\) The Connes algebra \(C_c^\infty (G)\) of (smooth) complex (or real) functions with compact support on G (see [2, 6, 7]), together with the base algebra \((C_c^\infty (G))_0\) defines a bialgebroid, where \((C_c^\infty (G))_0\) is the subalgebra of \(C_c^\infty (G)\) of functions with support in \(G_0\subset G.\) This subalgebra is commutative and may be identified with the commutative algebra \(C_c^\infty (G_0).\)

Considering the inclusion \(i:C_c^\infty (G_0)\hookrightarrow C_c^\infty (G)\) we take \(\alpha =\beta =i\). The product is the convolution

$$\begin{aligned} (aa')(g'')=\sum _{gg'=g''}a(g)a'(g'),\ \ \forall a,a'\in C_c^\infty (G),\ g''\in G, \end{aligned}$$

where the sum is over all possible decompositions of \(g''\in G\). We take the co-unit

$$\begin{aligned} \varepsilon :C_c^\infty (G)\rightarrow C_c^\infty (G_0) \end{aligned}$$

by

$$\begin{aligned} \varepsilon (a)(x)=\sum _{s(g)=x}a(g),\ \ \forall a\in C_c^\infty (G),\ x\in G_0. \end{aligned}$$

This sum is over all the elements g of G which satisfy \(s(g)=x\). The coproduct \(\Delta :C_c^\infty (G)\rightarrow C_c^\infty (G)\otimes _{C_c^\infty (G_0)}C_c^\infty (G)\) is defined as follows: Let \(d:G\rightarrow G^s\times _{G_0}^sG\) be the diagonal open embedding, i.e. \(d(g)=(g,g)\). This map gives the inclusion \(C_c^\infty (G)\rightarrow C_c^\infty (G\times _{G_0}G)\). We define \(\Delta\) as the composition of this inclusion with the inverse of the isomorphism

$$\begin{aligned}&\Omega :C_c^\infty (G)\otimes _{C_c^\infty (G_0)}C_c^\infty (G)\rightarrow C_c^\infty (G\times _{G_0}G) \\&\Omega (a\otimes a')(g,g')=a(g)a'(g'). \end{aligned}$$

(For the proof that \(\Omega\) is an isomorphism see [20]).

If \(a\in C_c^\infty (G)\) has the support in an open subset U of G which is so small that \(s|_U\) is injective, then \(\Delta (a)=a\otimes \xi =\xi \otimes a\) where \(\xi\) is any smooth function with compact support in U which constantly equals 1 on the support of a. The functions \(a\in C_c^\infty (G)\) which satisfy the above condition generate the linear space \(C_c^\infty (G)\).

Example 7

1) If A is any algebra then there is a bialgebroid structure on \(H=A\otimes A^{op}\) over A with

  1. 1.

    \(\Delta :H\rightarrow H\otimes _A H\) , \(a\otimes b\mapsto (a\otimes 1)\otimes (1\otimes b)\),

  2. 2.

    \(\varepsilon :H\rightarrow A\), \(a\otimes b\mapsto ab\),

  3. 3.

    \(\alpha :A\rightarrow H\), \(a\mapsto a\otimes 1\), \(\beta :A\rightarrow H\), \(a\mapsto 1\otimes a\).

2) For any finite dimensional algebra over k,  the algebra of k-linear maps from A to itself, i.e. \(H=End_k(A),\) has a bialgebroid structure over A,  see [15].

Definition 5

A morphism between two bialgebroids \((A,R,\alpha ,\beta ,\Delta ,\varepsilon )\) and \((A',R',{\alpha }',{\beta }',{\Delta }',{\varepsilon }')\) consists of an algebra morphism \(T:A\rightarrow A'\) and an algebra morphism \(t:R\rightarrow R'\) which commute with all the structure maps.

As for Hopf algebras we expect that a Hopf algebroid is a bialgebroid with antipode.

Definition 6

The antipode in a bialgebroid \((A,R,\alpha ,\beta ,\Delta ,\varepsilon )\) is a bijective map \(\tau :A\rightarrow A\) which has the following properties

  1. 1.

    \(\tau\) is an algebra anti-isomorphism for A.

  2. 2.

    \(\tau \beta =\alpha\)

  3. 3.

    \(\mu (\tau \otimes id)\Delta =\beta \varepsilon \tau :A\rightarrow A\), (it is the same as the definition of antipode in section 1.)

  4. 4.

    There exists a linear map \(\gamma :A {\otimes }_R A \rightarrow A\otimes A\) with:

    1. (a)

      \(\gamma\) is a section for the natural projection \(p:A\otimes A\rightarrow A{\otimes }_R A\)

    2. (b)

      \(\mu (id\otimes \tau )\gamma \Delta =\alpha \varepsilon :A\rightarrow A\).

Proposition 1

The maps \(\tau :A\rightarrow A\) and \(id:R\rightarrow R\) define a bialgebroid morphism.

Proof

In the above definition, condition (1) means that for all \(x,y\in A\), \(\tau (xy)=\tau (y)\tau (x)\) and condition (2) means that \(\tau (\beta (1_R))=\alpha (1_R)\) or \(\tau (1)=1.\) On the other hand \((\tau \otimes \tau )\Delta =\Delta ^{op}\tau\), and \(\varepsilon \circ \tau =\varepsilon.\) Moreover by condition (2) \(\tau\) commutes with structure maps. Clearly id is a morphism of algebra and it commutes with the structure maps. \(\square\)

Proposition 2

Bialgebras and bialgebroids are equivalent over a field k as a base algebra.

Proof

For any bialgebra \((A,\mu ,\eta ,\Delta ,\varepsilon )\) we can consider \(R=k\) , where k is the ground field for the algebra A,  and \(\alpha ,\beta :k\rightarrow A\) by \(\alpha (r)=r\) and \(\beta (r)=r\), \((r\in k)\). Hence, \((A,k,\alpha ,\beta ,\Delta ,\varepsilon )\) is a bialgebroid because for \(r,s\in k\) we have \(\alpha (rs)=rs=\alpha (r)\alpha (s)\) so \(\alpha\) is an algebra homomorphism, also \(\beta (rs)=rs=sr=\beta (s)\beta (r)\) so \(\beta\) is an anti-homomorphism, and finally \(\alpha (r)\beta (s)=rs=sr=\beta (s)\alpha (r)\) so the images of \(\alpha\) and \(\beta\) commute. Conversely, if R is the field k the definition of a bialgebroid over k is reduced to a bialgebra over k, see [15]. \(\square\)

Example 8

1. If L is a Lie algebra then the universal envelo** algebra of L,  that is U(L) can be considered as a bialgebroid by the above Proposition 2. Consider Hopf algebras \(SL_q(2)\) and \(U_q({\mathfrak {sl}}(2))\) as mentioned before (there is a duality between SL(2) and \(U({\mathfrak {sl}}(2)),\) see [12]). By the above proposition they are also examples of bialgebroids.

Let \((A,R,\alpha ,\beta ,\Delta ,\tau )\) be a Hopf algebroid over the field k of characteristic zero. And let \(End_k R\) be the algebra of linear endomorphisms of R over k. When R acts on \(End_k R\) from the left by left multiplication and acts from the right by right multiplication, we have \(End_k R\) as an (RR)-bimodule. Assume that R is a left A-module and moreover the representation \(\rho :A\rightarrow End_k R\) is an (RR)-bimodule map. We define \(\phi _{\alpha },\phi _{\beta }:(A{\otimes }_R A)\otimes R\rightarrow A\) by

$$\begin{aligned} \phi _{\alpha }(x\otimes _Ry\otimes a)= & {} \rho (x)(a).y\\ \phi _{\beta }(x\otimes _Ry\otimes a)= & {} x.\rho (y)(a) \end{aligned}$$

where \(x,y\in A, a\in R\).

Definition 7

Given a bialgebroid \((A,R,\alpha ,\beta ,\Delta ,\varepsilon )\), an anchor map is a representation \(\rho :A\rightarrow End_k R\) which is an (RR)-bimodule map such that

  1. 1.

    \(\phi _{\alpha }(\Delta x\otimes a)=x\alpha (a)\) and \(\phi _{\beta }(\Delta x\otimes a)=x\beta (a)\), \(\forall x\in A, a\in R\),

  2. 2.

    \(x(1_R)=\varepsilon (x)\), \(\forall x\in A\).

For a bialgebra \(R=k\) and \(End_kR\cong R,\) clearly we can take the co-unit as the anchor. In general, for a bialgebroid, the existence of an anchor map is stronger than the existence of a co-unit.

Example 9

Let P be a smooth manifold and D be the algebra of differential operators on P. Let R be the algebra of smooth functions on P. Then \((D,R,\alpha ,\beta ,\Delta ,\varepsilon )\) is a bialgebroid where \(\alpha =\beta\) is the embedding \(R\rightarrow D\) and \(\Delta :D\rightarrow D\otimes _R D\) is defined by

$$\begin{aligned} \Delta (d)(f,g)=d(fg), \forall d\in D \end{aligned}$$

and \(f,g\in R.\) Also, the usual action of differential operators on \(C^\infty (P)\) defines an anchor \(\mu :D\rightarrow End_kR\). The co-unit \(\varepsilon :D\rightarrow R\) is the natural projection to its 0-order part of a differential operator.

Below we see a construction of new bialgebroid from a given bialgebroid, which is called the twist construction. First we need the following proposition.

Proposition 3

(Xu, Proposition 4.2 in [23]) Let \((H, R, \alpha , \beta , m,\Delta , \varepsilon )\) be a Hopf algebroid with anchor \(\mu.\) Let \(F\in H\otimes _R H\) and define \(\alpha _F,\beta _F:R\rightarrow H\) by

$$\begin{aligned} \alpha _F(a)=\phi _{\alpha }(F\otimes a),\ \beta _F(a)=\phi _\beta (F\otimes a),\ \forall a\in R, \end{aligned}$$

and for any \(a,b\in R\), set \(a*_Fb=\alpha _F(a)(b)\). Now assume that F satisfies:

$$(\Delta \otimes _R id )FF^{12}=(id\otimes _R\Delta )FF^{23} \ \ in\ \ H\otimes _RH\otimes _RH,$$
(1)

and

$$(\varepsilon \otimes _Rid)F=1_H, (id\otimes _R\varepsilon )F=1_H$$
(2)

where \(F^{12}=F\otimes 1\in (H\otimes _R H)\otimes H),\ F^{23}=1\otimes F\in H\otimes (H\otimes _RH)\). Then

  • \((R,*_F)\) is an associative algebra, and \(1_R*_Fa=a*_F1_R=a,\ \forall a\in R\).

  • \(\alpha _F:R_F\rightarrow H\) is an algebra homomorphism, and \(\beta _F:R_F\rightarrow H\) is an algebra anti-homomorphism. Here \(R_F\) stands for the algebra \((R,*_F)\).

  • \((\alpha _Fa)(\beta _Fb)=(\beta _Fb)(\alpha _Fa),\ \forall a,b\in R\).

Definition 8

Let \(M_1\) and \(M_2\) be given left H-modules. If the linear map

$$\begin{aligned} F^{\#}:M_1\otimes _{R_F}M_2\rightarrow & {} M_1\otimes _RM_2\\ (m_1\otimes _{R_F}m_2)\rightarrow & {} F.(m_1\otimes m_2) \end{aligned}$$

(for \(m_1\in M_1\) and \(m_2\in M_2\)) is an isomorphism of vector spaces, we say that F is invertible. An element \(F\in H\otimes _RH\) is called a twistor if it is invertible and satisfies Eq. (1) and (2) in above proposition.

Theorem 1

(Xu, Theorem 4.14 in [23]) Assume that \((H,R,\alpha ,\beta ,\Delta ,\varepsilon )\) is a bialgebroid with anchor \(\mu\), and \(F\in H\otimes _R H\) is a twistor. Then \((H,R_F,\alpha _F,\beta _F,\Delta _F,\varepsilon )\) is a bialgebroid which still admits \(\mu\) as an anchor.

Example 10

([23]) Consider a smooth manifold P with the algebra D of differential operators on it, and \(R=C^{\infty }(P).\) We denote the space of formal power series in \(\hbar\) with coefficients in D by \(D[[\hbar ]].\) The Hopf algebroid structure on D naturally can be extended to a Hopf algebroid structure on \(D[[\hbar ]]\) over the base algebra \(R[[\hbar ]],\) and it admits a natural anchor map. One can check that

$$\begin{aligned} F=1\otimes _R1+\hbar B_1+\dots \in D\otimes _R D[[\hbar ]] (\cong D[[\hbar ]]\otimes _{R[[\hbar ]]} D[[\hbar ]]) \end{aligned}$$

as a formal power series of bidifferential operatos is a twistor if and only if the multiplication on \(R[[\hbar ]]\) defined by

$$\begin{aligned} f*_{\hbar }g=F(f,g), \ \ \forall f,g\in R[[\hbar ]] \end{aligned}$$

is associative with identity being the constant function 1,  in other words, \(*_{\hbar }\) is a star product on P. Hence, the bracket

$$\begin{aligned} \{f,g\}=B_1(f,g)-B_1(g,f), \ \ \forall f,g \in C^{\infty }(P), \end{aligned}$$

defines a Poisson structure on P and \(f*_{\hbar }g=F(f,g)\) is a deformation quantization of this Poisson structure.

Now, set \(D_{\hbar }=D[[\hbar ]]\) equipped with the usual multiplication, \(R_{\hbar }=R[[\hbar ]]\) equipped with the \(*\)-product defined above. Consider \(\alpha _{\hbar }: R_{\hbar }\rightarrow D_{\hbar }\) and \(\beta _{\hbar }:R_{\hbar }\rightarrow D_{\hbar }\) given by

$$\begin{aligned} \alpha _{\hbar }(f)g=f*_{\hbar }g, \ \ \ \beta _{\hbar }(f)g=g*_{\hbar }f, \ \ \ \forall f,g \in R. \end{aligned}$$

Moreover, consider the co-product \(\Delta _{\hbar }:D_{\hbar }\rightarrow D_{\hbar }\otimes _{R_{\hbar }} D_{\hbar }\) defined by \(\Delta _{\hbar }=F^{-1}\Delta F,\) and co-unit as the projection \(D_{\hbar }\rightarrow R_{\hbar }.\) By Theorem 1 we obtain twisted Hopf algebroid \((D_{\hbar }, R_{\hbar }, \alpha _{\hbar }, \beta _{\hbar }, m, \Delta _{\hbar }, \varepsilon ).\) This twisted Hopf algebroid is called the quantum groupoid associated to the star product \(*_{\hbar }.\)

3 Lie Bialgebras

Definition 9

Let G be a Lie group and \(\rho :G\rightarrow GL(V)\) be the representation of G on the vector space V. Also, let \(d\rho :\mathfrak g\rightarrow End(V)\) be the infinitesimal representation of the group representation. then

  • \(\varphi :G\rightarrow V\) is a 1-cocycle on G if \(\varphi (gh)=\varphi (g)+\rho (g)\varphi (h),\)

  • \(\phi :\mathfrak g\rightarrow V\) is a 1-cocycle on \(\mathfrak g\) if \(\phi ([u,v])=d\rho (u)\phi (v)-d\rho (v)\phi (u).\)

Note that if \(\varphi\) is 1-cocycle on G then \(\phi =d_e\varphi\) is 1-cocycle on \(\mathfrak g\). Also, if G is simply connected then the 1-cocycle on \(\mathfrak g\) can be integrated to 1-cocycle on G.

Definition 10

Let \(\mathfrak g\) be a Lie algebra. A Lie bialgebra structure on \(\mathfrak g\) is a linear map \(\delta _{\mathfrak g}:\mathfrak g\rightarrow \mathfrak g\otimes \mathfrak g\), called co-commutator, such that

  1. 1.

    \(\delta _{\mathfrak g}^*:\mathfrak g^*\otimes \mathfrak g^*\rightarrow \mathfrak g^*\) is a Lie bracket on \(\mathfrak g^*\)

  2. 2.

    \(\delta _{\mathfrak g}\) is a 1-cocycle of \(\mathfrak g\) with values in \(\mathfrak g\otimes \mathfrak g.\)

A homomorphism of Lie bialgebras \(\varphi :\mathfrak g\rightarrow \mathfrak h\) is a homomorphism of Lie algebras such that \((\varphi \otimes \varphi )\circ \delta _{\mathfrak g}=\delta _{\mathfrak h}\circ \varphi.\)

Definition 11

Let G be a Lie group. A Poisson Lie group is \((G,\Pi )\) where \(\Pi\) is a Poisson structure such that \(m:G\times G\rightarrow G\) is Poisson map. In this case \(\Pi\) is called multiplicative.

Remark 2

A Poisson structure \(\Pi\) is multiplicative if and only if \(\Pi _{gh}=(l_g)_*\Pi _h+(r_h)_*)\Pi _g\), \(g,h\in G\), where \(l_g\) and \(r_h\) are left and right translations, see [5].

Example 11

Any Lie group can be seen as a Poisson Lie group with \(\Pi =0\).

Example 12

Let \(\mathfrak g\) be a finite dimensional Lie algebra with Lie bracket \([.,.]_{\mathfrak g}\). Each \(\mu \in \mathfrak g^*\) is a function on \(\mathfrak g\). Given functions \(f,g\in C^\infty (\mathfrak g^*)\) the new function \(\{f,g\}\in C^\infty (\mathfrak g^*)\) evaluated at \(\mu\) is \(\{f,g\}(\mu )=\mu ([Df(\mu ),Dg(\mu )]_{\mathfrak g})\). Then \(\mathfrak g^*\) as an abelian Lie group (with addition) is a Poisson Lie group with this Poisson structure (Poisson bracket).

The following proposition gives an example of Lie bialgebras.

Proposition 4

Let \((G,\Pi )\) be a Poisson Lie group with the Lie algebra \(\mathfrak g\). Also, let \(F:\mathfrak g\rightarrow \mathfrak g\wedge \mathfrak g\) be 1-cocycle, and \(F^*\) Lie bracket then \((\mathfrak g,\mathfrak g^*)\) is a Lie bialgebra.

Proof

Let \((G,\Pi )\) be a Poisson Lie group, and \(F=\Pi ^{(1)}\) the linear part of \(\Pi\) at the point e. The map \(\Pi ^{(1)}:\mathfrak g\rightarrow \mathfrak g\wedge \mathfrak g\) is linear Poisson structure on \(\mathfrak g\) where \(\mathfrak g=T_eG\). Then \(F^*:\mathfrak g^*\wedge \mathfrak g^*\rightarrow \mathfrak g^*\) is a Lie algebra structure on \(\mathfrak g^*\). In fact, for \(\Pi :G\rightarrow \bigwedge ^2TG\), if we consider \(\tilde{\Pi }:G\rightarrow \bigwedge ^2\mathfrak g\) for which \(g\mapsto (r_g)_*\Pi _g\), then We can write \(\Pi ^{(1)}=d_e\tilde{\Pi }\). Note that, as we mentioned in the previous remark, \(\Pi\) is multiplicative if and only if \(\tilde{\Pi }\) satisfies \(\tilde{\Pi }(gh)=\tilde{\Pi }(g)+Ad_g\tilde{\Pi }(h).\) By above explanation, F satisfies \(F([u,v])=ad_uF(v)-ad_vF(u).\) So F is 1-cocycle on \(\mathfrak g\) with values in \(\bigwedge ^2\mathfrak g\). That is if \((G,\Pi )\) is a Poisson Lie group then \(F=d_e\tilde{\pi }:\mathfrak g\rightarrow \mathfrak g\wedge \mathfrak g\) such that

  1. 1.

    \(F^*:\mathfrak g^*\wedge \mathfrak g^*\rightarrow \mathfrak g^*\) is Lie bracket. (equivalently, \(\Pi\) is Poisson.)

  2. 2.

    F is a 1-cocycle. (equivalently, \(\Pi\) is multiplicative.)

\(\square\)

Lemma 1

Let \(\Lambda\) be a bivector in the Lie algebra, i.e. \(\Lambda \in \bigwedge ^2\mathfrak g\) and \(\Pi _g=(l_g)_*\Lambda -(r_g)_*\Lambda\). Then \(\Pi\) is multiplicative.

Proof

We have

$$\begin{aligned} (l_g)_*\Pi _h+(r_h)_*\Pi _g= & {} (l_g)_*((l_h)_*\Lambda -(r_h)_*\Lambda )+(r_h)_*((l_g)_*\Lambda -(r_g)_*\Lambda )\\= & {} (l_{gh})_*\Lambda -(l_g)_*(r_h)_*\Lambda +(r_h)_*(l_g)_*\Lambda -(r_{gh})_*\Lambda \\= & {} \Pi _{gh}. \end{aligned}$$

Hence, by Remark 2 in above, \(\Pi\) is multiplicative. \(\square\)

Proposition 5

\(\Pi\) is Poisson structure if and only if \([\Lambda ,\Lambda ]\in \bigwedge ^3\mathfrak g\) is Ad-invariant.

Proof

By the above lemma we write \(\Pi =\Lambda ^l-\Lambda ^r\). Moreover, a bivector field \(\Pi\) is a Poisson bivector field if and only if \([\Pi ,\Pi ]=0\). So, we have

$$\begin{aligned}{}[\Pi ,\Pi ]=0\Leftrightarrow & {} [\Lambda ^l-\Lambda ^r,\Lambda ^l-\Lambda ^r]=[\Lambda ^l,\Lambda ^l]+[\Lambda ^r,\Lambda ^r]=0\\\Leftrightarrow & {} [\Lambda ,\Lambda ]^l=[\Lambda ,\Lambda ]^r\\\Leftrightarrow & {} Ad_g[\Lambda ,\Lambda ]=[\Lambda ,\Lambda ]. \end{aligned}$$

For more details see [5] or [22]. \(\square\)

Definition 12

If \(\Pi =\Lambda ^l-\Lambda ^r\) then we call \((G,\Pi )\) a coboundary Poisson-Lie group.

If \(\Lambda\) is such that \([\Lambda ,\Lambda ]\) is Ad-invariant then we call it r-matrix. In the special case, \([\Lambda ,\Lambda ]=0\) is called the classical Yang-Baxter equation and \(\Lambda\) is called triangular r-matrix.

4 Lie Bialgebroids

Definition 13

A Lie algebroid over a manifold P is a vector bundle A over the manifold P together with a Lie algebra structure \([.,.]_A\) on the space \(\Gamma (A)\) of smooth sections of A,  and a bundle map \(\rho :A\rightarrow TP\) (called the anchor), such that

  • The induced map \(\Gamma (\rho ):\Gamma (A)\rightarrow \chi (P)\) is a Lie algebra homomorphism i.e., \(\rho ([X,Y])=[\rho (X),\rho (Y)],\ (X,Y\in \Gamma (A))\) and

  • For any \(f\in C^\infty (P)\) and for any smooth sections X and Y of A the Leibniz identity holds i.e., \([X,fY]_A=f[X,Y]_A+(\rho (X).f)Y\).

Remark 3

  1. 1)

    The map \(\Gamma (\rho )\) may be denoted simply by \(\rho\) and also called the anchor.

  2. 2)

    For every \(X\in \Gamma (A)\), we define A-Lie derivative operations on both \(\Gamma (A)\) and \(C^\infty (P)\) by \(L_XY=[X,Y]_A\), and \(L_Xf=\rho (X).f.\) Then the Leibniz identity is as a derivation

    $$\begin{aligned} L_X(fY)=f(L_XY)+(L_Xf)Y. \end{aligned}$$

Example 13

Let P be a Poisson manifold with Poisson Tensor \(\Pi\). Then \(T^*P\) carries a natural Lie algebroid structure, called the cotangent bundle Lie algebroid of the Poisson manifold P. The anchor map \(\Pi ^{\#}:T^*P\rightarrow TP\) is defined by

$$\begin{aligned} \Pi ^{\#}_p:T_p^*P\rightarrow T_pP:\Pi ^{\#}(\xi )(\eta )=\Pi (\xi ,\eta ) \end{aligned}$$

for each \(\xi ,\eta \in T^*_pP\) and the Lie bracket of 1-forms \(\alpha\) and \(\beta\) is given by

$$\begin{aligned}{}[\alpha ,\beta ]=L_{\Pi ^{\#}(\alpha )}\beta -L_{\Pi ^{\#}(\beta )}\alpha -d(\Pi (\alpha ,\beta )). \end{aligned}$$

Definition 14

Let \((A,\rho ,[.,.]_A)\) be a Lie algebroid over the manifold P,  and \(\bigwedge ^*A^*\) be the exterior algebra of its dual \(A^*\). Sections of \(\bigwedge ^*A^*\) are called A-differential forms on P, or A-forms on P.

If \(\omega \in \Gamma (\bigwedge ^kA^*)\), we say that \(\omega\) is homogeneous, and its degree is \(|\omega |=k\), and we call it as an A-k-form.

Definition 15

There is a differential operator which takes an A-k-form \(\omega\) to an A-\((k+1)\)-form \(d_A\omega\) as below

$$\begin{aligned} d_A\omega (v_1,...,v_{k+1})= & {} \sum _i(-1)^{i+1}\rho (v_i).\omega (v_1,...,\hat{v_i},...,v_{k+1})\\ & \quad + \sum _{i<j}(-1)^{i+j}\omega ([v_i,v_j]_A,v_1,...,\hat{v_i},...,\hat{v_j},...,v_{k+1}) \end{aligned}$$

where \(v_1,...,v_{k+1}\in \Gamma (A)\) are A-vector fields.

Remark 4

([4]) The Lie algebroid axioms for A implies that:

  1. 1.

    \(d_A\) is \(C^\infty (X)\)-multilinear,

  2. 2.

    \(d^2_A=0\),

  3. 3.

    \(d_A(\omega _1\wedge \omega _2)=d_A\omega _1\wedge \omega _2+(-1)^{|\omega _1|}\omega _1\wedge d_A\omega _2\).

The triple \((\Gamma \bigwedge ^*A^*,\wedge ,d_A)\) forms a differential graded algebra, the same as the usual algebra of differential forms.

Proposition 6

There is a one-to-one correspondence between Lie algebroid structures on A and differential operators on \(\Gamma (\bigwedge ^*A^*)\) satisfying properties 1-3 in above.

Proof

The anchor map is obtained from \(d_A\) on functions by

$$\begin{aligned} \rho (v).f=(d_Af)(v),\ v\in \Gamma (A),\ f\in C^\infty (P). \end{aligned}$$

The lie bracket \([.,.]_A\) is determined by

$$\begin{aligned} i_{[v,w]_A}\omega= & {} \rho (v).\omega (w)-\rho (w).\omega (v)-d_A\omega (v,w)\\= & {} i_v(d_A(i_w\omega ))-i_w(d_A(i_v\omega ))-i_{(v\wedge w)}d_A\omega , \end{aligned}$$

where \((v,w\in \Gamma (A),\ \omega \in \Gamma (A^*))\). \(\square\)

Definition 16

The exterior differential algebra \(\Gamma (\bigwedge ^*A^*),\wedge ,d_A)\) associated to a Lie algebroid \((A,\rho ,[.,.]_A)\) determines de Rham cohomology groups, called Lie algebroid cohomology of A or A-cohomology.

Example 14

Let \(A=\mathfrak g\) be a Lie algebra i.e., a Lie algebroid over a one-point space, the cohomology of the differential complex

$$\begin{aligned} \mathbb {R}\longrightarrow \mathfrak g^*\longrightarrow \mathfrak g^*\wedge \mathfrak g^*\longrightarrow... \end{aligned}$$

is the standard Lie algebra cohomology \((\bigwedge ^*\mathfrak g^*,\wedge ,d\mathfrak g)\).

Example 15

Let \(A=TP\) be a tangent bundle of a manifold P,  the cohomology computed by \((\Gamma (\bigwedge ^*A^*),\wedge ,d_A)=(\Omega ^*(P),\wedge ,d_{de Rham})\) is the usual de Rham cohomology.

Definition 17

Lie algebroid multivector fields or A-multivector fields are sections of the exterior algebra \(\bigwedge ^*A\) of a Lie algebroid \((A,\rho ,[.,.]_A)\). If \(v\in \Gamma (\bigwedge ^kA)\), then v is called homogeneous with degree \(|v|=k\).

Remark 5

([4])

  1. 1)

    The extension of \([.,.]_A\) to arbitrary A-multivector fields by setting it on homogeneous A-multivector fields vw is

    $$\begin{aligned} i_{[v,w]_A}\omega =(-1)^{(|v|-1)(|w|-1)}i_v(d_A(i_w\omega ))-i_w(d_A(i_v\omega ))-(-1)^{|v|-1}i_{(v\wedge w)}d_A\omega \end{aligned}$$
  2. 2)

    The bracket \([.,.]_A\) on A-multivector fields has the following properties.

  1. 1.

    \([.,.]_A\) allows us to extend the A-Lie derivative operation defined for A-vector fields in definition of Lie algebroids to arbitrary elements of \(v,w\in \Gamma (\bigwedge ^*A),\) i.e. \(L_vw:=[v,w]_A\).

  2. 2.

    \([.,.]_A\) is a graded Lie algebra structure i.e.

    $$\begin{aligned} {[}v,w]_A=-(-1)^{(|v|-1)(|w|-1)}[w,v]_A \end{aligned}$$
  3. 3.

    \([.,.]_A\) satisfies a super-Jacobi identity:

    $$\begin{aligned} {[}v,[w,y]_A]_A+(-1)^{(|y|-1)(|v|+|w|)}[y,[v,w]_A]_A+(-1)^{(|v|-1)(|w|+|y|)}[w,[y,v]_A]_A=0 \end{aligned}$$
  4. 4.

    \([.,.]_A\) satisfies a super-Leibniz identity:

    $$\begin{aligned} {[}v,w\wedge y]_A=[v,w]_A\wedge y+(-1)^{(|v|-1)|w|}w\wedge [v,y]_A. \end{aligned}$$

Definition 18

The triple \((\Gamma (\bigwedge ^*A),\wedge ,[.,.]_A)\) is called the Gerstenhaber algebra of the Lie algebroid \((A,\rho ,[.,.]_A)\), or just the A-Gerstenhaber algebra. We will refer to the bracket \([.,.]_A\) on \(\Gamma (\bigwedge ^*A)\) as the A-Gerstenhaber bracket.

In general,

Definition 19

A Gerstenhaber algebra \((\mathfrak a,\wedge ,[.,.])\) is a graded vector space \(\mathfrak a=\mathfrak a_0\oplus \mathfrak a_1\oplus...\) together with a super commutative associative multiplication of degree 0, \(\mathfrak a_i\wedge \mathfrak a_j\subseteq \mathfrak a_{i+j}\) and a super-Lie algebra structure of degree \(-1\), \([\mathfrak a_i,\mathfrak a_j]\subseteq \mathfrak a_{i+j-1}\), satisfying the super-Leibniz identity

$$\begin{aligned}{}[a,b\wedge c]=[a,b]\wedge c+(-1)^{(|a|-1)|b|}b\wedge [a,c]. \end{aligned}$$

Remark 6

From a Lie algebroid structure on A,  i.e. \((A,\rho ,[.,.]_A)\), we obtain a differential algebra structure on \(\Gamma (\bigwedge ^*A^*)\), i.e. \((\Gamma (\bigwedge ^*A^*),\wedge ,d_A)\), and from that we get a Gerstenhaber algebra structure on \(\Gamma (\bigwedge ^*A)\), i.e. \((\Gamma (\bigwedge ^*A),\wedge ,[.,.]_A)\).

Definition 20

For the tangent bundle Lie algebroid \((A,\rho ,[.,.]_A)=(TP,id,[.,.])\), \(d_A\) is the de Rham differential and A-Gerstenhaber bracket is usually called the Schouten-Nijenhuis bracket on multivector fields.

A bivector field \(\Pi \in \Gamma (\bigwedge ^2TP)\) is called a Poisson bivector field if and only if \([\Pi ,\Pi ]=0\). This is equivalent to \(d^2_{\Pi }=0\) for the differential operator \(d_{\Pi }:=[\Pi ,.]\).

The notion of Poisson structure naturally generalizes to arbitrary Lie algebroid as follows:

Let \((A,\rho ,[.,.]_A)\) be a Lie algebroid over P. An element \(\Pi \in \Gamma (\bigwedge ^2A)\) is called an A-Poisson bivector field when \([\Pi ,\Pi ]_A=0\), where \([.,.]_A\) is the A-Gerstenhaber bracket.

Example 16

Consider \(A=\mathfrak g\) as a Lie algebra, and a \(\mathfrak g\)-Poisson bivector field \(\Pi \in \mathfrak g\wedge \mathfrak g\) to a left invariant Poisson structure on the underlying Lie group G. The equation \([\Pi ,\Pi ]_{\mathfrak g}=0\) is called the classical Yang-Baxter equation.

Remark 7

([4]) The push-forward \(\rho _*\Pi\) of an A-Poisson bivector field \(\Pi\) by the anchor \(\rho :\Gamma (\bigwedge ^2A)\rightarrow \Gamma (\bigwedge ^2TP)\) defines an ordinary Poisson structure on the manifold P. By the Jacobi identity, an arbitrary (not necessarily Poisson) element \(\Theta \in \Gamma (\bigwedge ^2A)\) satisfies

$$\begin{aligned} d^2_{\Theta }+\left[\frac{1}{2}[\Theta ,\Theta ]_A,.\right]_A=0 \end{aligned}$$

which is similar to the equation for a flat connection.

In the first section we saw how one defines the universal envelo** algebra for a Lie algebra. Now we want to see the definition for Lie algebroids.

Definition 21

If \((A,\rho ,[.,.]_A)\) is a Lie algebroid over the manifold P,  then the \(C^\infty (P)\)-module \(C^\infty (P)\oplus \Gamma (A)\) is a Lie algebra over \(\mathbb R\) with the Lie bracket

$$\begin{aligned} {[}f+X,g+Y]=(\rho (X).g-\rho (Y).f)+[X,Y]. \end{aligned}$$

Considering the universal envelo** algebra of this Lie algebra as \(U=U(C^\infty (P)\oplus \Gamma (A))\), we denote \(f'\) and \(X'\) as the canonical images of \(f\in C^\infty (P)\) and \(X\in \Gamma (A)\) in U. If I is the two-sided ideal of U generated by all elements of the form \((fg)'-f'g'\) and \((fX)'-f'X'\) then the universal envelo** algebra of the Lie algebroid A is defined by \(U(A)=U/I\).

Theorem 2

The universal envelo** algebra U(A) of a Lie algebroid A admits a co-commutative bialgebroid structure.

Proof

Let \(R=C^\infty (P)\), and \(\alpha =\beta :R\rightarrow U(A)\) be the natural embedding. For the co-product define

$$\begin{aligned}&\Delta (f)=f\otimes _R1,\ \ \forall f\in R \\&\Delta (X)=X\otimes _R1+1\otimes _RX,\ \ \forall X\in \Gamma (A). \end{aligned}$$

This formula extends to a co-product \(\Delta :U(A)\rightarrow U(A)\otimes _RU(A)\) by the compatibility condition.

The co-unit map is the projection \(\varepsilon :U(A)\rightarrow R\). Also, the map \(\mu :U(A)\rightarrow End_kR\) defined by

$$\begin{aligned} (\mu x)(f)=(\rho x)(f),\ \ \forall x\in UA,\ f\in R \end{aligned}$$

is the anchor for our bialgebroid where \(\rho : U(A)\rightarrow TP\) is the algebra homomorphism extending the anchor of the Lie algebroid A. Hence, \((U(A),R,\alpha ,\beta ,m,\Delta ,\varepsilon )\) is a co-commutative bialgebroid with anchor \(\mu.\) \(\square\)

The notion of Lie bialgebroids is a natural generalization of Lie bialgebras.

Definition 22

A Lie bialgebroid is a dual pair \((A,A^*)\) of vector bundles equipped with Lie algebroid structures such that the differential d on \(\bigoplus _k\Gamma (\bigwedge ^kA^*)\) is defined by

$$\begin{aligned} d:\Gamma \left( \bigwedge ^kA^*\right) \rightarrow \Gamma \left( \bigwedge ^{k+1}A^*\right) \end{aligned}$$

and the differential \(d_*\) on \(\bigoplus _k\Gamma (\bigwedge ^k(A^*)^*)\cong \bigoplus _k\Gamma (\bigwedge ^kA)\) is defined by

$$\begin{aligned} d_*:\Gamma \left( \bigwedge ^kA\right) \rightarrow \Gamma \left( \bigwedge ^{k+1}A\right) \end{aligned}$$

coming from the structure on \(A^*\) is a derivation of the Schouten bracket on \(\bigoplus _k\Gamma (\bigwedge ^kA)\), equivalently, \(d_*\) is a derivation for sections of A i.e.,

$$\begin{aligned} d_*[X,Y]=[d_*X,Y]+[X,d_*Y],\ \ \forall X,Y\in \Gamma (A). \end{aligned}$$

In other words, \((\bigoplus _k\Gamma (\bigwedge ^kA),\wedge ,[.,.],d_*)\) is a differential Gerstenhaber algebra.

Example 17

For a Poisson manifold \((P,\Pi ),\) the dual pair \((TP,T^*P)\) where TP is the standard tangent bundle and \(T^*P\) is the cotangent Lie algebroid, together with

$$\begin{aligned} d_{\Pi }[X,Y]=[d_{\Pi }X,Y]+[X,d_{\Pi }Y] \end{aligned}$$

is a Lie bialgebroid. As before we saw, \(d_{\Pi }=[\Pi ,.]\) which is obtained from the graded Jacobi identity, see [16].

Proposition 7

In fact, a Lie bialgebroid is equivalent to a strong differential Gerstenhaber algebra structure on \(\bigoplus _k\Gamma (\bigwedge ^k(A)\). (See Proposition 2.3 in [24]).

Remark 8

([23]) In a Lie bialgebroid \((A,A^*)\), the base P has a natural Poisson structure as

\(\{f,g\}=<df,d_*g>,\ \ \forall f,g\in C^\infty (P)\) which satisfies \([df,dg]=d\{f,g\}\).

The same as Lie bialgebras, a useful method of constructing Lie bialgebroids is using r-matrices. Recall that an r-matrix is a section \(\Lambda \in \Gamma (\bigwedge ^2A)\) satisfying

$$\begin{aligned} L_X[\Lambda ,\Lambda ]=[X,[\Lambda ,\Lambda ]]=0,\ \ \forall X\in \Gamma (A). \end{aligned}$$

An r-matrix \(\Lambda\) defines a Lie bialgebroid, where the differential \(d_*:\Gamma (\bigwedge ^*A)\rightarrow \Gamma (\bigwedge ^{*+1}A)\) is \(d_*=[.,\Lambda ]\). The bracket on \(\Gamma (A^*)\) is

$$\begin{aligned} {[}\xi ,\eta ]=L_{\Lambda ^{\#}\xi }\eta -L_{\Lambda ^{\#}\eta }\xi -d[\Lambda (\xi ,\eta )]. \end{aligned}$$

The anchor is \(\rho \circ \Lambda ^{\#}:A^*\rightarrow TP\), where \(\Lambda ^{\#}\) is the bundle map \(A^*\rightarrow A\) with

$$\begin{aligned} \Lambda ^{\#}(\xi )(\eta )=\Lambda (\xi ,\eta ),\ \forall \xi ,\eta \in \Gamma (A^*). \end{aligned}$$

This Lie bialgebroid is called a coboundary Lie bialgebroid, and it is called a triangular Lie bialgebroid if \([\Lambda , \Lambda ]=0\). If \(\Lambda\) is also of constant rank then we call it as a regular triangular Lie bialgebroid.

Remark 9

([23]) When P has only one point, that means A is a Lie algebra, then \(L_X[\Lambda ,\Lambda ]=[X,[\Lambda ,\Lambda ]]=0\) is equivalent to that \([\Lambda ,\Lambda ]\) is Ad-invariant, so \(\Lambda\) is an r-matrix.

If A is the tangent bundle TP with the standard Lie algebroid structure, \(L_X[\Lambda ,\Lambda ]=[X,[\Lambda ,\Lambda ]]=0\) is equivalent to \([\Lambda ,\Lambda ]=0\) i.e., \(\Lambda\) is a Poisson tensor.

5 Deformation of Bialgebroids

We consider a topological bialgebra \((A,\mu ,\eta ,\Delta ,\varepsilon )\) where A is a module over the ring \(k[[\hbar ]]\) and \(\mu :A\otimes A\rightarrow A, \ \ \eta :k[[\hbar ]]\rightarrow A, \ \Delta :A\rightarrow A\otimes A\) and \(\varepsilon :A\rightarrow k[[\hbar ]]\) are \(k[[\hbar ]]\)-linear maps. Similarly, a topological bialgebroid is a bialgebroid which is a module over \(k[[\hbar ]]\) and all structure maps are \(k[[\hbar ]]\)-linear maps.

Remember from Sect. 2, using twist construction, in Theorem 1 one obtains a new bialgebroid from a given bialgebroid. Also remember the Example 10, in which using twisted construction we obtained twisted Hopf algebroid \((D_{\hbar }, R_{\hbar }, \alpha _{\hbar }, \beta _{\hbar }, m, \Delta _{\hbar }, \varepsilon )\) which is called the quantum groupoid associated to the star product \(*_{\hbar }.\)

Definition 23

A deformation of a bialgebroid \((A,R,\alpha ,\beta ,m,\Delta ,\varepsilon )\) over a field k is a topological bialgebroid \((A_{\hbar },R_{\hbar },\alpha _{\hbar },\beta _{\hbar },m_{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) over the ring \(k[[\hbar ]]\) of formal power series in \(\hbar\) such that

  1. 1.

    \(A_\hbar\) is isomorphic to \(A[[\hbar ]]\) as \(k[[\hbar ]]\)-module with identity \(1_A\), and \(R_{\hbar }\) is isomorphic to \(R[[\hbar ]]\) as \(k[[\hbar ]]\)-module with identity \(1_R\),

  2. 2.

    \(\alpha _{\hbar }=\alpha (mod \hbar ),\ \beta _{\hbar }=\beta (mod \hbar ),\ m_{\hbar }=m(mod \hbar ),\ \varepsilon _{\hbar }=\varepsilon (mod \hbar )\),

  3. 3.

    \(\Delta _{\hbar }=\Delta (mod \hbar )\).

In this case, we simply say that the quotient \(A_{\hbar }/\hbar A_{\hbar }\) is isomorphic to A as a bialgebroid.

Lemma 2

(Xu, [23]) By the conditions (1), (2) in the above definition, \(A_{\hbar }\otimes _{R_{\hbar }}A_{\hbar }/\hbar (A_{\hbar }\otimes _{R_{\hbar }}A_{\hbar })\) is isomorphic to \(A\otimes _RA\) as k-module.

Definition 24

A quantum universal envelo** algebroid (or QUE algebroid), also called a quantum groupoid is a deformation of the standard bialgebroid \((UA,R,\alpha ,\beta ,m,\Delta ,\varepsilon )\) of a Lie algebroid A.

Suppose \((U_{\hbar }A,R_{\hbar },\alpha _{\hbar },\beta _{\hbar },m_{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) is a quantum groupoid. Note that \(U_{\hbar }A=UA[[\hbar ]]\) and \(R_{\hbar }=R[[\hbar ]]\). Then \(R_{\hbar }\) defines a star product on the base manifold P for the Lie algebroid A,  so that

$$\begin{aligned} \{f,g\}=lim_{\hbar \rightarrow 0}\frac{1}{\hbar }(f*_{\hbar }g-g*_{\hbar }f) \end{aligned}$$

(with \(f,g\in R\)) is a Poisson structure on the base P. Define

  • \(\delta f:=lim_{\hbar \rightarrow 0}\frac{1}{\hbar }(\alpha _{\hbar }f-\beta _{\hbar }f)\in UA,\ \forall f\in R\),

  • \(\Delta ^1X:=lim_{\hbar \rightarrow 0}\frac{1}{\hbar }(\Delta _{\hbar }X-(1\otimes _{R_{\hbar }}X+X\otimes _{R_{\hbar }}1))\in UA\otimes _R UA,\ \forall X\in \Gamma (A)\), and

  • \(\delta X:=\Delta ^1X-(\Delta ^1X)_{21}\in UA\otimes _RUA.\)

For any \(f,g\in R,x,y\in UA\) write

$$\begin{aligned}&\alpha _{\hbar }f=f+\hbar \alpha _1f+\hbar ^2\alpha _2f+O(\hbar ^3), \\&\beta _{\hbar }f=f+\hbar \beta _1f+\hbar ^2\beta _2f+O(\hbar ^3), \\&f*_{\hbar }g=fg+\hbar B_1(f,g)+O(\hbar ^2), \\&x*_\hbar y=xy+\hbar m_1(x,y)+O(\hbar ^2) \end{aligned}$$

where \(\alpha _1f,\beta _1f,\alpha _2f,\beta _2f,m_1(x,y)\) are elements in U(A). So, \(\{f,g\}=B_1(f,g)-B_1(g,f)\) and \(\delta f=\alpha _1f-\beta _1f\).

Proposition 8

(Xu, [23]) For any \(f,g\in R\) and \(X\in \Gamma (A)\) we obtain

  1. 1.

    \(\delta f\in \Gamma (A)\) and \(\delta X\in \Gamma (\bigwedge ^2A)\),

  2. 2.

    \(\delta (fg)=f\delta g+g\delta f\),

  3. 3.

    \(\delta (fX)=f\delta X+\delta f\wedge X\),

  4. 4.

    \([\delta f,g]=\{f,g\}\),

  5. 5.

    \(\delta ^2f=0.\)

By properties (1)–(3) we can extend \(\delta\) to a well-defined degree 1 derivation \(\delta :\Gamma (\bigwedge ^*A)\rightarrow \Gamma (\bigwedge ^{*+1}A).\) Then the algebra \((\bigoplus \Gamma (\bigwedge ^*A),\wedge ,[.,.],\delta )\) is a strong differential Gerstenhaber algebra. Since, Considering the above proposition, it suffices to show that \(\delta\) is a derivation with respect to [., .], and \(\delta ^2=0\). The proof of these two are in Propositions 5.13 and 5.14 in [23].

By above proposition the dual pair \((A,A^*)\) is a Lie bialgebroid, which is called the classical limit of the quantum groupoid \(U_{\hbar }A\).

Therefore, the following theorem is obtained.

Theorem 3

(Xu, [23]) A quantum groupoid \((U_{\hbar }A,R_{\hbar },\alpha _{\hbar },\beta _{\hbar },m_{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) naturally induces a Lie bialgebroid \((A,A^*)\) as a classical limit. The induced Poisson structure of this Lie bialgebroid on the base manifold P coincides with the one obtained as the classical limit of the base \(*\)-algebra \(R_{\hbar }\).

In [10] it was shown that every Lie bialgebra is quantizable. Now, one can ask if every Lie bialgebroid is quantizable.

Definition 25

A Quantization of a Lie bialgebroid \((A,A^*)\) is a quantum groupoid \((U_{\hbar }A,R_{\hbar },\alpha _{\hbar },\beta _{\hbar },m_{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) whose classical limit is \((A,A^*)\).

Remark 10

In [13], Kontsevich has shown that given a Poisson structure \(\{.,.\}\) on a Poisson manifold P one can find a \(*\)-product such that

$$\begin{aligned} f*g=fg+\{f,g\}t+\dots \end{aligned}$$

So, we can say that the Lie bialgebroid \((TP,T^*P)\) associated to a Poisson manifold P is always quantizable.

In a special case, Xu in [23] showed that any regular triangular Lie bialgebroid is quantizable. In general case, in [3] it was shown that every Lie bialgebroid is quantizable.

6 The Lie Bialgebroid of Formal Pseudo-differential Operators

In this section we are going to indicate the Theorem 3 for formal pseudo-differential operators. In other words, we see how naturally the algebra of formal pseudo-differential operators can be considered as a Lie bialgebroid. First we recall some definitions. Here we consider the formal pseudo-differential operators in a general algebraic setting described for instance in [1].

Let A be a commutative k–algebra with unit 1,  where k is a field of the zero-characteristic. Assume that A is equipped with a derivation, i.e. there is a k–linear map \(D: A\rightarrow A\) satisfying the Leibnitz rule \(D(f.g)=(Df).g+f.(Dg)\) for all \(f,g\in A.\) We say that an element is constant if \(Df=0.\) Consider \(\xi\) as a formal variable not in A. The algebra of symbols over A is the vector space

$$\begin{aligned} \Psi _{\xi }(A)= \left\{ P_{\xi }=\sum _{\nu \in \mathbb {Z}} a_{\nu }\xi ^{\nu } \mid a_{\nu }\in A, a_{\nu }=0 \ \ for \ \ \nu \gg 0\right\} \end{aligned}$$

which is equipped with the associative multiplication \(\circ\) defined by

$$\begin{aligned} P_{\xi }\circ Q_{\xi }=\sum _{k\ge 0} \frac{1}{k!}\frac{\partial ^k}{\partial \xi ^k}D^kQ_{\xi }. \end{aligned}$$

Note that the algebra A is included in \(\Psi _{\xi }(A),\) and for \(f\in A\) and \(\nu \in Z\) the multiplication \(\circ\) is given by \(\xi ^{\nu }\circ f=\sum _{k\ge 0} \frac{\nu (\nu -1)\dots (\nu -k+1)}{k!}(D^kf)\xi ^{\nu -k}.\)

Definition 26

The algebra of formal pseudo-differential operators over A is the vector space

$$\begin{aligned} \Psi (A)=\left\{P=\sum _{\nu \in \mathbb {Z}} a_{\nu }D^{\nu } \mid a_{\nu }\in A, a_{\nu }=0 \ \ for \ \ \nu \gg 0\right\} \end{aligned}$$

which is equipped with the unique multiplication which makes the map \(\sum _{\nu \in \mathbb {Z}}a_{\nu }\xi ^{\nu }\mapsto \sum _{\nu \in \mathbb {Z}} a_{\nu }D^{\nu }\) an algebra homomorphism. The algebra \(\Psi (A)\) is associative but not commutative. It is a Lie algebra over k with the Lie bracket \([P,Q]:=PQ-QP.\)

Now, consider \(A=\mathbb {R}^n\) as the commutative \(\mathbb {R}\)–algebra with unit 1,  over the field \(\mathbb {R},\) equipped with the Hadamard product (entrywise product) which for \(n=1\) is the usual product on \(\mathbb {R}.\)

Theorem 4

There is a natural Lie bialgebroid structure on the algebra of formal pseudo-differential operators \(\Psi (\mathbb {R}^n)\) as the classical limit of the quantum groupoid associated to it.

Proof

Remember from above definition that the algebra A is included in \(\Psi (A).\) Moreover, if we consider \(\mathbb {R}^n\)as a manifold, then we can define the bundle map \(\pi : \Psi (\mathbb {R}^n)\rightarrow \mathbb R^n\) as the natural projection. Hence, \(\Psi (\mathbb {R}^n)\) is a Lie algebroid with the Lie bracket defined in above definition. By Definition (21), \(U(\Psi (\mathbb {R}^n))=U(\Gamma (\Psi (\mathbb {R}^n))\oplus C^\infty (\mathbb R^n))\) is the universal envelo** algebra of our Lie algebroid. Therefore, by Theorem 2, we have the bialgebroid \((U(\Psi (\mathbb {R}^n)),C^\infty (\mathbb R^n),\alpha ,\beta ,\Delta ,\varepsilon )\) in which \(\alpha =\beta =i:C^\infty (\mathbb R^n)\hookrightarrow U(\Psi (\mathbb {R}^n))\) are the natural embeddings, and \(\varepsilon :U(\Psi (\mathbb {R}^n))\rightarrow C^\infty (\mathbb R^n)\) is the projection. Moreover, we have

$$\begin{aligned} \Delta (f)= & {} f\otimes _{C^\infty (\mathbb R^n)}1, \ \ \ \forall f\in R=C^\infty (\mathbb R^n)\\ \Delta (X)= & {} X\otimes _{C^\infty (\mathbb R^n)}1+1\otimes _{C^\infty (\mathbb R^n)}X,\ \ \ \forall X\in \Gamma (U(\Psi (\mathbb {R}^n))). \end{aligned}$$

Now, consider \((U_{\hbar }(\Psi (\mathbb {R}^n)), C^\infty (\mathbb R^n)_{\hbar },\alpha _{\hbar },\beta _{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) as the quantum universal envelo** algebra of the above bialgebroid. Here we define \(\alpha _{\hbar },\beta _{\hbar }:C^\infty (\mathbb R^n)_{\hbar }\rightarrow U_{\hbar }(\Psi (\mathbb {R}^n))\) by

$$\begin{aligned} \alpha _{\hbar }f= & {} f+\hbar \sum _{i=1}^m\partial _i f+\hbar ^2\alpha _2f+O\left( \hbar ^3\right) \\ \beta _{\hbar }f= & {} f+\hbar \left( -\sum _{i=m}^n\partial _i f\right) +\hbar ^2\beta _2f+O\left( \hbar ^3\right) , \end{aligned}$$

where \(\partial _i\) means differentiation with respect to the i–th variable and m is fixed such that \(1\le m\le n.\)

Also, if we consider the coordinates \((q_i,p_i)\) on \(T\mathbb {R}^n\) we define

$$\begin{aligned} f*_{\hbar } g= & {} fg+\hbar \sum _{i=1}^n \frac{\partial f}{\partial q_i}\frac{\partial g}{\partial p_i}+O(\hbar ^2),\\ x*_{\hbar }y= & {} xy+\hbar m_1(x,y)+O(\hbar ^2), \end{aligned}$$

moreover, we consider \(\varepsilon _{\hbar }(x) = \varepsilon (x)+\hbar \varepsilon _1(x)+O(\hbar ^2),\) (\(\varepsilon _1(x)\in C^\infty (\mathbb R^n)\)), and \(\Delta _{\hbar } X\) with

$$\begin{aligned} \Delta ^1 X= & {} lim_{\hbar \rightarrow 0}\frac{1}{\hbar }(\Delta _{\hbar } X-(1\otimes _{C^\infty (\mathbb R^n)_{\hbar }}X+X\otimes _{C^\infty (\mathbb R^n)_{\hbar }}1))\\\in & {} U(\Psi (\mathbb {R}^n))\otimes _{C^\infty (\mathbb R^n)}U(\Psi (\mathbb {R}^n)),\ \forall X\in \Gamma (U(\Psi (\mathbb {R}^n))), \end{aligned}$$

for any \(f,g\in C^\infty (\mathbb R^n),\ x,y\in U(\Psi (\mathbb {R}^n))\), where \(\alpha _2f, \beta _2f,m_1(x,y)\) are elements in \(U(\Psi (\mathbb {R}^n))\).

Therefore, we have

$$\begin{aligned} \{f,g\}=\, & {} lim_{\hbar \rightarrow 0}\frac{1}{\hbar }(f*_{\hbar } g-g*_{\hbar }f)\\= \,& {} lim_{\hbar \rightarrow 0}\frac{1}{\hbar }\left( fg+\hbar \sum _{i=1}^n \frac{\partial f}{\partial q_i}\frac{\partial g}{\partial p_i}+O\left( \hbar ^2\right) -gf-\hbar \sum _{i=1}^n \frac{\partial g}{\partial q_i}\frac{\partial f}{\partial p_i}-O\left( \hbar ^2\right) \right) \\=\, & {} \sum _{i=1}^n \frac{\partial f}{\partial q_i}\frac{\partial g}{\partial p_i}-\frac{\partial g}{\partial q_i}\frac{\partial f}{\partial p_i} \end{aligned}$$

which is the standard Poisson bracket on the base manifold \(\mathbb R^n\). Moreover,

$$\begin{aligned} \delta f=\, & {} lim_{\hbar \rightarrow 0}\frac{1}{\hbar }(\alpha _{\hbar }f-\beta _{\hbar }f)\\= \,& {} lim_{\hbar \rightarrow 0}\frac{1}{\hbar }\left( \hbar \sum _{i=1}^m\partial _i f -\hbar \left( -\sum _{i=m}^n\partial _i f\right) +\hbar ^2\alpha _2f-\hbar ^2\beta _2f+O\left( \hbar ^3\right) -O\left( \hbar ^3\right) \right) \\=\, & {} \sum _{i=1}^m\partial _i f-\left( -\sum _{i=m}^n\partial _i f\right) \\= \, & {} df. \end{aligned}$$

Thus we obtain the ordinary derivation which clearly satisfies all the properties of \(\delta\) in Proposition 8. Therefore, by Theorem 3 the quantum groupoid \((U_{\hbar }(\Psi (\mathbb {R}^n),C^\infty (\mathbb R^n)_{\hbar },\alpha _{\hbar },\beta _{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) naturally induces the Lie bialgebroid \((\Psi (\mathbb {R}^n), (\Psi (\mathbb {R}^n))^*)\) as a classical limit.

With the bracket \([.,.]_{U(\Psi (\mathbb {R}^n)}\) and the exterior product \(\wedge\) we have \((\bigoplus \Gamma (\bigwedge ^*U(\Psi (\mathbb {R}^n)),\wedge ,[.,.],\delta )\) as a differential Gerstenhaber algebra, and

$$\begin{aligned}&\delta ^2=0, \\&\delta [X,Y]=[\delta X,Y]+[X,\delta Y]. \end{aligned}$$

\(\square\)

Remark 11

Note that in the proof of above theorem, we chose the algebra A as \(\mathbb {R}^n\) which is also considered as a manifold. We would like to mention that the above proof is applicable to a diffeological A,  that is a choice where A has its algebraic operations which are smooth in the diffeological sense, along the lines of [9]. In this context, \(*_{\hbar }\) needs to be smooth.

Remark 12

In fact for any algebra A,  if the algebra of pseudo-differential operators \(\Psi (A)\) admits a Lie agebroid structure then by Definition (21), \(U(\Psi (A))=U(\Gamma (\Psi (A))\oplus C^\infty (A))\) is the universal envelo** algebra of our Lie algebroid.Therefore, by Theorem 2, we have the bialgebroid \((U(\Psi (A)),C^\infty (A),\alpha ,\beta ,\Delta ,\varepsilon ).\) Now, consider \((U_{\hbar }(\Psi (A)), C^\infty (A)_{\hbar },\alpha _{\hbar },\beta _{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) as the quantum universal envelo** algebra of the above bialgebroid. By Theorem 3 and considering Remark 11, this quantum groupoid naturally induces a Lie bialgebroid \((\Psi (A),(\Psi (A))^*)\) as the classical limit. Then one can obtain non-commutative KP equations based on \(U_{\hbar }(\Psi (A))\) as we explain in the next section. The interesting application is the choice of A as Connes-Kreimer Hopf algebra, i.e. the algebra A is generated by rooted planar binary trees, as it is described in [8]. If \(\Psi (A)\) admits a Lie algebroid structure then by above explanation one can obtain non-commutative KP equations in terms of Connes-Kreimer Hopf algebra.

7 Non-Commutative Kadomtsev–Petviashvili Hierarchy

According to [11], the non-commutativity is arbitrarily introduced for the variables \((x_1 ,x_2,\dots )\) as the following equation.

$$\begin{aligned}{}[x_k, x_l]=i\theta ^{kl}, \end{aligned}$$

where the real constants \(\theta ^{kl}\) are called the non-commutative parameters.

Non-commutative field theories are obtained from given commutative field theories by using star-products instead of the ordinary products in the commutative field theories. For instance, on Euclidean spaces, a star-product is given by

$$\begin{aligned} f(x)\star g(x):= f(x)g(x)+\frac{i}{2}\theta ^{kl}\partial _kf(x)\partial _lg(x)+O(\theta ^2), \end{aligned}$$

which is known as Moyal product, where \(\partial _k=\frac{\partial }{\partial x_k}.\) The star-product is associative, i.e. \(f\star (g \star h)=(f\star g)\star h,\) and it turns to the ordinary product in the commutative limit \(\theta ^{kl}\rightarrow 0.\) Using the star product we obtain non-commutativity, i.e.

$$\begin{aligned}{}[x_k, x_l]_{\star }=x_k\star x_l-x_l\star x_k=i\theta ^{kl}. \end{aligned}$$

The usual definition of non-commutative KP hierarchy, i.e.

$$\begin{aligned} \partial _mL=[(L^m)_+,L]_{\star }, \end{aligned}$$

based on this non-commutative condition is given in [11].

Now, assume that our KP equations are defined on the algebra of formal pseudo-differential operators \(\Psi (\mathbb {R}^n).\) By Theorem 4, there is a Lie bialgebroid structure on \(\Psi (\mathbb {R}^n)\) as the classical limit of the quantum groupoid \((U_{\hbar }(\Psi (\mathbb {R}^n)),C^\infty (\mathbb R^n)_{\hbar },\alpha _{\hbar },\beta _{\hbar },\Delta _{\hbar },\varepsilon _{\hbar }).\) Therefore, we can naturally obtain non-commutative KP equations based on this quantum groupoid. More precisely, imitating the procedure given in [11], we obtain the non-commutative KP equations in terms of \(*_{\hbar }\) defined on the mentioned quantum groupoid (see definition of \(*_{\hbar }\) in the proof of Theorem 4) instead of Moyal product \(\star.\) Moreover, the Lie bracket on \(U_{\hbar }(\Psi (\mathbb {R}^n))\) is defined by

$$\begin{aligned}{}[x_k,x_l]_{\hbar }=(x_k *_{\hbar } x_l-x_l *_{\hbar } x_k)+(x_k *_{\hbar } x_l-x_l *_{\hbar } x_k)\hbar +(x_k *_{\hbar } x_l-x_l *_{\hbar } x_k)\hbar ^2+\dots , \end{aligned}$$

where in the classical limit \(\hbar \rightarrow 0,\) it turns to the usual Lie bracket on \(U(\Psi (\mathbb {R}^n)).\)

Consider an Nth order (monic) pseudo-differential operator T as

$$\begin{aligned} T=\partial _x^N+a_{N-1}\partial _x^{N-1}+\dots +a_0+a_{-1}\partial _x^{-1}+a_{-2}\partial _x^{-2}+\dots , \end{aligned}$$

and denote \(T_{\ge }:= \partial _x^N+a_{N-1}\partial _x^{N-1}+\dots +a_r\partial _x^r.\) A differential operator \(\partial _x^n\) acts formally on a multiplicity operator f as the following generalized Leibniz rule.

$$\begin{aligned} \partial _x ^n f:=\sum _{i\ge 0}\left( \!\!\! \begin{array}{c} n \\ i \end{array} \!\!\!\right) (\partial _x^ if)\partial ^{n-i}, \end{aligned}$$

where the binomial coefficient

$$\begin{aligned} \left( \!\!\! \begin{array}{c} n \\ i \end{array} \!\!\!\right) :=\frac{n(n-1)\dots (n-i+1)}{i(i-1)\dots 1} \end{aligned}$$

is applicable for negative n. For example,

$$\begin{aligned} \partial _x^{-1}f= & {} f\partial _x^{-1}-f'\partial _x^{-2}+f''\partial _x^{-3}-\dots ,\\ \partial _x^{-2}f= & {} f\partial _x^{-2}-2f'\partial _x^{-3}+3f''\partial _x^{-4}-\dots ,\\ \partial _x^{-3}f= & {} f\partial _x^{-3}-3f'\partial _x^{-4}+6f''\partial _x^{-5}-\dots , \end{aligned}$$

where \(f':=\frac{\partial f}{\partial x}, f'':=\frac{\partial ^2f}{\partial x^2}.\) Now, consider the following first-order pseudo-differential operator.

$$\begin{aligned} L=\partial _x+u_1+u_2\partial _x^{-1}+u_3\partial _x^{-2}+u_4\partial _x^{-3}+\dots , \end{aligned}$$

where the coefficients \(u_k, (k=1,2,\dots )\) are functions of infinite variables \((x_1,x_2,\dots )\) with \(x_1=x,\) i.e. \(u_k=u_k(x,x_2,\dots ).\) We consider the non-commutativity given by the star-product \(*_{\hbar },\) as described in the proof of Theorem 4. Then the Lax hierarchy is defined by

$$\begin{aligned} \partial _m L=[B_m,L]_{*_{\hbar }}, \ \ m=1,2,\dots , \end{aligned}$$

where applying \(\partial _m\) on the pseudo-differential operator L is considered coefficient-wise, i.e., \(\partial _m L:=[\partial _m,L]\) or \(\partial _m\partial _x^k=0.\) Moreover, the operator \(B_m\) is defined by \(B_m:=(L*_{\hbar }\dots *_{\hbar } L)_{\ge r}\) (for m times). When \(u_1=0 (r=0)\) the Lax hierarchy is the non-commutative KP hierarchy which includes the non-commutative equations, see [14,21]. The coefficients of each power of pseudo-differential operators in above Lax hierarchy give a series of infinite non-commutative evolution equations. For instance, for \(m=1\) the coefficient of \(\partial _x^{1-k}\) gives the equations

$$\begin{aligned} \partial _1u_k=u_k', \ \ \ \ k=2, 3, \dots \Rightarrow x_1=x. \end{aligned}$$

For \(m=2\) the coefficient of \(\partial _x^{-1}\) leads to

$$\begin{aligned} \partial _2u_2=u_2''+2u_3', \end{aligned}$$

and the coefficient of \(\partial _x^{-2}\) gives

$$\begin{aligned} \partial _2u_3 & = u_3''+2u_4'+2u_2 *_{\hbar } u_2'+2[u_2,u_3]_{*_{\hbar }}\\ & = u_3''+2u_4'+\Big (2u_2u_2'+m_1(2u_2,u_2')\hbar +O(\hbar ^2)\Big )\\ & \quad + 2\Big ((u_2u_3+m_1(u_2,u_3)\hbar -u_3u_2-m_1(u_3,u_2)\hbar )\\ & \quad + (u_2u_3+m_1(u_2,u_3)\hbar -u_3u_2-m_1(u_3,u_2)\hbar )\hbar +O(\hbar ^2)\Big ), \end{aligned}$$

where \(m_1(u_2,u_3)\) and \(m_1(u_3,u_2)\) are elements in \(U(\Psi (\mathbb {R}^n)).\)

Next, the coefficient of \(\partial _x^{-3}\) yields the equation

$$\begin{aligned} \partial _2u_4=u_4''+2u_5'+4u_3 *_{\hbar } u_2'-2u_2 *_{\hbar } u_2''+2[u_2,u_4]_{*_{\hbar }}, \end{aligned}$$

and the same as previous case we can expand it based on \(*_{\hbar }.\)

For \(m=3\) the coefficients of \(\partial _x^{-1}, \partial _x^{-2}, \partial _x^{-3},\) and \(\partial _x^{-4}\) respectively give the equations

$$\begin{aligned}&\partial _3 u_2=u_2'''+3u_3''+3u_4'+3u_2' *_{\hbar } u_2+3u_2 *_{\hbar } u_2',\\&\partial _3u_3=u_3'''+3u_4''+3u'_5+6u_2 *_{\hbar } u_3'+3u_2' *_{\hbar } u_3+3u_3 *_{\hbar } u_2'+3[u_2,u_4]_{*_{\hbar }},\\&\partial _3 u_4=u_4'''+3u_5''+3u_6'+3u_2' *_{\hbar } u_4+3u_2 *_{\hbar } u_4'+6u_4 *_{\hbar } u_2'-3u_2 *_{\hbar } u_3''\\& \quad -3u_3 *_{\hbar } u_2''+6u_3 *_{\hbar } u_3'+3[u_2,u_5]_{*_{\hbar }}+3[u_3,u_4]_{*_{\hbar }},\\&\partial _3 u_5=\dots. \end{aligned}$$

If we set \(2u_2\equiv u, x_2\equiv y, x_3\equiv t\) then we obtain the non-commutative KP equation

$$\begin{aligned} 3u_{yy}=\partial (4u_t-3(\partial u *_{\hbar } u+u *_{\hbar } \partial u)-\partial ^3u)+3\partial [u,\partial ^{-1} u_y]_{*_{\hbar }}. \end{aligned}$$

Note that this equation is analogous to the non-commutative KP equation in [11]. The difference between them is that here the star product is defined by \(*_{\hbar }\) instead of Moyal product.

Remark 13

To solve the above non-commutative KP equation, we consider the KP equation on \(U_{\hbar }(\Psi (\mathbb {R}^n))\) with the coefficient algebra equipped with \(*_{\hbar }.\) In [9] we showed the well-posedness of the KP hierarchy with arbitrary (diffeological) coefficient algebra. Hence, considering diffeological algebraic structures, and Remark 11 we can conclude that the non-commutative KP hierarchy (equipped with \(*_{\hbar }\)) is well-posed, see also [17].

8 Conclusion

The algebra of formal Pseudo-differential operators \(\Psi (\mathbb {R}^n)\) can be naturally equipped with a Lie bialgebroid structure \((\Psi (\mathbb {R}^n), (\Psi (\mathbb {R}^n))^*)\) as the classical limit of the quantum groupoid \((U_{\hbar }(\Psi (\mathbb {R}^n)),C^\infty (\mathbb R^n)_{\hbar },\alpha _{\hbar },\beta _{\hbar },\Delta _{\hbar },\varepsilon _{\hbar })\) associated to it. Then, the KP equations defined on Pseudo-differential operators \(\Psi (\mathbb {R}^n)\) are “classical limit” of non-commutative KP equations defined by the above quantum grupoid. Hence, we obtain non-commutative KP equations in a more general sense and not by using non-commutativity of the coordinates i.e., \([x_k, x_l]=i\theta ^{kl}.\) Up to now, in special cases, non-commutative KP equations have been solved or it has been shown the existence and uniqueness of solutions of them using Moyal product, see [18] and references there in. The algebraic procedure described during this paper provides the possibility of solving non-commutative KP equations very generally.