Standard bases for the universal associative conformal envelopes of Kac–Moody conformal algebras
Abstract.
We study the universal enveloping associative conformal algebra for the central extension of a current Lie conformal algebra at the locality level . A standard basis of defining relations for this algebra is explicitly calculated. As a corollary, we find a linear basis of the free commutative conformal algebra relative to the locality on the generators.
Key words and phrases:
conformal algebra, Gröbner–Shirshov basis2020 Mathematics Subject Classification:
17A61, 17B35, 17B691. Introduction
Conformal algebras also known as Lie vertex algebras were introduced in [16] as an algebraic tool to study the singular part of the operator product expansion (OPE) of chiral fields in 2-dimensional conformal field theory coming back to [5]. From the categorical point of view, a conformal algebra is just an algebra in the appropriate (pseudo-tensor) category of modules over the polynomial algebra in one variable [2]. The pseudo-tensor structure (see [4]) reflects the main features of multi-linear maps in the category of linear spaces: composition, identity, symmetric structure. These features are enough to define the basic notions like what is an algebra (associative, commutative, Lie, etc.), homomorphism, ideal, representation, module, cohomology. Therefore, the notion of a conformal algebra is a natural expansion of the notion of an “ordinary” algebra over to the pseudo-tensor category . Namely, as an ordinary algebra is a linear space equipped with a bilinear product, a conformal algebra is a -module equipped with a -bilinear map (pseudo-product)
A more convenient presentation for the operation uses the language of a -product or a family of -products for all integer ([16], see also Section 2).
Conformal algebras representing the singular part of OPE in vertex algebras are Lie algebras in the category , i.e., Lie conformal algebras. For example, if is a Lie algebra then the free module equipped with the pseudo-product , , is a Lie conformal algebra denoted (current conformal algebra). If is a bilinear symmetric invariant form on then has a 1-dimensional central extension defined by
where is a central element and . For example, in the Kac–Moody vertex algebra [12] the singular part of the OPE on the generating fields is described by this particular structure called a Kac–Moody conformal algebra.
As in the case of ordinary algebras, an associative conformal algebra turns into a Lie one with respect to the commutator , , where is the switching map on . However, not all Lie conformal algebras embeds into associative ones in this way [26]. This is an open problem whether every finite (i.e., finitely generated as a -module) Lie conformal algebra embeds into an associative conformal algebra with respect to conformal commutator. Even for the class of quadratic conformal algebras [27] (see also [15]) it remains unknown in general if every such Lie conformal algebra embeds into an appropriate associative one.
A routine way to solve this kind of problems is to construct a universal envelope. In general, such an algebra is defined by generators and relations. For a Lie conformal algebra, there exists a lattice of universal enveloping associative conformal algebras, each related to an (associative) locality bound on the generators ([26], see also Section 3.4). In order to prove (or disprove) the embedding of a Lie conformal algebra into its universal enveloping associative conformal algebra one needs to know the normal form of elements in the last algebra.
A general and powerful method for finding normal forms in an algebra defined by generators and relations is to calculate a standard (or Gröbner–Shirshov) basis of defining relations. The idea goes back to Newmann’s Diamond Lemma [23], see also [7, 6]. In the recent years, the Gröbner–Shirshov bases theory was developed to serve the problem of combinatorial analysis of various algebraic structures, see [9]. For associative conformal algebras it was initially invented in [8], later developed in [24] and [20]. In this paper, we use the last approach exposed in a form convenient for actual computation: we consider defining relations in a conformal algebra as rewriting rules on a module over an appropriate associative algebra (the Gröbner–Shirshov basis of the last algebra is known).
A series of particular observations made in [21], [22] shows that for all considered examples of quadratic Lie conformal algebras it is enough to consider universal associative conformal envelopes relative to the locality bound to get an injective mapping . This is one of the reasons why we focus on the locality bound for the envelopes of current conformal algebras as they are particular examples of quadratic conformal algebras.
The main purpose of this paper is to find a standard (Gröbner–Shirshov) basis of defining relations for the universal enveloping associative conformal algebra of a Kac–Moody conformal algebra at locality level . As a corollary, we get an analogue of the Poincaré–Birkhoff–Witt Theorem (PBW-Theorem) stating that the associated graded conformal algebra obtained from the universal envelope of a current Lie conformal algebra with respect to the natural filtration is isomorphic to the free commutative conformal algebra. Note that the classical PBW-Theorem may be interpreted as a conformal one at the locality level : for a Lie algebra , its “ordinary” universal envelope gives rise to the conformal algebra which is exactly the universal enveloping conformal algebra of with (here is the augmentation ideal of ).
There are several reasons for studying the universal envelopes of at higher locality than .
First, the envelope does not reflect the homological properties of . For example, if is a simple finite-dimensional Lie algebra then the second cohomology group is one-dimensional [3]. The corresponding central extension is the Kac–Moody conformal algebra representing the singular part of the Kac–Moody vertex algebra [12]. On the other hand, it is easy to find that the second Hochschild cohomology group of with coefficients in the trivial 1-dimensional module is zero: there are no nontrivial central extensions. Our results show that the universal enveloping associative conformal algebras for at locality level do have a nontrivial central extension which is exactly the universal envelope of .
The second reason is related with Poisson algebras. Assume is an ordinary commutative algebra with a Poisson bracket . Then may be considered as a Lie conformal algebra since is a Lie algebra relative to the Poisson bracket. There is a conformal representation of on itself given by the rule
The study of this representation provides a way to get new results in (quadratic) conformal algebras as well as in Poisson algebras [21, 22]. The conformal linear operators , , are local to each other, and the locality bound is . Indeed, according to the definition of a conformal representation [11, 16] we have
for . If then right-hand side is of degree 2 in that means in . Therefore, the corresponding associative envelope belongs to the class of envelopes with locality .
The third reason to study the case comes from the following relation between commutative conformal and Novikov algebras. Suppose is a commutative conformal algebra and is a subset of such that for all . Then generates an ordinary (nonassociative) subalgebra in the space considered relative to the single product . Indeed, all elements of are local to each other with locality bound . Moreover, the following relations hold:
for all . These identities are known to define the variety of Novikov algebras initially appeared in [1], [13]. In order to perform a systematic study of this relation, one needs to know the structure of the universal object in the category of commutative conformal algebras with locality bound on the generators.
For all these reasons, we study the universal enveloping associative conformal algebras for Kac–Moody conformal algebras relative to the locality level . The corollaries of the main result of the paper (Theorem 2) allow us to get the structure of the universal envelopes for current Lie conformal conformal algebras at and also describe the free commutative conformal algebra at the same locality level. Practically, we find a standard (Gröbner–Shirshov) basis of defining relations for these conformal algebras and derive an analogue of the PBW-Theorem.
2. Preliminaries in conformal algebras
The definition of a conformal algebra as an algebra in an appropriate pseudo-tensor category [2] corresponds to the convenient algebraic approach using -brackets [16] if it is presented in terms of operads associated with linear algebraic groups [18].
Let be a linear algebraic group over a field of characteristic zero, and let be the Hopf algebra of regular functions on . For every -module there is a non-symmetric operad (let us denote it ) defined as follows. Given , set
The condition of regularity means that may be presented by a polynomial function with coefficients in the space of -polylinear maps on , and the 3/2-linearity (sesqui-linearity) may be expressed as
for , , , (here is a variable ranging in ). In particular, is the space of all -linear transformations of thus contains the identity map .
The composition rule in is defined by the following partial composition. If , , then
acts as
(1) |
for , where in the right-hand side stands for the ordinary partial composition of polylinear maps. In particular, for the partial composition is equal to
It is easy to see that the resulting maps are indeed regular and 3/2-linear.
One may easily check that the partial composition in defined above meets the sequental, parallel, and unit axioms [10, Definition 3.2.2.3] and thus this is indeed an non-symmetric operad with a well-defined composition rule
Suppose the group is abelian. Then has a natural action of the symmetric group defined in the following way. If and , , is a transposition in then
for (here the action of in the right-hand side is just the permutation of arguments in a polylinear map). For , the definition is slightly more complicated: if is presented by a polynomial function
then is given by
where the each is the regular function .
The composition rule is equivariant (see, e.g., [10, Definition 5.2.1.1]) since the structure obtained is equivalent to the structure of an -module operad defined over a cocommutative Hopf algebra [2]. Namely, one may identify a map
with
Recall that if is a (symmetric) operad then a morphism defines an algebraic structure on the space . In the trivial case , , the operad coincides with the operad of polylinear maps on the linear space , and thus a morphism defines the ordinary notion of an -algebra over , a space equipped with a family of polylinear operations.
The classes of conformal [16] and -conformal [14] algebras naturally appear in the next step, if we choose to be a connected linear algebraic group of dimension 1 (the affine line and , respectively). For a non-connected group with the identity component denoted , the structure of a conformal algebra over is naturally interpreted as a -graded conformal algebra over [19]. If , (the variable is traditionally denoted by ), then a morphism defines a -conformal algebra structure on a -module .
For example, if is the operad governing the variety of associative algebras (generated by modulo the relation ) then an associative conformal algebra structure on a -module is given by an image of , a map which is 3/2-linear
for , and associative in the sense that
By (1), the latter means
(2) |
(to compute the right-hand side, put and in (1)).
According to the same scheme, a Lie conformal algebra structure on a -module is a morphism from the operad governing the variety of Lie algebras to . To define such a morphism, it is enough to fix a 3/2-linear map
such that and . The last two relations represent anti-commutativity and Jacobi identity, respectively:
.
In the sequel, we will use the notation for the operation on an associative conformal algebra and for Lie conformal algebras.
Since there is a morphism of operads sending to , every associative conformal algebra turns into a Lie conformal algebra relative to the operation
For an associative conformal algebra defined via a morphism of operads , let stand for the Lie conformal algebra obtained as a composition .
The property of a commutator to be a derivation on an associative algebra may also be expressed as a relation in . Being translated to conformal algebras it turns into the following identity on an associative conformal algebra :
(3) |
As in the case of ordinary algebras, is a functor from the category of associative conformal algebras to the category of Lie algebras. In contrast to the case of ordinary algebras, this functor does not have a left adjoint one when considered on the entire category of associative conformal algebras. However, if we restrict the class of associative conformal algebras by means of locality on the generators ([26], see Section 3.4 for details) then there is an analogue of the universal enveloping associative algebra for Lie conformal algebras.
In terms of “ordinary” algebraic operations, a conformal algebra is a linear space equipped with a linear operator , the generator of , and a series of bilinear operations , , given by
These operations are called -products. They have to satisfy the following properties:
-
(C1)
For every there exists such that for all ;
-
(C2)
;
-
(C3)
.
The property (C1) is known as the locality axiom, (C2) and (C3) represent 3/2-linearity. For every conformal algebra , the locality function is a map such that for every and .
A conformal algebra is associative if
for all and . In a similar way, one may rewrite the identities defining the class of Lie conformal algebras.
3. Gröbner–Shirshov bases for associative conformal algebras
3.1. Rewriting system and standard bases for associative algebras
In this section, we briefly describe the well-known technique of standard bases (Gröbner–Shirshov bases) in associative algebras in order to fix the notations. The usual exposition of this technique requires a proper ordering of the monomials. However, the core statements laying in the foundation of the approach do not need a monomial ordering.
Let be a set and let stand for the set of all words in (including the empty word). The free associative algebra (with a unit) over the field generated by is denoted . Suppose is a family of pairs called rewriting rules, where , . We will write a pair like this as since the family determines an oriented graph as follows. The vertices of are the elements of ; two vertices and are connected with an edge () if and only if there is a rewriting rule in and a summand of the form in (, ) such that
for some . In other words, is obtained from by replacing an occurrence of the subword with the polynomial .
The graph splits into connected components (in the non-oriented sense) which explicitly correspond to the elements of the quotient , where stands for the ideal in generated by all for . In some cases, there is a way to check algorithmically whether two vertices belong to the same connected component of , i.e., if the images of and are equal in .
An oriented graph is called a rewriting system if there are no infinite oriented paths (in particular, no oriented cycles). In a rewriting system, for every vertex there is a nonempty set of terminal vertices attached to , i.e., such that there is a path , but there are no edges originated at . A rewriting system is confluent if for every vertex the set contains a single vertex.
Definition 1.
A family of rewriting rules in the free associative algebra is a standard basis (Gröbner–Shirshov basis, GSB) if is a confluent rewriting system.
Obviously, if is a GSB then every connected component of has a unique terminal vertex which is a linear combination of terminal (reduced) words. This combination is called a normal form of an element in : two polynomials and in represent the same element of if and only if their normal forms coincide. Therefore, the images of terminal words form a linear basis of .
The most natural way to guarantee that is a rewriting system is to make the set well-ordered relative to an order such that implies and for all (i.e., is a monomial order), and for all (i.e., is greater than every monomial in ).
To check the confluence of a rewriting system one may apply the Diamond Lemma originated to [23]. The latter states that a rewriting system is confluent if an only if for every “fork” (a pair of edges , ) there exist a vertex and two oriented paths , . If the rewriting system is then it is enough to check the Diamond condition for the following two kinds of forks:
-
(1)
For , in , , consider , , and ;
-
(2)
For , in , , , is a nonempty word, consider , , .
In both cases, if there exit oriented paths and for an appropriate polynomial then we say that the composition of and relative to the word is confluent modulo . Denote the polynomial by .
Theorem 1 ([6, 7]).
Suppose a set of rewriting rules in the free associative algebra defines a rewriting system . If every composition of rewriting rules from is confluent modulo then is a confluent rewriting system, i.e., is a GSB.
Let respect a monomial order on . Then is a rewriting system and the confluence of a composition may be replaced with a more convenient condition.
Corollary 1 ([7]).
If for every rewriting rules , in having a composition relative to a word the polynomial may be presented as
(4) |
where in and , then is a GSB.
In the actual computation, we will often apply the following trick to show the confluence of a fork , : find some paths and and then present in the form (4).
3.2. Rewriting system for bimodules over associative algebras
Let be an associative algebra (with a unit) and let be a bimodule over . Suppose is generated by a subset as an algebra and is generated by a subset as an -module. Then is isomorphic to a quotient of the free associative algebra modulo an ideal generated by a set of defining relations , i.e.,
Similarly, is a quotient of the free -module generated by modulo a family of defining relations . One may identify an element of with a noncommutative polynomial in the variables which is linear in .
The split null extension is an associative algebra isomorphic to the quotient of the free algebra generated by modulo the ideal generated by the union of , , and
These relations reflect the properties of multiplication in : .
Remark 1.
To consider left modules, it is enough to add relations , , to reflect .
Suppose we may choose a monomial in each defining relation of (up to a scalar multiple) in such a way that the family of all rewriting rules defines a rewriting system . Note that the defining relations of are homogeneous relative to . All monomials that are of degree in belong to the same connected component as zero, so it is enough to consider only the relations of degree 0 and 1 in , these are exactly the defining relations of and of , respectively. Therefore, the confluence test needs to be applied to the forks started at a word which either belongs to or contains only one letter from . Hence, the compositions emerging in this rewriting system are exactly those described in [17].
3.3. Free associative conformal algebras
Recall the construction of a free associative conformal algebra generated by a set relative to a given locality function . From now on, denote by the polynomial algebra .
By definition, is an associative conformal algebra generated by which is universal in the class of all associative conformal algebras generated by such that the mutual locality of elements from in is bounded by . Namely, for every associative conformal algebra and for every map such that for all there exists unique homomorphism of conformal algebras such that for all .
Proposition 1 ([25]).
The free associative conformal algebra is a free -module with a basis
Remark 2.
The conformal algebra may be presented in a more convenient form as a (left) module over an appropriate associative algebra [20]. Given a set , let denote the associative algebra generated by the set
relative to the defining relations
(5) | ||||
(6) | ||||
(7) |
where , .
The free associative conformal algebra is a left module over if we define the action as follows:
for , , . Therefore, considered as a left -module is a homomorphic image the free left -module generated by the set . It is not hard to find explicitly the kernel of that homomorphism .
Fix a function and consider the quotient of relative to the -submodule generated by the following elements:
(8) | ||||
(9) |
where . Obviously, there is a homomorphism of -modules extending . This homomorphism is actually an isomorphism since (5) and (8) imply the following relations in :
(10) |
where , , , .
Consider the relations (5)–(10) as rewriting rules in such a way that the first monomial is always a principal one. The terminal words in of the rewriting system obtained are
The images of these words in are linearly independent by Proposition 1, hence we obtain the following
Corollary 2 ([20]).
The free associative conformal algebra is isomorphic to as an -module.
It follows from the definition of the action of on that every conformal ideal of is an -submodule and vice versa. Hence we may replace the study of conformal ideals with the study of “ordinary” submodules.
Example 1 ([8]).
Let us determine the structure of an associative conformal algebra generated by the set relative to with one defining relation .
The algebra is generated by , , and satisfying (5). Namely, consider these relations as rewriting rules
Similarly, define the free conformal algebra as a module over generated by a single element relative to the following rewriting rules (8):
The compositions (10) of these relations include
The defining relation is naturally written as
(11) |
Consider the composition of and (11) relative to . On the one hand,
on the other hand,
Hence, we should add a new rewriting rule
(12) |
The latter has a composition with (11) relative to :
Hence, we should add
Next, consider the composition of and (11) relative to . In a similar way, we obtain that and are connected by a (non-oriented) path, so add
(13) |
(the choice of the principal part is voluntary since we have not fixed an order on the words).
3.4. Universal associative conformal envelopes of Lie conformal algebras
Suppose is a Lie conformal algebra generated by a set . Thus is a quotient of an appropriate free Lie conformal algebra by the ideal generated by a set of defining relations stated in terms of Lie conformal operations . The structure of free Lie conformal algebras was described in [25].
For a given function , the universal enveloping associative conformal algebra of relative to the locality level on is defined as the quotient of relative to the same defining relations rewritten by the rules
where the upper limit of the summation is determined by the Dong Lemma.
The main purpose of this paper is to study universal enveloping associative conformal algebras for Kac–Moody conformal algebras. The latter are central extensions of current Lie conformal algebras. For this particular class of problems, the Gröbner–Shirshov bases method described above may be slightly modified. The main advantage of the modification is that the relations (10) become not necessary.
Suppose is a Lie conformal algebra with an -torsion such that the torsion-free is a free -module (for example, every finite Lie conformal algebra has that property). Assume , where is an -basis of and is a -basis of . Then the structure of is completely determined by relations
and
for appropriate . These relations describe the structure of as of a torsion -module, the multiplication table in the Lie conformal algebra , and the structure of the extension
Then, for a given function , the conformal algebra may be considered as an ordinary left module over the associative algebra generated by
relative to defining relations (5) (for ) along with the following ones:
(17) |
where is naturally understood as . The relations (17) reflect the property (3) of associative conformal algebras. So is a left module over generated by the entire set relative to the relations (8) (for ) together with
(18) | |||
(19) | |||
(20) |
Since the defining relations of already form a Gröbner–Shirshov basis, in order to determine the structure of one needs to find a confluent system of rewriting rules in this -module. In the next section, we solve this problem for a Kac–Moody conformal algebra.
4. The Poincaré–Birkhoff–Witt Theorem for Kac–Moody conformal algebras at
Let be a Lie algebra and let be a bilinear symmetric invariant form on (e.g., the Killing form). Then , where , equipped with
for every is a Lie conformal algebra with 1-dimensional torsion and the torsion-free image isomorphic to .
Let us fix a linear basis of . Then is a generating set of . The purpose of this section is to calculate the Gröbner–Shirshov basis for for and prove the Poincaré–Birkhoff–Witt Theorem for this universal enveloping associative conformal algebra.
According to the scheme described in the previous section, is a module over the associative algebra generated by the set modulo the relations
(21) | |||
(22) |
The set of generators of as of an -module is , and the defining relations of this module are
(23) | |||
(24) | |||
(25) | |||
(26) | |||
(27) |
for all .
In order to translate these defining relations into rewriting rules we need to choose a principal monomial in each relation. The choice of principal parts affects on the resulting system of rewriting rules obtained in a process of adding compositions similar to Example 11.
We will always choose a principal term in a rewriting rule as a leading monomial relative to an appropriate order on the monomials in the free -module generated by . Namely, suppose the set is linearly well ordered and . Induce an order on by the rule
assuming and iff , for (this ordering turns to be the most convenient for our purpose). Extend the order on the set of monomials in by the deg-lex principle, i.e., first compare the lengths and then lexicographically.
For two monomials and in , , , set iff is lexicographically less than .
Then, eliminating the monomials () in (25) and (26) by means of (27), we obtain the following set of rewriting rules defining (along with the confluent set of rewriting rules for ):
(28) |
(29) |
(30) |
Theorem 2.
Proof.
First, we will show how to derive the rules (31)–(37) as compositions of the initial relations. Next, we will check the triviality of compositions obtained in further iterations.
Since the calculations are routine, we will state them in details for several particular cases, other cases are essentially the same and may be processed in a similar way.
For the purpose of clarity, we will use a brief notation to point a rule applied for rewriting (e.g., stands for , for , etc).
The rule (31) for appears from the intersection of and . Then, by induction, the intersection with produces (31) for :
The next example of an intersection of and produces the rest of the required rules. On the one hand, we have
on the other hand,
(38) |
In order to apply we have to assume . However, the composition obtained by subtracting the right-hand sides of the two expressions above is
(39) |
it is (skew-)symmetric relative to the permutation of and . Hence, we may assume the relation (39) holds on for every .
For the composition is trivial due to the locality. For , apply the rules and to get the following:
(40) |
For a fixed order on , use if necessary to obtain (33) or (34).
Consider (39) for . For convenience of the exposition, let us split the polynomial into two summands and process the summands separately:
(41) |