Weight-finite modules over the quantum affine and double quantum affine algebras of type
Abstract.
We define the categories of weight-finite modules over the type quantum affine algebra and over the type double quantum affine algebra that we introduced in [MZ19]. In both cases, we classify the simple objects in those categories. In the quantum affine case, we prove that they coincide with the simple finite-dimensional -modules which were classified by Chari and Pressley in terms of their highest (rational and -dominant) -weights or, equivalently, by their Drinfel’d polynomials. In the double quantum affine case, we show that simple weight-finite modules are classified by their (-dominant) highest -weight spaces, a family of simple modules over the subalgebra of which is conjecturally isomorphic to a split extension of the elliptic Hall algebra. The proof of the classification, in the double quantum affine case, relies on the construction of a double quantum affine analogue of the evaluation modules that appear in the quantum affine setting.
1. Introduction
The representation theory of quantum affine algebras is a vast and extremely rich theory which is still the subject of an intense research activity after more than three decades. The recent discovery of its relevance to the monoidal categorification of cluster algebras provides one of the latest and most striking illustrations of it – see [HL19] for a review on that subject. Probably standing as one of the most significant breakthroughs in the early days of this research area, the classification of the simple finite-dimensional modules over the quantum affine algebra of type , , is due to Chari and Pressley [CP91]. It relies, on one hand, on a careful analysis of the -weight structure of those modules made possible by the existence of Drinfel’d’s presentation of – see [Dam93] for the proof that – and, on the other hand, on the existence of evaluation modules, proven earlier by Jimbo, [Jim86]. This seminal work paved the way for a more systematic study of the representation theory of quantum affine algebras of all Cartan types, leading to the development of powerful tools such as -characters, -characters and, consequently, to a much better understanding of the categories of their finite-dimensional modules that recently culminated with the realization that the Grothendieck rings of certain subcategories of the categories actually have the structure of a cluster algebra, [HL10].
By contrast, it is fair to say that the representation theory of quantum toroidal algebras, which were initially introduced in type by Ginzburg, Kapranov and Vasserot [GKV95] and later generalized to higher rank types, is significantly less well understood and remains, to this date, much more mysterious – although see [Her09] for a review and references therein. In our previous work, [MZ19], we constructed a new (topological) Hopf algebra , called double quantum affinization of type , and proved that its completion (in an appropriate topology) is bicontinuously isomorphic to (a corresponding completion) of the quantum toroidal algebra . Whereas is naturally graded over , where stands for the root lattice of the untwisted affine root system of type , is naturally graded over , where stands for the root lattice of the finite root system of type . Thus turns out to be to what is to , i.e. its Drinfel’d presentation. The latter, in the quantum affine case, has a natural triangular decomposition which allows one to define an adapted class of highest weight modules, namely highest -weight modules, in which finite-dimensional modules are singled out by the particular form of their highest -weights. Therefore, it is only natural to ask the question of whether plays a similar role for the representation theory of , leading, in particular, to a new notion of highest weight modules. We answer positively that question and introduce the corresponding notion of highest -weight modules. Schematically, whereas the transition from the classical Lie theoretic weights to -weights can be regarded as trading numbers for (rational) functions, the transition from -weights to -weights can be regarded as trading (rational) functions for entire modules over the non-commutative -subalgebra of . That substitution can be interpreted from the perspective of a conjecture in [MZ19], stating that is isomorphic to a split extension of the elliptic Hall algebra which was initially defined by Miki, in [Mik07], as a -analogue of the algebra and reappeared later on in different guises; the quantum continuous algebra in [FFJ+11], the Hall algebra of the category of coherent sheaves on some elliptic curve in [Sch12], or the quantum toroidal algebra associated with in [FJMM12] and in subsequent works by Feigin et al. Our conjecture is actually supported by the existence of an algebra homomorphism between and which we promote, in the present paper, to a (continuous) homomorphism of (topological) Hopf algebras. Intuitively, the weights adapted to our new triangular decomposition can therefore be regarded as representations of a quantized algebra of functions on a non-commutative 2-torus.
On the other hand, unless the value of some scalar depending on the deformation parameter is taken to be a root of unity, the question of the existence of finite-dimensional modules over quantum toroidal algebras of type was already answered negatively by Varagnolo and Vasserot in [VV96]. However, it is possible to push further the analogy with the quantum affine situation by defining another type of finiteness condition, namely weight-finiteness. It turns out that, in type , i.e. when the central charges act trivially, admits an infinite dimensional abelian subalgebra that, itself, admits as a subalgebra the Cartan subalgebra of the Drinfel’d-Jimbo quantum algebra of type . Hence, we can assign classical Lie theoretic weights to the -weight spaces of our modules and declare that a -module is weight-finite whenever it has only finitely many classical weights. The same notion is readily defined for modules over and we then focus on (resp. ), i.e. the full subcategory of the category (resp. ) of -modules (resp. -modules) whose modules are weight-finite. Of course, the widely studied category of finite-dimensional -modules is a full subcategory of . The main results of the present paper consist in showing that, on one hand, the simple objects in are all finite-dimensional and therefore coincide with the simple finite-dimensional -modules classified by Chari and Pressley, and, on the other hand, in classifying the simple objects in in terms of their highest -weight spaces. These results clearly establish as the natural quantum toroidal analogue of and suggest studying further its structure and, in particular, the structure of its Grothendieck ring. Another natural development at this point would be to generalize to the quantum toroidal setting the interesting classes of -modules outside of , for example by constructing a quantum toroidal analogue of category . We leave these questions for future work.
The present paper is organized as follows. In section 2, we briefly review classic results about the quantum affine algebra and its finite-dimensional modules. Then, we prove that simple objects in are actually finite-dimensional. In section 3, we review the main relevant results of [MZ19] and establish a few new results, as relevant for the subsequent sections. We define highest -weight modules in section 4 and, by thoroughly analyzing their structure, we establish one implication in our classification theorem, namely theorem 4.21. The opposite implication is established in section 5 by explicitly constructing a quantum toroidal analogue of the quantum affine evaluation modules. That construction is obtained after proving the existence of an evaluation homomorphism between and an evaluation algebra built as a double semi-direct product of with the completions of some Heisenberg algebras. The evaluation modules are then obtained by pulling back induced modules over the evaluation algebra along the evaluation homomorphism.
Notations and conventions
We let be the set of natural integers including . We denote by the set . For every , we denote by . We also let for every . For every , we let
denote the set of -compositions of , i.e. of compositions of having summands.
We let be defined by setting, for any ,
We assume throughout that is an algebraically closed field of characteristic and we let denote the field of rational functions over in the formal variable . As usual, we let and . Whenever we wish to evaluate to some element of , we shall always do so under the restriction that . For every , we define the following elements of
(1.1) |
We shall say that a polynomial is monic if . For every rational function , where and are relatively prime polynomials, we denote by
the Laurent series of at (resp. at ).
We shall let
for any symbols , , and provided the r.h.s of the above equations makes sense.
The Dynkin diagrams and correponding Cartan matrices of the root systems and are reminded in the following table.
Type | Dynkin diagram | Simple roots | Cartan matrix |
---|---|---|---|
2. Weight-finite modules over the quantum affine algebra
2.1. The quantum affine algebra
Definition 2.1.
The quantum affine algebra is the associative -algebra generated by
subject to the following relations
(2.1) |
(2.2) |
(2.3) |
(2.4) |
(2.5) |
(2.6) |
(2.7) |
(2.8) |
where we define the following -valued formal distributions
(2.9) |
(2.10) |
for every , we define the following -valued formal power series
(2.11) |
and
(2.12) |
is an -valued formal distribution. We denote by the subalgebra of generated by
We denote by the subalgebra of generated by
We let (resp. ) denote the subalgebra of generated by
(resp.
). We let denote the -algebra generated by the same generators as , subject to the relations (2.3 - 2.7) – i.e. we omit relation (2.8). We define the type quantum loop algebra as the quotient of by its two-sided ideal generated by . Similarly, we let and . We eventually set .
Obviously,
Proposition 2.2.
There exists a surjective -algebra homomorphism .
2.2. Finite dimensional -modules
Let be the category of -modules. We denote by the full subcategory of whose objects are finite-dimensional. Following [CP91], we make the following
Definition 2.3.
We shall say that a -module is:
-
•
a weight module if acts semisimply on ;
-
•
of type if it is a weight module and acts on as ;
-
•
highest -weight if it is of type and there exists such that
for some and . We shall refer to any such as a highest -weight vector and to as the corresponding highest -weight.
Clearly, type -modules coincide with -modules.
Definition 2.4.
For every , we construct a one-dimensional -module by setting
We then define the universal highest -weight -module with highest -weight by setting
as -modules. Let be the maximal -submodule of such that and set
By construction, is a simple highest -weight -module with highest -weight . It is unique up to isomorphisms.
The simple objects in were classified by Chari and Pressley in [CP91]. The main result is the following
Theorem 2.5 (Chari-Pressley).
The following hold:
-
i.
any simple finite-dimensional -module can be obtained by twisting a simple finite-dimensional -module of type with an algebra automorphism of ;
-
ii.
every simple finite dimensional -module of type is highest -weight;
-
iii.
the simple highest -weight module is finite-dimensional if and only if
for some monic polynomial called Drinfel’d polynomial of .
Proof.
The proof can be found in [CP91]. ∎
Up to isomorphisms, the simple objects in are uniquely parametrized by their Drinfel’d polynomials and we shall therefore denote by the (isomorphism class of the) simple -module with Drinfel’d polynomial . Note that the roles of and in the above constructions are clearly symmetrical and we could have equivalently considered lowest -weight modules. In particular, point iii of the above theorem immediately translates into
Proposition 2.6.
The simple lowest -weight module with lowest -weight is finite-dimensional if and only if
for some monic polynomial . In the latter case, we denote it by .
2.3. Weight-finite simple -modules
We now wish to consider a slightly broader family of modules over . In particular, we want to allow these modules to be infinite-dimensional, while retaining some of the nice features of finite dimensional -modules such as the fact that they decompose into -weight spaces. This is achieved by introducing the following notion.
Definition 2.7.
We shall say that a (not necessarily finite-dimensional) -module is -weight if there exists a countable set of indecomposable locally finite-dimensional -modules, called the -weight spaces of , such that, as -modules,
We shall say that is of type if acts on by .
Definition-Proposition 2.8.
Let be an -weight -module. Then:
-
i.
acts as over ;
-
ii.
for every -weight space , , of , there exists and such that
where we have set .
We let and refer to the formal power series
as the -weight of the -weight space .
Proof.
Let be an -weight space of and let . By definition, there exists a finite dimensional -submodule of such that . Over , must admit an eigenvector and, since is central, it follows that acts over by a scalar mutliple of . Assume for a contradiction that does not act by multiplication by zero. Then, it is possible to pull back into a finite-dimensional module over the Weyl algebra by the obvious algebra homomorphism . But the Weyl algebra is known to admit no finite-dimensional modules. A contradiction. It follows that acts as over . But this could be repeated for any non-zero vector in any -weight space of . i follows. As for ii, observe that, as a consequence of i and of the defining relations (2.3) and (2.4), acts by a family of commuting linear operators over . Thus ii follows from the decomposition of locally finite-dimensional vector spaces into the generalized eigenspaces of a commuting family of linear operators; the indecomposability of further imposing that it coincides with a single block in a single generalized eigenspace. ∎
Remark 2.9.
Definition 2.10.
We shall say that an -weight -module is weight-finite if is a finite set. We let denote the full subcategory of the category of -modules whose objects are weight-finite.
Clearly, finite dimensional -modules are objects in , but not every object in is in . However we have
Theorem 2.11.
The following hold:
-
i.
every simple -weight -module can be obtained by twisting a simple -weight -module of type with an algebra automorphism of ;
-
ii.
every weight-finite simple -module is highest -weight;
-
iii.
every weight-finite simple -module is finite dimensional.
Proof.
In view of definition-proposition 2.8, acts as over . Since the latter is simple and since is central, it is clear that acts over either as or as . In the former case, there is nothing to do; whereas in the latter, upon twisting as in the finite-dimensional case – see [CP91] –, we can ensure that acts as . This proves i. As for ii, the same proof as for part ii of theorem 2.5 can be used. So, we eventually prove iii. Let be a weight-finite simple -module. By ii it is highest -weight. Hence, there exists such that , and , for some with The triangular decomposition of implies that and, setting for every
(2.13) |
it is clear that
(2.14) |
is a spanning set of . The defining relations (2.5) and (2.6) of easily imply that, for every ,
(2.15) |
and, in particular,
Therefore, being weight-finite, there must exist an such that
(2.16) |
Making use of (2.8), one easily proves that, for every
where a hat over a variable indicates that that variable should be omitted. Combining (2.16) and (2.8), we get
Making use of (2.3) and (2.13), the above equation eventually yields
Acting on the l.h.s of the above equation with and making repeated use of (2.3), one easily shows that
(2.18) |
Since , its prefactor in the above equation must vanish. Now, it is clear that multiplication of the latter by annihilates all the summands with such that . Similarly, multiplication by annihilates all the summands with such that and . Repeating the argument finitely many times, we arrive at the fact that multiplication by annihilates all the summands with , so that, eventually,
Taking the zeroth order term in for , we get
Hence, the rows of the matrix on the r.h.s. of the above equation are linearly dependent and it follows that there exists a of degree at most , such that
(2.19) |
As a consequence, there clearly exists such that and
Now considering (2.3) with and multiplying it by obviously yields
(2.20) |
Set for very , . Then, (2.20), together with (2.15) for , implies that
is a strict submodule of the simple -module and it follows that for every . All the vectors in – see (2.14) – can therefore be expressed as linear combinations of the vectors in, say, and the linear span of turns out to be finite dimensional. Repeating that argument finitely many times for the linear spans of with eventually concludes the proof. ∎
Corollary 2.12.
Let be a weight-finite simple highest (resp. lowest) -weight -module. Then (resp. ), for some monic polynomial .
3. Double quantum affinization of type
3.1. Definition of
Definition 3.1.
The double quantum affinization of type is defined as the -algebra generated by
subject to the relations
and are central | (3.1) |
(3.2) |
(3.3) |
(3.4) |
(3.5) |
(3.6) |
(3.7) |
(3.8) |
(3.9) |
(3.10) |
(3.11) |
(3.12) | |||||
where , and we have set
(3.13) |
(3.14) |
and, for every and ,
(3.15) |
(3.16) |
In (3.12), we further assume that for every .
Definition 3.2.
We denote by the subalgebra of generated by
i.e. the subalgebra generated by all the generators of except and . We shall denote by
the natural injective algebra homomorphism.
Definition 3.3.
We denote by the subalgebra of generated by
and by the subalgebra of generated by
Similarly, we denote by the subalgebra of generated by . We eventually denote by (resp. ) the subalgebra of generated by
(resp.
)
Remark 3.4.
Obviously, is graded over whereas is graded over the root lattice of . is also graded over ;
where, for every , we let
Proposition 3.5.
The set generates a subalgebra of that is isomorphic to .
Proof.
This can be directly checked from the defining relations. Otherwise, it suffices to observe that the algebra isomorphism – see theorem 3.22 – restricts on that set to
∎
3.2. as a topological algebra
Because of relation (3.12), the definition of is not purely algebraic. Indeed, the r.h.s. of (3.12) involves two infinite series. One way to make sense of that relation is to equip – and, for later use, its tensor powers – with a topology, such that both series be convergent in the corresponding completion of . Making use of the natural -grading of the tensor algebras , , we let, for every ,
One easily checks that
Proposition 3.6.
The following hold true for every :
-
i.
For every , is a two-sided ideal of ;
-
ii.
For every , ;
-
iii.
;
-
iv.
;
-
v.
For every , ;
-
vi.
For every , .
Proof.
See [MZ19] for a proof in the case that can be transposed to the present situation. ∎
Definition-Proposition 3.7.
We endow with the topology whose open sets are either or nonempty subsets such that for every , for some . Similarly, we endow each tensor power with the topology induced by . These turn into a (separated) topological algebra. We then let denote its completion and we extend by continuity to all the (anti)-automorphisms defined over and its subalgebras in the previous section In particular, we extend into
Similarly, we denote with a hat the completion of any subalgebra of , like e.g. , and . We eventually denote by the corresponding completions of .
Proof.
This was proven in [MZ19]. ∎
Remark 3.8.
As was noted in [MZ19], the above defined topology is actually ultrametrizable.
3.3. The double quantum loop algebra
An alternative way to make sense of relations (3.12) consists in observing that is proalgebraic. Indeed, for every , let be the -algebra generated by
subject to relations ((3.1) – (3.12)), where, this time,
(3.17) |
Now clearly, each is algebraic since the sums on the r.h.s. of (3.12) are both finite – whenever is involved, just multiply through by to get an equivalent algebraic relation. Moreover, letting be the two-sided ideal of generated by (resp. ) for every (resp. for ), we obviously have a surjective algebra homomorphism
(3.18) |
and we can define as the inverse limit
of the system of algebras
Definition 3.9.
We shall refer to the quotient of by the two-sided ideal generated by as the double quantum loop algebra of type and denote it by . Correspondingly, we denote by and , the subalgebras of respectively generated by and
We denote by the subalgebra of generated by
It is worth emphasizing that is abelian.
3.4. Triangular decomposition of
In [MZ19], we proved that has a triangular decomposition in the following sense.
Definition 3.10.
Let be a complete topological algebra with closed subalgebras and . We shall say that is a triangular decomposition of if the multiplication induces a bicontinuous isomorphism of vector spaces .
Recalling the definitions of and from definition 3.1, we have
Proposition 3.11.
is a triangular decomposition of and is bicontinuously isomorphic to the algebra generated by subject to relation (3.11).
Proof.
See [MZ19].∎
3.5. The closed subalgebra as a topological Hopf algebra
Definition 3.12.
In , we define
(3.19) |
and for every ,
(3.20) |
(3.21) |
Then, we let be the subalgebra of generated by
and we let be its completion in the inherited topology.
Clearly, the closed subalgebra can be presented as in definition 3.3 or, equivalently, in terms of the generators in
In section 3.10, we will endow with a topological Hopf algebraic structure. It turns out that, for that structure, the closed subalgebra is not a closed Hopf subalgebra of . However, it is possible to endow with its own topological Hopf algebraic structure as follows.
Definition-Proposition 3.13.
We endow with:
-
i.
the comultiplication defined by
(3.22) (3.23) (3.24) (3.25) (3.26) for every , where and ,
-
ii.
the counit , , for every ,
-
iii.
and the antipode defined by
(3.27) (3.28) (3.29) (3.30) (3.31) where we have set, for every and every ,
and
for every .
With these operations, is a topological Hopf algebra.
Proof.
In that presentation, one readily checks that
Proposition 3.14.
is a closed Hopf subalgebra of .
Proof.
is a closed subalgebra of and it is clearly stable under and . ∎
3.6. The closed subalgebra and the elliptic Hall algebra
As emphasized in [MZ19], another remarkable feature of and, more particularly of its closed subalgebra , is the existence of an algebra homomorphism onto it, from the elliptic Hall algebra that we now define.
Definition 3.15.
Let be three (dependent) formal variables such that . The elliptic Hall algebra is the -algebra generated by , with invertible, subject to the relations
is central , | (3.32) |
(3.33) |
(3.34) |
(3.35) |
(3.36) |
(3.37) |
(3.38) |
(3.39) |
(3.40) |
where and we have introduced
(3.41) |
(3.42) |
(3.43) |
Remark 3.16.
The elliptic Hall algebra is -graded and can be equipped with a natural topology along the lines of what we did for in section 3.2. It then becomes a topological algebra and we denote by its completion. Similar topologies can be constructed on its tensor powers.
Definition-Proposition 3.17.
We endow with:
-
i.
the comultiplication defined by
(3.44) (3.45) (3.46) -
ii.
the counit defined by , ,
-
iii.
the antipode defined by
(3.47) (3.48) (3.49)
As usual, we have set and . With the above defined operations, is a topological Hopf algebra.
Proposition 3.18.
There exists a unique continuous Hopf algebra homomorphism such that
(3.50) |
(3.51) |
(3.52) |
(3.53) |
Proof.
In [MZ19], we proved that the assignment
defined an -algebra homomorphism. Hence, , which is obtained from the above assignment by rescaling the images of and , is obviously an -algebra homomorphism. Moreover, it suffices to write (3.24), (3.25) and (3.26) with , to get
as well as (3.29), (3.30) and (3.31), with , to get
and thus to prove that and as claimed. ∎
Remark 3.19.
Note that we have , meaning that descends to the quotient of by the two-sided ideal generated by . That quotient is actually Miki’s -analogue of the algebra [Mik07].
3.7. The quantum toroidal algebra
Let be a labeling of the nodes of the Dynkin diagram of type and let be a choice of simple roots for the corresponding root system. Let and let be the type root lattice.
Definition 3.20.
The quantum toroidal algebra is the associative -algebra generated by the generators
subject to the following relations
(3.54) |
(3.55) |
(3.56) |
(3.57) |
(3.58) |
(3.59) |
(3.60) |
(3.61) |
(3.62) |
where, for every , we define the following -valued formal distributions
(3.63) |
(3.64) |
for every , we define the following -valued formal power series
(3.65) |
is an -valued formal distribution,
Note that is invertible in with inverse , i.e.
(3.66) |
and that it can be viewed as the power series expansion of a rational function of as , which we shall denote as follows
(3.67) |
Observe furthermore that we have the following useful identity in
(3.68) |
Remark 3.21.
In type , , and we have an additional identity, namely .
is obviously a -graded algebra, i.e. we have
(3.69) |
It was proven in [Her05] to admit a triangular decomposition , where and are the subalgebras of respectively generated by and
Observe that admits a natural gradation over that we shall denote by
(3.70) |
Of course is graded over the root lattice . We finally remark that the two Dynkin diagram subalgebras and of generated by
with and respectively, are both isomorphic to , thus yielding two injective algebra homomorphisms . In [MZ19], making use of their natural -grading, and all its tensor powers were endowed with a topology along the lines of what we did in section 3.2 for and its tensor powers, and subsequently completed into and . The main result in [MZ19] is the following
Theorem 3.22.
There exists a unique bicontinuous -algebra isomorphism such that
Proof.
See [MZ19] for a proof. ∎
3.8. subalgebras of
Interestingly, admits countably many embeddings of the quantum affine algebra . This is the content of the following
Proposition 3.23.
For every , there exists a unique injective algebra homomorphism such that
(3.71) |
(3.72) |
(3.73) |
Proof.
See [MZ19]. ∎
We also have
Proposition 3.24.
For every , is an injective algebra homomorphism.
Proof.
This is obvious since is an isomorphism and is an injective algebra homomorphism. ∎
3.9. (Anti-)Automorphisms of
naturally inherits, through , all the continuous (anti-)automorphisms defined over .
Proposition 3.25.
Conjugation by clearly provides a group isomorphism . In particular, for every , we let .
As an example, consider the Cartan anti-involution of defined in [MZ19]. It extends by continuity into an anti-involution over which eventually yields, upon conjugation by , an anti-involution over . One can easily check – or take as a definition of the fact – that,
for every and every .
In addition to the above, also admits the following automorphisms that will prove useful in the study of its representation theory.
Proposition 3.26.
-
i.
There exists a unique -algebra automorphism of such that, for every and every ,
-
ii.
There exists a unique -algebra automorphism of such that
Proof.
It suffices to check the defining relations of . ∎
3.10. Topological Hopf algebra structure on
Definition 3.27.
We endow the topological -algebra with:
-
i.
the comultiplication defined by
(3.74) (3.75) (3.76) (3.77) where and ;
-
ii.
the counit , defined by , and;
-
iii.
the antipode , defined by , and
With these operations so defined and the topologies defined in section 3.7, is a topological Hopf algebra.
In view of theorem 3.22, it is clear that inherits that topological Hopf algebraic structure.
Definition-Proposition 3.28.
We define
(3.78) |
(3.79) |
(3.80) |
Equipped with the above comultiplication, antipode and counit, is a topological Hopf algebra.
Before we move on to introducing -weight -modules, we give the following
Lemma 3.29.
For every and every , we have
-
i.
;
-
ii.
;
where we have set and .
Proof.
We first prove i for upper choices of signs. Observe that (3.20) equivalently reads
for every . For every , let
In [MZ19] – see proposition-definition 4.9, definition 4.25 and eq. (4.66) –, we proved that and that, for every ,
where can be recursively defined by setting
(3.81) |
and
(3.82) | |||||
Hence, i for is clear. From (3.81) and definition 3.27, and making use of relations (3.58) and (3.59) as well as of the identity (3.68), we deduce that
where . Applying to the first two terms obviously yields . Since, on the other hand, – see theorem 3.22 –, applying to the third term yields an element of and it follows that i holds for and for upper choices of signs. Suppose it holds for upper choices of signs and for some . Then, making use of (3.82), one easily checks that i holds for and for upper choices of signs, which completes the proof of i for upper choices of signs. Now, i for lower choices of signs follows after applying and observing that, indeed,
As for ii, we let, for every ,
In [MZ19] – see definition 4.1 and proposition 4.8 –, we proved that could be defined recursively by setting and letting, for every ,
(3.83) |
and
(3.84) |
where – see proposition 4.3 in [MZ19]. Observing that , we clearly get
Now, applying to (3.76) in definition 3.27 clearly proves ii in the case . Assuming it holds for , it suffices to apply to (3.83) above to prove that it also holds for . Similarly, if ii holds for some , applying to (3.84) to prove that it also holds for . This concludes the proof. ∎
4. -weight -modules
4.1. -weight modules over
Remember that contains a subalgebra that is isomorphic to – see proposition 3.5. Hence, in view of remark 2.9, we can repeat for modules over what we did in section 2.3 for modules over . We thus make the following
Definition 4.1.
We shall say that a (topological) -module is -weight if there exists a countable set of indecomposable locally finite-dimensional -modules called -weight spaces of , such that, as -modules,
As in section 2.3, it follows that
Definition-Proposition 4.2.
Let be an -weight -module. Then:
-
i.
acts on by ;
-
ii.
for every -weight space , , of , there exist and sequences such that
(4.1) where we have set .
We let and we shall refer to
as the -weight of the -weight space . We shall say that is
-
•
of type if acts by over ;
-
•
of type for if it is of type and, for every , acts by multiplication over ;
-
•
of type if it is of type and acts by over .
Proof.
The proof follows the same arguments as the proof of definition-proposition 2.8. ∎
Proposition 4.3.
Let be a type -weight -module and let and be two -weight spaces of such that, for some and some , . Then, there exists a unique such that:
-
i.
the respective -weights and of and be related by
where and
(4.2) -
ii.
for some .
Proof.
There clearly exist two bases and of and respectively, in which
for some , with and , such that for every and for every .
Now, if , there must exist a largest nonempty subset such that, for every , . Let . Obviously, for every , there must exist a largest nonempty subset such that, for every and every , . Let and let for simplicity . Then, for every ,
for some . When needed, we shall extend by zero the definition of outside of the set of pairs . Making use of the relations in – namely (3.7) and (3.8) –, we get, for ,
The latter easily implies that, for every and every ,
(4.3) |
Taking and in the above equation immediately yields
The latter is equivalent to the fact that, for every ,
(4.4) |
where, as usual, we have set
Since , there must exist a such that . Assuming that , one easily obtains that, on one hand and that, on the other hand,
A contradiction. By similar arguments, one eventually proves that for every . But then dividing (4.1) by we get
where we have set, for every , and, consequently, . i follows. Moreover, we clearly have
for some . More generally, we claim that,
(4.5) |
for some and some . This is proven by a finite induction on and . Indeed, making use of (4.2), we can rewrite (4.1) as
(4.6) |
for every and every . Now, assume that (4.5) holds for every pair in
for some and some such that . Let be the smallest element of such that . It suffices to write (4.1) for and , to get
(4.7) |
Combining the recursion hypothesis and lemma A.1 from the appendix, one easily concludes that (4.5) holds for the pair . Repeating the argument finitely many times, we get that it actually holds for all the pairs in . Now, either and we are done; or and there exists a largest such that . Writing (4.1) for and , we get
Combining again the recursion hypothesis and lemma A.1, we easily get that (4.19) holds for . It is now clear that the claim holds for every and every . Letting , ii follows. Furthermore, for every and every , we obviously have , thus making the only element of satisfying ii. This concludes the proof. ∎
We let denote the fundamental weight of and we let be the corresponding weight lattice. In view of proposition 4.3, it is natural to make the following
Definition 4.4.
Let be a type -weight -module and let be the countable set of its -weight spaces. We shall say that is rational if, for every , there exist relatively prime monic polynomials , called Drinfel’d polynomials of , such that the -weight of be given by
With each rational -weight of a rational -module , we associate an integral weight , by setting
We shall say that is -dominant (resp. -anti-dominant) if it is rational and there exists such that, for every , and (resp. and ).
Remark 4.5.
The classical weight (resp. ) associated with any -dominant (resp. -anti-dominant) type -weight rational -module is a dominant (resp. anti-dominant) integral weight. Note that the converse need not be true.
Remark 4.6.
The data of the -weights of a rational -module is equivalent to the data of its Drinfel’d polynomials which, in turn, is equivalent to the data of their finite multisets of roots . The latter are finitely supported maps such that, for every ,
Note that, in the above formulae, since is finitely supported, the products only run through the finitely many numbers in the support of . Moreover, since and are relatively prime for every , we have . We denote by , the set of finitely supported -valued maps over . As is customary in the theory of -characters, we associate with every -weight given by a pair of Drinfel’d polynomials or, equivalently, by a pair with and , a monomial
Definition 4.7.
Let be an -dominant -module and let and be any two -weight spaces of with respective -weights
where are two monic polynomials. By proposition 4.3.i., if for some , then there exists a unique such that
where . We shall say that is -dominant if, under the same assumptions, we have, in addition, that
For every , we let be defined by
For every , we let and we define, for every , an operator by letting 111Although the definition of easily extends to , we will not make use of that extension and exclusively regard as a map ., for every ,
is obviously invertible, with inverse given by . Note that, for every , over . Given two finite multisets , we we shall say that they are equivalent and write iff
(4.8) |
for some , and some . In writing (4.8), it is assumed that, for every , . It is clear that is an equivalence relation and we denote by the equivalence class of in . Following remark 4.6, we naturally extend the action of to , by setting
The equivalence relation similarly extends from to . Note that, setting
for every and every , we have, for every
Corollary 4.8.
Let be a simple -dominant -module. Then there exists a multiset such that all the monomials associated with the -weights of be in the equivalence class of .
Proof.
By proposition 4.3, for any two -weight spaces, and , of an -dominant -module , with respective -weights
if for some and some , then we must have
(4.9) |
for some . Now, assuming , it is clear that:
-
-
for the upper choice of sign on the right hand side of the above equation, the last fraction line must completely cancel against factors in the first one, whereas the second one survives, eventually replacing the cancelled factors;
-
-
for the lower choice of sign, the second fraction line must cancel against factors in the first one, whereas the last one survives, eventually replacing the cancelled factors.
If on the other hand , since is -dominant, we have, by definition, that is a root of . In any case, denoting by (resp. ) the multiset of roots of (resp. ), it is clear that and hence . Since is simple, there can be no non-zero -weight space of such that for every -weight space of , every and every . ∎
In view of definition-proposition 4.2, we can make the following
Definition 4.9.
For every monic polynomial , denote by the one-dimensional -module such that
for every . There exists a universal -module that admits the -weight associated with . Denoting by the maximal -submodule of such that , we define the unique – up to isomorphisms – simple -module .
Proposition 4.10.
For every simple -dominant -module , there exists a monic polynomial such that .
Proof.
Obviously, for every , we have . Now since is -dominant, can be chosen as an -weight vector, i.e.
for some monic polynomial . ∎
Remark 4.11.
The above proof makes it clear that if is the set of Drinfel’d polynomials of a simple -dominant -module , then, for every , .
Theorem 4.12.
For every monic polynomial , is -dominant.
Proof.
We postpone the proof of this theorem until section 5, where we construct for every and directly check that it is indeed -dominant. ∎
Proposition 4.13.
Any topological -module pulls back to a module over the elliptic Hall algebra .
Proof.
Remark 4.14.
It is worth mentioning that, as an example of the above proposition, -anti-dominant -modules pullback to a family of -modules that were recently introduced in [DK19]. It might be interesting to investigate further the class of -modules obtained by pulling back other (rational) -modules.
We conclude the present subsection by proving the following
Lemma 4.15.
Let be an -dominant -module. Suppose that, for any two -weight spaces and of , with respective -weights and , such that , the unique such that , for every , and for some – see proposition 4.3 – also satisfies . Then is -dominant.
Proof.
Let be as above and let and be two -weight spaces of with respective -weights
Suppose that for some . If , writing down , we obtain equation (4.9) as in the proof of corollary 4.8. By the same discussion as the one following equation (4.9), we conclude that , as needed – see definition 4.7. Finally, if , writing down , we obtain
Then, it is clear that:
-
•
for the upper choice of sign on the right hand side of the above equation, the last fraction line must completely cancel against factors in the first one, whereas the second one survives, eventually replacing the cancelled factors;
-
•
for the lower choice of sign on the right hand side of the above equation, the second fraction line must completely cancel against factors in the first one, whereas the last one survives, eventually replacing the cancelled factors.
In any case, it follows that . But by our assumptions on , we also have that and the -dominance of follows – see definition 4.7. ∎
4.2. -weight -modules
Definition 4.16.
For every , we shall say that a (topological) module over is of type if:
-
i.
acts as on ;
-
ii.
acts by multiplication by on , for every .
We shall say that is of type if points i. and ii. above hold for every and, in addition, acts as on .
Remark 4.17.
Let . Then the -modules of type are in one-to-one correspondence with the -modules – see section 3.3 for a definition of . Obviously -modules of type descend to modules over the double quantum loop algebra of type , .
Definition 4.18.
We shall say that a (topological) -module is a -weight module if there exists a countable set of indecomposable -weight -modules, called -weight spaces of , such that, as (topological) -modules,
(4.10) |
We shall say that is weight-finite if, regarding it as a completely decomposable -module, its is finite – see definition-proposition 4.2 for the definition of . A vector is a highest -weight vector of if for some and, for every ,
(4.11) |
We shall say that is highest -weight if for some highest -weight vector .
Definition-Proposition 4.19.
Let be a -weight -module that admits a highest -weight vector . Denote by the -weight space of containing . Then and, for every ,
(4.12) |
We shall say that is a highest -weight space of . If in addition is simple, then it admits a unique – up to isomorphisms of -modules – highest -weight space .
Proof.
It is an easy consequence of the triangular decomposition of – see proposition 3.11 – and of the root grading of that, indeed, , for every . Now since is highest -weight, we have . By proposition 3.11, and it follows that . Now, assuming that is simple and that it admits highest -weight spaces and , we have that as -modules. In particular, as -modules. ∎
In view of the triangular decomposition of – see proposition 3.11 –, the above proposition implies that any highest -weight -modules is entirely determined as , by the data of its highest -weight space , a cyclic -weight -module. Now for any such that , let be the maximal -submodule of not containing and set 222 clearly does not depend on the chosen generator . Indeed, if contained a generator of , it would contain all the others, including . It follows that and hence are both independent of .. Then, by construction, is a simple -module such that, as -modules, . We therefore make the following
Definition 4.20.
We extend every simple (topological) -weight -module into a -module by setting for every . This being understood, we define the universal highest -weight -module with highest -weight space by setting
as -modules. Denoting by the maximal (closed) -submodule of such that , we define the simple highest -weight -module with highest -weight space by setting . It is unique up to isomorphisms.
Classifying simple highest -weight -modules therefore amounts to classifying those simple -weight -modules that appear as their highest -weight spaces. In the case of weight-finite -modules, this is achieved by the following
Theorem 4.21.
The following hold:
-
i.
Every weight-finite simple -module is highest -weight and can be obtained by twisting a type (1,0) weight-finite simple -module with an algebra automorphism from the subgroup of generated by the algebra automorphisms and of proposition 3.26.
-
ii.
The type (1,0) simple highest -weight -module is weight-finite if and only if its highest -weight space is a simple -dominant -module – see proposition-definition 4.4.
Proof.
Let be a weight-finite simple -weight -module and assume for a contradiction that, for every , there exist such that . Then, there must exist two sequences , such that
Choosing to be an eigenvector of with eigenvalue – see definition-proposition 4.2 for the existence of such a vector –, one easily sees from the relations that, for every , . It follows – see definition-proposition 4.2 – that . A contradiction with the weight-finiteness of . Thus, we conclude that there exists a highest -weight vector such that for some . Obviously, , for is a submodule of the simple -module . Thus is highest -weight. Denote by its highest -weight space. The latter is an -weight -module. As such, it completely decomposes into countably many locally finite-dimensional indecomposable -modules that constitute its -weight spaces. Over any of these, must admit an eigenvector. But since is simple and is central, the latter acts over by a scalar multiple of . It follows from definition-proposition 4.2 that C acts over by or . In the former case, there is nothing to do; whereas in the latter, it is quite clear from proposition 3.26 that, twisting the action on by , we can ensure that C acts by . It follows that acts by or . Again, in the former case, there is nothing to do; whereas in the latter, twisting by , we can ensure that acts by . Similarly, for every , must admit an eigenvector over any locally finite-dimensional -weight space of . But again, since is simple and is central, the latter must act over by a scalar multiple of .
In any case, in view of (3.7) and (3.8), commutes with all the other generators of and, since , we have for every . Moreover, turns out to be a type 1 -weight -module and, by definition-proposition 4.2,
for some countable set of sequences . By proposition 4.19,
(4.13) |
for every . Pulling back with and respectively, we can simultaneously regard as a -module for both of its Dynkin diagram subalgebras and – see section 3. Let be a simultaneous eigenvector of the pairwise commuting linear operators in . Equation (4.13) implies that . Thus is a highest (resp. lowest) -weight vector of (resp. ). The weight finiteness of now allows us to apply corollary 2.12 to prove that the respective simple quotients of and containing are both finite-dimensional and isomorphic to a unique simple highest (resp. lowest) -weight module (resp. ). As a consequence of theorem 2.5 and of proposition 2.6, we conclude that
for some monic polynomials and . On the other hand, pulling back with for every , we can regard as a -module in infinitely many independent ways. Again, for every , turns out to be a highest -weight vector for a unique simple weight finite, hence finite dimensional -module. As such, it satisfies
for some monic polynomial . Now since
and , we must have
(4.14) | |||||
for every . In the limit as , this implies for every and, consequently, . After obvious simplifications, (4.14) becomes
(4.15) |
for every . Now, is not a root of for any monic polynomial . Moreover, being a formal parameter – in case is regarded as a complex number, we shall assume that –, it follows that the map has no fixed points over the set of roots of a monic polynomial. Thus, for large enough, the respective sets of roots of and are disjoint. Similarly, for large enough, the respective sets of roots of and are disjoint. It follows that, for large enough, on the r.h.s. of (4.15), cancellations can only occur between factors on opposite sides of the same fraction line. Now, either or , in which case
for some such that and some -tuples , such that
But then, we should have, for large enough,
where, on the r.h.s., cancellations can only occur across the leftmost fraction line. A contradiction. i follows. As for part of ii, we shall prove it in section 5. ∎
Although we must postpone the proof of part ii of theorem 4.21, the proof above still makes it clear that
Proposition 4.22.
If a type (1,0) simple highest -weight -module is weight-finite, then its highest -weight space is a simple -dominant -module.
Proposition 4.23.
Let be a -weight -module and let and be two -weight spaces of such that, for some , . Then, there exists a unique such that:
-
i.
the respective -weights and of and be related by
(4.16) where and
-
ii.
for some .
Proof.
We keep the same notations as in the proof of proposition 4.3. More specifically, we have two bases and of and respectively, in which
for some , with and , such that for every and for every .
Now, if , there must exist a largest nonempty subset such that, for every , . Let . Obviously, for every , there must exist a largest nonempty subset such that, for every and every , . Let and let for simplicity . Then, for every ,
for some . When needed, we shall extend by zero the definition of outside of the set of pairs . Making use of the relations in – namely (3.9) and (3.10) –, we get, for every and every ,
The latter easily implies that, for every and every ,
(4.17) |
Taking and in the above equation immediately yields
The latter is equivalent to the fact that, for every ,
(4.18) |
where, as usual, we have set
Since , there exists at least one - such that . Assuming that , one easily derives a contradiction from (4.18) and, repeating the argument, one proves that for every . Dividing (4.18) by , one gets
where we have set, for every , . i. now follows. Moreover, we clearly have
for some . More generally, we claim that,
(4.19) |
for some and some . This is proven by a finite induction on and . Indeed, making use of (4.16), we can rewrite (4.17) as
(4.20) |
for every and every . Now, assume that (4.19) holds for every pair in
for some and some such that . Let be the smallest element of such that . It suffices to write (4.20) for and , to get
(4.21) | |||||
Combining the recursion hypothesis and lemma A.1 from the appendix, one easily concludes that (4.19) holds for the pair . Repeating the argument finitely many times, we get that it actually holds for all the pairs in . Now, either and we are done; or and there exists a largest such that . Writing (4.20) for and , we get
Combining again the recursion hypothesis and lemma A.1, we easily get that (4.19) holds for . It is now clear that the claim holds for every and every . Letting , ii. follows. Furthermore, for every and every , we obviously have , thus making the unique element of satisfying ii.. ∎
Corollary 4.24.
The -weights of any type weight-finite simple -module are all rational – see definition 4.4.
Proof.
Let be a type weight-finite simple -module. By proposition 4.22, its highest -weight space is an -dominant simple -module. Hence, and it easily follows by proposition 4.23 that all the -weights of are of the form
for some , some and
for some monic polynomial . Now, observe that
Hence, all the -weights of are of the form
(4.22) |
for some relatively prime monic polynomials , which concludes the proof. ∎
In view of remark 4.6, we can therefore associate with any weigh-finite simple -module a -character defined as the (formal) sum of the monomials corresponding to all its rational -weights.
Proposition 4.25.
Let and be two -dominant simple -modules such that be simple. Then:
-
i.
is a simple -dominant -module of type ;
-
ii.
there exists a short exact sequence of -modules
-
iii.
if, in addition, is simple, then
Proof.
and are both of type and (3.22) and (3.23) respectively imply that so is . Similarly, they are both -weight and -dominant. Combining eqs. (3.19), (3.20), (3.21), (3.24), (3.25) and (3.26), we easily prove that
(4.23) |
(4.24) |
for every . In particular, taking , we have . It follows that, if and are the countable sets of -weights of and respectively, with respective Drinfel’d polynomials and , then is the countable set of -weights of . Moreover, the latter is obviously -dominant since its Drinfel’d polynomials are in . Now let , and let , , and be the Drinfel’d polynomials of , , and respectively and assume that
(4.25) |
Then, writing (4.23) and (4.24) above with , we get
Since both and are -weight spaces, it follows that
Therefore, condition (4.25) holds only if the direct sum on the r.h.s. above has a non-vanishing intersection with . But since the latter is an -weight space, this happens only if either or . The -dominance of and implies that for the only such that , either or . In any case, and is -dominant. i follows. By lemma 3.29, it is clear that . Hence is a highest -weight space in . Let denote the largest closed -submodule of such that . ii obviously follows. iii is clear. ∎
5. An evaluation homomorphism and evaluation modules
In this section, we construct an evaluation algebra and an -algebra homomorphism , that we shall refer to as the evaluation homomorphism.
5.1. The quantum Heisenberg algebras and
Definition 5.1.
The quantum Heisenberg algebra is the Hopf algebra generated over by
subject to the relations,
for every , with comultiplication defined by setting
for every , antipode defined by setting
and counit defined by setting
Definition 5.2.
In , we let
Similarly, in , we let
Then, we have the following equivalent presentation of .
Proposition 5.3.
is the Hopf algebra generated over by
subject to the relations
where we have defined , by setting
Furthermore, we have
where, by definition,
and
Finally, .
Proof.
This is an easy consequence of the definition of . ∎
Remark 5.4.
Observe that and are not independent and that we actually have .
5.2. A PBW basis for
For every , we let denote the set of -partitions. We adopt the convention that reduces to the empty partition and we let be the set of all partitions.
Proposition 5.5.
Define, for every ,
(5.1) |
(5.2) |
with the convention that . Then,
(5.3) |
is a -basis for .
Proof.
The relations in read, for every ,
where, for every , can be obtained from
It is clear that any monomial in can therefore be rewritten as a linear combination with coefficients in of elements in . The independence of the latter is clear. ∎
A convenient way to encode the above basis elements is through -valued symmetric formal distributions. Let indeed, for every , every -tuple and every -tuple of formal variables,
where we have set
with the convention that if (resp. ), then (resp. ). It turns out that
Indeed, owing to the commutation relations in , the formal distribution is symmetric in each of its argument tuples, and respectively; i.e. it is invariant under the natural action of on its arguments. It is also clear that, for every and ,
where we have set
5.3. The dressing factors and
Definition 5.6.
For every , we let
(5.4) |
(5.5) |
It easily follows that
Proposition 5.7.
In , for every , we have
where we have set
Furthermore, we have, for every ,
It is worth emphasizing that the are not indepedent for all values of and that neither are the . Indeed, we have
Lemma 5.8.
For every ,
(5.6) |
(5.7) |
(5.8) |
(5.9) |
5.4. The algebra
Remember the Hopf algebra from definition 3.20. It has an invertible antipode and we denote by its coopposite Hopf algebra.
Proposition 5.9.
The quantum Heisenberg algebra (resp. ) is a left -module algebra (resp. a left -module algebra) with
for and where we have set
Proof.
One readily checks the compatibility with the defining relations of and . ∎
Proposition 5.10.
For every and every , we have
where we have set
and
Proof.
This is readily checked making use of definition 5.6, of the Hopf algebraic structures of and , of the -module algebra structures of and of the -module algebra structure of . ∎
Definition-Proposition 5.11.
We denote by the associative -algebra obtained by endowing with the multiplication given by setting, for every , every and every ,
– see proposition 5.10 for the definition of the -module algebra structure of and of the -module algebra structure of . In that algebra, generates a left ideal. The latter is actually a two-sided ideal since is central and, denoting it by , we can set .
Proof.
Making use of the coassociativity of the comultiplication , it is very easy to prove that, with the above defined multiplication, is actually an associative -algebra. ∎
Proposition 5.12.
Setting , for every , defines a unique injective -algebra homomorphism . Similarly, and define unique injective -algebra homomorphisms and respectively.
Remark 5.13.
We shall subsequently identify , and with their respective images in under the injective algebra homomorphisms of the above proposition.
Proposition 5.14.
In , for every and every , we have the following relations
(5.10) |
(5.11) |
(5.12) |
(5.13) |
(5.14) |
(5.15) |
(5.16) |
Proof.
Remark 5.15.
In addition to the above, we obviously have in , all the relations of its subalgebra and all the relations of its subalgebras and modulo .
Definition-Proposition 5.16.
Let be the left ideal of generated by
Then and is a two-sided ideal of . Set .
Proof.
Remark 5.17.
Thus, in addition to the relations in , we have, in ,
5.5. The completion of
Making use of its natural -grading, we endow with a topology, in the same way as we endowed with its topology in section 3. We denote by the corresponding completion. Consequently, its subalgebra inherits a topology and we denote by its corresponding completion in that topology.
5.6. The shift factors
Definition 5.18.
In , we define,
Similarly, for every , we let
Lemma 5.19.
For every ,
(5.17) |
(5.18) |
Proof.
Follows directly from the definition in the same way as lemma 5.8. ∎
Proposition 5.20.
In , we have, for every ,
where
Proof.
Proposition 5.21.
For every and every , we have
(5.19) |
(5.20) |
Proof.
The left -module algebra (resp. a left -module algebra) structure of (resp. ) – see proposition 5.9 – is extended by continuity to (resp. ) Then, it suffices to check that, for every and every ,
and that
∎
Corollary 5.22.
For every , every and every , we have
Proof.
It suffices to differentiate (5.19) times with respect to to obtain
The claim immediately follows. ∎
Proposition 5.23.
In , we have, for every ,
Proof.
This follows immediately from . ∎
5.7. The evaluation algebra
Definition-Proposition 5.24.
Let denote the closed left ideal of generated by
(5.21) |
Then, , making a closed two-sided ideal of , and we let .
Proof.
Proposition 5.25.
For every , the following relation holds in ,
(5.22) |
Proof.
Remark 5.26.
In addition to the above relation, obviously inherits the relations in modulo . In particular, all the relations in proposition 5.14 hold in .
5.8. The evaluation homomorphism
Proposition 5.27.
There exists a unique continuous -algebra homomorphism such that, for every and every ,
(5.23) |
(5.24) |
(5.25) |
(5.26) |
We shall refer to as the evaluation homomorphism. It is such that over .
Proof.
We have the following obvious
Corollary 5.28.
For every there exists an algebra homomorphism making the following diagram commutative.
We can furthermore define the algebra homomorphism by
5.9. Evaluation modules
Remember the surjective algebra homomorphism from proposition 2.2. It allows us to pull back any simple -module into a simple -module. With that construction in mind, we have
Proposition 5.29.
Let be a simple finite dimensional -module. Then,
-
i.
is a -module with the action defined by setting, for every , every , every and every ,
and extending by continuity.
-
ii.
descends to a -module.
-
iii.
is an -module. It pulls back along to a -module that we denote by .
-
iv.
As a -module, is weight-finite.
-
v.
For any highest -weight vector , the -module
is a highest -weight space of . We denote by the simple quotient of containing and we let .
-
vi.
is -dominant.
Proof.
i is readily checked. As for ii, it suffices to check that . But the latter is clear when is obtained by pulling back a -module over which the relation generating is automatically satisfied. iii is obvious. It easily follows from proposition 5.23 that, for every , . Hence, and the weight finiteness of follows from that of , which proves iv. It is clear that, for every , we have
v follows. Denote by the Drinfel’d polynomial associated with and let denote the multiset of its roots. Then,
(5.28) |
Moreover, the partial fraction decomposition
in which and the product and sum over are always finite since only has finitely many roots, allows us to write
Letting for every and every , it follows that, for every ,
(5.29) |
(5.30) |
Now, making use of (5.24), (5.28) and of corollary 5.22, one easily shows that, for every and every ,
thus proving that (resp. ) is an -weight vector in the -weight space (resp. ) of with -weight (resp. ), as expected from proposition 4.3.
On the other hand,
Thus, modulo , we have, for every ,
The above equation makes it clear that every such that is a zero of order at least of , unless . Hence, in view of (5.30), we have unless . But the latter implies that . A similar reasoning applies to any -weight vector in and is -dominant by lemma 4.15. Taking the quotient of to clearly preserves -dominance and vi follows. ∎
By the universality of – see definition 4.20 – and the above proposition, there must exist a surjective -module homomorphism . Restricting the latter to the (closed) -submodule of , we get the surjective -module homomorphism , whose image naturally injects as a -submodule in . The canonical short exact sequence involving , and the simple quotient – see definition 4.20 – allows us to define a surjective -module homomorphism to get the following commutative diagram,
where columns and rows are exact. It is obvious that is not identically zero and, by the simplicity of , we must have . Hence, is a -module isomorphism and we have constructed the simple weight-finite -modules as a quotient of the evaluation module . To see that all the simple weight-finite -modules can be obtained in this way, it suffices to observe that, by proposition 4.10, all the simple -dominant -modules are of the form for some monic polynomial and that, in the construction above, one can choose any , simply by choosing the corresponding simple finite-dimensional -module . Therefore, as a consequence of the above proposition, the highest -weight space of any simple weight-finite -modules is -dominant. This concludes the proof of part ii of theorem 4.21 as well as that of theorem 4.12.
Appendix A A lemma about formal distributions
In this short appendix, we prove the following
Lemma A.1.
Let and , let be a non-zero formal power series and let be a formal distribution such that
(A.1) |
for some non-zero scalar and some formal power series . Then,
for some scalars .
Proof.
Consider first the case where . Then, multiplying (A.1) by , we get
Since , we can specialize at a non-zero -mode and it follows that
for some scalars . Now consider the case where . It follows from (A.1) that
for some formal distribution . But specializing the l.h.s. of the above equation to any -mode of the form with , we immediately get that . We are thus back to the previous case. ∎
References
- [CP91] V. Chari and A. Pressley, Quantum affine algebras, Comm. Math. Phys. 142 (1991), 261–283.
- [Dam93] I. Damiani, A basis of type PBW for the quantum algebra of , Journal of Algebra 161 (1993), 291–310.
- [DK19] P. Di Francesco and R. Kedem, (t, q)-deformed Q-systems, DAHA and quantum toroidal algebras via generalized Macdonald operators, Comm. Math. Phys. 369 (2019), no. 3, 867–928.
- [FFJ+11] B. Feigin, E. Feigin, M. Jimbo, T. Miwa, and E. Mukhin, Quantum continuous : Semiinfinite construction of representations, Kyoto J. Math. 51 (2011), no. 2, 337–364.
- [FJMM12] B. Feigin, M. Jimbo, T. Miwa, and E. Mukhin, Quantum toroidal -algebra: Plane partitions, Kyoto J. Math. 52 (2012), no. 3, 621–659.
- [GKV95] V. Ginzburg, M. Kapranov, and E. Vasserot, Langlands reciprocity for algebraic surfaces, Mathelatical Research Letters 2 (1995), 147–160.
- [Her05] D. Hernandez, Representations of quantum affinizations and fusion product, Transformation groups 10 (2005), no. 2, 163–200.
- [Her09] by same author, Quantum toroidal algebras and their representations, Selecta Mathematica 14 (2009), 701–725.
- [HL10] D. Hernandez and B. Leclerc, Cluster algebras and quantum affine algebras, Duke Math. J. 154 (2010), no. 2, 265–341.
- [HL19] by same author, Quantum affine algebras and cluster algebras, arXiv:1902.01432 (2019).
- [Jim86] M. Jimbo, A q-analogue of , Hecke algebra, and the Yang-Baxter equation, Lett. Math. Phys. 11 (1986), no. 3, 247–252.
- [Mik07] K. Miki, A -analog of the algebra, J. Math. Phys. 48 (2007), 123520.
- [MZ19] E. Mounzer and R. Zegers, On double quantum affinization: 1. type , arXiv:1903.00418 (2019).
- [Sch12] O. Schiffmann, Drinfeld realization of the elliptic Hall algebra, J. Algebr. Comb. 35 (2012), 237–262.
- [VV96] M. Varagnolo and E. Vasserot, Schur duality in the toroidal setting, Comm. Math. Phys. 182 (1996), no. 2, 469–483.