K-theoretic Heisenberg algebras and permutation-equivariant Gromov–Witten theory
Abstract.
We found an interesting application of the K-theoretic Heisenberg algebras of Weiqiang Wang to the foundations of permutation equivariant K-theoretic Gromov–Witten theory. We also found an explicit formula for the genus 0 correlators in the permutation equivariant Gromov–Witten theory of the point. In the non-equivariant limit our formula reduces to a well known formula due to Y.P. Lee.
1. Introduction
K-theoretic Gromov–Witten (KGW) theory was introduced by Givental (see [4]) and Y.P. Lee (see [14]) as a generalization of the cohomological Gromov–Witten (GW) theory. One of the fundamental open problems in KGW theory is to compute the KGW invariants of the point. The genus-0 and genus-1 invariants were computed respectively in [13] and [12]. In the cohomological case, it is well known that the GW invariants of the point are governed by the KdV hierarchy. We expect that the KGW invariants of the point are also governed by an integrable hierarchy but so far even a conjectural description is missing. Let us point out however, that in the genus-0 case, the KGW invariants of the point are governed by a hierarchy which is a Miura transform of the dispersionless KdV hierarchy (see [15]). Therefore, it is quite possible that in higher-genus, the KGW invariants of the point are again solutions to the KdV hierarchy but after some complicated Miura transformation.
There is a very interesting recent development in KGW theory which to some extend shows that the definition of the KGW invariants must be extended. Namely, Givental observed that in the settings of toric geometry the so-called quantum Lefschetz hyperplane principle fails. In particular, many natural constructions in mirror symmetry, such as, constructing the mirror of a toric hypersurface, will fail in KGW theory too. In order to resolve these issues, Givental proposed an extension of KGW theory which is now called permutation-equivariant K-theoretic Gromov–Witten theory (see [5]). We believe that permutation-equivariant KGW invariants have better properties because one can use mirror symmetry to compute them. In particular, if we view the point as a hyperplane in , we can construct a 2-dimensional mirror model for the permutation-equivariant KGW invariants of the point. We would like to use this mirror model and ideas from the Eynard–Orantin recursion to compute the permutation-equivariant KGW invariants of the point. The first step in our project is to compute the genus-0 parmutation-equivariant KGW invariants of the point, that is, we would like to know the analogue of Y.P. Lee’s formula [13]. In some sense, this is the main motivation for this paper. Our computation is based on the techniques developed in [15]. Namely, we prove that the genus-0 permutation-equivariant KGW invariants are governed by an integrable hierarchy of hydrodynamic type. Using the differential equations we were able to find a closed formula for the genus-0 permutation-equivariant KGW invariants of the point (see Theorem 1.3.1). Let us point out that we do not work with the most general version of permutation-equivariant KGW theory, i.e., we do not allow descendants at the permutable marked points. Nevertheless, the generality that we consider should be enough for the applications that we have in mind.
1.1. K-theoretic Fock spaces
The definition of permutation-equivariant KGW invariants will be recalled later on (see Section 3.2). From the very beginning it is clear that we are dealing with representations of the symmetric groups. However, finding a good algebraic formalism to capture the entire information is a bit tricky. We refer to [5] for more details on the logic used by Givental. One of the surprises in the current paper (at least to the author) is that there is a slightly different way to organize the invariants which is equivalent to the choice made by Givental but proving the equivalence requires some non-trivial work.
Our logic is the following. Suppose that is a smooth projective variety. Let be the moduli space of degree stable maps from a genus- nodal Riemann surface equipped with marked points. We let the symmetric group act on the moduli space by permuting the last marked points. We will also refer to the last marked points as permutable marked points. Let
and
be the evaluation maps at respectively the first and the last marked points. The permutation-equivariant KGW invariants contain information about the -equivariant K-theoretic pushforward . Therefore, it is natural to introduce the following graded vector space:
where is the symmetric group and denotes the -equivariant topological K-ring with complex coefficients (see [17]). We will consider the case when in order to make the construction more transperant. The case when is similar. The vector space was investigated before by Weiqiang Wang (see [19]). Partially motivated by the work of Grojnowski (see [9]) and Nakajima (see [16]) and following ideas of Segal, Wang proved that is a Fock space, i.e., it is an irreducible representation of a Heisenberg algebra. Let us outline Wang’s construction with some minor modification. The vector space is naturally a commutative graded algebra with multiplication defined by the induction operation (see Section 2.2)
On the other hand, each graded piece is a module over the representation ring of the symmetric group. It is well known (see [2], Proposition 39.4) that has a virtual representation , that is, linear combination of irreducible representations with integer coefficients, such that, its character is
Note that can be viewed as a virtual -equivariant vector bundle on . Given , we denote by the tensor product of and the trivial bundle . Let be a basis of the topological K-ring . Let be the basis dual to the above one with respect to the Euler pairing, that is, where . Let .
Proposition 1.1.1 (Wang).
The Fock space , that is, generate freely as a commutative algebra.
We refer to [19], Proposition 3 for the proof of Proposition 1.1.1. In fact, the result of Wang is more general. He considered the case when is a -space where is a finite group. The corresponding Fock space could be viewed as the Fock space of the orbifold . From this point of view, the work of Wang should have a generalization to the case when is an orbifold which is not necessarily a global quotient. The orbifold case will be important for the applications to permutation-equivariant orbifold KGW theory.
We think of as the vacuum and of multiplication by as the creation operations. It turns out that the operators of differentiation by also have a natural K-theoretic interpretation. Namely, the following formula holds:
(1) |
where , is the projection map, and is the restriction functor (see Section 2.2). Here the K-theoretic pushforward yields a virtual -bundle whose coefficients are representations of and hence after taking the trace we obtain an element in . Let us point out that our construction of the Fock space is slightly different from the one in [19]. Namely, the multiplication operator of Wang involves the inverse of the Adam’s operations in while the differential operator is represented by a contraction operation that involves a choice in the dual vector space . The proof of (1) could be obtained directly from Wang’s results but for the sake of completeness we give a self-contained proof (see Theorem 2.3.1).
We will refer to as the K-theoretic Fock space of . We will think of , where , as formal parameters. The KGW invariants of with values in are defined as follows:
(2) |
where is the tautological line bundle formed by the cotangent lines at the -th marked point, is the virtual structure sheaf of the moduli space (see [14]), and the pushforward is the -equivariant -theoretic pushforward, that is,
We will usually drop the virtual structure sheaf in the above notation. Also, let us introduce formal parameters and introduce the following formal Laurent series:
Then the KGW invariants with values in can be organized into a set of formal power series in of the following form:
(3) |
The permutation-equivariant KGW invariants of Givental (see [7]) will be recalled in Section 3.2. Using the Heisenberg relations, that is, formula (1), we will prove in Section 3.2 that the correlators (3) are related to the correlators of Givental by a simple substitution (see (23)). Let us point out that the above definition (3) does not contain the entire information of permutation-equivariant KGW theory. We did not allow descendants at the permutable marked points. Our construction can be extended naturally by introducing the K-theoretic Fock space . We refer to Section 3.3 for more details. However, let us point out that in this paper the most general definition would not be needed.
1.2. Genus-0 integrable hierarchies
The standard identities in KGW theory, after some minor modifications, extend to permutation-equivariant KGW theory (see [6]). More precisely, the correlators (3) satisfy the string equation (see Section 3.4) and if the genus is 0, then we also have the dilaton equation (see Section 3.5) and the WDVV equations (see Section 3.6). These identities allow us to introduce permutation-equivariant K-theoretic quantum differential equations and a corresponding fundamental solution called the -matrix. Let be the matrix with entries
(4) |
where is the Euler pairing. Let be the entries of the inverse matrix . The -matrix is defined by the following identity:
(5) |
Note that is a formal power series in (and the Novikov variables ) whose coefficients are in . We refer to Proposition 3.6.1 for a list of properties of the -matrix. Let us define also the K-theoretic quantum cup product by
(6) |
Let us recall also the so-called -function:
Using the string equation, it is easy to prove that .
Suppose that is a formal power series in and with coefficients in . We will think of as parameters and quite often we suppress them in the notation. For example, we will write simply instead of . Furthermore, we identify , that is, and write . The following system of differential equations is the K-theoretic version of Dubrovin’s principal hierarchy (see [3]):
(7) |
where where . According to Milanov–Tonita (see [15], Theorem 1), the fact that is a solution to the quantum differential equations implies that the system of equations (7) is integrable, i.e., the system is compatible. The second goal of our paper is to construct a solution to (7) in terms of KGW invariants. Following Dubrovin, we will refer to this solution as the topological solution. The construction is the same as in [15], Theorem 2. Put
Note that the notation here is slightly different from the corresponding notation in (3).
Theorem 1.2.1.
Let be defined by
Let be a solution to the equation where we suppressed the dependence of and on . Then is a solution to the principal hierarchy (7).
1.3. Permutation-equivariant invariants of the point in genus 0
Suppose now that . The Fock space in this case is the representation ring of the symmetric group where is the virtual representation of introduced earlier (see Section 1.1).
Theorem 1.3.1.
The following formula holds:
for all
1.4. Acknowledgements
I am thankful to Yukinobu Toda for a very useful discussion on sheaf cohomology and rational singularities. The idea to consider the K-theoretic Fock space of Wang came after a talk by Timothy Logvinenko on the MS seminar at Kavli IPMU. I am thankful to him for giving an inspiring talk. This work is supported by the World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan and by JSPS Kakenhi Grant Number JP22K03265.
2. K-theoretic Fock space
Let us recall the background on K-theoretic Heisenberg algebras that will be needed in this paper. We give self-contained proofs mostly because we would like to give the reader the chance to grasp the techniques involved in the construction of the Fock space . We only focus on the results relevant for this paper. For more properties and further details we refer to [19].
2.1. Double cosets of the Young subgroups
Suppose that is a partition of , that is, a decreasing sequence of integers , such that, . Usually we write to denote that is a partition of . The number is called the length of . Put where . Given a partition we define the permutation
Note that is a complete set of representatives for the conjugacy classes of and that the conjugacy class consists of elements where is the number of elements in that commute with . Finally, let us denote by () the sequences , , etc. The subgroup of consisting of permutations that leave the sets invariant for all will be denoted by , that is,
The subgroups () are known as the Young subgroups of .
Suppose now that and are partitions of . We would like to recall the classification of the double coset classes (see [20], Appendix 3, Section A3.2). For a given parmutation , let us define
Note that are non-negative integers satisfying
(8) |
Lemma 2.1.1.
The numbers defined above depend only on the double coset of in .
The proof follows immediately from the definition (see [20], Appendix 3, Section A3.2). Suppose that we fix a matrix of size satisfyning the above conditions (8). Let us construct a permutation such that the corresponding matrix is . Split each set into subsets , such that, consists of the first elements of , – of the next elements, etc. We can do this because the number of elements in is Similarly, let us split each set into subsets consisting of respectively elements. We define to be the permutation that maps the elements of to and for the sake of definitness, let us require that preserves the order of the elements. Note that consists of elements.
Lemma 2.1.2.
Suppose that is a matrix of non-negative integers satisfying (8). Let be the permutation constructed above. Then the double coset consists of elements, where , , and
Proof.
Let us consider the map
The above formula defines a transitive action of on the double coset . The stabilizer of consists of pairs , such that, . Note that is uniquely determined by . We claim that leaves the subsets invariant for all and . Indeed, note that
On the other hand,
It remains only to notice that . Conversely, if preserves for all and , then leaves all invariant. In particular, . This proves that the stabilizer of is isomorphic to the subgroup of permutations in that leave the subsets invariant for all and . This subgroup is isomorphic to the direct product of all and hence it has elements. ∎
Lemma 2.1.3.
We have
where the sums is over all matrices of size with non-negative entries satisfying condition (8).
Proof.
Let us compute
in two different ways. First, using that
we get that the above expression coincides with
On the other hand, the coefficient in front of is . Comparing the two formulas, we get the identity that we wanted to prove. ∎
We proved the following proposition.
Proposition 2.1.1.
Suppose that and are partitions of of sizes respectively and . The double cosets in are parametrized by matrices of size whose entries are non-negative integers satisfying condition (8). Moreover, let us associate with each a permutation as explained above, then the resulting set of permutations gives a complete set of representatives for the double cosets in .
Note that in Proposition 2.1.1 the condition that and are partitions can be relaxed, that is, we do not have to require that and are decreasing sequences.
2.2. Induced vector bundles
We will assume that the reader is familiar with the notion of a -space and -vector bundle. For some background, we refer to [17]. Suppose that is a finite group, a subgroup of , and is a -space. For every -bundle on we define a -bundle on whose points are the set theoretic maps satisfying the following two conditions
-
(i)
The composition is -equivariant, that is, for all where is the projection map.
-
(ii)
The map is -equivariant, that is, for all and .
Note that when is a pont we have: is a representation of and is the induced representation. The structure of a -bundle on is defined as follows. First, the -action on is defined by . The structure projection is defined by . It is strightforward to check that is a -equivariant map. Suppose that is in the fiber . Then , that is, where denotes the fiber of at . Using the linear strucure on the fibers of we get that is naturally a linear vector space: and . Let be a complete set of representatives of the right coset classes in . Since we see that the map is uniquely determined by its values . This proves that is a finite dimensional vector space of dimension where is the index of in . Moreover, there is a set-theoretic isomorphism
(9) |
where is the unique index, such that, (note that acts on ) and is the map defined by the following Cartesian square:
where slightly abusing the notation we denote by the map defined by . Using the map (9), we equip with the structure of a vector bundle over . It is easy to check that the -action on defined above is continuous and therefore is a topological -vector bundle on .
We also have a restriction functor. Namely, if is a -space, is a subgroup, and is a -vector bundle, then is also an -vector bundle which will be denoted by . The Frobenius reciprocity rules take the following form.
Proposition 2.2.1 (Frobenius reciprocity).
Suppose that is a -space and that is a finite subgroup. Let be a -vector bundle and be an -vector bundle.
a) We have an isomorphism
uniquely determined by the following relation:
where and are maps that correspond to each other via the above isomorphism.
b) We have an isomorphism
uniquely determined by the following relation:
where and are maps that correspond to each other via the above isomorphism.
If is a point, Proposition 2.2.1 is well known in the representation theory of finite groups (see [2], Section 34). The proofs in general remains the same. The following proposition is Lemma 7 in [19].
Proposition 2.2.2.
Suppose that is a -space, and are finite subgroups of , and is a -vector bundle. For each , let . Using the projection on the -th factor , we view as a subgroup of for . We have the following relation in :
where the sum is over a set of elements which represent the double coset classes in .
In the case when is a point, the statement is well known (see [18], Proposition 22). Serre’s argument works in the general case too.
2.3. Anihilation operators
In order to define the anihilation operations, let us first recall the notion of a trace of a vector bundle. Suppose that is a complex vector bundle on some compact topological space and that is a finite order automorphism acting trivially on the base . To be more precise, we have for all and the induced map is a finite order linear map. In particular, the maps are diagonalizable and their eigenvalues are independent of . The trace of on is a virtual vector bundle on defined by
where only finitely many terms in the above sum are non-zero because must be an eigenvalue of and is a vector bundle because the eigenvalues of and their multiplicities are independent of . The following lemma is well known (see [1], Section 2).
Lemma 2.3.1.
Let be a complex vector bundle on , then is the Adam’s operation.
Proof.
The operation is compatible with pullback, that is, for all continuous maps . Recalling the splitting principle definition of the Adam’s operations, we get that it is sufficient to prove that if is a direct sum of line bundles, then . This is obvious because in the tensor product
all terms for which the sequence is not invariant under the cyclic permutation do not contribute to the trace. On the other hand, the sequence equals its cyclic permutation iff . ∎
Using the Chern character map, we get that the Adam’s operation is an isomorphism where recall that we work with the K-ring with complex coefficients. Therefore, for every , there exists an , such that, . We will need also the formula for the trace of an induced vector bundle.
Proposition 2.3.1.
Suppose that is a finite group, is a subgroup, and is a trivial -space. For every -vector bundle on we have the following formula:
where is a complete set of representatives for the right coset classes .
Proof.
Since acts trivially on , we have where the isomorphism is given by . Let be the permutation defined by and let be defined by . If is an orbit of , then the subspace is -invariant. Therefore,
where the sum is over all orbits of in . Note that the action of on is represented by a matrix with entries in of the following type
where and is the operator representing the action of . Note that
where . Since for , we get that only the one-point orbits of contribute to the trace of . It remains only to notice that the one point orbits , that is, the fixed points of , correspond precisely to those for which and that . ∎
Suppose that is a virtual vector bundle on . Let us define the contraction operation
where is the contraction map and is the K-theoretic pushforward, that is,
In the next lemma we will use the Lefschetz trace formula. We refer to Givental’s work (see [7], Section 2) for a proof of the Lefscetz trace formula based on the Kawasaki’s Riemann–Roch formula (see [10]).
Lemma 2.3.2.
Suppose that is a -bundle on . Then
where is such that where is the diagonal of , that is, the submnaifold .
Proof.
Put . According to the Lefschetz trace formula, coinicdes with the Euler characteristics of the following vector bundle
(12) |
where is the conormal bundle to in and
The numerator in (12) is precisely . The denominator can be expressed in terms of the K-theoretic Chern roots of . Indeed, we have
We get , where is identified with the hyperplane in and the action of on corresponds to permuting cyclically the coordinates . If are the K-theoretic Chern roots of , then we get
where is the eigenvector of with eigenvalue where . The denominator in (12) takes the form
Recalling the Hierzerbruch–Riemann–Roch formula, we get that the Euler characterisics of (12) is
Let be the complex degree operator, that is, for . Note that and that . Therefore, the above formula coincides with
Corollary 2.3.1.
We have .
Proof.
The contraction operation extends to the entire Fock space. Namely, let us define
by composing the restriction and the anihilation operations as follows:
Theorem 2.3.1.
Under the isomorphism , we have .
Proof.
Let us compute the composition
where . By definition
where we put for the ease of notation. We would like to use Proposition 2.2.2 with , and . The double cosets in , according to Proposition 2.1.1, are parametrized by the set of matrices
where . Moreover, if is the representative of the double coset corresponding to a matrix of the above form (see Proposition 2.1.1), then
and
In other words, is the permutation that fixes the first and the last numbers in , that is,
and that switches the two blocks and , that is,
According to Proposition 2.2.2 we have
(17) |
where the sum is over all matrices of the above type representing the double cosets in . Let us examine the restriction
(18) |
Since is non-zero only if , we get that the above restriction is non-zero only in two cases: , or , . Let us analyze these two cases separately.
First, suppose that and . Then and . We get
The contribution to (17) becomes
Applying to the above expression the contraction operation we get .
The second case is and . Then and . The restriction (18) becomes
The contribution to (17) becomes (note that now )
Note that when we apply the contraction operation we will get a sum of terms that have a factor of the form where is a representation of , that is, a -vector bundle on for which acts trivially on . This trace is for . Indeed, recalling the formula for the trace of an induced vector bundle (see Proposition 2.3.1), we get that the trace must be a sum over traces of elements in of the form . The latter is a cycle of length . However, the group contains such elements only if . Therefore, the contribution in this case is non-zero only if , that is,
Recall that and let us specalize . Then the above computations imply the following commutation relations:
The statement in the theorem follows. ∎
3. Permutation-equivariant K-theoretic Gromov–Witten theory
Suppose now that is a smooth projective variety. We would like to recall the background on permutation-equivariant KGW theory. Let
be the -theoretic Fock space of .
3.1. Translation symmetry
The correlators (3) have the following symmetry:
(19) |
The derivative on the LHS is equivalent to the operation . The correlator has the form . Its derivative becomes
where is the projection map and we used that
Note that and .
Using formula (19), we get that no information is lost if we set for all . Indeed, if we do this, then in order to recover the dependence on we simply have to translate .
3.2. Givental’s permutation-equivariant correlators
We would like to compare our definition to the permutation-equivariant KGW invariants of Givental. The definition evolved on 3 stages and the most general one appears in [7], Section 1. Givental’s correlators depend on the choice of a -algebra . The inputs of the correlators are parametrized by a sequence of formal variables , where the index set is defined by and and . Let be a sequence such that and for only finitely many . Givental introduced correlators of the following form:
(20) |
where is repeated times, that is, the correlator has inputs. Put and let us fix a permutation whose cycle type is determined by , that is, the action of on has exactly orbits of length for all . Here, for a given element and let be the minimal positive integer, such that, . We refer to the sequence as the orbit of through and to the number as the length of the orbit. The set of all orbits of will be denoted by . The correlator is by definition the following sum of super-traces
(21) |
where the sum is over all sequences and consisting of integers and – one for each orbit , is the number of elements in the orbit , and is the -th Adam’s operation in . Note that the definition of the correlator (20) is independent of the choice of , that is, it depends only on the cycle type of . Finally, the genus- potential of permutation-equivariant theory is defined by
In order to compare to our definition, let us put for all and , that is, let us consider only invariants that do not involve descendants at the permutable marked points. Also, let us restrict (3) to . The super-trace in (21) becomes
(22) |
where (resp. ) denotes the set of orbits of of length (resp. ), , and is the contraction map. Without loss of generality, we may assume that . Recalling Theorem 2.3.1 we get that the trace (22) coincides with
where , , and is the coefficient in front of the monomial in the correlator . Note that and that the number of sequences and in the sum in (21), such that,
is precisely because the values of and are fixed for the one-point orbits and for the more than one-point-orbits, we have and we have to choose orbits in for which will be assigned value . Now it is clear that the coefficient in front of in (21) is . Therefore, under the substitutions
(23) |
where , we get that the genus- potential coincides with
Therefore, definition (3) contains all the information for permutation-equivariant KGW invariants that do not involve descendants at the permutable marked points.
3.3. Descendants at the permutable entries
Let us sketch the necessary extension of the Fock space which will alow us to keep track of descendants. The reader not interested in the construction can skip this section as the results here would not be used in the rest of the paper. The main point is that we have to introduce the K-theoretic Fock space of where is equipped with the discrete topology. Note that this is a topological space with infinitely many connected components , . The K-theoretic Fock space is defined by
where the inclusion is defined via the pushforward with respect to the natural inclusion . Theorem 2.3.1 implies that
(24) |
where the vacuum is the same as before, that is, and the creation operator is defined by the induction operation and exterior tensor product by where is a virtual vector bundle on whose restriction to the connected component is for and if .
The following isomorphism will be needed in the definition of KGW invariants with values in . Suppose that has finitely many connected components. We have an isomorphism
(25) |
where the direct sum is over all monotonely increasing sequences satisfying and . The isomorphism is induced by restricting -bundles on to . Note that the inverse of this isomorphism, that is, constructing an -bundle from an -bundles, is given by an operation that in some sense generalizes the induction operation of vector bundles: first we have to make an -space from an -space by adding extra connected components and then extend the -bundle to an -bundle via the usual induction operation.
Applying the isomorphism (25) to our settings we get
(26) |
where the direct sum is over all monotonely increasing sequences of integer numbers , is the (Young) subgroup of preserving the sequence , and is the -th connected component of .
Now we are in position to define KGW invariant with values in . The definition (2) should be modified as follows. We replace the sum over with a sum over all monotonely increasing sequences of integers , that is, , define the evaluation map
and replace the pushforward by
The above pushforward takes value in . Recalling the isomorphisms (26) and (24) we get that the KGW invariants take value in a certain completion of the polynomial ring . Similarly to (19), we have
(27) |
Using the Heisenberg structure of , the same argument as in Section 3.2 proves that under the substitution
the permutation-equivariant genus-g potential takes the form
3.4. String equation
The following formula holds:
where for and
where and . The above formula is known as the string equation. The standard technique, based on the map forgetting the last marked point, works in our settings too (see [6], Proposition 1). We would like to work out only the correction terms .
Suppose that and where is the number of permutable marked points and is the number of the remaining ones. There are 3 cases. First, if and , we get
The second case is when and : the moduli space and the corresponding contribution to the KGW invariant is
where we used that the topological -ring is embedded in the Fock space via Finally, if and : the evaluation map coincides with the diagonal embedding . We have to compute . Let us compute the derivatives of with respect to and .
where is the projection on the second factor. Similarly,
where is the constant map and is the -theoretic pushforward. We get
3.5. Dilaton equation
The following formula is the permutation-equivariant K-theoretic version of the so called dilaton equation:
where . Using formula (19) we can rewrite the RHS of the above formula as follows:
The standard argument proving the dilaton equation in cohomological Gromov–Witten theory is based on pushing forward along the forgetful map . The K-theoretic version of the proof works in our settings too. We refer to [6] for more details. We would like only to work out the exceptional term corresponding to the case when and , that is, the case when the forgetful map does not exist. If , then is trivial as an -bundle, so the exceptional term is . The only non-trivial contribution comes when and . We have to compute , where is the diagonal embedding and is the alternating representation of . We have
Similar computation, since , proves that . Therefore,
3.6. WDVV equations and the -matrix
The WDVV equations say that the expression
(28) |
is symmetric in the pairs (). For example, by exchanging and , we get
(29) |
The equality of (28) and (29) determines the remaining identities. The standard proof based on the forgetful map works in our settings too (see [6], Proposition 3). The only new feature, compared to cohomological GW theory, is the metric (4) – see [4] for more details.
Let us recall the operator depending on the complex number and the parameters – see formula (5).
Proposition 3.6.1.
a) The following formula holds:
b) The following formula holds: where is defined by the identity .
c) The matrix satisfies the following differential equations:
where is the permutation-equivariant quantum K-product defined by (6).
Proof.
a) Let us apply the WDVV equations with , . For brevity, put , and , . We get
where we use Einstein’s convention to sum over repeating upper and lower indexes. Using the string equation, we get . The RHS takes the form
where we used the string equation and the identity
The RHS of the WDVV equation takes the form
Similarly, using the string equation we get that
Now it is clear that
and this is precisely the identity that we had to prove.
b) is a direct consequence of a).
c) Let us apply the WDVV equations with
We get
Again, using the string equation, we get that and that the RHS becomes
On the other hand, the LHS (see the computation in part a)) after removing via the string equation and recalling the definition of the quantum -product, becomes
Therefore,
Recalling b), we get and the formula that we have to prove becomes a direct consequence of the above formula. ∎
4. Genus-0 integrable hierarchies
The goal of this section is to prove Theorem 1.2.1. The argument is essentially the same as in [15], Section 4.3.
4.1. From descendants to ancestors
The ancestor KGW invariants are defined in the same way as the descendent KGW invariants, that is, by formula (2), except that we replace with the pullback where is the map that forgets the stable map and the last marked points. According to Givental (see [6]), the ancestor invariants can be be expressed in terms of the descendent ones and the -matrix. Let us recall Givental’s formula.
Let be the field of all rational functions on . Using elementary fraction decomposition, we have a natural projection map:
(30) |
It is easy to check that the above residue truncates all terms in the elementary fraction decomposition of that have poles in . The matrix has entries that are formal power series in , , and the Novikov variables whose coefficients are rational functions. In particular, we can define .
Lemma 4.1.1.
a) The following formula holds:
b) The following formula holds:
Proof.
a) Recalling the definition of the -matrix we get
Recalling Cauchy’s theorem we get
It remains only to recall the definition of and to note that
Part b) follows from a) and the string equation
The relation between descendants and ancestors is given by the following formula:
(31) |
where and must satisfy .
Let us choose a formal power series with coefficients in , such that,
(32) |
This is a fixed-point problem for and we can construct a formal solution via the iterations:
The sequence of formal series has a limit as which provides a solution to our fixed-point problem.
Lemma 4.1.2.
The following formula holds:
where we suppressed the dependence of on , that is, .
Proof.
The quadratic terms in clearly match. The equality of the linear terms is equivalent to
The above equality is precisely the dilaton equation . The equality of the remaining terms is equivalent to
The RHS can be written as
(33) |
Using the dilaton equation we get
and
Substituting the above two formulas in (33) we get the identity that we had to prove.∎
Proposition 4.1.1.
Proof.
Let us prove the first formula only. The argument for the remaining two is similar. Using the translation invariance (19), we can rewrite as follows:
(34) |
where is an arbitrary formal parameter and we wrote instead of in order to make the notation less cumbersome. Let us apply the formula expressing descendants in terms of ancestors (31). We get
where and we used Lemma 4.1.1 to compute . Let us specialize . By definition which implies that is proportional to for all . However, the product on because is an -dimensional manifold. We conclude that after the substitution all terms in the sum in (34) with must vanish. The sum of the remaining terms, according to Lemma 4.1.2, must be
4.2. Proof of Theorem 1.2.1
Let be the solution to the fixed-point problem (32). We will prove Theorem 1.2.1 by showing that and that is a solution to the integrable hierarchy (7). Let us denote the partial derivative by .
Let us prove that , that is,
(35) |
where is the same as in Proposition 4.1.1 and recall that . According to Proposition 4.1.1, the second order partial derivative in the above formula is
where we used the string equation. The formula that we want to prove follows.
Let us prove that is a solution to (7). Differentiating (35) with respect to we get
(36) |
where we suppressed the dependence of the two-point correlator on and we used Proposition 4.1.1. Using the projection map (30) we transform the two-point correlator into
Note that , or equivalently . Let us change into and substitute the result in formula (36). We get
(37) |
Using that is a solution to the quantum differential equations and that we get
Similarly, we find that is
Substituting these formulas in (37) and cancelling we get
where the residue at was dropped because does not have a pole at . Comparing with (7), we get that is a solution to the principal hierarchy. ∎
5. The genus-0 K-theoretic Gromov–Witten invariants of the point
The goal of this section is to prove Theorem 1.3.1.
5.1. Genus-0 moduli spaces of curves
Note that the vector space is naturally a -space where acts by permuting the coordinates . We have an isomorphism defined as follows. Given and we construct a cotangent vector where the meromorphic 1-form on is uniquely determined by the following conditions:
-
(i)
is holomorphic except for at most first order poles at the nodes and the first marked points of .
-
(ii)
If is a node and and are the two irreducible components of at , then where and are local coordinates on respectively and near .
-
(iii)
.
Holomorphic forms satisfying conditions (i) and (ii) are by definitions sections of the sheaf where is the dualizing sheaf of . Condition (iii), uniquely determines the meromorphic 1-form because is rational and hence the only obstruction to the existence of is that the sum of the residues at all poles is , that is, . This description also shows that which is a standard fact.
Recall that every line bundle, provided it has sufficiently many global sections, determines a holomorphic map to a projective space (see [8], Chapter I, Section 4). In our case, the line bundle determines a holomorphic map , such that, . More precisely, we have
(38) |
where is a local coordinate on at . The class of in is independent of the choice of local coordinate .
Lemma 5.1.1.
Suppose that the action on is induced from where and . The map (38) is -equivariant, that is, where the action of on is defined by .
Proof.
Let be the meromorphic form on corresponding to . We have or equivalently
Therefore, and the equivariance of is clear. ∎
The map can be described more explicitly as follows. Suppose that we have a point . Let , we call it the central component, be the irreducible component of that carries the marked point . Let us denote by the nodal points of that are on . Put . Note that by removing from we get a curve consisting of connected components . Therefore, the remaining marked points split into pairwise disjoint groups , that is, . Since the isomorphism is determined up to the action of , we can arrange that and where is the number of elements in . Note that
Let be the coordinate near . We have
Restricting we get that
On the other hand, the projective space where are the homogeneous coordinates on and the isomorphism is given by mapping to the linear functional . Under this identification, the action of on is given by and
(39) |
where the homogeneous coordinates on the RHS of the above formula should be ordered as follows: () is placed on the -th position and a copy of is placed on the -th position for every . Using formula (39), let us describe the fibers of . Given we define an equivalence relation in by saying that iff . This equivalence relation splits into several equivalence classes. The ones consisting of a single element we denote by (). The remaining ones, consisting of at least two elements, we denote be . If is in the fiber then the central component must be with marked points () and nodal points () where for each we choose one . In other words, the isomorphism class of the central component is uniquely fixed. The remaining components can be fixed in ways, that is, the fiber
(40) |
The map is a birational equivalence. More precisely, let be the analytic subvariety consisting of points , such that, there exist 3 pairwise different , and , such that, . This is clearly a complex co-dimension 2 algebraic subvariety and induces an isomorphism . Since is non-singular, the map is in particular a rational resolution. The general theory of rational singularities (see [11], Theorem 5.10) yields the following lemma.
Lemma 5.1.2.
The higher direct images of vanish, that is, for .
5.2. J-function
The -function of a point was computed by Givental (see the Theorem in [5]). We have the following formula:
Let us outline a proof of the above formula which is slightly different than the one given in [5]. Let be the contraction map. Using Lemma 5.1.2 we compute . Therefore,
The -function takes the form
On the other hand, we have as -modules where is the trivial -module. Note that for and where we assume that . Therefore,
where in the last equality we exchanged the summations over and . Using the above formula we get that the -function is
It remains only to prove that
Let us denote the LHS of the above equality by . Let us compute the derivatives by using the Heisenberg algebra structure of the Fock space. Recalling Theorem 2.3.1, we get
where we take , otherwise the derivative is for degree reasons. In order to compute this trace, let us identify with the space of homogeneous polynomials in of degree . This space has a basis consisting of monomials , such that, . Since the permutation permutes these monomials, the trace is equal to the number of monomials that remain fixed. However, there is at most one monomial fixed by , that is, the monomial with . Such monomial exists only if divides . In other words, the trace in the above formula is if and otherwise. Summing over all , we get
Summing over all , we get . Solving this differential equation with the initial condition yields the formula that we had to prove.
5.3. Proof of Theorem 1.3.1
Since we get
and
Put
and let be defined as the solution to the equation
Recalling Theorem 1.2.1, we get that is a solution to the following system of differential equations:
where and . On the other hand, since , we get that the insertion of is obtained from via the action of the differential operator
that is, the correlator on the LHS in Theorem 1.3.1 is .
Let us examine the action of on
where is an arbitrary sequence of complex numbers. We have
Recalling the differential equation for , we get
The residues in the above formula can be interpreted analytically as follows. We expand the exponential as a formal power series in , then each coefficient is a rational function in with poles on the unit circle. Therefore, we may and we will think of each residue as where is a sufficiently big real number and the orientation of the contour is clockwise around the center . Let us choose so small that . The sum over of the geometric series is uniformly convergent in to and we get
Recalling the Cauchy residue theorem we get that the above integral is the sum of the residues at and . The residue at is a constant which does not contribute at the end because we have to apply to it the derivation . We get
Note that the above formula is equivalent to
Using the above formula we get
Note that if we set , then
Therefore, and the formula that we had to prove follows. ∎
References
- [1] Michael Atiyah. Power operations in K-theory . The Quarterly Journal of Mathematics, 17(1):165–193, 1966.
- [2] Daniel Bump. Lie Groups, volume 225 of Graduate Texts in Mathematics. Springer, 2014.
- [3] Boris Dubrovin. Geometry of D topological field theories. In Integrable systems and quantum groups (Montecatini Terme, 1993), volume 1620 of Lecture Notes in Math., pages 120–348. Springer, Berlin, 1996.
- [4] Alexander Givental. On the WDVV equation in quantum K-theory. Michigan Math. J., 48:295–304, Jan 2000.
- [5] Alexander Givental. Permutation-equivariant quantum K-theory I. arXiv:1508.02690, Aug 2015.
- [6] Alexander Givental. Permutation-equivariant quantum K-theory VII. arXiv:1510.03076, Oct 2015.
- [7] Alexander Givental. Permutation-equivariant quantum K-theory IX. arXiv:1709.03180, Sep 2017.
- [8] P. Griffiths and J. Harris. Principles of Algebraic Geometry. Wiley Classics Library. Wiley, 1994.
- [9] I. Grojnowski. Instantons and affine algebras I: The Hilbert scheme and vertex operators. Mathematical Research Letters, 3:275–291, 1996.
- [10] Tetsuro Kawasaki. The Riemann-Roch theorem for complex -manifolds. Osaka Math. J., 16(1):151–159, 1979.
- [11] Janos Kollár and Shigefumi Mori. Birational Geometry of Algebraic Varieties, volume 134 of Cambridge Tracts in Mathematics. Cambridge University Press, 1998.
- [12] Y.-P. Lee and F. Qu. Euler characteristics of universal cotangent line bundles on . Proceedings of the American Mathematical Society, 142(2):429–440, 2014.
- [13] Yuan Pin Lee. A formula for Euler characteristics of tautological line bundles on the Deligne-Mumford moduli spaces. International Mathematics Research Notices, 1997(8):393–400, 01 1997.
- [14] Yuan Pin Lee. Quantum -theory, I: Foundations. Duke Mathematical Journal, 121(3):389 – 424, 2004.
- [15] Todor Milanov and Valentin Tonita. K-theoretic Gromov–Witten invariants in genus 0 and integrable hierarchies. Lett. Math. Phys., 108:1455–1483, 2018.
- [16] Hiraku Nakajima. Heisenberg algebra and Hilbert schemes of points on projective surfaces. Annals of Mathematics, 145(2):379–388, 1997.
- [17] Gream Segal. Equivariant K-theory. Publications Mathématiques de l’Institut des Hautes Scientifiques, 34:129–151, 1968.
- [18] Jean-Pierre Serre. Linear Representations of Finite Groups, volume 42 of Graduate Texts in Mathematics. Springer New York, NY, 2012.
- [19] Weiqiang Wang. Equivariant -theory, wreath products, and Heisenberg algebra. Duke Mathematical Journal, 103(1):1 – 23, 2000.
- [20] Andrey Zelevinsky. Representations of Finite Classical Groups. Lecture Notes in Mathematics. Springer Berlin, Heidelberg, 1981.