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

K-theoretic Heisenberg algebras and permutation-equivariant Gromov–Witten theory

Todor Milanov Kavli IPMU (WPI), UTIAS, The University of Tokyo, Kashiwa, Chiba 277-8583, Japan [email protected]
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.

Key words and phrases: Gromov–Witten theory, finite groups, integrable systems

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 1\mathbb{P}^{1}, 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 XX is a smooth projective variety. Let ¯g,n+k(X,d)\overline{\mathcal{M}}_{g,n+k}(X,d) be the moduli space of degree dd stable maps from a genus-gg nodal Riemann surface equipped with n+kn+k marked points. We let the symmetric group SkS_{k} act on the moduli space by permuting the last kk marked points. We will also refer to the last kk marked points as permutable marked points. Let

ev(n):¯g,n+k(X,d)Xn\displaystyle\operatorname{ev}_{(n)}:\overline{\mathcal{M}}_{g,n+k}(X,d)\to X^{n}

and

ev(k):¯g,n+k(X,d)Xk\displaystyle\operatorname{ev}^{(k)}:\overline{\mathcal{M}}_{g,n+k}(X,d)\to X^{k}

be the evaluation maps at respectively the first nn and the last kk marked points. The permutation-equivariant KGW invariants contain information about the SkS_{k}-equivariant K-theoretic pushforward ev(k)\operatorname{ev}^{(k)}_{*}. Therefore, it is natural to introduce the following graded vector space:

(X)=n0KSn(Xn),KS0(X0):=,\displaystyle\mathcal{F}(X)=\bigoplus_{n\geq 0}K_{S_{n}}(X^{n}),\quad K_{S_{0}}(X^{0}):=\mathbb{C},

where SnS_{n} is the symmetric group and KSn(Xn)K_{S_{n}}(X^{n}) denotes the SnS_{n}-equivariant topological K-ring with complex coefficients (see [17]). We will consider the case when K1(X)=0K^{1}(X)=0 in order to make the construction more transperant. The case when K1(X)0K^{1}(X)\neq 0 is similar. The vector space (X)\mathcal{F}(X) 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 (X)\mathcal{F}(X) 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 (X)\mathcal{F}(X) is naturally a commutative graded algebra with multiplication defined by the induction operation (see Section 2.2)

KSn(Xn)×KSm(Xm)KSn+m(Xm+n),(E,F)IndSn×SmSn+m(EF).\displaystyle K_{S_{n}}(X^{n})\times K_{S_{m}}(X^{m})\to K_{S_{n+m}}(X^{m+n}),\quad(E,F)\mapsto\operatorname{Ind}_{S_{n}\times S_{m}}^{S_{n+m}}(E\boxtimes F).

On the other hand, each graded piece KSn(Xn)K_{S_{n}}(X^{n}) is a module over the representation ring R(Sn)R(S_{n}) of the symmetric group. It is well known (see [2], Proposition 39.4) that R(Sn)R(S_{n}) has a virtual representation pnp_{n}, that is, linear combination of irreducible representations with integer coefficients, such that, its character is

χpn(g)={n if g=(1,2,,n),0 otherwise .\displaystyle\chi_{p_{n}}(g)=\begin{cases}n&\mbox{ if }g=(1,2,\dots,n),\\ 0&\mbox{ otherwise }.\end{cases}

Note that Xn×pnX^{n}\times p_{n} can be viewed as a virtual SnS_{n}-equivariant vector bundle on XnX^{n}. Given EKSn(Xn)E\in K_{S_{n}}(X^{n}), we denote by EpnKSn(Xn)E\otimes p_{n}\in K_{S_{n}}(X^{n}) the tensor product of EE and the trivial bundle Xn×pnX^{n}\times p_{n}. Let {Φα}α=1N\{\Phi_{\alpha}\}_{\alpha=1}^{N} be a basis of the topological K-ring K(X)K(X). Let {Φα}α=1N\{\Phi^{\alpha}\}_{\alpha=1}^{N} be the basis dual to the above one with respect to the Euler pairing, that is, (Φα,Φβ)=δα,β(\Phi^{\alpha},\Phi_{\beta})=\delta_{\alpha,\beta} where (E,F):=χ(EF)(E,F):=\chi(E\otimes F). Let νn,α:=(Φα)npn\nu_{n,\alpha}:=(\Phi^{\alpha})^{\boxtimes n}\otimes p_{n}.

Proposition 1.1.1 (Wang).

The Fock space (X)=[νn,α(1αN,n0)]\mathcal{F}(X)=\mathbb{C}[\nu_{n,\alpha}(1\leq\alpha\leq N,n\geq 0)], that is, νn,α\nu_{n,\alpha} generate freely (X)\mathcal{F}(X) 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 XX is a GG-space where GG is a finite group. The corresponding Fock space G(X)\mathcal{F}_{G}(X) could be viewed as the Fock space of the orbifold [X/G][X/G]. From this point of view, the work of Wang should have a generalization to the case when XX 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 1(X)1\in\mathcal{F}(X) as the vacuum and of multiplication by νn,α\nu_{n,\alpha} as the creation operations. It turns out that the operators of differentiation by rνr,αr\tfrac{\partial}{\partial\nu_{r,\alpha}} also have a natural K-theoretic interpretation. Namely, the following formula holds:

(1) rνr,α(E)=tr(1,2,,r)π(nr)(Φαr1(nr)ResSr×SnrSn(E)),r\partial_{\nu_{r,\alpha}}(E)=\operatorname{tr}_{(1,2,\dots,r)}\pi^{(n-r)}_{*}\Big{(}\Phi_{\alpha}^{\boxtimes r}\boxtimes 1^{\boxtimes(n-r)}\otimes\operatorname{Res}^{S_{n}}_{S_{r}\times S_{n-r}}(E)\Big{)},

where EKSn(Xn)E\in K_{S_{n}}(X^{n}), π(nr):Xn=Xr×XnrXnr\pi^{(n-r)}:X^{n}=X^{r}\times X^{n-r}\to X^{n-r} is the projection map, and ResSr×SnrSn\operatorname{Res}^{S_{n}}_{S_{r}\times S_{n-r}} is the restriction functor (see Section 2.2). Here the K-theoretic pushforward π(nr)\pi^{(n-r)}_{*} yields a virtual SnrS_{n-r}-bundle whose coefficients are representations of SrS_{r} and hence after taking the trace tr(1,2,,r)\operatorname{tr}_{(1,2,\dots,r)} we obtain an element in KSnr(Xnr)K_{S_{n-r}}(X^{n-r}). 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 K(X)K(X) while the differential operator rνr,αr\partial_{\nu_{r,\alpha}} is represented by a contraction operation that involves a choice in the dual vector space K(X)K(X)^{\vee}. 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 (X)\mathcal{F}(X) as the K-theoretic Fock space of XX. We will think of ν=(ν1,ν2,)\nu=(\nu_{1},\nu_{2},\dots), where νr=(νr,a)1aN\nu_{r}=(\nu_{r,a})_{1\leq a\leq N}, as formal parameters. The KGW invariants of XX with values in (X)\mathcal{F}(X) are defined as follows:

(2) Φa1L1i1,,ΦanLning,n(ν):=k=0dQdev(k)(𝒪g,n+k,dL1i1Lninev(n)(Φa1Φan)),\langle\Phi_{a_{1}}L_{1}^{i_{1}},\dots,\Phi_{a_{n}}L_{n}^{i_{n}}\rangle_{g,n}(\nu):=\sum_{k=0}^{\infty}\sum_{d}Q^{d}\operatorname{ev}^{(k)}_{*}\left(\mathcal{O}_{g,n+k,d}\otimes L_{1}^{i_{1}}\otimes\cdots\otimes L_{n}^{i_{n}}\operatorname{ev}_{(n)}^{*}(\Phi_{a_{1}}\boxtimes\cdots\boxtimes\Phi_{a_{n}})\right),

where LiL_{i} is the tautological line bundle formed by the cotangent lines at the ii-th marked point, 𝒪g,n+k,d\mathcal{O}_{g,n+k,d} is the virtual structure sheaf of the moduli space ¯g,n+k(X,d)\overline{\mathcal{M}}_{g,n+k}(X,d) (see [14]), and the pushforward is the SkS_{k}-equivariant KK-theoretic pushforward, that is,

ev(k):KSk(¯g,n+k(X,d))KSk(Xk)(X).\displaystyle\operatorname{ev}^{(k)}_{*}:K_{S_{k}}(\overline{\mathcal{M}}_{g,n+k}(X,d))\to K_{S_{k}}(X^{k})\subset\mathcal{F}(X).

We will usually drop the virtual structure sheaf in the above notation. Also, let us introduce formal parameters ti,at_{i,a} and introduce the following formal Laurent series:

𝐭(q)=ia=1Nti,aΦaqi.\displaystyle\mathbf{t}(q)=\sum_{i\in\mathbb{Z}}\sum_{a=1}^{N}t_{i,a}\Phi_{a}q^{i}.

Then the KGW invariants with values in (X)\mathcal{F}(X) can be organized into a set of formal power series in 𝐭\mathbf{t} of the following form:

(3) 𝐭(L1),,𝐭(Ln)g,n(ν)=i1,,ina1,,an=1Nti1,a1tin,anΦa1L1i1,,ΦanLning,n(ν).\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{g,n}(\nu)=\sum_{i_{1},\dots,i_{n}\in\mathbb{Z}}\sum_{a_{1},\dots,a_{n}=1}^{N}t_{i_{1},a_{1}}\cdots t_{i_{n},a_{n}}\,\langle\Phi_{a_{1}}L_{1}^{i_{1}},\dots,\Phi_{a_{n}}L_{n}^{i_{n}}\rangle_{g,n}(\nu).

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 (X×)\mathcal{F}(X\times\mathbb{Z}). 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 SS-matrix. Let GG be the N×NN\times N matrix with entries

(4) Gab(ν)=(Φa,Φb)+Φa,Φb0,2(ν),G_{ab}(\nu)=(\Phi_{a},\Phi_{b})+\langle\Phi_{a},\Phi_{b}\rangle_{0,2}(\nu),

where (Φa,Φb)=χ(ΦaΦb)(\Phi_{a},\Phi_{b})=\chi(\Phi_{a}\otimes\Phi_{b}) is the Euler pairing. Let Gab(ν)G^{ab}(\nu) be the entries of the inverse matrix G1G^{-1}. The SS-matrix is defined by the following identity:

(5) G(S(ν,q)Φa,Φb):=(Φa,Φb)+Φa1q1L,Φb0,2(ν).G(S(\nu,q)\Phi_{a},\Phi_{b}):=(\Phi_{a},\Phi_{b})+\Big{\langle}\frac{\Phi_{a}}{1-q^{-1}L},\Phi_{b}\Big{\rangle}_{0,2}(\nu).

Note that S(ν,q)S(\nu,q) is a formal power series in ν\nu (and the Novikov variables QQ) whose coefficients are in End(K(X))(q)\operatorname{End}(K(X))(q). We refer to Proposition 3.6.1 for a list of properties of the SS-matrix. Let us define also the K-theoretic quantum cup product by

(6) G(ΦiΦj,Φk):=Φi,Φj,Φk0,3(ν),1i,j,kN.G(\Phi_{i}\bullet\Phi_{j},\Phi_{k}):=\langle\Phi_{i},\Phi_{j},\Phi_{k}\rangle_{0,3}(\nu),\quad 1\leq i,j,k\leq N.

Let us recall also the so-called JJ-function:

J(ν,q):=1q+ν1+a=1NΦaΦa1qL0,1(ν).\displaystyle J(\nu,q):=1-q+\nu_{1}+\sum_{a=1}^{N}\Phi^{a}\,\Big{\langle}\frac{\Phi_{a}}{1-qL}\Big{\rangle}_{0,1}(\nu).

Using the string equation, it is easy to prove that J(ν,q)=(1q)S(ν,q)11J(\nu,q)=(1-q)S(\nu,q)^{-1}1.

Suppose that v(𝐭,ν2,ν3,)=α=1Nvα(𝐭,ν2,ν3,)Φαv(\mathbf{t},\nu_{2},\nu_{3},\dots)=\sum_{\alpha=1}^{N}v_{\alpha}(\mathbf{t},\nu_{2},\nu_{3},\dots)\Phi_{\alpha} is a formal power series in 𝐭=(tk,α)\mathbf{t}=(t_{k,\alpha}) and ν2,ν3,\nu_{2},\nu_{3},\dots with coefficients in K(X)K(X). We will think of ν2,ν3,\nu_{2},\nu_{3},\dots as parameters and quite often we suppress them in the notation. For example, we will write simply v(𝐭)v(\mathbf{t}) instead of v(𝐭,ν2,ν3,)v(\mathbf{t},\nu_{2},\nu_{3},\dots). Furthermore, we identify ν1:=v(𝐭)\nu_{1}:=v(\mathbf{t}), that is, ν1,α:=vα(𝐭)\nu_{1,\alpha}:=v_{\alpha}(\mathbf{t}) and write S(v(𝐭),q):=S(ν,q)|ν1=v(𝐭)S(v(\mathbf{t}),q):=S(\nu,q)|_{\nu_{1}=v(\mathbf{t})}. The following system of differential equations is the K-theoretic version of Dubrovin’s principal hierarchy (see [3]):

(7) tn,αv(𝐭)=Resq=dq(q1)n1v(𝐭)S(v(𝐭),q)Φα(n0, 1αN),\partial_{t_{n,\alpha}}v(\mathbf{t})=-\operatorname{Res}_{q=\infty}dq(q-1)^{n-1}\partial v(\mathbf{t})\bullet S(v(\mathbf{t}),q)\Phi_{\alpha}\quad(n\geq 0,\ 1\leq\alpha\leq N),

where =t0,1\partial=\partial_{t_{0,1}} where Φ1:=1K(X)\Phi_{1}:=1\in K(X). According to Milanov–Tonita (see [15], Theorem 1), the fact that S(ν,q)S(\nu,q) 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

𝐭(q)=k=0α=1Ntk,αΦα(q1)k.\displaystyle\mathbf{t}(q)=\sum_{k=0}^{\infty}\sum_{\alpha=1}^{N}t_{k,\alpha}\Phi_{\alpha}(q-1)^{k}.

Note that the notation tk,αt_{k,\alpha} here is slightly different from the corresponding notation in (3).

Theorem 1.2.1.

Let w(𝐭,ν2,ν3,)=α=1Nwα(𝐭,ν2,ν3,)Φαw(\mathbf{t},\nu_{2},\nu_{3},\dots)=\sum_{\alpha=1}^{N}w_{\alpha}(\mathbf{t},\nu_{2},\nu_{3},\dots)\Phi^{\alpha} be defined by

wα(𝐭,ν2,ν3,)=t0,αt0,1(0)(𝐭)|ν1=0=n=01n!1,Φα,𝐭(L),,𝐭(L)0,2+n(0,ν2,ν3,).\displaystyle\left.w_{\alpha}(\mathbf{t},\nu_{2},\nu_{3},\dots)=\partial_{t_{0,\alpha}}\partial_{t_{0,1}}\mathcal{F}^{(0)}(\mathbf{t})\right|_{\nu_{1}=0}=\sum_{n=0}^{\infty}\frac{1}{n!}\,\langle 1,\Phi_{\alpha},\mathbf{t}(L),\dots,\mathbf{t}(L)\rangle_{0,2+n}(0,\nu_{2},\nu_{3},\dots).

Let v(𝐭)v(\mathbf{t}) be a solution to the equation J(v(𝐭),0)=1+w(𝐭)J(v(\mathbf{t}),0)=1+w(\mathbf{t}) where we suppressed the dependence of vv and ww on ν2,ν3,\nu_{2},\nu_{3},\dots. Then v(𝐭)v(\mathbf{t}) is a solution to the principal hierarchy (7).

1.3. Permutation-equivariant invariants of the point in genus 0

Suppose now that X=ptX=\operatorname{pt}. The Fock space in this case (pt)=k=0R(Sk)=[ν1,ν2,]\mathcal{F}(\operatorname{pt})=\bigoplus_{k=0}^{\infty}R(S_{k})\otimes\mathbb{C}=\mathbb{C}[\nu_{1},\nu_{2},\dots] is the representation ring of the symmetric group where νn=pn\nu_{n}=p_{n} is the virtual representation of SnS_{n} introduced earlier (see Section 1.1).

Theorem 1.3.1.

The following formula holds:

11q1L,,11qnL,1,10,n+2(ν)=\displaystyle\Big{\langle}\frac{1}{1-q_{1}L},\dots,\frac{1}{1-q_{n}L},1,1\Big{\rangle}_{0,n+2}(\nu)=
1(1q1)(1qn)(1+1q111++1qn11)n1exp(r=1νrr(1+1q1r1++1qnr1))\displaystyle\frac{1}{(1-q_{1})\cdots(1-q_{n})}\Big{(}1+\frac{1}{q_{1}^{-1}-1}+\cdots+\frac{1}{q_{n}^{-1}-1}\Big{)}^{n-1}\ \exp\left(\sum_{r=1}^{\infty}\frac{\nu_{r}}{r}\Big{(}1+\frac{1}{q_{1}^{-r}-1}+\cdots+\frac{1}{q_{n}^{-r}-1}\Big{)}\right)

for all n1.n\geq 1.

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 (X)\mathcal{F}(X). 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 λ=(λ1,,λr)\lambda=(\lambda_{1},\dots,\lambda_{r}) is a partition of nn, that is, a decreasing sequence of integers λ1λ2λr>0\lambda_{1}\geq\lambda_{2}\geq\cdots\geq\lambda_{r}>0, such that, λ1++λr=n\lambda_{1}+\cdots+\lambda_{r}=n. Usually we write λn\lambda\vdash n to denote that λ\lambda is a partition of nn. The number (λ):=r\ell(\lambda):=r is called the length of λ\lambda. Put k(λ):=card{i|λi=k}\ell_{k}(\lambda):=\operatorname{card}\,\{i\ |\ \lambda_{i}=k\} where k=1,2,k=1,2,\dots. Given a partition λ\lambda we define the permutation

σ(λ):=(1,2,,λ1)(λ1+1,λ1+2,,λ1+λ2)(λ1++λr1+1,,λ1++λr1+λr).\displaystyle\sigma(\lambda):=(1,2,\dots,\lambda_{1})\,(\lambda_{1}+1,\lambda_{1}+2,\dots,\lambda_{1}+\lambda_{2})\,\cdots\,(\lambda_{1}+\cdots+\lambda_{r-1}+1,\dots,\lambda_{1}+\cdots+\lambda_{r-1}+\lambda_{r}).

Note that {σ(λ)}λn\{\sigma(\lambda)\}_{\lambda\vdash n} is a complete set of representatives for the conjugacy classes of SnS_{n} and that the conjugacy class Cλ:={gσ(λ)g1|gSn}C_{\lambda}:=\{g\sigma(\lambda)g^{-1}\ |\ g\in S_{n}\} consists of n!zλ1n!z_{\lambda}^{-1} elements where zλ:=kkk(λ)k(λ)!z_{\lambda}:=\prod_{k}k^{\ell_{k}(\lambda)}\ell_{k}(\lambda)! is the number of elements in SnS_{n} that commute with σ(λ)\sigma(\lambda). Finally, let us denote by Yi(λ)Y_{i}(\lambda) (1i(λ)1\leq i\leq\ell(\lambda)) the sequences Y1(λ):={1,2,,λ1}Y_{1}(\lambda):=\{1,2,\dots,\lambda_{1}\}, Y2(λ):={λ1+1,λ1+2,,λ1+λ2}Y_{2}(\lambda):=\{\lambda_{1}+1,\lambda_{1}+2,\dots,\lambda_{1}+\lambda_{2}\}, etc. The subgroup of SnS_{n} consisting of permutations that leave the sets Yi(λ)Y_{i}(\lambda) invariant for all ii will be denoted by SλS_{\lambda}, that is,

Sλ:={σSn|σ(Yi(λ))=Yi(λ)}Sλ1××Sλr.\displaystyle S_{\lambda}:=\{\sigma\in S_{n}\ |\ \sigma(Y_{i}(\lambda))=Y_{i}(\lambda)\}\cong S_{\lambda_{1}}\times\cdots\times S_{\lambda_{r}}.

The subgroups SλS_{\lambda} (λn\lambda\vdash n) are known as the Young subgroups of SnS_{n}.

Suppose now that λ\lambda and μ\mu are partitions of nn. We would like to recall the classification of the double coset classes Sμ\Sn/SλS_{\mu}\backslash S_{n}/S_{\lambda} (see [20], Appendix 3, Section A3.2). For a given parmutation σSn\sigma\in S_{n}, let us define

γij:=card(σ(Yi(λ))Yj(μ)),1i(λ)=:r,1j(μ)=:s.\displaystyle\gamma_{ij}:=\operatorname{card}(\sigma(Y_{i}(\lambda))\cap Y_{j}(\mu)),\quad 1\leq i\leq\ell(\lambda)=:r,\quad 1\leq j\leq\ell(\mu)=:s.

Note that γij\gamma_{ij} are non-negative integers satisfying

(8) j=1sγij=λi,i=1rγij=μj.\sum_{j=1}^{s}\gamma_{ij}=\lambda_{i},\quad\sum_{i=1}^{r}\gamma_{ij}=\mu_{j}.
Lemma 2.1.1.

The numbers γij\gamma_{ij} defined above depend only on the double coset of σ\sigma in Sμ\Sn/SλS_{\mu}\backslash S_{n}/S_{\lambda}.

The proof follows immediately from the definition (see [20], Appendix 3, Section A3.2). Suppose that we fix a matrix γ=(γij)\gamma=(\gamma_{ij}) of size r×sr\times s satisfyning the above conditions (8). Let us construct a permutation σSn\sigma\in S_{n} such that the corresponding matrix is γ\gamma. Split each set Yi(λ)Y_{i}(\lambda) into subsets Yi1(λ),,Yis(λ)Y_{i1}(\lambda),\dots,Y_{is}(\lambda), such that, Yi1(λ)Y_{i1}(\lambda) consists of the first γi1\gamma_{i1} elements of Yi(λ)Y_{i}(\lambda), Yi2(λ)Y_{i2}(\lambda) – of the next γi2\gamma_{i2} elements, etc. We can do this because the number of elements in Yi(λ)Y_{i}(\lambda) is λi=γi1+γi2++γis.\lambda_{i}=\gamma_{i1}+\gamma_{i2}+\cdots+\gamma_{is}. Similarly, let us split each set Yj(μ)Y_{j}(\mu) into subsets Y1j(μ),,Yrj(μ)Y_{1j}(\mu),\dots,Y_{rj}(\mu) consisting of respectively γ1j,γrj\gamma_{1j},\dots\gamma_{rj} elements. We define σ\sigma to be the permutation that maps the elements of Yij(λ)Y_{ij}(\lambda) to Yij(μ)Y_{ij}(\mu) and for the sake of definitness, let us require that σ\sigma preserves the order of the elements. Note that σ(Yi(λ))Yj(μ)=Yij(μ)\sigma(Y_{i}(\lambda))\cap Y_{j}(\mu)=Y_{ij}(\mu) consists of γij\gamma_{ij} elements.

Lemma 2.1.2.

Suppose that γ=(γij)\gamma=(\gamma_{ij}) is a matrix of non-negative integers satisfying (8). Let σSn\sigma\in S_{n} be the permutation constructed above. Then the double coset SμσSλS_{\mu}\sigma S_{\lambda} consists of λ!μ!/γ!\lambda!\mu!/\gamma! elements, where λ!=λ1!λr!\lambda!=\lambda_{1}!\cdots\lambda_{r}!, μ!=μ1!μs!\mu!=\mu_{1}!\cdots\mu_{s}!, and γ!=i,jγij!.\gamma!=\prod_{i,j}\gamma_{ij}!.

Proof.

Let us consider the map

Sμ×SλSμσSλ,(τ1,τ2)τ1στ21.\displaystyle S_{\mu}\times S_{\lambda}\to S_{\mu}\sigma S_{\lambda},\quad(\tau_{1},\tau_{2})\mapsto\tau_{1}\sigma\tau_{2}^{-1}.

The above formula defines a transitive action of Sμ×SλS_{\mu}\times S_{\lambda} on the double coset SμσSλS_{\mu}\sigma S_{\lambda}. The stabilizer of σSμσSλ\sigma\in S_{\mu}\sigma S_{\lambda} consists of pairs (τ1,τ2)(\tau_{1},\tau_{2}), such that, τ1=στ2σ1\tau_{1}=\sigma\tau_{2}\sigma^{-1}. Note that τ1\tau_{1} is uniquely determined by τ2\tau_{2}. We claim that τ2\tau_{2} leaves the subsets Yij(λ)Y_{ij}(\lambda) invariant for all ii and jj. Indeed, note that

τ2(Yij(λ))τ2(Yi(λ))=Yi(λ).\displaystyle\tau_{2}(Y_{ij}(\lambda))\subset\tau_{2}(Y_{i}(\lambda))=Y_{i}(\lambda).

On the other hand,

σ1τ1σ(Yij(λ))=σ1τ1(Yij(μ))σ1(Yj(μ)).\displaystyle\sigma^{-1}\tau_{1}\sigma(Y_{ij}(\lambda))=\sigma^{-1}\tau_{1}(Y_{ij}(\mu))\subset\sigma^{-1}(Y_{j}(\mu)).

It remains only to notice that Yi(λ)σ1(Yj(μ))=Yij(λ)Y_{i}(\lambda)\cap\sigma^{-1}(Y_{j}(\mu))=Y_{ij}(\lambda). Conversely, if τ2\tau_{2} preserves Yij(λ)Y_{ij}(\lambda) for all ii and jj, then τ1=στ2σ1\tau_{1}=\sigma\tau_{2}\sigma^{-1} leaves all Yij(μ)Y_{ij}(\mu) invariant. In particular, τ1Sμ\tau_{1}\in S_{\mu}. This proves that the stabilizer of σ\sigma is isomorphic to the subgroup of permutations in SnS_{n} that leave the subsets Yij(λ)Y_{ij}(\lambda) invariant for all ii and jj. This subgroup is isomorphic to the direct product of all SγijS_{\gamma_{ij}} and hence it has γ!\gamma! elements. ∎

Lemma 2.1.3.

We have

γλ!μ!γ!=n!,\sum_{\gamma}\frac{\lambda!\mu!}{\gamma!}=n!,

where the sums is over all matrices γ\gamma of size r×sr\times s with non-negative entries satisfying condition (8).

Proof.

Let us compute

(x1++xs)n=(x1++xs)λ1(x1++xs)λr\displaystyle(x_{1}+\cdots+x_{s})^{n}=(x_{1}+\cdots+x_{s})^{\lambda_{1}}\cdots(x_{1}+\cdots+x_{s})^{\lambda_{r}}

in two different ways. First, using that

(x1++xs)λi=γi1++γis=λiλi!γi1!γis!x1γi1xsγis\displaystyle(x_{1}+\cdots+x_{s})^{\lambda_{i}}=\sum_{\gamma_{i1}+\cdots+\gamma_{is}=\lambda_{i}}\frac{\lambda_{i}!}{\gamma_{i1}!\cdots\gamma_{is}!}\,x_{1}^{\gamma_{i1}}\cdots x_{s}^{\gamma_{is}}

we get that the above expression coincides with

μ1++μs=nγλ!γ!x1μ1xsμs.\displaystyle\sum_{\mu_{1}+\cdots+\mu_{s}=n}\sum_{\gamma}\frac{\lambda!}{\gamma!}x_{1}^{\mu_{1}}\cdots x_{s}^{\mu_{s}}.

On the other hand, the coefficient in front of xμ:=x1μ1xsμsx^{\mu}:=x_{1}^{\mu_{1}}\cdots x_{s}^{\mu_{s}} is n!/μ!n!/\mu!. Comparing the two formulas, we get the identity that we wanted to prove. ∎

We proved the following proposition.

Proposition 2.1.1.

Suppose that λ\lambda and μ\mu are partitions of nn of sizes respectively rr and ss. The double cosets in Sμ\Sn/SλS_{\mu}\backslash S_{n}/S_{\lambda} are parametrized by matrices γ=(γij)\gamma=(\gamma_{ij}) of size r×sr\times s whose entries are non-negative integers satisfying condition (8). Moreover, let us associate with each γ\gamma a permutation σ\sigma as explained above, then the resulting set of permutations gives a complete set of representatives for the double cosets in Sμ\Sn/SλS_{\mu}\backslash S_{n}/S_{\lambda}.

Note that in Proposition 2.1.1 the condition that λ\lambda and μ\mu are partitions can be relaxed, that is, we do not have to require that λ=(λ1,,λr)\lambda=(\lambda_{1},\dots,\lambda_{r}) and μ=(μ1,,μs)\mu=(\mu_{1},\dots,\mu_{s}) are decreasing sequences.

2.2. Induced vector bundles

We will assume that the reader is familiar with the notion of a GG-space and GG-vector bundle. For some background, we refer to [17]. Suppose that GG is a finite group, HH a subgroup of GG, and XX is a GG-space. For every HH-bundle on XX we define a GG-bundle IndHG(E)\operatorname{Ind}_{H}^{G}(E) on XX whose points are the set theoretic maps f:GEf:G\to E satisfying the following two conditions

  1. (i)

    The composition πf:GX\pi\circ f:G\to X is GG-equivariant, that is, π(f(g))=gπ(f(1))\pi(f(g))=g\pi(f(1)) for all gGg\in G where π:EX\pi:E\to X is the projection map.

  2. (ii)

    The map ff is HH-equivariant, that is, f(hg)=hf(g)f(hg)=hf(g) for all hHh\in H and gGg\in G.

Note that when XX is a pont we have: EE is a representation of HH and IndHG(E)\operatorname{Ind}_{H}^{G}(E) is the induced representation. The structure of a GG-bundle on IndHG(E)\operatorname{Ind}_{H}^{G}(E) is defined as follows. First, the GG-action on IndHG(E)\operatorname{Ind}_{H}^{G}(E) is defined by (gf)(g):=f(gg)(gf)(g^{\prime}):=f(g^{\prime}g). The structure projection p:IndHG(E)Xp:\operatorname{Ind}_{H}^{G}(E)\to X is defined by fπ(f(1))f\mapsto\pi(f(1)). It is strightforward to check that pp is a GG-equivariant map. Suppose that f:GEf:G\to E is in the fiber p1(x)p^{-1}(x). Then π(f(g))=gπ(f(1))=gp(f)=gx\pi(f(g))=g\pi(f(1))=gp(f)=gx, that is, f(g)Egxf(g)\in E_{gx} where EyE_{y} denotes the fiber of EE at yy. Using the linear strucure on the fibers of EE we get that p1(x)p^{-1}(x) is naturally a linear vector space: (f1+f2)(g):=f1(g)+f2(g)(f_{1}+f_{2})(g):=f_{1}(g)+f_{2}(g) and (cf)(g):=cf(g)(cf)(g):=cf(g). Let g1,g2,,gsg_{1},g_{2},\dots,g_{s} be a complete set of representatives of the right coset classes in H\GH\backslash G. Since f(hg)=hf(g)f(hg)=hf(g) we see that the map ff is uniquely determined by its values f(gi)Egixf(g_{i})\in E_{g_{i}x}. This proves that p1(x)p^{-1}(x) is a finite dimensional vector space of dimension rk(E)|G:H|\operatorname{rk}(E)|G:H| where |G:H||G:H| is the index of HH in GG. Moreover, there is a set-theoretic isomorphism

(9) ϕ:j=1sgjEIndHG(E),ϕ(v1,,vs)(g)=ggi1g~i(vi),\phi:\bigoplus_{j=1}^{s}g_{j}^{*}E\to\operatorname{Ind}_{H}^{G}(E),\quad\phi(v_{1},\dots,v_{s})(g)=gg_{i}^{-1}\cdot\widetilde{g}_{i}(v_{i}),

where i=i(g)i=i(g) is the unique index, such that, gHgig\in Hg_{i} (note that ggi1Hgg_{i}^{-1}\in H acts on EE) and g~i\widetilde{g}_{i} is the map defined by the following Cartesian square:

where slightly abusing the notation we denote by gi:XXg_{i}:X\to X the map defined by xgixx\mapsto g_{i}x. Using the map (9), we equip IndHG(E)\operatorname{Ind}_{H}^{G}(E) with the structure of a vector bundle over XX. It is easy to check that the GG-action on IndHG(E)\operatorname{Ind}_{H}^{G}(E) defined above is continuous and therefore IndHG(E)\operatorname{Ind}_{H}^{G}(E) is a topological GG-vector bundle on XX.

We also have a restriction functor. Namely, if XX is a GG-space, HGH\leq G is a subgroup, and EE is a GG-vector bundle, then EE is also an HH-vector bundle which will be denoted by ResHG(E)\operatorname{Res}^{G}_{H}(E). The Frobenius reciprocity rules take the following form.

Proposition 2.2.1 (Frobenius reciprocity).

Suppose that XX is a GG-space and that HGH\leq G is a finite subgroup. Let EE be a GG-vector bundle and FF be an HH-vector bundle.

a) We have an isomorphism

HomG(E,IndHG(F))HomH(ResHG(E),F)\displaystyle\operatorname{Hom}_{G}(E,\operatorname{Ind}_{H}^{G}(F))\cong\operatorname{Hom}_{H}(\operatorname{Res}^{G}_{H}(E),F)

uniquely determined by the following relation:

J(e)(g)=j(ge),eE,gG,\displaystyle J(e)(g)=j(ge),\quad e\in E,\quad g\in G,

where JHomG(E,IndHG(F))J\in\operatorname{Hom}_{G}(E,\operatorname{Ind}_{H}^{G}(F)) and jHomH(ResHG(E),F)j\in\operatorname{Hom}_{H}(\operatorname{Res}^{G}_{H}(E),F) are maps that correspond to each other via the above isomorphism.

b) We have an isomorphism

HomG(IndHG(F),E)HomH(F,ResHG(E))\displaystyle\operatorname{Hom}_{G}(\operatorname{Ind}_{H}^{G}(F),E)\cong\operatorname{Hom}_{H}(F,\operatorname{Res}^{G}_{H}(E))

uniquely determined by the following relation:

J(f)=γG/Hγjf(γ1),fIndHG(F),\displaystyle J(f)=\sum_{\gamma\in G/H}\gamma\,j\circ f(\gamma^{-1}),\quad f\in\operatorname{Ind}_{H}^{G}(F),

where JHomG(IndHG(F),E)J\in\operatorname{Hom}_{G}(\operatorname{Ind}_{H}^{G}(F),E) and jHomH(F,ResHG(E))j\in\operatorname{Hom}_{H}(F,\operatorname{Res}^{G}_{H}(E)) are maps that correspond to each other via the above isomorphism.

If XX 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 XX is a GG-space, H1H_{1} and H2H_{2} are finite subgroups of GG, and WW is a H1H_{1}-vector bundle. For each sGs\in G, let Ks:={(h1,h2)H1×H2|h1s=sh2}K_{s}:=\{(h_{1},h_{2})\in H_{1}\times H_{2}\ |\ h_{1}s=sh_{2}\}. Using the projection on the ii-th factor H1×H2HiH_{1}\times H_{2}\to H_{i}, we view KsK_{s} as a subgroup of HiH_{i} for i=1,2i=1,2. We have the following relation in KH2(X)K_{H_{2}}(X):

ResH2GIndH1G(W)=sH1\G/H2IndKsH2sResKsH1(W),\displaystyle\operatorname{Res}_{H_{2}}^{G}\,\operatorname{Ind}_{H_{1}}^{G}(W)=\sum_{s\in H_{1}\backslash G/H_{2}}\operatorname{Ind}_{K_{s}}^{H_{2}}\,s^{*}\,\operatorname{Res}^{H_{1}}_{K_{s}}(W),

where the sum is over a set of elements sGs\in G which represent the double coset classes in H1\G/H2H_{1}\backslash G/H_{2}.

In the case when XX 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 EE is a complex vector bundle on some compact topological space YY and that g:EEg:E\to E is a finite order automorphism acting trivially on the base YY. To be more precise, we have g(y)=yg(y)=y for all yYy\in Y and the induced map gy:EyEyg_{y}:E_{y}\to E_{y} is a finite order linear map. In particular, the maps gyg_{y} are diagonalizable and their eigenvalues are independent of yy. The trace of gg on EE is a virtual vector bundle on YY defined by

Trg(E)=λλEλ,Eλ:=Ker(λg:EE),\displaystyle\operatorname{Tr}_{g}(E)=\sum_{\lambda\in\mathbb{C}}\lambda E_{\lambda},\quad E_{\lambda}:=\operatorname{Ker}(\lambda-g:E\to E),

where only finitely many terms in the above sum are non-zero because λ\lambda must be an eigenvalue of gg and EλE_{\lambda} is a vector bundle because the eigenvalues of gyg_{y} and their multiplicities are independent of yy. The following lemma is well known (see [1], Section 2).

Lemma 2.3.1.

Let EE be a complex vector bundle on XX, then ψn(E):=Tr(1,2,,n)(En)\psi^{n}(E):=\operatorname{Tr}_{(1,2,\dots,n)}(E^{\otimes n}) is the Adam’s operation.

Proof.

The operation is compatible with pullback, that is, ψn(fE)=fψn(E)\psi^{n}(f^{*}E)=f^{*}\psi^{n}(E) for all continuous maps f:XYf:X\to Y. Recalling the splitting principle definition of the Adam’s operations, we get that it is sufficient to prove that if E=L1++LrE=L_{1}+\cdots+L_{r} is a direct sum of line bundles, then ψn(E)=L1n++Lrn\psi^{n}(E)=L_{1}^{n}+\cdots+L_{r}^{n}. This is obvious because in the tensor product

En=1i1,,inrLi1Lin\displaystyle E^{\otimes n}=\oplus_{1\leq i_{1},\cdots,i_{n}\leq r}L_{i_{1}}\otimes\cdots\otimes L_{i_{n}}

all terms for which the sequence (i1,,in1,in)(i_{1},\dots,i_{n-1},i_{n}) is not invariant under the cyclic permutation do not contribute to the trace. On the other hand, the sequence equals its cyclic permutation (i2,,in,i1)(i_{2},\dots,i_{n},i_{1}) iff i1=i2==ini_{1}=i_{2}=\dots=i_{n}. ∎

Using the Chern character map, we get that the Adam’s operation ψm:K(X)K(X)\psi^{m}:K(X)\to K(X) is an isomorphism where recall that we work with the K-ring with complex coefficients. Therefore, for every FK(X)F\in K(X), there exists an EK(X)E\in K(X), such that, ψm(E)=F\psi^{m}(E)=F. We will need also the formula for the trace of an induced vector bundle.

Proposition 2.3.1.

Suppose that GG is a finite group, HGH\leq G is a subgroup, and YY is a trivial GG-space. For every HH-vector bundle on YY we have the following formula:

Trg(IndHG(E))=i:giggi1HTrgiggi1(E),\displaystyle\operatorname{Tr}_{g}(\operatorname{Ind}_{H}^{G}(E))=\sum_{i:g_{i}gg_{i}^{-1}\in H}\operatorname{Tr}_{g_{i}gg_{i}^{-1}}(E),

where g1,,gkGg_{1},\dots,g_{k}\in G is a complete set of representatives for the right coset classes H\GH\backslash G.

Proof.

Since GG acts trivially on YY, we have IndHG(E)i=1kE\operatorname{Ind}_{H}^{G}(E)\cong\oplus_{i=1}^{k}E where the isomorphism is given by f(f(g1),,f(gk))f\mapsto(f(g_{1}),\dots,f(g_{k})). Let σSk\sigma\in S_{k} be the permutation defined by Hgig=Hgσ(i)Hg_{i}g=Hg_{\sigma(i)} and let hiHh_{i}\in H be defined by gig=higσ(i)g_{i}g=h_{i}g_{\sigma(i)}. If I={i1,,ir}I=\{i_{1},\dots,i_{r}\} is an orbit of σ\sigma, then the subspace iIE\oplus_{i\in I}E is gg-invariant. Therefore,

Trg(IndHG(E))=ITrg(iIE),\displaystyle\operatorname{Tr}_{g}(\operatorname{Ind}_{H}^{G}(E))=\sum_{I}\operatorname{Tr}_{g}\Big{(}\bigoplus_{i\in I}E\Big{)},

where the sum is over all orbits II of σ\sigma in {1,2,,k}\{1,2,\dots,k\}. Note that the action of gg on iIE\oplus_{i\in I}E is represented by a I×II\times I matrix HIH_{I} with entries in End(E)\operatorname{End}(E) of the following type

HI:=[0Hi10000Hi20000Hir1Hir000],\displaystyle H_{I}:=\begin{bmatrix}0&H_{i_{1}}&0&\cdots&0\\ 0&0&H_{i_{2}}&\cdots&0\\ \vdots&\vdots&\ddots&\ddots&\vdots\\ 0&0&\cdots&0&H_{i_{r-1}}\\ H_{i_{r}}&0&\cdots&0&0\end{bmatrix},

where {i1,,ir}:=I\{i_{1},\dots,i_{r}\}:=I and HiEnd(E)H_{i}\in\operatorname{End}(E) is the operator representing the action of hi=giggσ(i)1h_{i}=g_{i}gg_{\sigma(i)}^{-1}. Note that

Trg(iIE)=TrhI(E),\displaystyle\operatorname{Tr}_{g}\Big{(}\oplus_{i\in I}E\Big{)}=\operatorname{Tr}_{h_{I}}(E),

where hI:=Tr(HI)End(E)h_{I}:=\operatorname{Tr}(H_{I})\in\operatorname{End}(E). Since hI=0h_{I}=0 for r>1r>1, we get that only the one-point orbits I={i}I=\{i\} of σ\sigma contribute to the trace of gg. It remains only to notice that the one point orbits I={i}I=\{i\}, that is, the fixed points of σ\sigma, correspond precisely to those ii for which giggi1Hg_{i}gg_{i}^{-1}\in H and that hI=giggi1h_{I}=g_{i}gg_{i}^{-1}. ∎

Suppose that WK(X)W\in K(X) is a virtual vector bundle on XX. Let us define the contraction operation

ιm(W):KSm(Xm),Etr(1,2,,m)π(EWm),\displaystyle\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 40.40298pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-40.40298pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\iota_{-m}(W):K_{S_{m}}(X^{m})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 64.40298pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 64.40298pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathbb{C},}$}}}}}}}\ignorespaces}}}}\ignorespaces\quad E\mapsto\operatorname{tr}_{(1,2,\dots,m)}\pi_{*}(E\otimes W^{\boxtimes m}),

where π:Xmpt\pi:X^{m}\to\operatorname{pt} is the contraction map and π\pi_{*} is the K-theoretic pushforward, that is,

ιm(W)(E):=i=0(1)itr(1,2,,m)Hi(Xm,EWm).\displaystyle\iota_{-m}(W)(E):=\sum_{i=0}^{\infty}(-1)^{i}\operatorname{tr}_{(1,2,\dots,m)}H^{i}(X^{m},E\otimes W^{\boxtimes m}).

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 EE is a SmS_{m}-bundle on XmX^{m}. Then

ιm(W)(E)=χ(FW),\displaystyle\iota_{-m}(W)(E)=\chi(F\otimes W),

where FK(X)F\in K(X) is such that ψm(F)=Tr(1,2,,m)(E|X)\psi^{m}(F)=\operatorname{Tr}_{(1,2,\dots,m)}(E|_{X}) where XXmX\subset X^{m} is the diagonal of XmX^{m}, that is, the submnaifold {xXm|x1==xm}\{x\in X^{m}\ |\ x_{1}=\cdots=x_{m}\}.

Proof.

Put c=(1,2,,m)Smc=(1,2,\dots,m)\in S_{m}. According to the Lefschetz trace formula, ιm(W)(E)\iota_{-m}(W)(E) coinicdes with the Euler characteristics of the following vector bundle

(12) Trc(E|XWm)strc((NXm|X)),\frac{\operatorname{Tr}_{c}(E|_{X}\otimes W^{\otimes m})}{\operatorname{str}_{c}(\wedge^{\bullet}(N^{\vee}_{X^{m}|X}))},

where NXm|XN^{\vee}_{X^{m}|X} is the conormal bundle to XX in XmX^{m} and

strc((NXm|X))=i=0(1)iTrci(NXm|X).\displaystyle\operatorname{str}_{c}(\wedge^{\bullet}(N^{\vee}_{X^{m}|X}))=\sum_{i=0}^{\infty}(-1)^{i}\operatorname{Tr}_{c}\wedge^{i}(N^{\vee}_{X^{m}|X}).

The numerator in (12) is precisely ψm(FW)\psi^{m}(F\otimes W). The denominator can be expressed in terms of the K-theoretic Chern roots of TXTX. Indeed, we have

We get NXm|X=TXm1N_{X^{m}|X}=TX\otimes\mathbb{C}^{m-1}, where m1\mathbb{C}^{m-1} is identified with the hyperplane x1++xm=0x_{1}+\cdots+x_{m}=0 in m\mathbb{C}^{m} and the action of cc on NXm|XN_{X^{m}|X} corresponds to permuting cyclically the coordinates (x1,,xm)m1(x_{1},\dots,x_{m})\in\mathbb{C}^{m-1}. If L1,,LDL_{1},\dots,L_{D} are the K-theoretic Chern roots of TXTX, then we get

NXm|X=i=1Dj=1m1Li1vj,\displaystyle N^{\vee}_{X^{m}|X}=\bigoplus_{i=1}^{D}\bigoplus_{j=1}^{m-1}L_{i}^{-1}\otimes v_{j},

where vjm1v_{j}\in\mathbb{C}^{m-1} is the eigenvector of cc with eigenvalue ηj\eta^{j} where η=e2π𝐢/m\eta=e^{2\pi\mathbf{i}/m}. The denominator in (12) takes the form

strc((NXm|X))=i=1Dj=1m1(1Li1ηj).\displaystyle\operatorname{str}_{c}(\wedge^{\bullet}(N^{\vee}_{X^{m}|X}))=\prod_{i=1}^{D}\prod_{j=1}^{m-1}(1-L^{-1}_{i}\eta^{j}).

Recalling the Hierzerbruch–Riemann–Roch formula, we get that the Euler characterisics of (12) is

Xi=1Dxi1exich(ψm(FW))i=1Dj=1m1(1exiηj)=Xi=1Dxi1emxich(ψm(FW)).\displaystyle\int_{X}\prod_{i=1}^{D}\frac{x_{i}}{1-e^{-x_{i}}}\,\frac{\operatorname{ch}(\psi^{m}(F\otimes W))}{\prod_{i=1}^{D}\prod_{j=1}^{m-1}(1-e^{-x_{i}}\eta^{j})}=\int_{X}\prod_{i=1}^{D}\frac{x_{i}}{1-e^{-mx_{i}}}\,\operatorname{ch}(\psi^{m}(F\otimes W)).

Let deg:H(X)H(X)\operatorname{deg}:H^{*}(X)\to H^{*}(X) be the complex degree operator, that is, deg(ϕ)=iϕ\operatorname{deg}(\phi)=i\phi for ϕH2i(X)\phi\in H^{2i}(X). Note that chψm=mdegch\operatorname{ch}\circ\psi^{m}=m^{\operatorname{deg}}\circ\operatorname{ch} and that Xα=mDXmdegα\int_{X}\alpha=m^{D}\int_{X}m^{-\operatorname{deg}}\alpha. Therefore, the above formula coincides with

mDXi=1Dm1xi1exich(FW)=χ(FW).\displaystyle m^{D}\int_{X}\prod_{i=1}^{D}\frac{m^{-1}x_{i}}{1-e^{-x_{i}}}\,\operatorname{ch}(F\otimes W)=\chi(F\otimes W).\qed
Corollary 2.3.1.

We have ιm(W)(νm,α)=mχ(WΦα)\iota_{-m}(W)(\nu_{m,\alpha})=m\,\chi(W\otimes\Phi^{\alpha}).

Proof.

We have

Trc((Φα)mpm)=Trc((Φα)mTrc(pm)=mψm(Φα)=ψm(mΦα).\displaystyle\operatorname{Tr}_{c}((\Phi^{\alpha})^{\otimes m}\otimes p_{m})=\operatorname{Tr}_{c}((\Phi^{\alpha})^{\otimes m}\otimes\operatorname{Tr}_{c}(p_{m})=m\psi^{m}(\Phi^{\alpha})=\psi^{m}(m\Phi^{\alpha}).

It remains only to recall Lemma 2.3.2. ∎

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 (X)[νm,α(1αm,n0)]\mathcal{F}(X)\cong\mathbb{C}[\nu_{m,\alpha}(1\leq\alpha\leq m,n\geq 0)], we have ιm(Φα)=m/νm,α\iota_{-m}(\Phi_{\alpha})=m\,\partial/\partial\nu_{m,\alpha}.

Proof.

Let us compute the composition

where k:=n+l=m+n′′k:=n^{\prime}+l=m+n^{\prime\prime}. By definition

ιm(W)(νl,αE)=ιm(W)idResSm×Sn′′SkIndSl×SnSk(VlplE),\displaystyle\iota_{-m}(W)(\nu_{l,\alpha}E)=\iota_{-m}(W)\otimes\operatorname{id}\,\circ\,\operatorname{Res}^{S_{k}}_{S_{m}\times S_{n^{\prime\prime}}}\circ\operatorname{Ind}^{S_{k}}_{S_{l}\times S_{n^{\prime}}}(V^{\boxtimes l}\otimes p_{l}\boxtimes E),

where we put V=ΦαV=\Phi^{\alpha} for the ease of notation. We would like to use Proposition 2.2.2 with G=SkG=S_{k}, H1:=Sl×SnH_{1}:=S_{l}\times S_{n^{\prime}} and H2=Sm×Sn′′H_{2}=S_{m}\times S_{n^{\prime\prime}}. The double cosets in Sl×Sn\G/Sm×Sn′′S_{l}\times S_{n^{\prime}}\backslash G/S_{m}\times S_{n^{\prime\prime}}, according to Proposition 2.1.1, are parametrized by the set of 2×22\times 2 matrices

[abcd],a+b=ma+c=l,c+d=n′′b+d=n,\displaystyle\begin{bmatrix}a&b\\ c&d\end{bmatrix},\quad\begin{matrix}a+b=m&a+c=l,\\ c+d=n^{\prime\prime}&b+d=n^{\prime},\end{matrix}

where a,b,c,d0a,b,c,d\in\mathbb{Z}_{\geq 0}. Moreover, if sSks\in S_{k} is the representative of the double coset Sl×Sn\G/Sm×Sn′′S_{l}\times S_{n^{\prime}}\backslash G/S_{m}\times S_{n^{\prime\prime}} corresponding to a matrix of the above form (see Proposition 2.1.1), then

K1,s=H1sH2s1=Sa×Sc×Sb×Sd\displaystyle K_{1,s}=H_{1}\cap sH_{2}s^{-1}=S_{a}\times S_{c}\times S_{b}\times S_{d}

and

K2,s=H2s1H1s=Sa×Sb×Sc×Sd.\displaystyle K_{2,s}=H_{2}\cap s^{-1}H_{1}s=S_{a}\times S_{b}\times S_{c}\times S_{d}.

In other words, sSks\in S_{k} is the permutation that fixes the first aa and the last dd numbers in 1,2,,k1,2,\dots,k, that is,

s(i)=i(1ia),s(a+b+c+j)=a+b+c+j(1jd),\displaystyle s(i)=i\ (1\leq i\leq a),\quad s(a+b+c+j)=a+b+c+j\ (1\leq j\leq d),

and that switches the two blocks a+1,,a+ba+1,\dots,a+b and a+b+1,,a+b+ca+b+1,\dots,a+b+c, that is,

s(a+i)=a+c+i(1ib),s(a+b+j)=a+j(1jc).\displaystyle s(a+i)=a+c+i\ (1\leq i\leq b),\quad s(a+b+j)=a+j\ (1\leq j\leq c).

According to Proposition 2.2.2 we have

(17) ResSm×Sn′′SkIndSl×SnSk(VlplE)=IndSa×Sb×Sc×SdSm×Sn′′sResSa×Sc×Sb×SdSl×Sn(VlplE),\operatorname{Res}^{S_{k}}_{S_{m}\times S_{n^{\prime\prime}}}\circ\operatorname{Ind}^{S_{k}}_{S_{l}\times S_{n^{\prime}}}(V^{\boxtimes l}\otimes p_{l}\boxtimes E)=\sum\operatorname{Ind}^{S_{m}\times S_{n^{\prime\prime}}}_{S_{a}\times S_{b}\times S_{c}\times S_{d}}\circ s^{*}\circ\operatorname{Res}^{S_{l}\times S_{n^{\prime}}}_{S_{a}\times S_{c}\times S_{b}\times S_{d}}(V^{\boxtimes l}\otimes p_{l}\boxtimes E),

where the sum is over all 2×22\times 2 matrices of the above type representing the double cosets in H1\G/H2H_{1}\backslash G/H_{2}. Let us examine the restriction

(18) ResSa×Sc×Sb×SdSl×Sn(VlplE).\operatorname{Res}^{S_{l}\times S_{n^{\prime}}}_{S_{a}\times S_{c}\times S_{b}\times S_{d}}(V^{\boxtimes l}\otimes p_{l}\boxtimes E).

Since ResSa×ScSl(pl)\operatorname{Res}^{S_{l}}_{S_{a}\times S_{c}}(p_{l}) is non-zero only if Sl=Sa×ScS_{l}=S_{a}\times S_{c}, we get that the above restriction is non-zero only in two cases: a=0a=0, c=lc=l or a=la=l, c=0c=0. Let us analyze these two cases separately.

First, suppose that a=0a=0 and c=lc=l. Then b=mb=m and d=nm=n′′ld=n^{\prime}-m=n^{\prime\prime}-l. We get

ResSa×Sc×Sb×SdSl×Sn(VlplE))=VlplResSnSm×Snm(E).\displaystyle\operatorname{Res}^{S_{l}\times S_{n^{\prime}}}_{S_{a}\times S_{c}\times S_{b}\times S_{d}}(V^{\boxtimes l}\otimes p_{l}\boxtimes E))=V^{\boxtimes l}\otimes p_{l}\boxtimes\operatorname{Res}^{S_{n^{\prime}}}_{S_{m}\times S_{n^{\prime}-m}}(E).

The contribution to (17) becomes

IndSm×Sl×Sn′′lSm×Sn′′s(VlplResSm×SnmSn(E)).\displaystyle\operatorname{Ind}^{S_{m}\times S_{n^{\prime\prime}}}_{S_{m}\times S_{l}\times S_{n^{\prime\prime}-l}}\circ s^{*}(V^{\boxtimes l}\otimes p_{l}\boxtimes\operatorname{Res}^{S_{n^{\prime}}}_{S_{m}\times S_{n^{\prime}-m}}(E)).

Applying to the above expression the contraction operation ιm(W)\iota_{-m}(W) we get νl,αιm(W)(E)\nu_{l,\alpha}\,\iota_{-m}(W)(E).

The second case is a=la=l and c=0c=0. Then b=ml=nn′′b=m-l=n^{\prime}-n^{\prime\prime} and d=n′′d=n^{\prime\prime}. The restriction (18) becomes

ResSa×Sc×Sb×SdSl×Sn(VlplE))=VlplResSnSnn′′×Sn′′(E).\displaystyle\operatorname{Res}^{S_{l}\times S_{n^{\prime}}}_{S_{a}\times S_{c}\times S_{b}\times S_{d}}(V^{\boxtimes l}\otimes p_{l}\boxtimes E))=V^{\boxtimes l}\otimes p_{l}\boxtimes\operatorname{Res}^{S_{n^{\prime}}}_{S_{n^{\prime}-n^{\prime\prime}}\times S_{n^{\prime\prime}}}(E).

The contribution to (17) becomes (note that now s=ids=\operatorname{id})

IndSl×Sml×Sn′′Sm×Sn′′(VlplResSml×Sn′′Sn(E)).\displaystyle\operatorname{Ind}^{S_{m}\times S_{n^{\prime\prime}}}_{S_{l}\times S_{m-l}\times S_{n^{\prime\prime}}}\Big{(}V^{\boxtimes l}\otimes p_{l}\boxtimes\operatorname{Res}^{S_{n^{\prime}}}_{S_{m-l}\times S_{n^{\prime\prime}}}(E)\Big{)}.

Note that when we apply the contraction operation ιm(W)\iota_{-m}(W) we will get a sum of terms that have a factor of the form Tr(1,2,,m)IndSl×SmlSm(VlplF)\operatorname{Tr}_{(1,2,\dots,m)}\operatorname{Ind}^{S_{m}}_{S_{l}\times S_{m-l}}(V^{\otimes l}\otimes p_{l}\otimes F) where FF is a representation of SmlS_{m-l}, that is, a SmlS_{m-l}-vector bundle on XX for which SmlS_{m-l} acts trivially on XX. This trace is 0 for mlm\neq l. 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 Sl×SmlS_{l}\times S_{m-l} of the form x(1,2,,m)x1x(1,2,\dots,m)x^{-1}. The latter is a cycle of length mm. However, the group Sl×SmlS_{l}\times S_{m-l} contains such elements only if m=lm=l. Therefore, the contribution in this case is non-zero only if m=lm=l, that is,

(ιm(W)id)(VmpmE)=mχ(WV)E.\displaystyle(\iota_{-m}(W)\otimes\operatorname{id})\,(V^{\boxtimes m}\otimes p_{m}\boxtimes E)=m\,\chi(W\otimes V)\,E.

Recall that V=ΦαV=\Phi^{\alpha} and let us specalize W=ΦβW=\Phi_{\beta}. Then the above computations imply the following commutation relations:

[ιm(Φβ),νl,α]=mδm,lδα,β.\displaystyle[\iota_{-m}(\Phi_{\beta}),\nu_{l,\alpha}]=m\delta_{m,l}\delta_{\alpha,\beta}.

The statement in the theorem follows. ∎

3. Permutation-equivariant K-theoretic Gromov–Witten theory

Suppose now that XX is a smooth projective variety. We would like to recall the background on permutation-equivariant KGW theory. Let

(X)[νm,α(1αN,m1)]\displaystyle\mathcal{F}(X)\cong\mathbb{C}[\nu_{m,\alpha}(1\leq\alpha\leq N,m\geq 1)]

be the KK-theoretic Fock space of XX.

3.1. Translation symmetry

The correlators (3) have the following symmetry:

(19) ν1,α𝐭(L1),,𝐭(Ln)0,n(ν)=𝐭(L1),,𝐭(Ln),Φα0,n+1(ν).\frac{\partial}{\partial\nu_{1,\alpha}}\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{0,n}(\nu)=\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n}),\Phi_{\alpha}\rangle_{0,n+1}(\nu).

The derivative on the LHS is equivalent to the operation ι1(Φα)\iota_{-1}(\Phi_{\alpha}). The correlator has the form ev(k)(Aev(n)B)\operatorname{ev}^{(k)}_{*}(A\otimes\operatorname{ev}_{(n)}^{*}B). Its derivative becomes

π(Φα1k1ev(k)(Aev(n)B))=π(ev(k)(Aev(n)Bevn+1Φα))\displaystyle\pi_{*}\left(\Phi_{\alpha}\boxtimes 1^{\boxtimes k-1}\otimes\operatorname{ev}^{(k)}_{*}(A\otimes\operatorname{ev}_{(n)}^{*}B)\right)=\pi_{*}\left(\operatorname{ev}^{(k)}_{*}(A\otimes\operatorname{ev}_{(n)}^{*}B\otimes\operatorname{ev}_{n+1}^{*}\Phi_{\alpha})\right)

where π:Xk=X×Xk1Xk1\pi:X^{k}=X\times X^{k-1}\to X^{k-1} is the projection map and we used that

(ev(k))(Φα1k1)=evn+1Φα.\displaystyle(\operatorname{ev}^{(k)})^{*}(\Phi_{\alpha}\boxtimes 1^{\boxtimes k-1})=\operatorname{ev}_{n+1}^{*}\Phi_{\alpha}.

Note that πev(k)=ev(k1)\pi\circ\operatorname{ev}^{(k)}=\operatorname{ev}^{(k-1)} and ev(n)Bevn+1Φα=ev(n+1)(BΦα)\operatorname{ev}_{(n)}^{*}B\otimes\operatorname{ev}_{n+1}^{*}\Phi_{\alpha}=\operatorname{ev}_{(n+1)}^{*}(B\boxtimes\Phi_{\alpha}).

Using formula (19), we get that no information is lost if we set ν1,α=0\nu_{1,\alpha}=0 for all α\alpha. Indeed, if we do this, then in order to recover the dependence on ν1,α\nu_{1,\alpha} we simply have to translate t0,αt0,α+ν1,αt_{0,\alpha}\mapsto t_{0,\alpha}+\nu_{1,\alpha}.

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 λ\lambda-algebra Λ\Lambda. The inputs of the correlators are parametrized by a sequence of formal variables 𝐭=(𝐭1,𝐭2,)\mathbf{t}=(\mathbf{t}_{1},\mathbf{t}_{2},\dots), 𝐭r={tr,k,α}\mathbf{t}_{r}=\{t_{r,k,\alpha}\} where the index set is defined by r1,k,r\geq 1,k\in\mathbb{Z}, and 1αN1\leq\alpha\leq N and tr,k,αΛt_{r,k,\alpha}\in\Lambda. Let 𝐥=(l1,l2,)\mathbf{l}=(l_{1},l_{2},\dots) be a sequence such that li0l_{i}\geq 0 and li>0l_{i}>0 for only finitely many ii. Givental introduced correlators of the following form:

(20) 𝐭1,,𝐭1l1;𝐭2,,𝐭2l2;;𝐭r,,𝐭rlr;g,𝐥,d,\langle\underbrace{\mathbf{t}_{1},\dots,\mathbf{t}_{1}}_{l_{1}};\underbrace{\mathbf{t}_{2},\dots,\mathbf{t}_{2}}_{l_{2}};\dots;\underbrace{\mathbf{t}_{r},\dots,\mathbf{t}_{r}}_{l_{r}};\dots\rangle_{g,\mathbf{l},d},

where 𝐭i\mathbf{t}_{i} (i=1,2,)(i=1,2,\dots) is repeated lil_{i} times, that is, the correlator has l1+l2+l_{1}+l_{2}+\cdots inputs. Put n=r=1rlrn=\sum_{r=1}^{\infty}rl_{r} and let us fix a permutation hSnh\in S_{n} whose cycle type is determined by 𝐥\mathbf{l}, that is, the action of hh on {1,2,,n}\{1,2,\dots,n\} has exactly lrl_{r} orbits of length rr for all r=1,2,r=1,2,\dots. Here, for a given element hSnh\in S_{n} and i{1,2,,n}i\in\{1,2,\dots,n\} let r>0r>0 be the minimal positive integer, such that, hr(i)=ih^{r}(i)=i. We refer to the sequence (i,h(i),h2(i),,hr1(i))(i,h(i),h^{2}(i),\dots,h^{r-1}(i)) as the orbit of hh through ii and to the number rr as the length of the orbit. The set of all orbits of hh will be denoted by O(h)O(h). The correlator is by definition the following sum of super-traces

(21) α=(αI),β=(βI)strhH(¯g,n(X,d),𝒪g,n,dIO(h)iI(LiβIeviΦαI))IO(h)1rIΨrI(trI,βI,αI),\sum_{\alpha=(\alpha_{I}),\beta=(\beta_{I})}\operatorname{str}_{h}\ H^{*}\Big{(}\overline{\mathcal{M}}_{g,n}(X,d),\mathcal{O}_{g,n,d}\otimes\otimes_{I\in O(h)}\otimes_{i\in I}(L_{i}^{\beta_{I}}\operatorname{ev}_{i}^{*}\Phi_{\alpha_{I}})\Big{)}\ \prod_{I\in O(h)}\frac{1}{r_{I}}\Psi^{r_{I}}(t_{r_{I},\beta_{I},\alpha_{I}}),

where the sum is over all sequences α=(αI)\alpha=(\alpha_{I}) and β=(βI)\beta=(\beta_{I}) consisting of integers 1αIN1\leq\alpha_{I}\leq N and βI\beta_{I}\in\mathbb{Z} – one for each orbit IO(h)I\in O(h), rIr_{I} is the number of elements in the orbit IO(h)I\in O(h), and Ψr\Psi^{r} is the rr-th Adam’s operation in Λ\Lambda. Note that the definition of the correlator (20) is independent of the choice of hh, that is, it depends only on the cycle type 𝐥\mathbf{l} of hh. Finally, the genus-gg potential of permutation-equivariant theory is defined by

(g)(𝐭)=dQd𝐥1rlr!𝐭1,,𝐭1l1;𝐭2,,𝐭2l2;;𝐭r,,𝐭rlr;g,𝐥,d.\displaystyle\mathcal{F}^{(g)}(\mathbf{t})=\sum_{d}Q^{d}\sum_{\mathbf{l}}\frac{1}{\prod_{r}l_{r}!}\,\langle\underbrace{\mathbf{t}_{1},\dots,\mathbf{t}_{1}}_{l_{1}};\underbrace{\mathbf{t}_{2},\dots,\mathbf{t}_{2}}_{l_{2}};\dots;\underbrace{\mathbf{t}_{r},\dots,\mathbf{t}_{r}}_{l_{r}};\dots\rangle_{g,\mathbf{l},d}.

In order to compare to our definition, let us put tr,k,α=0t_{r,k,\alpha}=0 for all r>1r>1 and k0k\neq 0, that is, let us consider only invariants that do not involve descendants at the permutable marked points. Also, let us restrict (3) to ν1,α=0\nu_{1,\alpha}=0. The super-trace in (21) becomes

(22) trhπ(IO>1(h)ΦαIrIev(k)(iO1(h)(Liβiev(Φαi))),\operatorname{tr}_{h}\,\operatorname{\pi}_{*}\left(\boxtimes_{I\in O_{>1}(h)}\Phi_{\alpha_{I}}^{\boxtimes r_{I}}\operatorname{ev}^{(k)}_{*}\Big{(}\prod_{i\in O_{1}(h)}(L_{i}^{\beta_{i}}\operatorname{ev}^{*}(\Phi_{\alpha_{i}})\Big{)}\right),

where O>1(h)O_{>1}(h) (resp. O1(h)O_{1}(h)) denotes the set of orbits of hh of length >1>1 (resp. 11), k=2l2+3l3+k=2l_{2}+3l_{3}+\cdots, and π:Xkpt\pi:X^{k}\to\operatorname{pt} is the contraction map. Without loss of generality, we may assume that O1(h)={(1),(2),,(l1)}O_{1}(h)=\{(1),(2),\dots,(l_{1})\}. Recalling Theorem 2.3.1 we get that the trace (22) coincides with

IO>1(h)(rIνrI,αI)Φα1Lβ1,,Φαl1Lβl1g,l1,d(ν)|ν=0=C𝐦(α1,β1,,αl1,βl1)r>1α=1Nrmr,αmr,α!\displaystyle\left.\prod_{I\in O_{>1}(h)}(r_{I}\partial_{\nu_{r_{I},\alpha_{I}}})\,\langle\Phi_{\alpha_{1}}L^{\beta_{1}},\dots,\Phi_{\alpha_{l_{1}}}L^{\beta_{l_{1}}}\rangle_{g,l_{1},d}(\nu)\right|_{\nu=0}=C_{\mathbf{m}}(\alpha_{1},\beta_{1},\dots,\alpha_{l_{1}},\beta_{l_{1}})\,\prod_{r>1}\prod_{\alpha=1}^{N}r^{m_{r,\alpha}}\,m_{r,\alpha}!

where 𝐦=(mr,α)\mathbf{m}=(m_{r,\alpha}), mr,α:=card{IOr(h)|αI=α}m_{r,\alpha}:=\operatorname{card}\{I\in O_{r}(h)\ |\ \alpha_{I}=\alpha\}, and C𝐦(α1,β1,,αl1,βl1)C_{\mathbf{m}}(\alpha_{1},\beta_{1},\dots,\alpha_{l_{1}},\beta_{l_{1}}) is the coefficient in front of the monomial r>1,ανr,αmr,α\prod_{r>1,\alpha}\nu_{r,\alpha}^{m_{r,\alpha}} in the correlator Φα1Lβ1,,Φαl1Lβl1g,l1,d(ν)\langle\Phi_{\alpha_{1}}L^{\beta_{1}},\dots,\Phi_{\alpha_{l_{1}}}L^{\beta_{l_{1}}}\rangle_{g,l_{1},d}(\nu). Note that α=1Nmr,α=lr\sum_{\alpha=1}^{N}m_{r,\alpha}=l_{r} and that the number of sequences α=(αI)\alpha=(\alpha_{I}) and β=(βI)\beta=(\beta_{I}) in the sum in (21), such that,

IO(h)ΨrI(tr,βI,αI)=t1,β1,α1t1,βl1,αl1r>1α=1NΨr(tr,0,α)mr,α\displaystyle\prod_{I\in O(h)}\Psi^{r_{I}}(t_{r,\beta_{I},\alpha_{I}})=t_{1,\beta_{1},\alpha_{1}}\cdots t_{1,\beta_{l_{1}},\alpha_{l_{1}}}\prod_{r>1}\prod_{\alpha=1}^{N}\Psi^{r}(t_{r,0,\alpha})^{m_{r,\alpha}}

is precisely r>1lr!mr,1!mr,N!\prod_{r>1}\tfrac{l_{r}!}{m_{r,1}!\cdots m_{r,N}!} because the values of βI\beta_{I} and αI\alpha_{I} are fixed for the one-point orbits IO1(h)I\in O_{1}(h) and for the more than one-point-orbits, we have βI=0\beta_{I}=0 and we have to choose mr,αm_{r,\alpha} orbits II in Or(h)O_{r}(h) for which αI\alpha_{I} will be assigned value α\alpha. Now it is clear that the coefficient in front of tβ1,α1tβl1,αl1r>1α=1NΨr(tr,0,α)mr,αt_{\beta_{1},\alpha_{1}}\cdots t_{\beta_{l_{1}},\alpha_{l_{1}}}\prod_{r>1}\prod_{\alpha=1}^{N}\Psi^{r}(t_{r,0,\alpha})^{m_{r,\alpha}} in (21) is C𝐦(α1,β1,,αl1,βl1)r>1lr!C_{\mathbf{m}}(\alpha_{1},\beta_{1},\dots,\alpha_{l_{1}},\beta_{l_{1}})\prod_{r>1}l_{r}!. Therefore, under the substitutions

(23) ν1,α=0,νr,α:=Ψr(tr,0,α)(r>1),tk,α:=t1,k,α(k),tr,k,α:=0(r>1,k0),\nu_{1,\alpha}=0,\quad\nu_{r,\alpha}:=\Psi^{r}(t_{r,0,\alpha})\ (r>1),\quad t_{k,\alpha}:=t_{1,k,\alpha}\ (k\in\mathbb{Z}),\quad t_{r,k,\alpha}:=0\ (r>1,k\neq 0),

where 1αN1\leq\alpha\leq N, we get that the genus-gg potential (g)(𝐭)\mathcal{F}^{(g)}(\mathbf{t}) coincides with

n=01n!𝐭(L1),,𝐭(Ln)g,n(ν).\displaystyle\sum_{n=0}^{\infty}\frac{1}{n!}\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{g,n}(\nu).

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 X×X\times\mathbb{Z} where \mathbb{Z} is equipped with the discrete topology. Note that this is a topological space with infinitely many connected components Xm:=X×{m}X_{m}:=X\times\{m\}, mm\in\mathbb{Z}. The K-theoretic Fock space is defined by

(X×):=limk(X×[k,k])=k=0(X×[k,k]),\displaystyle\mathcal{F}(X\times\mathbb{Z}):=\varinjlim_{k}\ \mathcal{F}(X\times[-k,k])=\bigcup_{k=0}^{\infty}\mathcal{F}(X\times[-k,k]),

where the inclusion (X×[k,k])(X×[k1,k+1])\mathcal{F}(X\times[-k,k])\subset\mathcal{F}(X\times[-k-1,k+1]) is defined via the pushforward with respect to the natural inclusion X×[k,k]X×[k1,k+1]X\times[-k,k]\subset X\times[-k-1,k+1]. Theorem 2.3.1 implies that

(24) (X×)[νr,k,α(r1,k,1αN)],\mathcal{F}(X\times\mathbb{Z})\cong\mathbb{C}[\nu_{r,k,\alpha}\ (r\geq 1,k\in\mathbb{Z},1\leq\alpha\leq N)],

where the vacuum is the same as before, that is, 1(X)=(X×{0})(X×)1\in\mathcal{F}(X)=\mathcal{F}(X\times\{0\})\subset\mathcal{F}(X\times\mathbb{Z}) and the creation operator νr,k,α\nu_{r,k,\alpha} is defined by the induction operation and exterior tensor product by (Φkα)rpr(\Phi^{\alpha}_{k})^{\boxtimes r}\otimes p_{r} where Φkα\Phi^{\alpha}_{k} is a virtual vector bundle on X×X\times\mathbb{Z} whose restriction to the connected component XmX_{m} is Φα\Phi^{\alpha} for m=km=k and 0 if mkm\neq k.

The following isomorphism will be needed in the definition of KGW invariants with values in (X×)\mathcal{F}(X\times\mathbb{Z}). Suppose that Y=Y1YnY=Y_{1}\sqcup\dots\sqcup Y_{n} has finitely many connected components. We have an isomorphism

(25) KSk(Yk)b=(b1,,bk)KSb(Yb1××Ybk),K_{S_{k}}(Y^{k})\cong\bigoplus_{b=(b_{1},\dots,b_{k})}K_{S_{b}}(Y_{b_{1}}\times\cdots\times Y_{b_{k}}),

where the direct sum is over all monotonely increasing sequences satisfying 1b1b2bkn1\leq b_{1}\leq b_{2}\leq\cdots\leq b_{k}\leq n and Sb={σSn|bσ(i)=bii}S_{b}=\{\sigma\in S_{n}\ |\ b_{\sigma(i)}=b_{i}\ \forall i\}. The isomorphism is induced by restricting SkS_{k}-bundles on YkY^{k} to Yb1××YbkY_{b_{1}}\times\cdots\times Y_{b_{k}}. Note that the inverse of this isomorphism, that is, constructing an SkS_{k}-bundle from an SbS_{b}-bundles, is given by an operation that in some sense generalizes the induction operation of vector bundles: first we have to make an SkS_{k}-space from an SbS_{b}-space by adding extra connected components and then extend the SbS_{b}-bundle to an SkS_{k}-bundle via the usual induction operation.

Applying the isomorphism (25) to our settings we get

(26) (X×)b=(b1,,bk)KSb(Xb1××Xbk),\mathcal{F}(X\times\mathbb{Z})\cong\bigoplus_{b=(b_{1},\dots,b_{k})}K_{S_{b}}(X_{b_{1}}\times\cdots\times X_{b_{k}}),

where the direct sum is over all monotonely increasing sequences of integer numbers b1b2bkb_{1}\leq b_{2}\leq\cdots\leq b_{k}, SbS_{b} is the (Young) subgroup of SkS_{k} preserving the sequence bb, and Xbi=X×{bi}X_{b_{i}}=X\times\{b_{i}\} is the bib_{i}-th connected component of X×X\times\mathbb{Z}.

Now we are in position to define KGW invariant with values in (X×)\mathcal{F}(X\times\mathbb{Z}). The definition (2) should be modified as follows. We replace the sum over kk with a sum over all monotonely increasing sequences of integers b=(b1,,bk)b=(b_{1},\dots,b_{k}), that is, b1bkb_{1}\leq\cdots\leq b_{k}, define the evaluation map

ev(b):¯g,n+k(X,d)Xb1××Xbk,(C,p1,,pn+k;f)((f(pn+1),b1),,(f(pn+k),bk)),\displaystyle\operatorname{ev}^{(b)}:\overline{\mathcal{M}}_{g,n+k}(X,d)\to X_{b_{1}}\times\cdots\times X_{b_{k}},\quad(C,p_{1},\dots,p_{n+k};f)\mapsto((f(p_{n+1}),b_{1}),\dots,(f(p_{n+k}),b_{k})),

and replace the pushforward by

ev(b)(𝒪g,n+k,dL1i1LninLn+1b1Ln+kbkev(n)(Φa1Φan)).\displaystyle\operatorname{ev}^{(b)}_{*}\left(\mathcal{O}_{g,n+k,d}\otimes L_{1}^{i_{1}}\otimes\cdots\otimes L_{n}^{i_{n}}\otimes L_{n+1}^{b_{1}}\otimes\cdots\otimes L_{n+k}^{b_{k}}\otimes\operatorname{ev}_{(n)}^{*}(\Phi_{a_{1}}\boxtimes\cdots\boxtimes\Phi_{a_{n}})\right).

The above pushforward takes value in KSb(Xb1××Xbk)K_{S_{b}}(X_{b_{1}}\times\cdots\times X_{b_{k}}). Recalling the isomorphisms (26) and (24) we get that the KGW invariants take value in a certain completion of the polynomial ring [νr,k,α(r1,k,1αN)]\mathbb{C}[\nu_{r,k,\alpha}(r\geq 1,k\in\mathbb{Z},1\leq\alpha\leq N)]. Similarly to (19), we have

(27) ν1,k,α𝐭(L1),,𝐭(Ln)0,n(ν)=𝐭(L1),,𝐭(Ln),Ln+1kΦα0,n+1(ν).\frac{\partial}{\partial\nu_{1,k,\alpha}}\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{0,n}(\nu)=\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n}),L_{n+1}^{k}\Phi_{\alpha}\rangle_{0,n+1}(\nu).

Using the Heisenberg structure of (X×)\mathcal{F}(X\times\mathbb{Z}), the same argument as in Section 3.2 proves that under the substitution

ν1,k,α=t1,k,αtk,α,νr,k,α=Ψr(tr,k,α)(r>1)\displaystyle\nu_{1,k,\alpha}=t_{1,k,\alpha}-t_{k,\alpha},\quad\nu_{r,k,\alpha}=\Psi^{r}(t_{r,k,\alpha})\ (r>1)

the permutation-equivariant genus-g potential takes the form

(g)(𝐭)=n=01n!𝐭(ψ1),,𝐭(ψn)g,n(ν).\displaystyle\mathcal{F}^{(g)}(\mathbf{t})=\sum_{n=0}^{\infty}\frac{1}{n!}\langle\mathbf{t}(\psi_{1}),\dots,\mathbf{t}(\psi_{n})\rangle_{g,n}(\nu).

3.4. String equation

The following formula holds:

𝐭(L1),,𝐭(Ln),10,n+1(ν)=𝐭(L1),,𝐭(Ln)0,n(ν)+\displaystyle\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n}),1\rangle_{0,n+1}(\nu)=\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{0,n}(\nu)+
i=1n𝐭(L1),,𝐭(Li)𝐭(1)Li1,𝐭(Ln)0,n(ν)+Qn(𝐭,ν),\displaystyle\sum_{i=1}^{n}\langle\mathbf{t}(L_{1}),\dots,\frac{\mathbf{t}(L_{i})-\mathbf{t}(1)}{L_{i}-1},\dots\mathbf{t}(L_{n})\rangle_{0,n}(\nu)+Q_{n}(\mathbf{t},\nu),

where Qn(𝐭,ν)=0Q_{n}(\mathbf{t},\nu)=0 for n>2n>2 and

Q2(𝐭,ν)\displaystyle Q_{2}(\mathbf{t},\nu) =(𝐭(1),𝐭(1))\displaystyle=(\mathbf{t}(1),\mathbf{t}(1))
Q1(𝐭,ν)\displaystyle Q_{1}(\mathbf{t},\nu) =(𝐭(1),ν1)\displaystyle=(\mathbf{t}(1),\nu_{1})
Q0(𝐭,ν)\displaystyle Q_{0}(\mathbf{t},\nu) =12(ν1,ν1)+12(ψ2(ν2),1)\displaystyle=\frac{1}{2}(\nu_{1},\nu_{1})+\frac{1}{2}(\psi^{2}(\nu_{2}),1)

where νk=α=1Nνk,αΦα\nu_{k}=\sum_{\alpha=1}^{N}\nu_{k,\alpha}\Phi_{\alpha} and ψm(νk):=α=1Nνk,αψm(Φα)\psi^{m}(\nu_{k}):=\sum_{\alpha=1}^{N}\nu_{k,\alpha}\psi^{m}(\Phi_{\alpha}). 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 Qn(𝐭,ν)Q_{n}(\mathbf{t},\nu).

Suppose that d=0d=0 and n+k=2n+k=2 where kk is the number of permutable marked points and nn is the number of the remaining ones. There are 3 cases. First, if n=2n=2 and k=0k=0, we get

𝐭(L1),𝐭(L2),10,3,0=(𝐭(1),𝐭(1))=Q2(𝐭,ν).\displaystyle\langle\mathbf{t}(L_{1}),\mathbf{t}(L_{2}),1\rangle_{0,3,0}=(\mathbf{t}(1),\mathbf{t}(1))=Q_{2}(\mathbf{t},\nu).

The second case is when n=1n=1 and k=1k=1: the moduli space ¯0,3(X,0)=¯0,3×X=X\overline{\mathcal{M}}_{0,3}(X,0)=\overline{\mathcal{M}}_{0,3}\times X=X and the corresponding contribution to the KGW invariant is

ev(1)(ev(1)(𝐭(1)))=𝐭(1)=Q1(𝐭,ν),\displaystyle\operatorname{ev}^{(1)}_{*}(\operatorname{ev}_{(1)}^{*}(\mathbf{t}(1)))=\mathbf{t}(1)=Q_{1}(\mathbf{t},\nu),

where we used that the topological KK-ring K(X)K(X) is embedded in the Fock space (X)\mathcal{F}(X) via vα=1N(v,Φα)ν1,α=(v,ν1).v\mapsto\sum_{\alpha=1}^{N}(v,\Phi_{\alpha})\nu_{1,\alpha}=(v,\nu_{1}). Finally, if n=0n=0 and k=2k=2: the evaluation map ev(2)\operatorname{ev}^{(2)} coincides with the diagonal embedding Δ:XX×X\Delta:X\to X\times X. We have to compute f(ν1,ν2):=Δ(𝒪X)f(\nu_{1},\nu_{2}):=\Delta_{*}(\mathcal{O}_{X}). Let us compute the derivatives of ff with respect to ν1\nu_{1} and ν2\nu_{2}.

ν1,αf=π(2)((Φα1)Δ(𝒪X))=π(2)Δ(Φα)=Φα=β=1N(Φα,Φβ)ν1,β,\displaystyle\partial_{\nu_{1,\alpha}}f=\pi^{(2)}_{*}((\Phi_{\alpha}\boxtimes 1)\otimes\Delta_{*}(\mathcal{O}_{X}))=\pi^{(2)}_{*}\Delta_{*}(\Phi_{\alpha})=\Phi_{\alpha}=\sum_{\beta=1}^{N}(\Phi_{\alpha},\Phi_{\beta})\nu_{1,\beta},

where π(2):X×XX\pi^{(2)}:X\times X\to X is the projection on the second factor. Similarly,

2ν2,αf=tr(12)π((ΦαΦα)Δ(𝒪X))=tr(12)πΔ(Φα2)=πΔTr(12)(Φα2)=(ψ2(Φα),1),\displaystyle 2\partial_{\nu_{2,\alpha}}f=\operatorname{tr}_{(12)}\pi_{*}((\Phi_{\alpha}\boxtimes\Phi_{\alpha})\otimes\Delta_{*}(\mathcal{O}_{X}))=\operatorname{tr}_{(12)}\pi_{*}\Delta_{*}(\Phi_{\alpha}^{\otimes 2})=\pi_{*}\Delta_{*}\operatorname{Tr}_{(12)}(\Phi_{\alpha}^{\otimes 2})=(\psi^{2}(\Phi_{\alpha}),1),

where π:X×Xpt\pi:X\times X\to{\rm pt} is the constant map and π\pi_{*} is the KK-theoretic pushforward. We get

Δ(𝒪X)=12α,β=1Nχ(ΦαΦβ)ν1,αν1,β+12α=1Nχ(ψ2(Φα))ν2,α=Q0(𝐭,ν).\displaystyle\Delta_{*}(\mathcal{O}_{X})=\frac{1}{2}\sum_{\alpha,\beta=1}^{N}\chi(\Phi_{\alpha}\otimes\Phi_{\beta})\,\nu_{1,\alpha}\nu_{1,\beta}+\frac{1}{2}\sum_{\alpha=1}^{N}\chi(\psi^{2}(\Phi_{\alpha}))\,\nu_{2,\alpha}=Q_{0}(\mathbf{t},\nu).

3.5. Dilaton equation

The following formula is the permutation-equivariant K-theoretic version of the so called dilaton equation:

𝐭(L1),,𝐭(Ln),Ln+110,n+1(ν)=(n2+b=1Nν1,bν1,b)𝐭(L1),,𝐭(Ln)0,n(ν)δn,0(ψ2(ν2),1),\displaystyle\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n}),L_{n+1}-1\rangle_{0,n+1}(\nu)=\Big{(}n-2+\sum_{b=1}^{N}\nu_{1,b}\partial_{\nu_{1,b}}\Big{)}\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{0,n}(\nu)-\delta_{n,0}(\psi^{2}(\nu_{2}),1),

where ψ2(ν2)=α=1Nν2,αψ2(Φα)\psi^{2}(\nu_{2})=\sum_{\alpha=1}^{N}\nu_{2,\alpha}\psi^{2}(\Phi_{\alpha}). Using formula (19) we can rewrite the RHS of the above formula as follows:

(n2)𝐭(L1),,𝐭(Ln)0,n(ν)+𝐭(L1),,𝐭(Ln),ν10,n+1(ν)δn,0(ψ2(ν2),1).\displaystyle(n-2)\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{0,n}(\nu)+\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n}),\nu_{1}\rangle_{0,n+1}(\nu)-\delta_{n,0}(\psi^{2}(\nu_{2}),1).

The standard argument proving the dilaton equation in cohomological Gromov–Witten theory is based on pushing forward along the forgetful map ¯0,n+1+k(X,d)¯0,n+k(X,d)\overline{\mathcal{M}}_{0,n+1+k}(X,d)\to\overline{\mathcal{M}}_{0,n+k}(X,d). 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 d=0d=0 and n+k=2n+k=2, that is, the case when the forgetful map does not exist. If k<2k<2, then Ln+1L_{n+1} is trivial as an SkS_{k}-bundle, so the exceptional term is 0. The only non-trivial contribution comes when n=0n=0 and k=2k=2. We have to compute f(ν):=ev(2)(L11)=Δ(𝒪X(ϵ1))f(\nu):=\operatorname{ev}^{(2)}_{*}(L_{1}-1)=\Delta_{*}(\mathcal{O}_{X}\otimes(\epsilon-1)), where Δ:XX×X\Delta:X\to X\times X is the diagonal embedding and ϵ=L1\epsilon=L_{1} is the alternating representation of S2S_{2}. We have

2να,2f=tr(12)π(Φα2Δ(𝒪X(ϵ1)))=(πΔ)(Tr(12)(Φα2(ϵ1)))=2(ψ2(Φα),1).\displaystyle 2\partial_{\nu_{\alpha,2}}f=\operatorname{tr}_{(12)}\pi_{*}\left(\Phi_{\alpha}^{\boxtimes 2}\Delta_{*}(\mathcal{O}_{X}\otimes(\epsilon-1))\right)=(\pi\circ\Delta)_{*}(\operatorname{Tr}_{(12)}(\Phi_{\alpha}^{\otimes 2}\otimes(\epsilon-1)))=-2(\psi^{2}(\Phi_{\alpha}),1).

Similar computation, since Tr1(ϵ1)=0\operatorname{Tr}_{1}(\epsilon-1)=0, proves that νa,1f=0\partial_{\nu_{a,1}}f=0. Therefore,

f(ν)=a=1Nνa,2(ψ2(Φa),1)=(ψ2(ν2),1).\displaystyle f(\nu)=-\sum_{a=1}^{N}\nu_{a,2}(\psi^{2}(\Phi_{a}),1)=-(\psi^{2}(\nu_{2}),1).

3.6. WDVV equations and the SS-matrix

The WDVV equations say that the expression

(28) a,b=1NGab(ν)ϕ11x1L,ϕ21x2L,Φa0,3(ν)ϕ31x3L,ϕ41x4L,Φb0,3(ν)\sum_{a,b=1}^{N}G^{ab}(\nu)\left\langle\frac{\phi_{1}}{1-x_{1}L},\frac{\phi_{2}}{1-x_{2}L},\Phi_{a}\right\rangle_{0,3}(\nu)\left\langle\frac{\phi_{3}}{1-x_{3}L},\frac{\phi_{4}}{1-x_{4}L},\Phi_{b}\right\rangle_{0,3}(\nu)

is symmetric in the pairs (ϕi,xi)(\phi_{i},x_{i}) (1i41\leq i\leq 4). For example, by exchanging (ϕ2,x2)(\phi_{2},x_{2}) and (ϕ3,x3)(\phi_{3},x_{3}), we get

(29) a,b=1NGab(ν)ϕ11x1L,ϕ31x3L,Φa0,3(ν)ϕ21x2L,ϕ41x4L,Φb0,3(ν).\sum_{a,b=1}^{N}G^{ab}(\nu)\left\langle\frac{\phi_{1}}{1-x_{1}L},\frac{\phi_{3}}{1-x_{3}L},\Phi_{a}\right\rangle_{0,3}(\nu)\left\langle\frac{\phi_{2}}{1-x_{2}L},\frac{\phi_{4}}{1-x_{4}L},\Phi_{b}\right\rangle_{0,3}(\nu).

The equality of (28) and (29) determines the remaining identities. The standard proof based on the forgetful map ¯0,4+k(X,d)¯0,4=1\overline{\mathcal{M}}_{0,4+k}(X,d)\mapsto\overline{\mathcal{M}}_{0,4}=\mathbb{P}^{1} 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 S(ν,q)End(K(X))S(\nu,q)\in\operatorname{End}(K(X)) depending on the complex number qq\in\mathbb{C}^{*} and the parameters ν=(νr,α)\nu=(\nu_{r,\alpha}) – see formula (5).

Proposition 3.6.1.

a) The following formula holds:

G(S(ν,q1)Φa,S(ν,q2)Φb)=(Φa,Φb)+(1q11q21)Φa1q11L,Φb1q21L0,2(ν).\displaystyle G(S(\nu,q_{1})\Phi_{a},S(\nu,q_{2})\Phi_{b})=(\Phi_{a},\Phi_{b})+(1-q_{1}^{-1}q_{2}^{-1})\Big{\langle}\frac{\Phi_{a}}{1-q_{1}^{-1}L},\frac{\Phi_{b}}{1-q_{2}^{-1}L}\Big{\rangle}_{0,2}(\nu).

b) The following formula holds: ST(ν,q1)S(ν,q)=1S^{T}(\nu,q^{-1})S(\nu,q)=1 where STS^{T} is defined by the identity G(Sa,b)=(a,STb)G(Sa,b)=(a,S^{T}b).

c) The matrix SS satisfies the following differential equations:

(q1)ν1,iS(ν,q)=ΦiS(ν,q),1iN,\displaystyle(q-1)\,\partial_{\nu_{1,i}}S(\nu,q)=\Phi_{i}\bullet S(\nu,q),\quad 1\leq i\leq N,

where \bullet is the permutation-equivariant quantum K-product defined by (6).

Proof.

a) Let us apply the WDVV equations with ϕ1=ϕ3=1\phi_{1}=\phi_{3}=1, x1=x3=0x_{1}=x_{3}=0. For brevity, put ϕ=ϕ2\phi=\phi_{2}, x=x2x=x_{2} and ψ=ϕ4\psi=\phi_{4}, y=x4y=x_{4}. We get

1,ϕ1xL,Φa0,31,ψ1yL,Φb0,3Gab=1,1,Φa0,3ϕ1xL,ψ1yL,Φb0,3Gab,\displaystyle\Big{\langle}1,\frac{\phi}{1-xL},\Phi_{a}\Big{\rangle}_{0,3}\Big{\langle}1,\frac{\psi}{1-yL},\Phi_{b}\Big{\rangle}_{0,3}\,G^{ab}=\Big{\langle}1,1,\Phi_{a}\Big{\rangle}_{0,3}\Big{\langle}\frac{\phi}{1-xL},\frac{\psi}{1-yL},\Phi_{b}\Big{\rangle}_{0,3}\,G^{ab},

where we use Einstein’s convention to sum over repeating upper and lower indexes. Using the string equation, we get 1,1,Φa0,3=G(1,Φa)\Big{\langle}1,1,\Phi_{a}\Big{\rangle}_{0,3}=G(1,\Phi_{a}). The RHS takes the form

ϕ1xL,ψ1yL,10,3=(1+x1x+y1y)ϕ1xL,ψ1yL0,2+(ϕ,ψ)(1x)(1y),\displaystyle\Big{\langle}\frac{\phi}{1-xL},\frac{\psi}{1-yL},1\Big{\rangle}_{0,3}=\Big{(}1+\frac{x}{1-x}+\frac{y}{1-y}\Big{)}\,\Big{\langle}\frac{\phi}{1-xL},\frac{\psi}{1-yL}\Big{\rangle}_{0,2}+\frac{(\phi,\psi)}{(1-x)(1-y)},

where we used the string equation and the identity

1L1(11xL11x)=x1x11xL.\displaystyle\frac{1}{L-1}\Big{(}\frac{1}{1-xL}-\frac{1}{1-x}\Big{)}=\frac{x}{1-x}\,\frac{1}{1-xL}.

The RHS of the WDVV equation takes the form

1(1x)(1y)((ϕ,ψ)+(1xy)ϕ1xL,ψ1yL0,2).\displaystyle\frac{1}{(1-x)(1-y)}\Big{(}(\phi,\psi)+(1-xy)\Big{\langle}\frac{\phi}{1-xL},\frac{\psi}{1-yL}\Big{\rangle}_{0,2}\Big{)}.

Similarly, using the string equation we get that

1,ϕ1xL,Φa0,3=11x(ϕ1xL,Φa0,3+(ϕ,Φa))=11xG(S(ν,x1)ϕ,Φa).\displaystyle\Big{\langle}1,\frac{\phi}{1-xL},\Phi_{a}\Big{\rangle}_{0,3}=\frac{1}{1-x}\,\Big{(}\Big{\langle}\frac{\phi}{1-xL},\Phi_{a}\Big{\rangle}_{0,3}+(\phi,\Phi_{a})\Big{)}=\frac{1}{1-x}\,G(S(\nu,x^{-1})\phi,\Phi_{a}).

Now it is clear that

1(1x)(1y)G(S(ν,x1)ϕ,S(ν,y1)ψ)=1(1x)(1y)((ϕ,ψ)+(1xy)ϕ1xL,ψ1yL0,2)\displaystyle\frac{1}{(1-x)(1-y)}\,G(S(\nu,x^{-1})\phi,S(\nu,y^{-1})\psi)=\frac{1}{(1-x)(1-y)}\Big{(}(\phi,\psi)+(1-xy)\Big{\langle}\frac{\phi}{1-xL},\frac{\psi}{1-yL}\Big{\rangle}_{0,2}\Big{)}

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

ϕ1=1,x1=0,ϕ2=ϕ,x2=x,ϕ3=Φi,x3=0,ϕ4=Φj,x4=0.\displaystyle\phi_{1}=1,\ x_{1}=0,\quad\phi_{2}=\phi,\ x_{2}=x,\quad\phi_{3}=\Phi_{i},\ x_{3}=0,\quad\phi_{4}=\Phi_{j},\ x_{4}=0.

We get

1,ϕ1xL,ΦaΦi,Φj,ΦbGab=1,Φi,Φaϕ1xL,Φj,ΦbGab\displaystyle\Big{\langle}1,\frac{\phi}{1-xL},\Phi_{a}\Big{\rangle}\,\langle\Phi_{i},\Phi_{j},\Phi_{b}\rangle\ G^{ab}=\langle 1,\Phi_{i},\Phi_{a}\rangle\,\Big{\langle}\frac{\phi}{1-xL},\Phi_{j},\Phi_{b}\Big{\rangle}\ G^{ab}

Again, using the string equation, we get that 1,Φi,Φa0,3=Gia\langle 1,\Phi_{i},\Phi_{a}\rangle_{0,3}=G_{ia} and that the RHS becomes

ϕ1xL,Φj,Φi0,3=ν1,iG(S(ν,x1)ϕ,Φj)=(ϕ,ν1,iST(ν,x1)Φj).\displaystyle\Big{\langle}\frac{\phi}{1-xL},\Phi_{j},\Phi_{i}\Big{\rangle}_{0,3}=\partial_{\nu_{1,i}}G(S(\nu,x^{-1})\phi,\Phi_{j})=(\phi,\partial_{\nu_{1,i}}\,S^{T}(\nu,x^{-1})\Phi_{j}).

On the other hand, the LHS (see the computation in part a)) after removing 11 via the string equation and recalling the definition of the quantum KK-product, becomes

11xG(S(ν,x1)ϕ,ΦiΦj)=11x(ϕ,ST(ν,x1)ΦiΦj).\displaystyle\frac{1}{1-x}G(S(\nu,x^{-1})\phi,\Phi_{i}\bullet\Phi_{j})=\frac{1}{1-x}(\phi,S^{T}(\nu,x^{-1})\Phi_{i}\bullet\Phi_{j}).

Therefore,

ν1,iST(ν,x1)=11xST(ν,x1)Φi.\displaystyle\partial_{\nu_{1,i}}\,S^{T}(\nu,x^{-1})=\frac{1}{1-x}\,S^{T}(\nu,x^{-1})\Phi_{i}\bullet.

Recalling b), we get ST(ν,x1)=S(ν,x)1S^{T}(\nu,x^{-1})=S(\nu,x)^{-1} 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 𝐭(L¯1),,𝐭(L¯n)g,n(ν)\langle\mathbf{t}(\overline{L}_{1}),\dots,\mathbf{t}(\overline{L}_{n})\rangle_{g,n}(\nu) are defined in the same way as the descendent KGW invariants, that is, by formula (2), except that we replace LiL_{i} with the pullback L¯i=ftLi\overline{L}_{i}=\operatorname{ft}^{*}L_{i} where ft:¯g,n+k(X,d)¯g,n\operatorname{ft}:\overline{\mathcal{M}}_{g,n+k}(X,d)\to\overline{\mathcal{M}}_{g,n} is the map that forgets the stable map and the last kk marked points. According to Givental (see [6]), the ancestor invariants can be be expressed in terms of the descendent ones and the SS-matrix. Let us recall Givental’s formula.

Let (q)\mathbb{C}(q) be the field of all rational functions on \mathbb{C}. Using elementary fraction decomposition, we have a natural projection map:

(30) []+:(q)[q,q1],f(q)[f(q)]+:=Resw=0,f(w)dwwq.[\ ]_{+}:\mathbb{C}(q)\to\mathbb{C}[q,q^{-1}],\quad f(q)\mapsto[f(q)]_{+}:=-\operatorname{Res}_{w=0,\infty}\ \frac{f(w)dw}{w-q}.

It is easy to check that the above residue truncates all terms in the elementary fraction decomposition of f(q)f(q) that have poles in ={0}\mathbb{C}^{*}=\mathbb{C}\setminus{\{0\}}. The matrix S(ν,q)S(\nu,q) has entries that are formal power series in 𝐭\mathbf{t}, ν\nu, and the Novikov variables QQ whose coefficients are rational functions. In particular, we can define [S𝐭]+(ν,q):=[S(ν,q)𝐭(q)]+[S\mathbf{t}]_{+}(\nu,q):=[S(\nu,q)\mathbf{t}(q)]_{+}.

Lemma 4.1.1.

a) The following formula holds:

[S𝐭]+(ν,q)=𝐭(q)+α,β=1N𝐭(L)𝐭(q)LqL,Φα0,2(ν)Gαβ(ν)Φβ.\displaystyle[S\mathbf{t}]_{+}(\nu,q)=\mathbf{t}(q)+\sum_{\alpha,\beta=1}^{N}\Big{\langle}\frac{\mathbf{t}(L)-\mathbf{t}(q)}{L-q}\,L,\Phi_{\alpha}\Big{\rangle}_{0,2}(\nu)\,G^{\alpha\beta}(\nu)\,\Phi_{\beta}.

b) The following formula holds:

[S𝐭]+(ν,1)=α,β=1N1,Φα,𝐭(L)0,3(ν)Gαβ(ν)Φβ.\displaystyle[S\mathbf{t}]_{+}(\nu,1)=\sum_{\alpha,\beta=1}^{N}\Big{\langle}1,\Phi_{\alpha},\mathbf{t}(L)\Big{\rangle}_{0,3}(\nu)\,G^{\alpha\beta}(\nu)\,\Phi_{\beta}.
Proof.

a) Recalling the definition of the SS-matrix we get

G([S𝐭]+(ν,q),Φα)=(𝐭(q),Φα)+Resw=0,𝐭(w)dw(1w1L)(wq),Φα0,2(ν).\displaystyle G([S\mathbf{t}]_{+}(\nu,q),\Phi_{\alpha})=(\mathbf{t}(q),\Phi_{\alpha})+\Big{\langle}-\operatorname{Res}_{w=0,\infty}\frac{\mathbf{t}(w)dw}{(1-w^{-1}L)(w-q)},\Phi_{\alpha}\Big{\rangle}_{0,2}(\nu).

Recalling Cauchy’s theorem we get

Resw=0,𝐭(w)dw(1w1L)(wq)=Resw=L,q𝐭(w)dw(1w1L)(wq)=𝐭(L)L𝐭(q)qLq=𝐭(L)𝐭(q)LqL+𝐭(q).\displaystyle-\operatorname{Res}_{w=0,\infty}\frac{\mathbf{t}(w)dw}{(1-w^{-1}L)(w-q)}=\operatorname{Res}_{w=L,q}\frac{\mathbf{t}(w)dw}{(1-w^{-1}L)(w-q)}=\frac{\mathbf{t}(L)L-\mathbf{t}(q)q}{L-q}=\frac{\mathbf{t}(L)-\mathbf{t}(q)}{L-q}\,L+\mathbf{t}(q).

It remains only to recall the definition of GG and to note that

(𝐭(q),Φα)+𝐭(q),Φα0,2(ν)=G(𝐭(q),Φα).\displaystyle(\mathbf{t}(q),\Phi_{\alpha})+\langle\mathbf{t}(q),\Phi_{\alpha}\rangle_{0,2}(\nu)=G(\mathbf{t}(q),\Phi_{\alpha}).

Part b) follows from a) and the string equation

1,Φα,𝐭(L)0,3(ν)=𝐭(L),Φα0,2(ν)+𝐭(L)𝐭(1)L1,Φα0,2(ν)+(𝐭(1),Φα).\displaystyle\Big{\langle}1,\Phi_{\alpha},\mathbf{t}(L)\Big{\rangle}_{0,3}(\nu)=\langle\mathbf{t}(L),\Phi_{\alpha}\rangle_{0,2}(\nu)+\Big{\langle}\frac{\mathbf{t}(L)-\mathbf{t}(1)}{L-1},\Phi_{\alpha}\Big{\rangle}_{0,2}(\nu)+(\mathbf{t}(1),\Phi_{\alpha}).\qed

The relation between descendants and ancestors is given by the following formula:

(31) 𝐭(L1),,𝐭(Ln)g,n(ν)=[S𝐭]+(ν,L¯1),,[S𝐭]+(ν,L¯n)g,n(ν),\langle\mathbf{t}(L_{1}),\dots,\mathbf{t}(L_{n})\rangle_{g,n}(\nu)=\langle[S\mathbf{t}]_{+}(\nu,\overline{L}_{1}),\dots,[S\mathbf{t}]_{+}(\nu,\overline{L}_{n})\rangle_{g,n}(\nu),

where gg and nn must satisfy 2g2+n>02g-2+n>0.

Let us choose a formal power series τ(𝐭,ν2,ν3,)\tau(\mathbf{t},\nu_{2},\nu_{3},\dots) with coefficients in K(X)K(X), such that,

(32) [S𝐭]+(τ,ν2,ν3,,1)=τ.[S\mathbf{t}]_{+}(\tau,\nu_{2},\nu_{3},\dots,1)=\tau.

This is a fixed-point problem for τ\tau and we can construct a formal solution via the iterations:

τ(0):=0,τ(n+1):=[S𝐭]+(τ(n),ν2,ν3,,1).\displaystyle\tau^{(0)}:=0,\quad\tau^{(n+1)}:=[S\mathbf{t}]_{+}(\tau^{(n)},\nu_{2},\nu_{3},\dots,1).

The sequence τ(n)\tau^{(n)} of formal series has a limit as nn\to\infty which provides a solution to our fixed-point problem.

Lemma 4.1.2.

The following formula holds:

0,0(ν)+12(ψ2(ν2),1)+𝐭0,1(ν)+12𝐭,𝐭0,2(ν)=12𝐭+1L+ν1,𝐭+1L+ν10,2(ν),\displaystyle\langle\ \rangle_{0,0}(\nu)+\frac{1}{2}(\psi^{2}(\nu_{2}),1)+\langle\mathbf{t}\rangle_{0,1}(\nu)+\frac{1}{2}\langle\mathbf{t},\mathbf{t}\rangle_{0,2}(\nu)=\frac{1}{2}\langle\mathbf{t}+1-L+\nu_{1},\mathbf{t}+1-L+\nu_{1}\rangle_{0,2}(\nu),

where we suppressed the dependence of 𝐭(L)\mathbf{t}(L) on LL, that is, 𝐭=𝐭(L)\mathbf{t}=\mathbf{t}(L).

Proof.

The quadratic terms in 𝐭\mathbf{t} clearly match. The equality of the linear terms is equivalent to

𝐭0,1(ν)=𝐭,1L+ν10,2(ν).\displaystyle\langle\mathbf{t}\rangle_{0,1}(\nu)=\langle\mathbf{t},1-L+\nu_{1}\rangle_{0,2}(\nu).

The above equality is precisely the dilaton equation 𝐭,L10,2(ν)=𝐭0,1(ν)+𝐭,ν10,2(ν)\langle\mathbf{t},L-1\rangle_{0,2}(\nu)=-\langle\mathbf{t}\rangle_{0,1}(\nu)+\langle\mathbf{t},\nu_{1}\rangle_{0,2}(\nu). The equality of the remaining terms is equivalent to

0,0(ν)+12(ψ2(ν2),1)=121L+ν1,1L+ν10,2(ν).\displaystyle\langle\ \rangle_{0,0}(\nu)+\frac{1}{2}(\psi^{2}(\nu_{2}),1)=\frac{1}{2}\langle 1-L+\nu_{1},1-L+\nu_{1}\rangle_{0,2}(\nu).

The RHS can be written as

(33) 12L1,L10,2(ν)L1,ν10,2(ν)+12ν1,ν10,2.\frac{1}{2}\langle L-1,L-1\rangle_{0,2}(\nu)-\langle L-1,\nu_{1}\rangle_{0,2}(\nu)+\frac{1}{2}\langle\nu_{1},\nu_{1}\rangle_{0,2}.

Using the dilaton equation we get

L1,ν10,2(ν)=ν10,1(ν)+ν1,ν10,2(ν)\displaystyle\langle L-1,\nu_{1}\rangle_{0,2}(\nu)=-\langle\nu_{1}\rangle_{0,1}(\nu)+\langle\nu_{1},\nu_{1}\rangle_{0,2}(\nu)

and

L1,L10,2(ν)=L10,1(ν)+L1,ν10,2(ν)=20,0(ν)2ν10,1(ν)+(ψ2(ν2),1)+ν1,ν10,2(ν).\displaystyle\langle L-1,L-1\rangle_{0,2}(\nu)=-\langle L-1\rangle_{0,1}(\nu)+\langle L-1,\nu_{1}\rangle_{0,2}(\nu)=2\langle\ \rangle_{0,0}(\nu)-2\langle\nu_{1}\rangle_{0,1}(\nu)+(\psi^{2}(\nu_{2}),1)+\langle\nu_{1},\nu_{1}\rangle_{0,2}(\nu).

Substituting the above two formulas in (33) we get the identity that we had to prove.∎

Proposition 4.1.1.

Let

(𝐭):=12(ψ2(ν2),1)+n=01n!𝐭(L),,𝐭(L)0,n(0,ν2,ν3,)\displaystyle\mathcal{F}(\mathbf{t}):=\frac{1}{2}(\psi^{2}(\nu_{2}),1)+\sum_{n=0}^{\infty}\frac{1}{n!}\langle\mathbf{t}(L),\dots,\mathbf{t}(L)\rangle_{0,n}(0,\nu_{2},\nu_{3},\dots)

and let τ(𝐭,ν2,ν3,)\tau(\mathbf{t},\nu_{2},\nu_{3},\dots) be a solution to the fixed-point problem (32). Then the following formulas hold:

(𝐭)=12𝐭(L)+1L,𝐭(L)+1L0,2(τ,ν2,ν3,),\displaystyle\mathcal{F}(\mathbf{t})=\frac{1}{2}\langle\mathbf{t}(L)+1-L,\mathbf{t}(L)+1-L\rangle_{0,2}(\tau,\nu_{2},\nu_{3},\dots),
tm,α(𝐭)=12Φα(L1)m,𝐭(L)+1L0,2(τ,ν2,ν3,),\displaystyle\partial_{t_{m,\alpha}}\mathcal{F}(\mathbf{t})=\frac{1}{2}\langle\Phi_{\alpha}(L-1)^{m},\mathbf{t}(L)+1-L\rangle_{0,2}(\tau,\nu_{2},\nu_{3},\dots),

and

tm,αtn,β(𝐭)=12Φα(L1)m,Φβ(L1)n0,2(τ,ν2,ν3,).\displaystyle\partial_{t_{m,\alpha}}\partial_{t_{n,\beta}}\mathcal{F}(\mathbf{t})=\frac{1}{2}\langle\Phi_{\alpha}(L-1)^{m},\Phi_{\beta}(L-1)^{n}\rangle_{0,2}(\tau,\nu_{2},\nu_{3},\dots).
Proof.

Let us prove the first formula only. The argument for the remaining two is similar. Using the translation invariance (19), we can rewrite \mathcal{F} as follows:

(34) (𝐭):=12(ψ2(ν2),1)+n=01n!𝐭ν1,,𝐭ν10,n(ν1,ν2,ν3,)\mathcal{F}(\mathbf{t}):=\frac{1}{2}(\psi^{2}(\nu_{2}),1)+\sum_{n=0}^{\infty}\frac{1}{n!}\langle\mathbf{t}-\nu_{1},\dots,\mathbf{t}-\nu_{1}\rangle_{0,n}(\nu_{1},\nu_{2},\nu_{3},\dots)

where ν1\nu_{1} is an arbitrary formal parameter and we wrote 𝐭\mathbf{t} instead of 𝐭(L)\mathbf{t}(L) in order to make the notation less cumbersome. Let us apply the formula expressing descendants in terms of ancestors (31). We get

𝐭ν1,,𝐭ν10,n(ν1,ν2,ν3,)=[S𝐭]+(ν,L¯1)ν1,,[S𝐭]+(ν,L¯n)ν10,n(ν1,ν2,ν3,),\displaystyle\langle\mathbf{t}-\nu_{1},\dots,\mathbf{t}-\nu_{1}\rangle_{0,n}(\nu_{1},\nu_{2},\nu_{3},\dots)=\langle[S\mathbf{t}]_{+}(\nu,\overline{L}_{1})-\nu_{1},\dots,[S\mathbf{t}]_{+}(\nu,\overline{L}_{n})-\nu_{1}\rangle_{0,n}(\nu_{1},\nu_{2},\nu_{3},\dots),

where n3n\geq 3 and we used Lemma 4.1.1 to compute [Sν1]+=ν1[S\nu_{1}]_{+}=\nu_{1}. Let us specialize ν1=τ(𝐭,ν2,ν3,)\nu_{1}=\tau(\mathbf{t},\nu_{2},\nu_{3},\dots). By definition [S𝐭]+(ν,1)ν1=0[S\mathbf{t}]_{+}(\nu,1)-\nu_{1}=0 which implies that [S𝐭]+(ν,L¯i)ν1[S\mathbf{t}]_{+}(\nu,\overline{L}_{i})-\nu_{1} is proportional to L¯i1=ft(Li1)\overline{L}_{i}-1=\operatorname{ft}^{*}(L_{i}-1) for all 1in1\leq i\leq n. However, the product (L11)(Ln1)=0(L_{1}-1)\cdots(L_{n}-1)=0 on ¯0,n\overline{\mathcal{M}}_{0,n} because ¯0,n\overline{\mathcal{M}}_{0,n} is an (n3)(n-3)-dimensional manifold. We conclude that after the substitution ν1=τ(𝐭,ν2,ν3,)\nu_{1}=\tau(\mathbf{t},\nu_{2},\nu_{3},\dots) all terms in the sum in (34) with n3n\geq 3 must vanish. The sum of the remaining terms, according to Lemma 4.1.2, must be

12𝐭ν1+1L+ν1,𝐭ν1+1L+ν10,2(ν)=12𝐭+1L,𝐭+1L0,2(ν).\displaystyle\frac{1}{2}\langle\mathbf{t}-\nu_{1}+1-L+\nu_{1},\mathbf{t}-\nu_{1}+1-L+\nu_{1}\rangle_{0,2}(\nu)=\frac{1}{2}\langle\mathbf{t}+1-L,\mathbf{t}+1-L\rangle_{0,2}(\nu).\qed

4.2. Proof of Theorem 1.2.1

Let τ(𝐭,ν2,ν3,)\tau(\mathbf{t},\nu_{2},\nu_{3},\dots) be the solution to the fixed-point problem (32). We will prove Theorem 1.2.1 by showing that v(𝐭)=τ(𝐭,ν2,ν3,)v(\mathbf{t})=\tau(\mathbf{t},\nu_{2},\nu_{3},\dots) and that τ(𝐭,ν2,ν3,)\tau(\mathbf{t},\nu_{2},\nu_{3},\dots) is a solution to the integrable hierarchy (7). Let us denote the partial derivative tm,α\partial_{t_{m,\alpha}} by m,α\partial_{m,\alpha}.

Let us prove that v(𝐭)=τ(𝐭,ν2,ν3,)v(\mathbf{t})=\tau(\mathbf{t},\nu_{2},\nu_{3},\dots), that is,

(35) J(τ,0)=1+α=1N0,α0,1(𝐭)Φα,J(\tau,0)=1+\sum_{\alpha=1}^{N}\partial_{{0,\alpha}}\partial_{{0,1}}\mathcal{F}(\mathbf{t})\Phi^{\alpha},

where (𝐭)\mathcal{F}(\mathbf{t}) is the same as in Proposition 4.1.1 and recall that Φ1=1\Phi_{1}=1. According to Proposition 4.1.1, the second order partial derivative in the above formula is

1,Φα0,2(τ,ν2,ν3,)=Φα0,1(τ,ν2,ν3,)+(Φα,τ),\displaystyle\langle 1,\Phi_{\alpha}\rangle_{0,2}(\tau,\nu_{2},\nu_{3},\dots)=\langle\Phi_{\alpha}\rangle_{0,1}(\tau,\nu_{2},\nu_{3},\dots)+(\Phi_{\alpha},\tau),

where we used the string equation. The formula that we want to prove follows.

Let us prove that τ\tau is a solution to (7). Differentiating (35) with respect to tn,βt_{n,\beta} we get

(36) n,βJ(τ,0)=0,1α=1NΦα,Φβ(L1)n0,2(τ)Φα,\partial_{n,\beta}J(\tau,0)=\partial_{0,1}\sum_{\alpha=1}^{N}\langle\Phi_{\alpha},\Phi_{\beta}(L-1)^{n}\rangle_{0,2}(\tau)\,\Phi^{\alpha},

where we suppressed the dependence of the two-point correlator on ν2,ν3,\nu_{2},\nu_{3},\dots and we used Proposition 4.1.1. Using the projection map (30) we transform the two-point correlator 0,1Φα,Φβ(L1)n0,2(τ)\partial_{0,1}\langle\Phi_{\alpha},\Phi_{\beta}(L-1)^{n}\rangle_{0,2}(\tau) into

0,1Resq=0,dq(q1)nΦα,ΦβqL(τ)=0,1Resq=0,dqq1(q1)nG(S(τ,q)Φβ,Φα).\displaystyle-\partial_{0,1}\operatorname{Res}_{q=0,\infty}dq(q-1)^{n}\Big{\langle}\Phi_{\alpha},\frac{\Phi_{\beta}}{q-L}\Big{\rangle}(\tau)=-\partial_{0,1}\operatorname{Res}_{q=0,\infty}dqq^{-1}(q-1)^{n}G(S(\tau,q)\Phi_{\beta},\Phi_{\alpha}).

Note that G(a,S(τ,0)b)=(a,b)G(a,S(\tau,0)b)=(a,b), or equivalently G(a,b)=(a,S(τ,0)1b)G(a,b)=(a,S(\tau,0)^{-1}b). Let us change G(S(τ,q)Φβ,Φα)G(S(\tau,q)\Phi_{\beta},\Phi_{\alpha}) into (S(τ,0)1S(τ,q)Φβ,Φα)(S(\tau,0)^{-1}S(\tau,q)\Phi_{\beta},\Phi_{\alpha}) and substitute the result in formula (36). We get

(37) n,βJ(τ,0)=0,1Resq=0,dqq1(q1)nS(τ,0)1S(τ,q)Φβ.\partial_{n,\beta}J(\tau,0)=-\partial_{0,1}\operatorname{Res}_{q=0,\infty}dqq^{-1}(q-1)^{n}S(\tau,0)^{-1}S(\tau,q)\Phi_{\beta}.

Using that S(ν,q)S(\nu,q) is a solution to the quantum differential equations and that J(ν,q)=(1q)S(ν,q)11J(\nu,q)=(1-q)S(\nu,q)^{-1}1 we get

n,βJ(τ,0)=S(τ,0)1(n,βτS(τ,0))S(τ,0)11=S(τ,0)1n,βτ.\displaystyle\partial_{n,\beta}J(\tau,0)=-S(\tau,0)^{-1}(-\partial_{n,\beta}\tau\bullet S(\tau,0))S(\tau,0)^{-1}1=S(\tau,0)^{-1}\partial_{n,\beta}\tau.

Similarly, we find that 0,1S(τ,0)1S(τ,q)\partial_{0,1}S(\tau,0)^{-1}S(\tau,q) is

S(τ,0)10,1τS(τ,q)+S(τ,0)1(q1)10,1τS(τ,q)=qq1S(τ,0)10,1τS(τ,q).\displaystyle S(\tau,0)^{-1}\partial_{0,1}\tau\bullet S(\tau,q)+S(\tau,0)^{-1}(q-1)^{-1}\partial_{0,1}\tau\bullet S(\tau,q)=\frac{q}{q-1}S(\tau,0)^{-1}\partial_{0,1}\tau\bullet S(\tau,q).

Substituting these formulas in (37) and cancelling S(τ,0)1S(\tau,0)^{-1} we get

n,βτ=Resq=dq(q1)n10,1τS(τ,q)Φβ,\displaystyle\partial_{n,\beta}\tau=-\operatorname{Res}_{q=\infty}dq(q-1)^{n-1}\partial_{0,1}\tau\bullet S(\tau,q)\Phi_{\beta},

where the residue at q=0q=0 was dropped because S(τ,q)S(\tau,q) does not have a pole at q=0q=0. Comparing with (7), we get that τ(𝐭,ν2,ν3,)\tau(\mathbf{t},\nu_{2},\nu_{3},\dots) 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 Vn={an|a1++an=0}V_{n}=\{a\in\mathbb{C}^{n}\ |\ a_{1}+\cdots+a_{n}=0\} is naturally a SnS_{n}-space where SnS_{n} acts by permuting the coordinates σ(a1,,an):=(aσ1(1),,aσ1(n))\sigma\cdot(a_{1},\dots,a_{n}):=(a_{\sigma^{-1}(1)},\dots,a_{\sigma^{-1}(n)}). We have an isomorphism H0(¯0,n+1,Ln+1)VnH^{0}(\overline{\mathcal{M}}_{0,n+1},L_{n+1})\cong V_{n} defined as follows. Given aVna\in V_{n} and (C,p1,,pn+1)¯0,n+1(C,p_{1},\dots,p_{n+1})\in\overline{\mathcal{M}}_{0,n+1} we construct a cotangent vector φa(pn+1)Tpn+1C\varphi_{a}(p_{n+1})\in T^{*}_{p_{n+1}}C where the meromorphic 1-form φa\varphi_{a} on CC is uniquely determined by the following conditions:

  1. (i)

    φa\varphi_{a} is holomorphic except for at most first order poles at the nodes and the first nn marked points of CC.

  2. (ii)

    If qCq\in C is a node and CC^{\prime} and C′′C^{\prime\prime} are the two irreducible components of CC at qq, then Resp=q(φa(p))+Resp′′=q(φa(p′′))=0\operatorname{Res}_{p^{\prime}=q}(\varphi_{a}(p^{\prime}))+\operatorname{Res}_{p^{\prime\prime}=q}(\varphi_{a}(p^{\prime\prime}))=0 where pp^{\prime} and p′′p^{\prime\prime} are local coordinates on respectively CC^{\prime} and C′′C^{\prime\prime} near qq.

  3. (iii)

    Resp=piφa(p)=ai\operatorname{Res}_{p=p_{i}}\varphi_{a}(p)=a_{i}.

Holomorphic forms satisfying conditions (i) and (ii) are by definitions sections of the sheaf ωC(p1++pn)\omega_{C}(p_{1}+\cdots+p_{n}) where ωC\omega_{C} is the dualizing sheaf of CC. Condition (iii), uniquely determines the meromorphic 1-form φa\varphi_{a} because CC is rational and hence the only obstruction to the existence of φa\varphi_{a} is that the sum of the residues at all poles is 0, that is, aVna\in V_{n}. This description also shows that Ln+1ωC(p1++pn)L_{n+1}\cong\omega_{C}(p_{1}+\cdots+p_{n}) 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 Ln+1L_{n+1} determines a holomorphic map π:¯0,n+1(Vn)\pi:\overline{\mathcal{M}}_{0,n+1}\to\mathbb{P}(V_{n}^{*}), such that, π𝒪(1)=Ln+1\pi^{*}\mathcal{O}(1)=L_{n+1}. More precisely, we have

(38) π(C,p1,,pn+1)(a):=φa(pn+1)/dw(pn+1),\pi(C,p_{1},\dots,p_{n+1})(a):=\varphi_{a}(p_{n+1})/dw(p_{n+1}),

where ww is a local coordinate on CC at pn+1p_{n+1}. The class of π(C,p1,,pn+1)\pi(C,p_{1},\dots,p_{n+1}) in (Vn)\mathbb{P}(V_{n}^{*}) is independent of the choice of local coordinate ww.

Lemma 5.1.1.

Suppose that the action SnS_{n} on (Vn)\mathbb{P}(V_{n}^{*}) is induced from (σξ)(a):=ξ(σ1a)(\sigma\xi)(a):=\xi(\sigma^{-1}a) where ξVn\xi\in V_{n}^{*} and aVna\in V_{n}. The map (38) is SnS_{n}-equivariant, that is, π(σ(C,p1,,pn))=σπ(C,p1,,pn),\pi(\sigma(C,p_{1},\dots,p_{n}))=\sigma\pi(C,p_{1},\dots,p_{n}), where the action of SnS_{n} on ¯0,n+1\overline{\mathcal{M}}_{0,n+1} is defined by σ(C,p1,,pn+1):=(C,pσ1(1),,pσ1(n),pn+1)\sigma(C,p_{1},\dots,p_{n+1}):=(C,p_{\sigma^{-1}(1)},\dots,p_{\sigma^{-1}(n)},p_{n+1}).

Proof.

Let ψa\psi_{a} be the meromorphic form on CC corresponding to (C,pσ1(1),,pσ1(n),pn+1)(C,p_{\sigma^{-1}(1)},\dots,p_{\sigma^{-1}(n)},p_{n+1}). We have Resp=pσ1(i)φa(p)=ai\operatorname{Res}_{p=p_{\sigma^{-1}(i)}}\varphi_{a}(p)=a_{i} or equivalently

Resp=pjψa(p)=aσ(j)=(σ1a)j=Resp=pjφσ1a(p).\displaystyle\operatorname{Res}_{p=p_{j}}\psi_{a}(p)=a_{\sigma(j)}=(\sigma^{-1}a)_{j}=\operatorname{Res}_{p=p_{j}}\varphi_{\sigma^{-1}a}(p).

Therefore, ψa=φσ1a\psi_{a}=\varphi_{\sigma^{-1}a} and the equivariance of π\pi is clear. ∎

The map π\pi can be described more explicitly as follows. Suppose that we have a point (C,p1,,pn+1)¯0,n+1(C,p_{1},\dots,p_{n+1})\in\overline{\mathcal{M}}_{0,n+1}. Let C01C_{0}\cong\mathbb{P}^{1}, we call it the central component, be the irreducible component of CC that carries the marked point pn+1p_{n+1}. Let us denote by q1,qlq_{1},\dots q_{l} the nodal points of CC that are on C0C_{0}. Put I:={i{1,2,,n}|piC0}I:=\{i\in\{1,2,\dots,n\}\ |\ p_{i}\in C_{0}\}. Note that by removing C0C_{0} from CC we get a curve consisting of ll connected components C1,,ClC_{1},\dots,C_{l}. Therefore, the remaining marked points {1,2,,n}I\{1,2,\dots,n\}\setminus I split into ll pairwise disjoint groups J1,J2,,JlJ_{1},J_{2},\dots,J_{l}, that is, Jk={j|pjCk}J_{k}=\{j\ |\ p_{j}\in C_{k}\}. Since the isomorphism C01C_{0}\cong\mathbb{P}^{1} is determined up to the action of PSL(2,)\operatorname{PSL}(2,\mathbb{C}), we can arrange that pn+1=p_{n+1}=\infty and |J1|q1++|Jl|ql+iIpi=0|J_{1}|q_{1}+\dots+|J_{l}|q_{l}+\sum_{i\in I}p_{i}=0 where |Jk||J_{k}| is the number of elements in JkJ_{k}. Note that

φa|C0=k=1ljJkajdzzqk+iIaidzzpi.\displaystyle\varphi_{a}|_{C_{0}}=\sum_{k=1}^{l}\sum_{j\in J_{k}}\frac{a_{j}dz}{z-q_{k}}+\sum_{i\in I}\frac{a_{i}dz}{z-p_{i}}.

Let w=1/zw=1/z be the coordinate near =pn+1\infty=p_{n+1}. We have

φa(w)dw=k=1ljJkajqk1wqkiIaipi1wpi.\displaystyle\frac{\varphi_{a}(w)}{dw}=-\sum_{k=1}^{l}\sum_{j\in J_{k}}\frac{a_{j}q_{k}}{1-wq_{k}}-\sum_{i\in I}\frac{a_{i}p_{i}}{1-wp_{i}}.

Restricting w=0w=0 we get that

π(C,p1,,pn+1)(a)=k=1ljJkajqkiIaipi.\displaystyle\pi(C,p_{1},\dots,p_{n+1})(a)=-\sum_{k=1}^{l}\sum_{j\in J_{k}}a_{j}q_{k}-\sum_{i\in I}a_{i}p_{i}.

On the other hand, the projective space (Vn){ξn1|ξ1++ξn=0}\mathbb{P}(V_{n}^{*})\cong\{\xi\in\mathbb{P}^{n-1}\ |\ \xi_{1}+\cdots+\xi_{n}=0\} where (ξ1:ξ2::ξn)(\xi_{1}:\xi_{2}:\cdots:\xi_{n}) are the homogeneous coordinates on n1\mathbb{P}^{n-1} and the isomorphism is given by mapping (ξ1:ξ2::ξn)(\xi_{1}:\xi_{2}:\cdots:\xi_{n}) to the linear functional aξ1a1++ξnana\mapsto\xi_{1}a_{1}+\cdots+\xi_{n}a_{n}. Under this identification, the action of SnS_{n} on n1\mathbb{P}^{n-1} is given by σ(ξ1:ξ2::ξn)=(ξσ1(1):ξσ1(2)::ξσ1(n))\sigma(\xi_{1}:\xi_{2}:\cdots:\xi_{n})=(\xi_{\sigma^{-1}(1)}:\xi_{\sigma^{-1}(2)}:\cdots:\xi_{\sigma^{-1}(n)}) and

(39) π(C,p1,,pn+1)=(pi(iI):qk::qk|Jk|(1kl)),\pi(C,p_{1},\dots,p_{n+1})=(p_{i}(i\in I):\underbrace{q_{k}:\cdots:q_{k}}_{|J_{k}|}(1\leq k\leq l)),

where the homogeneous coordinates on the RHS of the above formula should be ordered as follows: pip_{i} (iIi\in I) is placed on the ii-th position and a copy of qkq_{k} is placed on the jj-th position for every jJkj\in J_{k}. Using formula (39), let us describe the fibers of π\pi. Given ξ=(ξ1::ξn)\xi=(\xi_{1}:\cdots:\xi_{n}) we define an equivalence relation in {1,2,,n}\{1,2,\dots,n\} by saying that iji\sim j iff ξi=ξj\xi_{i}=\xi_{j}. This equivalence relation splits {1,2,,n}\{1,2,\dots,n\} into several equivalence classes. The ones consisting of a single element we denote by {i}\{i\} (iIi\in I). The remaining ones, consisting of at least two elements, we denote be J1,,JlJ_{1},\dots,J_{l}. If (C,p1,,pn+1)(C,p_{1},\dots,p_{n+1}) is in the fiber π1(ξ)\pi^{-1}(\xi) then the central component C0C_{0} must be 1\mathbb{P}^{1} with marked points pip_{i} (iIi\in I) and nodal points qk=ξjq_{k}=\xi_{j} (1kl1\leq k\leq l) where for each kk we choose one jJkj\in J_{k}. In other words, the isomorphism class of the central component C0C_{0} is uniquely fixed. The remaining ll components can be fixed in ¯0,|J1|+1×ׯ0,|Jl|+1\overline{\mathcal{M}}_{0,|J_{1}|+1}\times\cdots\times\overline{\mathcal{M}}_{0,|J_{l}|+1} ways, that is, the fiber

(40) π1(ξ)¯0,|J1|+1×ׯ0,|Jl|+1.\pi^{-1}(\xi)\cong\overline{\mathcal{M}}_{0,|J_{1}|+1}\times\cdots\times\overline{\mathcal{M}}_{0,|J_{l}|+1}.

The map π\pi is a birational equivalence. More precisely, let Σn2\Sigma\subset\mathbb{P}^{n-2} be the analytic subvariety consisting of points ξ=(ξ1::ξn)\xi=(\xi_{1}:\cdots:\xi_{n}), such that, there exist 3 pairwise different i,ji,j, and kk, such that, ξi=ξj=ξk\xi_{i}=\xi_{j}=\xi_{k}. This is clearly a complex co-dimension 2 algebraic subvariety and π\pi induces an isomorphism π1(n2Σ)n2Σ\pi^{-1}(\mathbb{P}^{n-2}\setminus{\Sigma})\cong\mathbb{P}^{n-2}\setminus{\Sigma}. Since n2\mathbb{P}^{n-2} is non-singular, the map π\pi 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 π\pi vanish, that is, Riπ𝒪=0R^{i}\pi_{*}\mathcal{O}=0 for i>0i>0.

5.2. J-function

The JJ-function of a point was computed by Givental (see the Theorem in [5]). We have the following formula:

J(ν,q)=1q+ν1+11qL0,1(ν)=(1q)exp(k=1νkk(1qk)).\displaystyle J(\nu,q)=1-q+\nu_{1}+\Big{\langle}\frac{1}{1-qL}\Big{\rangle}_{0,1}(\nu)=(1-q)\exp\Big{(}\sum_{k=1}^{\infty}\frac{\nu_{k}}{k(1-q^{k})}\Big{)}.

Let us outline a proof of the above formula which is slightly different than the one given in [5]. Let ev:¯0,n+1pt\operatorname{ev}:\overline{\mathcal{M}}_{0,n+1}\to\operatorname{pt} be the contraction map. Using Lemma 5.1.2 we compute πLn+1k=ππ𝒪(k)=𝒪(k)π𝒪=𝒪(k)\pi_{*}L_{n+1}^{k}=\pi_{*}\pi^{*}\mathcal{O}(k)=\mathcal{O}(k)\otimes\pi_{*}\mathcal{O}=\mathcal{O}(k). Therefore,

ev(Ln+1k)=H0((Vn),𝒪(k))=Symk(Vn),Sym0(Vn):=.\displaystyle\operatorname{ev}_{*}(L_{n+1}^{k})=H^{0}(\mathbb{P}(V_{n}^{*}),\mathcal{O}(k))=\operatorname{Sym}^{k}(V_{n}),\quad\operatorname{Sym}^{0}(V_{n}):=\mathbb{C}.

The JJ-function takes the form

J(ν,q)=1q+ν1+n=2k=0Symk(Vn)qk.\displaystyle J(\nu,q)=1-q+\nu_{1}+\sum_{n=2}^{\infty}\sum_{k=0}^{\infty}\operatorname{Sym}^{k}(V_{n})q^{k}.

On the other hand, we have n=Vn\mathbb{C}^{n}=V_{n}\oplus\mathbb{C} as SnS_{n}-modules where \mathbb{C} is the trivial SnS_{n}-module. Note that Symk(n)=l=0kSyml(Vn)\operatorname{Sym}^{k}(\mathbb{C}^{n})=\sum_{l=0}^{k}\operatorname{Sym}^{l}(V_{n}) for n1n\geq 1 and k0k\geq 0 where we assume that V1=0V_{1}=0. Therefore,

n=1k=0Symk(n)qk=k=0ν1qk+n=2k=0l=0kSyml(Vn)qk=ν11q+n=2l=0Syml(Vn)qll=kqkl,\displaystyle\sum_{n=1}^{\infty}\sum_{k=0}^{\infty}\operatorname{Sym}^{k}(\mathbb{C}^{n})q^{k}=\sum_{k=0}^{\infty}\nu_{1}q^{k}+\sum_{n=2}^{\infty}\sum_{k=0}^{\infty}\sum_{l=0}^{k}\operatorname{Sym}^{l}(V_{n})q^{k}=\frac{\nu_{1}}{1-q}+\sum_{n=2}^{\infty}\sum_{l=0}^{\infty}\operatorname{Sym}^{l}(V_{n})q^{l}\sum_{l=k}^{\infty}q^{k-l},

where in the last equality we exchanged the summations over kk and ll. Using the above formula we get that the JJ-function is

J(ν,q)=(1q)(1+n=1k=0Symk(n)qk).\displaystyle J(\nu,q)=(1-q)\Big{(}1+\sum_{n=1}^{\infty}\sum_{k=0}^{\infty}\operatorname{Sym}^{k}(\mathbb{C}^{n})q^{k}\Big{)}.

It remains only to prove that

1+n=1k=0Symk(n)qk=exp(r=1νrr(1qr)).\displaystyle 1+\sum_{n=1}^{\infty}\sum_{k=0}^{\infty}\operatorname{Sym}^{k}(\mathbb{C}^{n})q^{k}=\exp\Big{(}\sum_{r=1}^{\infty}\frac{\nu_{r}}{r(1-q^{r})}\Big{)}.

Let us denote the LHS of the above equality by f(ν,q)f(\nu,q). Let us compute the derivatives rνrf(ν,q)r\partial_{\nu_{r}}f(\nu,q) by using the Heisenberg algebra structure of the Fock space. Recalling Theorem 2.3.1, we get

rνrSymk(n)qk=tr(1,2,,r)ResSr×SnrSnSymk(n)qk=a+b=ktr(1,2,,r)Syma(r)qaSymb(nr)qb,\displaystyle r\partial_{\nu_{r}}\operatorname{Sym}^{k}(\mathbb{C}^{n})q^{k}=\operatorname{tr}_{(1,2,\dots,r)}\operatorname{Res}^{S_{n}}_{S_{r}\times S_{n-r}}\,\operatorname{Sym}^{k}(\mathbb{C}^{n})q^{k}=\bigoplus_{a+b=k}\operatorname{tr}_{(1,2,\dots,r)}\operatorname{Sym}^{a}(\mathbb{C}^{r})q^{a}\otimes\operatorname{Sym}^{b}(\mathbb{C}^{n-r})q^{b},

where we take nrn\geq r, otherwise the derivative is 0 for degree reasons. In order to compute this trace, let us identify Syma(r)\operatorname{Sym}^{a}(\mathbb{C}^{r}) with the space of homogeneous polynomials in [x1,,xr]\mathbb{C}[x_{1},\dots,x_{r}] of degree aa. This space has a basis consisting of monomials x1d1x2d2xrdrx_{1}^{d_{1}}x_{2}^{d_{2}}\cdots x_{r}^{d_{r}}, such that, d1+d2++dr=ad_{1}+d_{2}+\cdots+d_{r}=a. Since the permutation (1,2,,r)(1,2,\dots,r) permutes these monomials, the trace is equal to the number of monomials that remain fixed. However, there is at most one monomial fixed by (1,2,,r)(1,2,\dots,r), that is, the monomial with d1==drd_{1}=\cdots=d_{r}. Such monomial exists only if rr divides aa. In other words, the trace in the above formula is 11 if r|ar|a and 0 otherwise. Summing over all kk, we get

rνrk=0Symk(n)qk=11qrb=0Symb(nr)qb.\displaystyle r\partial_{\nu_{r}}\sum_{k=0}^{\infty}\operatorname{Sym}^{k}(\mathbb{C}^{n})q^{k}=\frac{1}{1-q^{r}}\sum_{b=0}^{\infty}\operatorname{Sym}^{b}(\mathbb{C}^{n-r})q^{b}.

Summing over all n1n\geq 1, we get rνrf(ν,q)=11qrf(ν,q)r\partial_{\nu_{r}}f(\nu,q)=\frac{1}{1-q^{r}}f(\nu,q). Solving this differential equation with the initial condition f(0,q)=1f(0,q)=1 yields the formula that we had to prove.

5.3. Proof of Theorem 1.3.1

Since J(ν,q)=(1q)S(ν,q)11,J(\nu,q)=(1-q)S(\nu,q)^{-1}1, we get

S(ν,q)=exp(ν1q1+k=2νkk(qk1))\displaystyle S(\nu,q)=\exp\Big{(}\frac{\nu_{1}}{q-1}+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k(q^{k}-1)}\Big{)}

and

G(ν)=1+1,10,2(ν)=ν1J(ν,0)=exp(ν1+k=2νkk).\displaystyle G(\nu)=1+\langle 1,1\rangle_{0,2}(\nu)=\partial_{\nu_{1}}J(\nu,0)=\exp\Big{(}\nu_{1}+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k}\Big{)}.

Put

w(𝐭):=n=01n!1,1,𝐭,,𝐭0,n+2(0,ν2,ν3,)\displaystyle w(\mathbf{t}):=\sum_{n=0}^{\infty}\frac{1}{n!}\langle 1,1,\mathbf{t},\dots,\mathbf{t}\rangle_{0,n+2}(0,\nu_{2},\nu_{3},\dots)

and let v(𝐭)v(\mathbf{t}) be defined as the solution to the equation

w(𝐭)=J(v(𝐭),0)1=exp(v(𝐭)+k=2νkk)1.\displaystyle w(\mathbf{t})=J(v(\mathbf{t}),0)-1=\exp\Big{(}v(\mathbf{t})+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k}\Big{)}-1.

Recalling Theorem 1.2.1, we get that v(𝐭)v(\mathbf{t}) is a solution to the following system of differential equations:

nv=Resq=dq(q1)nexp(vq1+k=2νkk(qk1)),\displaystyle\partial_{n}v=-\operatorname{Res}_{q=\infty}dq(q-1)^{n}\partial\,\exp\Big{(}\frac{v}{q-1}+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k(q^{k}-1)}\Big{)},

where n:=tn\partial_{n}:=\tfrac{\partial}{\partial t_{n}} and =0\partial=\partial_{0}. On the other hand, since 11qL=11qk=0qk(1q)k(L1)k\frac{1}{1-qL}=\frac{1}{1-q}\,\sum_{k=0}^{\infty}\frac{q^{k}}{(1-q)^{k}}\,(L-1)^{k}, we get that the insertion of 11qL\frac{1}{1-qL} is obtained from w(𝐭)w(\mathbf{t}) via the action of the differential operator

P(q)=11qk=0qk(1q)kk=11qk=01(q11)kk,\displaystyle P(q)=\frac{1}{1-q}\,\sum_{k=0}^{\infty}\frac{q^{k}}{(1-q)^{k}}\,\partial_{k}=\frac{1}{1-q}\,\sum_{k=0}^{\infty}\frac{1}{(q^{-1}-1)^{k}}\,\partial_{k},

that is, the correlator on the LHS in Theorem 1.3.1 is P(q1)P(qn)w(𝐭)|t0=ν1,t1=t2==0P(q_{1})\cdots P(q_{n})w(\mathbf{t})|_{t_{0}=\nu_{1},t_{1}=t_{2}=\cdots=0}.

Let us examine the action of P(q)P(q) on

wa(𝐭)=exp(a1v(𝐭)+k=2akνkk)\displaystyle w_{a}(\mathbf{t})=\exp\Big{(}a_{1}v(\mathbf{t})+\sum_{k=2}^{\infty}a_{k}\frac{\nu_{k}}{k}\Big{)}

where a=(a1,a2,a3,)a=(a_{1},a_{2},a_{3},\dots) is an arbitrary sequence of complex numbers. We have

(1q)P(q)wa(𝐭)=n=0a1(q11)nnvexp(a1v(𝐭)+k=2akνk/k).\displaystyle(1-q)P(q)w_{a}(\mathbf{t})=\sum_{n=0}^{\infty}\frac{a_{1}}{(q^{-1}-1)^{n}}\,\partial_{n}v\,\exp\Big{(}a_{1}v(\mathbf{t})+\sum_{k=2}^{\infty}a_{k}\nu_{k}/k\Big{)}.

Recalling the differential equation for vv, we get

(1q)P(q)wa(𝐭)\displaystyle(1-q)P(q)w_{a}(\mathbf{t}) =\displaystyle= a1exp(a1v(𝐭)+k=2akνk/k)×\displaystyle-a_{1}\exp\Big{(}a_{1}v(\mathbf{t})+\sum_{k=2}^{\infty}a_{k}\nu_{k}/k\Big{)}\times
×n=0Resu=du(u1)n(q11)nexp(vu1+k=2νkk(uk1)).\displaystyle\times\sum_{n=0}^{\infty}\operatorname{Res}_{u=\infty}du\frac{(u-1)^{n}}{(q^{-1}-1)^{n}}\,\partial\,\exp\Big{(}\frac{v}{u-1}+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k(u^{k}-1)}\Big{)}.

The residues in the above formula can be interpreted analytically as follows. We expand the exponential as a formal power series in 𝐭,ν2,ν3,\mathbf{t},\nu_{2},\nu_{3},\dots, then each coefficient is a rational function in uu with poles on the unit circle. Therefore, we may and we will think of each residue as 12π𝐢|u1|=R\tfrac{1}{2\pi\mathbf{i}}\oint_{|u-1|=R} where R1R\gg 1 is a sufficiently big real number and the orientation of the contour is clockwise around the center 11. Let us choose qq so small that |u1|<|q11||u-1|<|q^{-1}-1|. The sum over nn of the geometric series is uniformly convergent in uu to 1q1qu\tfrac{1-q}{1-qu} and we get

P(q)wa(𝐭)=a1exp(a1v(𝐭)+k=2akνk/k)2π𝐢|u1|=Rexp(vu1+k=2νkk(uk1))du1qu.\displaystyle P(q)w_{a}(\mathbf{t})=-a_{1}\exp\Big{(}a_{1}v(\mathbf{t})+\sum_{k=2}^{\infty}a_{k}\nu_{k}/k\Big{)}\,\frac{\partial}{2\pi\mathbf{i}}\,\oint_{|u-1|=R}\exp\Big{(}\frac{v}{u-1}+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k(u^{k}-1)}\Big{)}\frac{du}{1-qu}.

Recalling the Cauchy residue theorem we get that the above integral is the sum of the residues at u=q1u=q^{-1} and u=u=\infty. The residue at u=u=\infty is a constant which does not contribute at the end because we have to apply to it the derivation \partial. We get

P(q)wa(𝐭)=a1q1exp(a1v(𝐭)+k=2akνk/k)exp(vq11+k=2νkk(qk1)).\displaystyle P(q)w_{a}(\mathbf{t})=a_{1}q^{-1}\exp\Big{(}a_{1}v(\mathbf{t})+\sum_{k=2}^{\infty}a_{k}\nu_{k}/k\Big{)}\,\partial\,\exp\Big{(}\frac{v}{q^{-1}-1}+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k(q^{-k}-1)}\Big{)}.

Note that the above formula is equivalent to

P(q)wa(𝐭)=11qa1a1+1q11exp((a1+1q11)v+k=2(ak+1qk1)νkk).\displaystyle P(q)w_{a}(\mathbf{t})=\frac{1}{1-q}\,\frac{a_{1}}{a_{1}+\frac{1}{q^{-1}-1}}\,\partial\,\exp\left(\Big{(}a_{1}+\frac{1}{q^{-1}-1}\Big{)}v+\sum_{k=2}^{\infty}\Big{(}a_{k}+\frac{1}{q^{-k}-1}\Big{)}\frac{\nu_{k}}{k}\right).

Using the above formula we get

P(q1)P(qn)w(𝐭)=1(1q1)(1qn)11+1q111++1qn11×\displaystyle P(q_{1})\cdots P(q_{n})w(\mathbf{t})=\frac{1}{(1-q_{1})\cdots(1-q_{n})}\,\frac{1}{1+\frac{1}{q_{1}^{-1}-1}+\cdots+\frac{1}{q_{n}^{-1}-1}}\times
×nexp((1+1q111++1qn11)v(𝐭)+k=2(1+1q1k1++1qnk1)νkk).\displaystyle\times\partial^{n}\,\exp\left(\Big{(}1+\frac{1}{q_{1}^{-1}-1}+\cdots+\frac{1}{q_{n}^{-1}-1}\Big{)}v(\mathbf{t})+\sum_{k=2}^{\infty}\Big{(}1+\frac{1}{q_{1}^{-k}-1}+\cdots+\frac{1}{q_{n}^{-k}-1}\Big{)}\frac{\nu_{k}}{k}\right).

Note that if we set t1=t2==0t_{1}=t_{2}=\cdots=0, then

w(t0,0,0,)=G(t0,ν2,ν3,)1=et0+k=2νkk1.\displaystyle w(t_{0},0,0,\dots)=G(t_{0},\nu_{2},\nu_{3},\dots)-1=e^{t_{0}+\sum_{k=2}^{\infty}\frac{\nu_{k}}{k}}-1.

Therefore, v(t0,0,0,)=t0v(t_{0},0,0,\dots)=t_{0} 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 22D 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 VV-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 ¯1,n\overline{\mathcal{M}}_{1,n} . 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 KK-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 KK-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.