1 Introduction

The Toda lattice hierarchy, which contains the Toda lattice equation

$$\begin{aligned} \ddot{\sigma }(n) \; = \; e^{\sigma (n-1)-\sigma (n)} \, - \, e^{\sigma (n)-\sigma (n+1)}, \end{aligned}$$
(1)

is an important integrable hierarchy of nonlinear differential–difference equations [18, 19, 22, 27]. In this paper, following the idea of [13], we derive new formulae for generating series of k-point correlation functions for the Toda lattice hierarchy by using the matrix resolvent approach [10] and by introducing a pair of wave functions.

1.1 Toda lattice hierarchy and tau-function

Let

$$\begin{aligned} {{\mathcal {A}}}\, := \, {{\mathbb {Z}}}\, [ v_0,w_0,v_{\pm 1}, w_{\pm 1}, v_{\pm 2}, w_{\pm 2}, \ldots ] \end{aligned}$$
(2)

be the polynomial ring. Define the shift operator \(\Lambda :{{\mathcal {A}}}\rightarrow {{\mathcal {A}}}\) via

$$\begin{aligned} \Lambda (1)\; = \;1, \quad \Lambda (v_i) \; = \; v_{i+1}, \quad \Lambda (w_i) \; = \; w_{i+1}, \quad \Lambda (f g) \; = \; \Lambda (f) \, \Lambda (g)\end{aligned}$$

\(\forall \, i\in {{\mathbb {Z}}}\) and \(f,g\in {{\mathcal {A}}}\). Denote by \(\Lambda ^{-1}\) the inverse of \(\Lambda \) satisfying \(\Lambda ^{-1} (v_i) = v_{i-1}\), \(\Lambda ^{-1} (w_i) = w_{i-1}\), and \(\Lambda ^{-1} (f g) = \Lambda ^{-1} (f) \, \Lambda ^{-1} (g)\). For a difference operator P on \({{\mathcal {A}}}\), we mean an operator of the form \(P = \sum _{m\in {{\mathbb {Z}}}} P_m \, \Lambda ^m \), where \(P_m \in {{\mathcal {A}}}\). Denote \(P_+:=\sum _{m\ge 0} P_m \, \Lambda ^m\), \(P_-:=\sum _{m< 0} P_m \, \Lambda ^m\), \({\mathrm{Coef}}(P,m):=P_m\). A linear operator \(D:{{\mathcal {A}}}\rightarrow {{\mathcal {A}}}\) is called a derivation on \({{\mathcal {A}}}\), if

$$\begin{aligned} D(fg) \; = \; D(f) \, g \; + \; f \, D(g), \quad \forall \,f,g\in {{\mathcal {A}}}. \end{aligned}$$

The derivation D is called admissible if it commutes with \(\Lambda \). Clearly, every admissible derivation D is uniquely determined by the values \(D(v_0)\) and \(D(w_0)\). Let

$$\begin{aligned} L\, := \, \Lambda \; + \; v_0 \; + \; w_0 \, \Lambda ^{-1} \end{aligned}$$
(3)

be a difference operator, and define a sequence of difference operators \(A_k\)\(k\ge 0\) by

$$\begin{aligned} A_k \, := \, \bigl (L^{k+1}\bigr )_+. \end{aligned}$$
(4)

We associate with \(A_k\) a sequence of admissible derivations \(D_k:{{\mathcal {A}}}\rightarrow {{\mathcal {A}}}\) defined via

$$\begin{aligned} D_k(v_0) \, := \, {\mathrm{Coef}} \bigl ([A_k,L], 0\bigr ), \quad D_k(w_0) \, := \, {\mathrm{Coef}} \bigl ([A_k,L], -1\bigr ), \qquad k\ge 0. \end{aligned}$$
(5)

The first few \(D_k(v_0)\) and \(D_k(w_0)\) are \(D_0(v_0)=w_1-w_0\), \(D_0(w_0)=w_0 \, (v_0-v_{-1})\); \(D_1(v_0)=w_1(v_1+v_0)-w_0(v_0+v_{-1})\), \(D_1(w_0)=w_0\bigl (w_1-w_{-1}+v_0^2-v_{-1}^2\bigr )\), etc.

Lemma 1

The operators \(D_k\), \(k\ge 0\) pairwise commute.

This lemma was known. We call \(D_k\) the Toda lattice derivations, and (5) the abstract Toda lattice hierarchy.

A tau-structure associated with the derivations \((D_k)_{k\ge 0}\) is a collection of polynomials \(\bigl (\Omega _{p,q}, S_p\bigr )_{p,q\ge 0}\) in \({\mathcal {A}}\) satisfying

$$\begin{aligned}&\Omega _{p,q} \; = \; \Omega _{q,p}, \quad D_r \bigl (\Omega _{p,q}\bigr ) \; = \; D_q \bigl (\Omega _{p,r}\bigr ), \end{aligned}$$
(6)
$$\begin{aligned}&(\Lambda -1 ) \, \bigl (\Omega _{p,q}\bigr ) \; = \; D_q \bigl (S_p\bigr ) , \end{aligned}$$
(7)
$$\begin{aligned}&w_0 \, \bigl (1 \,-\, \Lambda ^{-1} \bigr ) \bigl (S_p\bigr ) \; = \; D_p (w_0) \, \end{aligned}$$
(8)

for all \(p,q,r\ge 0\). It can be shown (e.g., [10]) that the tau-structure exists and is unique up to replacing \(\Omega _{p,q},S_p\) by \(\Omega _{p,q}+c_{p,q}\) and \(S_p+a_p\) respectively, where \(c_{p,q}=c_{q,p}\) and \(a_ p\) are arbitrary constants. The tau-structure \(\Omega _{p,q},S_p\) is called canonical if

$$\begin{aligned} \Omega _{p,q}\big |_{v_i=0, \, w_i=0, \, i\in {{\mathbb {Z}}}} \; = \; 0, \quad S_{p} \big |_{v_i=0, \, w_i=0, \, i\in {{\mathbb {Z}}}} \; = \; 0. \end{aligned}$$

Let us take \(\Omega _{p,q},S_p\) the canonical tau-structure. For \(m\ge 3\), define

$$\begin{aligned} \Omega _{p_1,\ldots ,p_m} \, := \, D_{p_1} \cdots D_{p_{m-2}} \, \bigl (\Omega _{p_{m-1} p_m} \bigr ) \in {{\mathcal {A}}}, \qquad p_1,\ldots ,p_m\ge 0. \end{aligned}$$
(9)

By (6), we know that the \(\Omega _{p_1,\ldots ,p_m}\), \(m\ge 2\), are totally symmetric with respect to permutations of the indices \(p_1,\ldots ,p_m\). The first few of these polynomials are

$$\begin{aligned}&S_0 \; = \; v_0, \quad S_1 \; = \; w_1+w_0 +v_0^2, \end{aligned}$$
(10)
$$\begin{aligned}&\Omega _{0,0} \; = \; w_0, \quad \Omega _{0,1}\; = \;\Omega _{1,0}\; = \;w_1(v_1+v_0). \end{aligned}$$
(11)

If we think of \(v_0\), \(w_0\) as two functions v(n), w(n) of n, respectively, and \(v_i\), \(w_i\) as \(v(n+i)\), \(w(n+i)\), then the Toda lattice derivations \(D_k\) lead to a hierarchy of evolutionary differential–difference equations, called the Toda lattice hierarchy, given by

$$\begin{aligned}&\frac{\partial v(n)}{\partial t_k} \; = \; D_k(v_0) (n), \qquad \frac{\partial w(n)}{\partial t_k} \; = \; D_k(w_0) (n), \end{aligned}$$
(12)

where \(k\ge 0\), and the \(D_k(v_0) (n), D_k(w_0) (n)\) are defined as \(D_k(v_0), D_k(w_0)\) with \(v_i\), \(w_i\) replaced by \(v(n+i)\), \(w(n+i)\), respectively. Lemma 1 implies that the flows (12) all commute. So we can solve the whole Toda lattice hierarchy (12) together, which yields solutions of the form \((v=v(n,\mathbf{t}),w=w(n,\mathbf{t}))\). Here, \(\mathbf{t}:=(t_0,t_1,\ldots )\) denotes the infinite time vector. Note that the \(k=0\) equations read

$$\begin{aligned}&\dot{v}(n) \; = \; w(n+1) \,- \, w(n), \qquad \dot{w}(n) \; = \; w(n) \, \bigl (v(n)-v(n-1)\bigr ), \end{aligned}$$
(13)

which are equivalent to Eq. (1) via the transformation

$$\begin{aligned} w(n) \; = \; e^{\sigma (n-1)-\sigma (n)}, \qquad v(n) \; = \; - {{\dot{\sigma }}}(n). \end{aligned}$$

Here, dot, “ \(\dot{}\) ”, is identified with \(\partial /\partial t_0\).

Let V be a ring of functions of n closed under shifting n by \(\pm 1\). For two given \(f(n),g(n)\in V\), consider the initial value problem for (12) with the initial condition:

$$\begin{aligned} v(n,\mathbf{0}) \; = \;f(n), \quad w(n,\mathbf{0})\; = \;g(n). \end{aligned}$$
(14)

The solution \((v(n,\mathbf{t}), w(n,\mathbf{t}))\in V[[\mathbf{t}]]^2\) exists and is unique, which gives the following 1–1 correspondence:

$$\begin{aligned} \bigl \{\hbox {solution } (v,w)\; \hbox {of } (12) \hbox { in } V[[\mathbf{t}]]^2\bigr \} \; \longleftrightarrow \; \bigl \{\hbox {initial data } (f,g) \bigr \}. \end{aligned}$$
(15)

Example 1

\( f(n)=0,~g(n)=n\). (For this case, one can take \(V={{\mathbb {Q}}}[n]\).) The corresponding unique solution governs the enumerations of ribbon graphs in all genera.

Example 2

\(f(n)= (n+\frac{1}{2})\epsilon , ~ g(n)=1\). (For this case, one can take \(V={{\mathbb {Q}}}[\epsilon ][n]\).) The corresponding unique solution governs the Gromov–Witten invariants of \({\mathbb {P}}^1\) in the stationary sector in all genera and all degrees.

Let \((v,w)\in V[[\mathbf{t}]]^2\) be an arbitrary solution to the Toda lattice hierarchy (12). Write \(\Omega _{p,q}(n,\mathbf{t})\) and \(S_p(n,\mathbf{t})\) as the images of \(\Omega _{p,q}\) and \(S_p\) under the substitutions

$$\begin{aligned} v_i\mapsto v(n+i,\mathbf{t}),\quad w_i\mapsto w(n+i,\mathbf{t}),\qquad i\in {{\mathbb {Z}}}, \end{aligned}$$
(16)

respectively. (Similar notations will be used for other elements of \({{\mathcal {A}}}\).) Equalities (6) then imply the existence of a function \(\tau =\tau (n,\mathbf{t})\) such that for \(p,q\ge 0\),

$$\begin{aligned}&\Omega _{p,q}(n,\mathbf{t}) \; = \; \frac{\partial ^2 \log \tau (n,\mathbf{t})}{\partial t_p \partial t_q }, \end{aligned}$$
(17)
$$\begin{aligned}&S_p(n,\mathbf{t}) \; = \; \frac{\partial }{\partial t_p}\log \frac{\tau (n+1,\mathbf{t})}{\tau (n,\mathbf{t})}, \end{aligned}$$
(18)
$$\begin{aligned}&w(n,\mathbf{t}) \; = \; \frac{\tau (n+1,\mathbf{t}) \, \tau (n-1,\mathbf{t})}{\tau (n,\mathbf{t})^2} . \end{aligned}$$
(19)

We call \(\tau (n,\mathbf{t})\) the Dubrovin–Zhang (DZ)-type tau-function [10, 15] of the solution (vw), in short the tau-function of the solution. The symmetry in (9) is more obvious: the image \(\Omega _{p_1,\ldots ,p_m}(n,\mathbf{t})\) of \(\Omega _{p_1,\ldots ,p_m}\) under (16) satisfies

$$\begin{aligned} \Omega _{p_1,\ldots ,p_m}(n,\mathbf{t}) \; = \; \frac{\partial ^m \log \tau (n,\mathbf{t})}{\partial t_{p_1} \ldots \partial t_{p_m}}, \; \qquad m\ge 2,\; p_1,\ldots ,p_m\ge 0. \end{aligned}$$
(20)

Define \(\Omega _p(n,\mathbf{t})=\partial _{t_p} \log \tau (n,\mathbf{t})\), \(p\ge 0\). These logarithmic derivatives of \(\tau (n,\mathbf{t})\) are called correlation functions of the solution (vw). The specializations \(\Omega _{p_1,\ldots ,p_m}(n,\mathbf{0})\) are called m-point partial correlation functions of (vw).

Remark 1

The tau-function \(\tau (n,\mathbf{t})\) of the solution (vw) is unique up to multiplying it by the exponential of a linear function of \(n,t_0,t_1,t_2,\ldots \).

1.2 Matrix resolvent

The matrix resolvent (MR) method for computing correlation functions for integrable hierarchies was introduced in [1,2,3], and was extended to the discrete case in [10] (in particular to the Toda lattice hierarchy). Denote

$$\begin{aligned} U(\lambda )\, := \, \begin{pmatrix} v_0-\lambda &{}\quad w_0\\ -1 &{}\quad 0\end{pmatrix}. \end{aligned}$$

The following lemma for the Toda lattice hierarchy was proven in [10].

Lemma 2

[10] There exists a unique series \(R(\lambda )\in {\mathrm{Mat}}\left( 2,{{\mathcal {A}}}\left[ \left[ \lambda ^{-1}\right] \right] \right) \) satisfying

$$\begin{aligned}&\Lambda \bigl (R(\lambda )\bigr ) \, U(\lambda ) \,-\,U(\lambda ) \, R(\lambda ) \; = \; 0, \end{aligned}$$
(21)
$$\begin{aligned}&{\mathrm{Tr}} \, R (\lambda ) \; = \; 1, \quad \det R(\lambda ) \; = \; 0, \end{aligned}$$
(22)
$$\begin{aligned}&R(\lambda ) \,-\, \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{} \quad 0 \end{pmatrix} \; \in \; {\mathrm{Mat}}\left( 2,{{\mathcal {A}}}\left[ \left[ \lambda ^{-1}\right] \right] \lambda ^{-1}\right) . \end{aligned}$$
(23)

The unique series \(R(\lambda )\) in Lemma 2 is called the basic matrix resolvent. The first few terms of \(R(\lambda )\) are given by

$$\begin{aligned}&R(\lambda )\; = \; \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad 0\end{pmatrix} \; + \; \begin{pmatrix} 0 &{}\quad -w_0 \\ 1 &{}\quad 0\end{pmatrix} \frac{1}{\lambda } \; + \; \begin{pmatrix} w_0 &{}\quad -v_0w_0 \\ v_{-1} &{}\quad -w_0\end{pmatrix} \frac{1}{\lambda ^2} \nonumber \\&\qquad \qquad \qquad \; + \; \begin{pmatrix} w_0(v_0+v_{-1}) &{}\quad -w_0(w_0+w_1+v_0^2) \\ w_0+w_{-1}+v_{-1}^2 &{}\quad -w_0(v_0+v_{-1})\end{pmatrix} \frac{1}{\lambda ^3} \; + \;\cdots . \end{aligned}$$
(24)

Proposition 1

[10] For any \(k\ge 2\), the following formula holds true:

$$\begin{aligned} \sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}} \; = \; - \sum _{\pi \in {\mathcal {S}}_k/C_k} \frac{{\mathrm{tr}}\, \prod _{j=1}^k R\bigl (\lambda _{\pi (j)}\bigr )}{\prod _{j=1}^k \bigl (\lambda _{\pi {(j)}}-\lambda _{\pi {(j+1)}}\bigr )} \,-\, \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2}, \end{aligned}$$
(25)

where \({\mathcal {S}}_k\) denotes the symmetry group and \(C_k\) the cyclic group, and \(\pi (k+1):=\pi (1)\).

The meaning of (25) is the following: For any fixed permutation \((j_1,\ldots ,j_k)\) of \((1,\ldots ,k)\), expanding the right-hand side with respect to \(|\lambda _{j_1}|>\cdots>|\lambda _{j_k}|>>0\) gives identical formal power series with the left-hand side. This is because, after the summation over the \({\mathcal {S}}_k/C_k\) and subtracting \(\frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2}\), the poles in the diagonal cancel (cf. Proposition 2 of [12] for a straightforward proof of this point). We note that, as formal power series, the coefficients of the both sides of (25) are in \({{\mathcal {A}}}\). We give in Sect. 2 a new proof of (25), where we keep all derivations with coefficients in \({{\mathcal {A}}}\).

1.3 From wave functions to correlation functions

In [13], we introduced the notion of a tuple of wave functions (in many cases a pair) to the study of tau-function without using the Sato theory. Let us generalize it to the Toda lattice hierarchy. Our definition of a pair will be based on the standard construction of wave functions for the Toda lattice hierarchy [5, 6, 27]. For given (f(n), g(n)) a pair of arbitrary elements in V, let L be the linear difference operator \(L=\Lambda +f(n)+g(n) \Lambda ^{-1}\). Denote

$$\begin{aligned} s(n) \, := \, - \bigl (1-\Lambda ^{-1}\bigr )^{-1} \bigl (\log g(n)\bigr ). \end{aligned}$$
(26)

The function s(n) is in a certain extension  \({\widehat{V}}\) of V and is uniquely determined by \(\log g(n)\) up to a constant. Below we fix a choice of s(n). An element \(\psi _A(\lambda ,n)=\bigl (1+{\mathrm{O}}\left( \lambda ^{-1}\right) \bigr ) \,\lambda ^n\) in the module \({{\widetilde{V}}}\left[ \left[ \lambda ^{-1} \right] \right] \lambda ^n\) is called a (formal) wave function of type A associated with f(n), g(n), if \(L \bigl ( \psi _A(\lambda ,n)\bigr )= \lambda \psi _A(\lambda ,n)\). Here, \(\widetilde{V}\) is a ring of functions of n satisfying

$$\begin{aligned} V\subset (\Lambda -1)\bigl ({\widetilde{V}}\bigr ) \subset {\widetilde{V}}. \end{aligned}$$

An element \(\psi _B(\lambda ,n)=\bigl (1+{\mathrm{O}}\bigl (\lambda ^{-1}\bigr )\bigr ) \, e^{-s(n)} \lambda ^{-n}\) in the module \({\widetilde{V}} \left[ \left[ \lambda ^{-1} \right] \right] e^{-s(n)} \lambda ^{-n}\) is called a (formal) wave function of type B, if \(L\bigl ( \psi _B(\lambda ,n) \bigr ) = \lambda \psi _B(\lambda ,n) \). Let \( \psi _A \in {\widetilde{V}}\left[ \left[ \lambda ^{-1}\right] \right] \, \lambda ^n\) and \(\psi _B \in {\widetilde{V}} \left[ \left[ \lambda ^{-1} \right] \right] e^{-s(n)} \lambda ^{-n}\) be two wave functions of type A and of type B associated with (f(n), g(n)), respectively. Define

$$\begin{aligned} d(\lambda ,n) \, := \, \psi _A(\lambda ,n) \, \psi _B(\lambda ,n-1) - \psi _B(\lambda ,n) \, \psi _A(\lambda ,n-1). \end{aligned}$$
(27)

We call \(\psi _A,\psi _B\) form a pair if the following normalization condition holds:

$$\begin{aligned} e^{s(n-1)} d(\lambda ,n) \; = \; \lambda . \end{aligned}$$
(28)

The existence of a pair of wave functions is proved in Sect. 3.

Denote by \((v(n,\mathbf{t}),w(n,\mathbf{t}))\) the unique solution in \(V[[\mathbf{t}]]^2\) to the Toda lattice hierarchy with (f(n), g(n)) as its initial value, by \(\psi _A(\lambda ,n)\) and \(\psi _B(\lambda ,n)\) a pair of wave functions associated with (f(n), g(n)) and by \(\tau (n,\mathbf{t})\) the DZ-type tau-function of \((v(n,\mathbf{t}),w(n,\mathbf{t}))\). Introduce

$$\begin{aligned} D(\lambda ,\mu ,n) \, := \, \frac{\psi _A(\lambda ,n) \, \psi _B(\mu ,n-1) \,-\, \psi _A(\lambda ,n-1) \,\psi _B(\mu ,n)}{\lambda -\mu }. \end{aligned}$$
(29)

Theorem 1

Fix \(k\ge 2\) being an integer. The generating series of k-point partial correlation functions has the following expression:

$$\begin{aligned}&\sum _{i_1,\ldots ,i_k\ge 0} \frac{\partial ^k \log \tau }{\partial t_{i_1} \ldots \partial t_{i_k}} (n,\mathbf{0}) \, \frac{1}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}} \nonumber \\&\qquad \qquad \; = \; (-1)^{k-1} \frac{e^{k s(n-1)}}{\prod _{j=1}^k \lambda _{j}} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k D(\lambda _{\pi (j)},\lambda _{\pi (j+1)},n) \,-\, \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2} . \end{aligned}$$
(30)

Theorem 1 gives an algorithm with the initial value (f(n), g(n)) as the only input for computing the \(k_{{\mathrm{th}}}\)-order logarithmic derivatives of the tau-function \(\tau (n,\mathbf{t})\) evaluated at \(\mathbf{t}=\mathbf{0}\) for \(k\ge 2\). Indeed, by solving the spectral problem \(L(\psi )=\lambda \psi \) with \(L=\Lambda +f(n)+g(n)\Lambda ^{-1}\) and with the normalization condition (28), one constructs a pair of wave functions; the coefficients in the \({\mathbf{t}}\)-expansion of \(\log \tau (n,\mathbf{t})\) are then obtained through algebraic manipulations by using (85). (Recall that in the inverse scattering method (cf., e.g., [18, 19]), an additional integral equation needs to be solved.) Two applications of Theorem 1 are given in Sect. 5. For a certain class of bispectral solutions (cf. [20]), it would be possible to give a canonical way of constructing a pair of wave functions, which was briefly mentioned in [13] for the KdV hierarchy; we plan to do this for KdV and for Toda lattice in a future publication.

Organization of the paper In Sect. 2, we review the MR method of studying tau-structure for the Toda lattice hierarchy. In Sect. 3, we prove the existence of a pair of wave functions. In Sect. 4, we prove Theorem 1 and several other theorems. Applications to the computations of GUE correlators and Gromov–Witten invariants of \({\mathbb {P}}^1\) are given in Sect. 5. In Appendix A, we give an extension of \({{\mathcal {A}}}\), define a pair of abstract pre-wave functions, and prove an abstract version for Theorem 1.

2 Matrix resolvent and tau-structure

We continue in this section with more details in reviewing the MR method [10] to the Toda lattice hierarchy. Denote by \({\mathcal {L}}\) the matrix Lax operator for the Toda lattice:

$$\begin{aligned} {\mathcal {L}}\, := \, \begin{pmatrix} \Lambda &{}\quad 0\\ 0 &{}\quad \Lambda \end{pmatrix} \; + \; \begin{pmatrix} v_0-\lambda &{}\quad w_0\\ -1 &{}\quad 0\end{pmatrix} \; = \; \Lambda \; + \; U(\lambda ). \end{aligned}$$

Let \(R(\lambda )\) be the basic matrix resolvent (of \({\mathcal {L}}\)). Write

$$\begin{aligned}&R(\lambda ) \; = \; \begin{pmatrix} 1+\alpha (\lambda ) &{}\quad \beta (\lambda ) \\ \gamma (\lambda ) &{}\quad -\alpha (\lambda ) \end{pmatrix}, \end{aligned}$$
(31)
$$\begin{aligned}&\alpha (\lambda ) \; = \; \sum _{i\ge 0} \frac{a_i}{\lambda ^{i+1}}, \quad \beta (\lambda ) \; = \; \sum _{i\ge 0} \frac{b_i}{\lambda ^{i+1}}, \quad \gamma (\lambda ) \; = \; \sum _{i\ge 0} \frac{c_i}{\lambda ^{i+1}}, \end{aligned}$$
(32)

where \(a_i,b_i,c_i\in {{\mathcal {A}}}\). From the defining Eqs. (21)–(23), we see that the series \(\alpha ,\beta ,\gamma \) satisfy the equations

$$\begin{aligned}&\beta (\lambda ) \; = \; -w_0 \, \Lambda \bigl (\gamma (\lambda )\bigr ), \end{aligned}$$
(33)
$$\begin{aligned}&\gamma (\lambda ) \; = \; \frac{1\; + \;\alpha (\lambda ) \; + \; \Lambda ^{-1} \bigl (\alpha (\lambda )\bigr )}{\lambda - v_{-1}}, \end{aligned}$$
(34)
$$\begin{aligned}&\Bigl (\alpha (\lambda )- \Lambda \bigl (\alpha (\lambda )\bigr )\Bigr )(\lambda -v_0) \,-\, w_0 \, \frac{1+\alpha (\lambda )+\Lambda ^{-1}\bigl (\alpha (\lambda )\bigr )}{\lambda -v_{-1}} \nonumber \\&\qquad \qquad \qquad \qquad \qquad \quad \qquad \quad \quad + w_1 \, \frac{1+\Lambda \bigl (\alpha (\lambda )\bigr )+\Lambda ^2 \bigl (\alpha (\lambda )\bigr )}{\lambda -v_1} \; = \;0, \end{aligned}$$
(35)
$$\begin{aligned}&\alpha (\lambda )+\alpha (\lambda )^2 \; + \; \beta (\lambda ) \gamma (\lambda ) \; = \; 0. \end{aligned}$$
(36)

These equalities give rise to the following recursion relation for \(a_i,b_i,c_i\):

$$\begin{aligned}&b_j\; = \; -w_0 \, \Lambda (c_{j}), \qquad c_{j+1} \; = \; v_{-1} \, c_{j} \; + \; \bigl (1 \; + \; \Lambda ^{-1}\bigr ) \, (a_j),\end{aligned}$$
(37)
$$\begin{aligned}&\bigl (1-\Lambda \bigr ) \, (a_{j+1}) \; + \; v_0 \, \bigl (\Lambda -1\bigr ) \, (a_j) \; + \; w_1 \, \Lambda ^2 (c_{j}) \,-\, w_0 \, c_{j} \; = \; 0, \end{aligned}$$
(38)
$$\begin{aligned}&a_{\ell }= \sum _{i+j=\ell -1} \Bigl (w_0 \, c_{i} \, \Lambda (c_j) \,-\, a_{i} \, a_{j} \Bigr ) \end{aligned}$$
(39)

along with

$$\begin{aligned} a_0\; = \;0, \qquad c_0\; = \;1. \end{aligned}$$
(40)

Equations (37)–(40) are called the matrix resolvent recursion relation.

It was proven [10] that the abstract Toda lattice hierarchy (5) can be equivalently written as

$$\begin{aligned}&D_j \, (v_0) \; = \; \bigl (\Lambda -1\bigr ) \, (a_{j+1}), \\&D_j \, (w_0) \; = \; w_0 \, \bigl (\Lambda -1\bigr )\,(c_{j+1}), \end{aligned}$$

where \(j\ge 0\). Define an operator \(\nabla (\lambda )\) by

$$\begin{aligned} \nabla (\lambda ) \, := \, \sum _{j\ge 0} \frac{D_j}{\lambda ^{j+2}}. \end{aligned}$$
(41)

We have

$$\begin{aligned}&\nabla (\lambda ) \, ( v_0 ) \; = \; \bigl (\Lambda -1\bigr ) \, \bigl (\alpha (\lambda )\bigr ), \end{aligned}$$
(42)
$$\begin{aligned}&\nabla (\lambda ) \, ( w_0 ) \; = \; w_0 \, \bigl (\Lambda -1\bigr ) \, \bigl (\gamma (\lambda )-1\bigr ). \end{aligned}$$
(43)

Lemma 3

There exists a unique element \(W(\lambda ,\mu )\) in \({{\mathcal {A}}}\otimes {\mathrm{sl}}_2({{\mathbb {C}}}) \left[ \left[ \lambda ^{-1},\mu ^{-1}\right] \right] \lambda ^{-1} \mu ^{-1}\) of the form

$$\begin{aligned} W(\lambda ,\mu ) \; = \; \begin{pmatrix} X(\lambda ,\mu ) &{}\quad Y(\lambda ,\mu ) \\ Z(\lambda ,\mu ) &{}\quad -X(\lambda ,\mu )\end{pmatrix} \end{aligned}$$

satisfying the following linear inhomogeneous equations for the entries of W:

$$\begin{aligned}&\Lambda \bigl (W(\lambda ,\mu )\bigr ) \, U(\lambda ) \,-\, U(\lambda ) \, W(\lambda ,\mu ) \; + \; \Lambda \bigl (R(\lambda )\bigr ) \, \nabla (\mu ) \bigl (U(\lambda )\bigr ) \nonumber \\&\quad \,-\, \nabla (\mu ) \bigl (U(\lambda )\bigr ) \, R(\lambda ) \; = \;0, \end{aligned}$$
(44)
$$\begin{aligned}&X(\lambda ,\mu ) \; + \; 2 \alpha (\lambda ) \, X(\lambda ,\mu ) \; + \; \gamma (\lambda ) \, Y(\lambda ,\mu ) \; + \; \beta (\lambda ) \, Z(\lambda ,\mu ) \; = \; 0. \end{aligned}$$
(45)

Proof

The existence part of this lemma follows from Lemma 2. Indeed, if we define

$$\begin{aligned} W(\lambda ,\mu ) \, := \, \nabla (\mu ) \bigl (R(\lambda )\bigr ), \end{aligned}$$

then \(W(\lambda ,\mu )\) satisfies (44)–(45). To see the uniqueness part, we first note that the (1,2)-entry and the (2,1)-entry of the matrix equation (44) imply that Y and Z can be uniquely expressed in terms of X. Indeed, we have

$$\begin{aligned}&Z(\lambda ,\mu ) \; = \; \tfrac{(1 + \Lambda ^{-1}) (X(\lambda ,\mu ))}{\lambda - v_{-1}} \; + \; \gamma (\lambda ) \tfrac{\Lambda ^{-1} \circ \nabla (\mu ) (v_0)}{\lambda - v_{-1}}, \end{aligned}$$
(46)
$$\begin{aligned}&Y(\lambda ,\mu ) \; = \; - \nabla (\mu ) (w_0) \, \tfrac{1\; + \;\alpha (\lambda ) \; + \; \Lambda (\alpha (\lambda )) }{\lambda - v_{0}} \,- \, w_0 \, \tfrac{(1+ \Lambda ) (X(\lambda ,\mu ))}{\lambda - v_{0}} \; + \; \beta (\lambda ) \tfrac{\nabla (\mu ) (v_0)}{\lambda - v_{0}} . \end{aligned}$$
(47)

Substituting these two expressions in (45), we obtain the following linear inhomogeneous difference equation for X:

$$\begin{aligned}&\Bigl (1 + 2 \alpha (\lambda ) + \tfrac{\beta (\lambda )}{\lambda - v_{-1}} - \tfrac{w_0 \gamma (\lambda )}{\lambda - v_{0}} \Bigr ) X(\lambda ,\mu ) \,- \, \tfrac{ w_0 \gamma (\lambda ) }{\lambda - v_{0}} \Lambda \bigl (X(\lambda ,\mu )\bigr ) \nonumber \\&\quad + \tfrac{\beta (\lambda )}{\lambda - v_{-1}} \Lambda ^{-1} \bigl (X(\lambda ,\mu )\bigr ) \nonumber \\&\quad \; = \; \Bigl (1+\alpha (\lambda ) + \Lambda \bigl (\alpha (\lambda )\bigr )\Bigr ) \gamma (\lambda ) \tfrac{ \nabla (\mu ) (w_0) }{\lambda - v_{0}} \,-\, \beta (\lambda ) \gamma (\lambda ) \bigl (1+\Lambda ^{-1}\bigr ) \Bigl (\tfrac{\nabla (\mu ) (v_0)}{\lambda - v_{0}} \Bigr ). \end{aligned}$$
(48)

Suppose this equation has two solutions \(X_1,X_2\) in \({{\mathcal {A}}}\left[ \left[ \lambda ^{-1},\mu ^{-1}\right] \right] \lambda ^{-1} \mu ^{-1}\). Let \(X_0=X_1-X_2\), then \(X_0\in {{\mathcal {A}}}\left[ \left[ \lambda ^{-1},\mu ^{-1}\right] \right] \lambda ^{-1} \mu ^{-1}\), and it satisfies the following equation:

$$\begin{aligned}&\Bigl (1 + 2 \alpha (\lambda ) + \tfrac{\beta (\lambda )}{\lambda - v_{-1}} - \tfrac{w_0 \gamma (\lambda )}{\lambda - v_{0}} \Bigr ) X_0(\lambda ,\mu ) \,- \, \tfrac{ w_0 \gamma (\lambda ) }{\lambda - v_{0}} \Lambda \bigl (X_0(\lambda ,\mu )\bigr ) \nonumber \\&\quad \; + \; \tfrac{\beta (\lambda )}{\lambda - v_{-1}} \, \Lambda ^{-1} \bigl (X_0(\lambda ,\mu )\bigr ) \; = \; 0. \end{aligned}$$
(49)

It follows that \(X_0\) vanishes. Indeed, write \(X_0=\sum _{j\ge 0} X_{0,j}(\mu ) \lambda ^{-(j+1)}\). Observe that

$$\begin{aligned} \tfrac{1}{\lambda -v_{m}} \; = \; \tfrac{1}{\lambda }\; + \; \tfrac{v_m}{\lambda ^2} \; + \; \cdots \; \in \; {{\mathcal {A}}}\left[ \left[ \lambda ^{-1}\right] \right] \lambda ^{-1}, \qquad m=-1,0, \end{aligned}$$

and recall that \(\alpha (\lambda ),\beta (\lambda ),\gamma (\lambda )\in {{\mathcal {A}}}\left[ \left[ \lambda ^{-1}\right] \right] \lambda ^{-1}\). Then, by comparing the coefficients of powers of \(\lambda ^{-1}\) consecutively, we find that \(X_{0,0}(\mu )=0\), \(X_{0,1}(\mu )=0\), \(X_{0,2}(\mu )=0\), \(\ldots \). So \(X_0=0\). Hence, \(X_1=X_2\). The lemma is proved. \(\square \)

Based on this lemma, we now give a new proof for the following proposition.

Proposition 2

[10] The following equation holds true:

$$\begin{aligned} \nabla (\mu ) \, R(\lambda ) \; = \; \frac{1}{\mu -\lambda } \bigl [ R(\mu ), R(\lambda ) \bigr ] + \bigl [Q(\mu ),R(\lambda )\bigr ], \end{aligned}$$
(50)

where

$$\begin{aligned} Q(\mu ) \, := \, -\frac{{\mathrm{id}}}{\mu } \; + \; \begin{pmatrix} 0 &{} \quad 0 \\ 0 &{} \quad \gamma (\mu ) \\ \end{pmatrix}. \end{aligned}$$

Proof

Define \(W^*\) as the right-hand side of (50), i.e.,

$$\begin{aligned} W^*\, := \, \frac{1}{\mu -\lambda } \bigl [ R(\mu ), R(\lambda ) \bigr ] + \bigl [Q(\mu ),R(\lambda )\bigr ]. \end{aligned}$$

More precisely, the entries of \(W^*\) have the expressions:

$$\begin{aligned} X^*&= \tfrac{w_0}{\mu -\lambda } \Bigl (\tfrac{(\alpha (\lambda )+\Lambda (\alpha (\lambda ))+1) (\Lambda ^{-1}(\alpha (\mu ))+\alpha (\mu )+1)}{(\lambda -v_0) (\mu -v_{-1})} - \tfrac{(\Lambda ^{-1}(\alpha (\lambda ))+\alpha (\lambda )+1) (\alpha (\mu )+\Lambda (\alpha (\mu ))+1)}{(\lambda -v_{-1}) (\mu -v_0 )}\Bigr ),\end{aligned}$$
(51)
$$\begin{aligned} Y^*&= \tfrac{w_0}{\lambda -\mu } \Bigl (\tfrac{(\alpha (\lambda )+\Lambda (\alpha (\lambda ))+1) (\Lambda ^{-1}(\alpha (\mu )) (\lambda -\mu )+\alpha (\mu ) (\lambda +\mu -2 v_{-1})+\lambda -v_{-1})}{(\lambda -v_0) (\mu -v_{-1})}\nonumber \\&\quad +\tfrac{(2 \alpha (\lambda )+1) (\alpha (\mu )+\Lambda (\alpha (\mu ))+1)}{v_0-\mu }\Bigr ),\end{aligned}$$
(52)
$$\begin{aligned} Z^*&= \tfrac{1}{\lambda -\mu } \Bigl ( \tfrac{(\Lambda ^{-1}(\alpha (\lambda ))+\alpha (\lambda )+1) (\Lambda ^{-1}(\alpha (\mu ))-\alpha (\mu ))}{v_{-1}-\lambda } +\tfrac{(\Lambda ^{-1}(\alpha (\lambda ))-\alpha (\lambda )) (\Lambda ^{-1}(\alpha (\mu ))+\alpha (\mu )+1)}{\mu -v_{-1}} \Bigr ). \end{aligned}$$
(53)

We can then verify that \(W^* \in {{\mathcal {A}}}\otimes {\mathrm{sl}}_2({{\mathbb {C}}}) \left[ \left[ \lambda ^{-1},\mu ^{-1}\right] \right] \lambda ^{-1} \mu ^{-1}\), as well as that \(W:=W^*\) satisfies Eqs. (44), (45). The latter is done by a lengthy but straightforward calculation. The proposition is proved due to Lemma 3. \(\square \)

If we define \({{\widetilde{\Omega }}}_{i,j}, {\widetilde{S}}_i\) by

$$\begin{aligned}&\sum _{i,j\ge 0} \frac{{{\widetilde{\Omega }}}_{i,j}}{\lambda ^{i+2} \mu ^{j+2}} \; = \; \frac{{\mathrm{Tr}}\, \bigl (R(\lambda ) R(\mu ) \bigr )}{(\lambda -\mu )^2} \,-\, \frac{1}{(\lambda _1-\lambda _2)^2},\end{aligned}$$
(54)
$$\begin{aligned}&\Lambda \bigl (\gamma (\lambda )\bigr ) \; = \; \lambda ^{-1}\; + \;\sum _{i\ge 0} {\widetilde{S}}_i \, \lambda ^{-i-2}, \end{aligned}$$
(55)

then according to [10], \({{\widetilde{\Omega }}}_{i,j}, {\widetilde{S}}_i\) gives the canonical tau-structure for the Toda lattice, i.e.,

$$\begin{aligned} {{\widetilde{\Omega }}}_{i,j}\; = \;\Omega _{i,j}, \quad {\widetilde{S}}_i\; = \;S_i. \end{aligned}$$

These equalities together with Proposition 2 lead to Proposition 1; see [10] for the detailed proof of Proposition 1.

Before ending this section, we will make two remarks. The first remark is that all the entries of \(R(\lambda )\) can be expressed by the canonical tau-structure. Indeed, we have

$$\begin{aligned}&\alpha (\lambda ) \; = \; \sum _{p\ge 0} \Omega _{p,0} \, \lambda ^{-p-2}, \quad \beta (\lambda )=-w_0 \, \Lambda \bigl (\gamma (\lambda )\bigr ), \end{aligned}$$
(56)
$$\begin{aligned}&\Lambda \bigl (\gamma (\lambda )\bigr ) \; = \; \lambda ^{-1}\; + \;\sum _{p\ge 0} S_p \, \lambda ^{-p-2}. \end{aligned}$$
(57)

The proof was in [10]. The second remark is that existence of a tau-structure in general implies Lemma 1, and note that the proof in [10] of the fact that \({{\widetilde{\Omega }}}_{i,j}, {\widetilde{S}}_i\) is a tau-structure does not use the commutativity of the abstract Toda lattice hierarchy, so as a by-product of the matrix resolvent method we get a new proof of Lemma 1 together with a simple construction of the Toda lattice hierarchy. Similar idea was in [3].

3 Pair of wave functions

As in the Introduction, we start with the linear operator \(L(n)= \Lambda +f(n) + g(n) \, \Lambda ^{-1}\), where f(n) and g(n) are two given arbitrary elements in V. We show in this section the existence of pairs of wave functions associated with (f(n), g(n)). Let us write

$$\begin{aligned}&\psi _A(\lambda ,n) \; = \; e^{(\Lambda -1)^{-1}y(\lambda ,n)} \lambda ^n , \quad y(\lambda ,n)\, := \, \sum _{i\ge 1} \frac{y_i(n)}{\lambda ^i}, \end{aligned}$$
(58)
$$\begin{aligned}&\psi _B(\lambda ,n) \; = \; e^{(\Lambda -1)^{-1}z(\lambda ,n)} e^{-s(n)} \, \lambda ^{-n}, \quad z(\lambda ,n)\, := \, \sum _{i\ge 1} \frac{z_i(n)}{\lambda ^i}. \end{aligned}$$
(59)

Then, the spectral problems \(L(n) \bigl (\psi (\lambda ,n)\bigr ) = \lambda \psi (\lambda ,n)\) for \(\psi =\psi _A\) and for \(\psi =\psi _B\) recast into the following equations:

$$\begin{aligned}&\lambda \, e^{y(\lambda ,n)} \; + \; f(n) \,-\, \lambda \; + \; g(n) \, \lambda ^{-1} e^{-y(\lambda ,n-1)} \; = \; 0, \end{aligned}$$
(60)
$$\begin{aligned}&\lambda \, e^{-z(\lambda ,n-1)} \; + \; f(n) \,-\, \lambda \; + \; g(n+1) \, \lambda ^{-1} e^{z(\lambda ,n)} \; = \; 0, \end{aligned}$$
(61)

yielding recursions of the form (as equivalent conditions to (60)–(61))

$$\begin{aligned} y_{k+1}(n)&= - \sum _{\begin{array}{c} m_1,\ldots ,m_k\ge 0\\ \sum _{i=1}^{k} im_i=k+1 \end{array}} \frac{\prod _{i=1}^k y_i(n)^{m_i}}{\prod _{i=1}^k m_i!} -f(n)\delta _{k,0} \nonumber \\&\quad -g(n) \sum _{\begin{array}{c} m_1,\ldots ,m_{k-1}\ge 0\\ \sum _{i=1}^{k-1} im_i=k-1 \end{array}} \frac{\prod _{i=1}^{k-1} (-1)^{m_i} y_i(n-1)^{m_i}}{\prod _{i=1}^{k-1} m_i!}, \end{aligned}$$
(62)
$$\begin{aligned} z_{k+1}(n)&= \sum _{\begin{array}{c} m_1,\ldots ,m_k\ge 0\\ \sum _{i=1}^{k} im_i=k+1 \end{array}} \frac{\prod _{i=1}^k (-1)^{m_i} z_i(n)^{m_i}}{\prod _{i=1}^k m_i!} \; + \; f(n+1)\delta _{k,0} \nonumber \\&\quad + g(n+2) \sum _{\begin{array}{c} m_1,\ldots ,m_{k-1}\ge 0\\ \sum _{i=1}^{k-1} im_i=k-1 \end{array}} \frac{\prod _{i=1}^{k-1} z_i(n+1)^{m_i}}{\prod _{i=1}^{k-1} m_i!}, \end{aligned}$$
(63)

where \(k\ge 0\). From these recursions, it easily follows that \(y_k,z_k\in V\), \(k\ge 0\). This proves the existence of wave functions of type A and of type B meeting the definitions in Sect. 1.3. Clearly, \(\psi _A\) and \(\psi _B\) are unique up to multiplying by arbitrary series \(G(\lambda )\) and \(E(\lambda )\) of \(\lambda ^{-1}\) with constant coefficient of the form \(G(\lambda )\in 1+ {{\mathbb {C}}}\left[ \left[ \lambda ^{-1}\right] \right] \lambda ^{-1}\) and \(E(\lambda ) \in 1+ {{\mathbb {C}}}\left[ \left[ \lambda ^{-1}\right] \right] \lambda ^{-1}\). Since \(\psi _A(\lambda ,n)=\bigl (1+{\mathrm{O}}(\lambda ^{-1})\bigr ) \,\lambda ^n\) and since \(\psi _B(\lambda ,n)=\left( 1+{\mathrm{O}}\left( \lambda ^{-1}\right) \right) \, e^{-s(n)} \lambda ^{-n}\), we find that the \(d(\lambda ,n)\) defined in (27) must have the form

$$\begin{aligned} d(\lambda ,n)\; = \; \lambda \, e^{-s(n-1)} \, e^{\sum _{k\ge 1} d_k(n) \, \lambda ^{-k}}. \end{aligned}$$

Then, by using the definitions of wave functions and of s(n), one easily derives that

$$\begin{aligned} e^{s(n)} \, d(\lambda ,n+1)\; = \; e^{s(n-1)} \, d(\lambda ,n). \end{aligned}$$
(64)

It follows that all \(d_k(n)\), \(k\ge 1\), are constants. Therefore, for any fixed choice of \(\psi _A\), we can suitably choose the factor \(E(\lambda )\) for \(\psi _B\) such that \(\psi _A,\psi _B\) form a pair. This proves the existence of pair of wave functions associated with f(n), g(n).

We proceed with the time dependence. Let \((v(n,\mathbf{t}),w(n,\mathbf{t}))\) be the unique solution in \(V[[\mathbf{t}]]^2\) to the Toda lattice hierarchy satisfying the initial condition \(v(n,\mathbf{0})=f(n)\), \(w(n,\mathbf{0})=g(n)\). Let \(L(n,\mathbf{t}) := \Lambda +v(n,\mathbf{t}) + w(n,\mathbf{t}) \, \Lambda ^{-1}\). Define \(\sigma (n,\mathbf{t})\) as the unique up to a constant function satisfying the following equations:

$$\begin{aligned}&w(n,\mathbf{t}) \; = \; e^{\sigma (n-1,\mathbf{t})-\sigma (n,\mathbf{t})}, \end{aligned}$$
(65)
$$\begin{aligned}&\frac{\partial \sigma (n,\mathbf{t})}{\partial t_p} \; = \; - S_p (n,\mathbf{t}),\quad p\ge 0. \end{aligned}$$
(66)

An element \(\psi _A(n,\mathbf{t},\lambda )= \bigl (1+\mathrm{O}\bigl (\lambda ^{-1}\bigr )\bigr ) \, \lambda ^n \, e^{\sum _{k\ge 0} t_k\lambda ^{k+1}}\) in \({\widetilde{V}}\left[ \left[ \mathbf{t}, \lambda ^{-1}\right] \right] \, \lambda ^n e^{\sum _{k\ge 0} t_k\lambda ^{k+1}}\) is called a wave function of type A associated with \((v(n,\mathbf{t}),w(n,\mathbf{t}))\) if

$$\begin{aligned}&L(n,\mathbf{t}) \, \bigl (\psi _A(\lambda ,n,\mathbf{t})\bigr ) \; = \; \lambda \, \psi _A(\lambda ,n,\mathbf{t}) , \quad \frac{\partial \psi _A}{\partial t_k} \; = \; \bigl (L^{k+1}\bigr )_+ \bigl (\psi _A\bigr ). \end{aligned}$$
(67)

An element \(\psi _B(n,\mathbf{t},\lambda ) = \bigl (1+\mathrm{O}\bigl (\lambda ^{-1}\bigr )\bigr ) \lambda ^{-n} e^{-\sum _{k\ge 0} t_k\lambda ^{k+1}}\) in \({\widetilde{V}}\left[ \left[ \mathbf{t},\lambda ^{-1}\right] \right] e^{-\sigma (n,\mathbf{t})} \lambda ^{-n} e^{-\sum _{k\ge 0} t_k\lambda ^{k+1}}\) is called a wave function of type B associated with \((v(n,\mathbf{t}),w(n,\mathbf{t}))\) if

$$\begin{aligned}&L(n,\mathbf{t}) \, \bigl (\psi _B(\lambda ,n,\mathbf{t})\bigr ) \; = \; \lambda \, \psi _B(\lambda ,n,\mathbf{t}) , \quad \frac{\partial \psi _B}{\partial t_k} \; = \; - \bigl (L^{k+1}\bigr )_- \bigl (\psi _B\bigr ). \end{aligned}$$
(68)

The existence of wave functions \(\psi _A\) and \(\psi _B\) of type A and of type B associated with \((v(n,\mathbf{t}),w(n,\mathbf{t}))\) is a standard result in the theory of integrable systems (cf. [5, 6, 13, 27]); therefore, we omit its details. Denote

$$\begin{aligned} d(\lambda ,n,\mathbf{t})\, := \, \psi _A(\lambda ,n,\mathbf{t}) \, \psi _B(\lambda ,n-1,\mathbf{t}) \,-\, \psi _B(\lambda ,n,\mathbf{t}) \, \psi _A(\lambda ,n-1,\mathbf{t}), \end{aligned}$$
(69)

and introduce

$$\begin{aligned} m(\mu , \lambda , n,\mathbf{t}) \, := \, \frac{R(\mu ,n,\mathbf{t})}{\mu -\lambda } \; + \; Q(\mu ,n,\mathbf{t}), \end{aligned}$$
(70)

where \(Q(\mu ,n,\mathbf{t}) := -\frac{{\mathrm{id}}}{\mu } \; + \; \begin{pmatrix} 0 &{}\quad 0 \\ 0 &{}\quad \gamma (\mu ,n,\mathbf{t}) \\ \end{pmatrix}\). We know from, e.g., [10] that the wave function \(\psi _A(\lambda ,n,\mathbf{t})\) satisfies

$$\begin{aligned}&\nabla (\mu ) \, \begin{pmatrix} \psi _A(\lambda ,n,\mathbf{t}) \\ \psi _A(\lambda ,n-1,\mathbf{t}) \end{pmatrix} \; = \; m(\mu ,\lambda ,n,\mathbf{t}) \, \begin{pmatrix} \psi _A(\lambda ,n,\mathbf{t}) \\ \psi _A(\lambda ,n-1,\mathbf{t}) \end{pmatrix}. \end{aligned}$$
(71)

Similarly, the wave function \(\psi _B(\lambda ,n,\mathbf{t})\) satisfies

$$\begin{aligned}&\nabla (\mu ) \, \begin{pmatrix} \psi _B(\lambda ,n,\mathbf{t}) \\ \psi _B(\lambda ,n-1,\mathbf{t}) \end{pmatrix} \; = \; \biggl (m(\mu ,\lambda ,n,\mathbf{t}) \, - \, \frac{\lambda }{\mu (\mu -\lambda )} I \biggr )\, \begin{pmatrix} \psi _B(\lambda ,n,\mathbf{t}) \\ \psi _B(\lambda ,n-1,\mathbf{t}) \end{pmatrix}. \end{aligned}$$
(72)

Here, I denotes the \(2\times 2\) identity matrix.

Lemma 4

The following formula holds true:

$$\begin{aligned} \nabla (\mu ) \, \bigl (d(\lambda ,n,\mathbf{t})\bigr ) \; = \; \biggl ( - \frac{1}{\mu } + \gamma (\mu ,n,\mathbf{t}) \biggr ) \, d(\lambda ,n,\mathbf{t}). \end{aligned}$$
(73)

Proof

Recalling definition (69) for d and using (71)–(72), we find

$$\begin{aligned} \nabla (\mu ) \, \bigl ( d(\lambda ,n,\mathbf{t}) \bigr )&\; = \; \biggl ( {\mathrm{tr}}\bigl (m(\mu ,\lambda ,n,\mathbf{t})\bigr ) - \frac{\lambda }{\mu (\mu -\lambda )} \biggr ) \, d(\lambda ,n,\mathbf{t}). \end{aligned}$$
(74)

The lemma is then proved via a straightforward computation. \(\square \)

Definition 1

We say \(\psi _A,\psi _B\) form a pair if \( e^{\sigma (n-1,\mathbf{t})} d(\lambda ,n,\mathbf{t})=\lambda \).

The next lemma shows the existence of a pair.

Lemma 5

There exist a pair of wave functions \(\psi _A,\psi _B\) associated with \((v(n,\mathbf{t}),w(n,\mathbf{t}))\). Moreover, the freedom of the pair is characterized by a factor \(G(\lambda )\) via

$$\begin{aligned}&\displaystyle \psi _A(\lambda ,n,\mathbf{t}) \; \mapsto \; G(\lambda ) \, \psi _A(\lambda ,n,\mathbf{t}), \quad \psi _B(\lambda ,n,\mathbf{t}) \; \mapsto \; \frac{1}{G(\lambda )} \, \psi _B(\lambda ,n,\mathbf{t}), \end{aligned}$$
(75)
$$\begin{aligned}&\displaystyle G(\lambda ) \; = \; \sum _{j\ge 0} G_j \lambda ^{-j}, \quad G_0\; = \;1 \end{aligned}$$
(76)

with \(G_j\), \(j\ge 1\) being arbitrary constants.

Proof

Firstly, the freedom of a wave function \(\psi _A\) associated with (vw) is characterized by the multiplication by a factor \(G(\lambda )\) of the form (76). Fix an arbitrary choice of \(\psi _A\). For \(\psi _B\) being a wave function of type B associated with (vw), from (69) and the definitions of wave functions, we know \(e^{\sigma (n-1,\mathbf{t})}d(\lambda ,n,\mathbf{t})\) must have the form

$$\begin{aligned} e^{\sigma (n-1,\mathbf{t})}d(\lambda ,n,\mathbf{t}) \; = \; \lambda \, e^{\sum _{k\ge 1} d_k(n,\mathbf{t}) \, \lambda ^{-k}} \end{aligned}$$
(77)

for some \(d_k(n,\mathbf{t})\), \(k\ge 1\). By using (67), (68), (69), we find

$$\begin{aligned}&d(\lambda ,n+1,\mathbf{t}) \; = \; w(n,\mathbf{t}) \, d(\lambda ,n,\mathbf{t}) \; = \; e^{\sigma (n-1,\mathbf{t})-\sigma (n,\mathbf{t})} \, d(\lambda ,n,\mathbf{t}) , \end{aligned}$$

i.e.,

$$\begin{aligned} e^{\sigma (n,\mathbf{t})} d(\lambda ,n+1,\mathbf{t})\; = \; e^{\sigma (n-1,\mathbf{t})} \, d(\lambda ,n,\mathbf{t}) , \end{aligned}$$
(78)

Using Lemma 4 and (66), we have

$$\begin{aligned}&\nabla (\mu ) \Bigl (e^{\sigma (n-1,\mathbf{t})}d(\lambda ,n,\mathbf{t})\Bigr ) \\&\quad \; = \; e^{\sigma (n-1,\mathbf{t})} \nabla (\mu ) \bigl (\sigma (n-1,\mathbf{t})\bigr ) d(\lambda ,n,\mathbf{t}) \; + \; e^{\sigma (n-1,\mathbf{t})} \nabla (\mu ) \bigl (d(\lambda ,n,\mathbf{t})\bigr ) \\&\quad \; = \; - e^{\sigma (n-1,\mathbf{t})}\sum _{p\ge 0} \frac{S_p(n-1,\mathbf{t})}{\mu ^{p+2}} d(\lambda ,n,\mathbf{t}) \\&\qquad \qquad + e^{\sigma (n-1,\mathbf{t})} d(\lambda ,n,\mathbf{t}) \biggl ( - \frac{1}{\mu } \; + \; \gamma (\mu ,n,\mathbf{t}) \biggr ) \; = \; 0. \end{aligned}$$

So we have

$$\begin{aligned} \frac{\partial (e^{\sigma (n-1,\mathbf{t})}d(\lambda ,n,\mathbf{t}))}{\partial t_p} \; = \; 0, \quad \forall \, p\ge 0. \end{aligned}$$
(79)

We deduce from (77), (78), (79) that \(d_k(n,\mathbf{t})\), \(k\ge 1\), are all constants. Therefore, there exists a unique choice of \(\psi _B\) such that \(\psi _A,\psi _B\) form a pair. The lemma is proved. \(\square \)

4 The k-point generating series

Let \((v,w)=(v(n,\mathbf{t}),w(n,\mathbf{t}))\in V[[\mathbf{t}]]^2\) be the unique solution to the Toda lattice hierarchy with the initial value \((v(n,\mathbf{0}),w(n,\mathbf{0}))=(f(n),g(n))\), and \((\psi _A,\psi _B)\) a pair of wave functions associated with (vw). Define

$$\begin{aligned} \Psi _{{\mathrm{pair}}}(\lambda ,n,\mathbf{t}) \; = \; \begin{pmatrix} \psi _A (\lambda ,n,\mathbf{t}) &{}\quad \psi _B(\lambda ,n,\mathbf{t}) \\ \psi _A(\lambda ,n-1,\mathbf{t}) &{}\quad \psi _B(\lambda ,n-1,\mathbf{t}) \end{pmatrix}. \end{aligned}$$
(80)

Proposition 3

The following identity holds true:

$$\begin{aligned} R(\lambda , n, \mathbf{t}) \; \equiv \; \Psi _{{\mathrm{pair}}}(\lambda ,n,\mathbf{t}) \, \begin{pmatrix} 1 &{} \quad 0 \\ 0 &{} \quad 0 \end{pmatrix} \, \Psi _{{\mathrm{pair}}}^{-1}(\lambda ,n,\mathbf{t}) . \end{aligned}$$
(81)

Proof

Define

$$\begin{aligned}M\; = \;M(\lambda ,n,\mathbf{t})\, := \, \Psi _{{\mathrm{pair}}}(\lambda ,n,\mathbf{t}) \, \begin{pmatrix} 1 &{} \quad 0 \\ 0 &{}\quad 0 \end{pmatrix} \, \Psi _{{\mathrm{pair}}}^{-1}(\lambda ,n,\mathbf{t}). \end{aligned}$$

It is easy to verify that M satisfies

$$\begin{aligned} \bigl [{\mathcal {L}},M\bigr ] \, \bigl (\Psi _{{\mathrm{pair}}}\bigr ) \; = \; 0 , \qquad \det \, M \; = \; 0. \end{aligned}$$

The entries of M in terms of the pair of wave functions read

$$\begin{aligned} M=\frac{1}{d(\lambda ,n,\mathbf{t})} \begin{pmatrix} \psi _A(\lambda ,n,\mathbf{t}) \, \psi _B(\lambda ,n-1,\mathbf{t}) &{}\quad -\psi _A(\lambda ,n,\mathbf{t}) \, \psi _B(\lambda ,n,\mathbf{t}) \\ \psi _A(\lambda ,n-1,\mathbf{t}) \, \psi _B(\lambda ,n-1,\mathbf{t}) &{}\quad -\psi _A(\lambda ,n-1,\mathbf{t}) \, \psi _B(\lambda ,n,\mathbf{t}) \end{pmatrix} , \end{aligned}$$
(82)

where we recall that \(d(\lambda ,n,\mathbf{t}) = \psi _A(\lambda ,n,\mathbf{t}) \, \psi _B(\lambda ,n-1,\mathbf{t}) - \psi _B(\lambda ,n,\mathbf{t}) \, \psi _A(\lambda ,n-1,\mathbf{t})\), which coincides with the determinant of \(\Psi (\lambda ,n,\mathbf{t})\). It follows from \(\psi _A(\lambda ,n,\mathbf{t})= \bigl (1+\mathrm{O}\bigl (\lambda ^{-1}\bigr )\bigr ) \, \lambda ^n \, e^{\sum _{k\ge 0} t_k\lambda ^{k+1}}\) and \(\psi _B(\lambda ,n,\mathbf{t}) = \bigl (1+\mathrm{O}\bigl (\lambda ^{-1}\bigr )\bigr ) \, e^{-\sigma (n,\mathbf{t})} \lambda ^{-n} \, e^{-\sum _{k\ge 0} t_k\lambda ^{k+1}}\) that

$$\begin{aligned} M(\lambda ) \,-\, \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad 0 \end{pmatrix} \; \in \; {\mathrm{Mat}}\left( 2,{\widetilde{V}}\left[ \left[ \mathbf{t},\lambda ^{-1}\right] \right] \lambda ^{-1}\right) . \end{aligned}$$
(83)

The proposition then follows from the uniqueness theorem proved in Sect. 2. \(\square \)

Define

$$\begin{aligned} D(\lambda ,\mu ,n,\mathbf{t}) \, := \, \frac{\psi _A(\lambda ,n,\mathbf{t}) \, \psi _B(\mu ,n-1,\mathbf{t}) \,-\, \psi _A(\lambda ,n-1,\mathbf{t}) \,\psi _B(\mu ,n,\mathbf{t})}{\lambda -\mu }.\nonumber \\ \end{aligned}$$
(84)

Theorem 2

Fix \(k\ge 2\) being an integer. The generating series of k-point correlation functions of the solution \((v(n,\mathbf{t}),w(n,\mathbf{t}))\) has the following expression:

$$\begin{aligned}&\sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}(n,\mathbf{t})}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}} \nonumber \\&\quad \; = \; (-1)^{k-1} \frac{e^{k \sigma (n-1,\mathbf{t})}}{\prod _{j=1}^k \lambda _{j}} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k D(\lambda _{\pi (j)},\lambda _{\pi (j+1)},n,\mathbf{t}) \,-\, \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2} . \end{aligned}$$
(85)

Proof

It follows from (81) that

$$\begin{aligned} R(\lambda , n, \mathbf{t}) \; = \; \frac{r_1(\lambda ,n,\mathbf{t})^T r_2(\lambda ,n,\mathbf{t})}{d(\lambda ,n,\mathbf{t})} , \end{aligned}$$
(86)

where \(r_1(\lambda ,n,\mathbf{t}):=(\psi _A(\lambda ,n,\mathbf{t}), \psi _A(\lambda ,n-1,\mathbf{t}))\), \(r_2(\lambda ,n,\mathbf{t}):=(\psi _B(\lambda ,n-1,\mathbf{t}),-\psi _B(\lambda ,n,\mathbf{t}))\). Substituting this expression into the identity

$$\begin{aligned}&\sum _{i_1,i_2\ge 0} \frac{\Omega _{i_1,i_2}(n,\mathbf{t})}{\lambda _1^{i_1+2} \lambda _2^{i_2+2}} \; = \; \frac{{\mathrm{Tr}}\, \bigl (R_1(\lambda _1,n,\mathbf{t}) \, R_2(\lambda _2,n,\mathbf{t})\bigr )}{(\lambda _1-\lambda _2)^2} \,-\, \frac{1}{(\lambda _1-\lambda _2)^2}, \end{aligned}$$
(87)

we obtain

$$\begin{aligned} \sum _{i_1,i_2\ge 0} \frac{\Omega _{i_1,i_2}(n,\mathbf{t})}{\lambda _1^{i_1+2} \lambda _2^{i_2+2}}&\; = \; \frac{{\mathrm{Tr}}\, \bigl (r_1(\lambda _1,n,\mathbf{t})^T r_2(\lambda _1,n,\mathbf{t}) \, r_1(\lambda _2,n,\mathbf{t})^T r_2(\lambda _2,n,\mathbf{t})\bigr )}{d(\lambda _1,n,\mathbf{t}) d(\lambda _2,n,\mathbf{t}) (\lambda _1-\lambda _2)^2}\nonumber \\&\qquad \,-\, \frac{1}{(\lambda _1-\lambda _2)^2} \nonumber \\&\; = \; \frac{\bigl (r_2(\lambda _2,n,\mathbf{t}) \, r_1(\lambda _1,n,\mathbf{t})^T \bigr ) \, \bigl (r_2(\lambda _1,n,\mathbf{t}) \, r_1(\lambda _2,n,\mathbf{t})^T \bigr )}{ d(\lambda _1,n,\mathbf{t}) d(\lambda _2,n,\mathbf{t}) (\lambda _1-\lambda _2)^2}\nonumber \\&\qquad \,-\, \frac{1}{ (\lambda _1-\lambda _2)^2} \nonumber \\&\; = \; - \frac{D(\lambda _1,\lambda _2,n,\mathbf{t}) \, D(\lambda _2,\lambda _1,n,\mathbf{t}) }{ \lambda _1 \lambda _2 \, e^{-2\sigma (n-1,\mathbf{t})} } \,-\, \frac{1}{(\lambda _1-\lambda _2)^2}, \end{aligned}$$
(88)

where we used definition (84) and

$$\begin{aligned}&\frac{\psi _A(\lambda ,n,\mathbf{t}) \, \psi _B(\mu ,n-1,\mathbf{t}) \,-\, \psi _A(\lambda ,n-1,\mathbf{t}) \,\psi _B(\mu ,n,\mathbf{t})}{\lambda -\mu } \nonumber \\&\quad \; = \; \frac{r_2(\mu ,n,\mathbf{t}) \, r_1(\lambda ,n,\mathbf{t})^T}{\lambda -\mu }. \end{aligned}$$

This proves the \(k=2\) case of (85). For \(k\ge 3\), the proof is similar. Indeed,

$$\begin{aligned}&\sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}(n,\mathbf{t})}{\lambda _1^{i_1+1} \ldots \lambda _k^{i_k+1}} \nonumber \\&\qquad \; = \; - \sum _{\pi \in {\mathcal {S}}_k/C_k} \frac{{\mathrm{Tr}}\, \Bigl (\prod _{j=1}^k r_1\bigl (\lambda _{\pi (j)},n,\mathbf{t}\bigr )^T r_2\bigl (\lambda _{\pi (j)},n,\mathbf{t}\bigr ) \Bigr )}{e^{-k \, \sigma (n-1,\mathbf{t})} \, \prod _{j=1}^k \bigl (\lambda _{\pi {(j)}}-\lambda _{\pi {(j+1)}}\bigr )} \nonumber \\&\qquad \; = \; - \sum _{\pi \in {\mathcal {S}}_k/C_k} \frac{r_2\bigl (\lambda _{\pi (k)},n,\mathbf{t}\bigr ) \, r_1\bigl (\lambda _{\pi (1)},n,\mathbf{t}\bigr )^T \, \ldots \, r_2\bigl (\lambda _{\pi (k-1)},n,\mathbf{t}\bigr ) \, r_1\bigl (\lambda _{\pi (k)},n,\mathbf{t}\bigr )^T}{e^{-k \, \sigma (n-1,\mathbf{t})} \, \prod _{j=1}^k \bigl (\lambda _{\pi {(j)}}-\lambda _{\pi {(j+1)}}\bigr )} \nonumber \\&\qquad \; = \; - \frac{(-1)^k}{e^{-k \, \sigma (n-1,\mathbf{t})}} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k D\bigl (\lambda _{\pi (j)}, \lambda _{\pi (j+1)},n,\mathbf{t}\bigr ). \end{aligned}$$
(89)

This proves the \(k\ge 3\) case of (85). The theorem is proved. \(\square \)

Remark 2

In (85) or (30), the freedom (75) affects the \(D(\lambda , \mu )\) through multiplying it by a factor of the form \(\frac{G(\lambda )}{G(\mu )}\), but the product \(\prod _{j=1}^k D(\lambda _{\pi (j)},\lambda _{\pi (j+1)})\) remains unchanged.

In Appendix A, the abstract form of (85) is obtained, where a pair of abstract pre-wave functions are introduced.

Proof of Theorem 1

Taking \(\mathbf{t}=\mathbf{0}\) on the both sides of (85) gives (30). \(\square \)

Write

$$\begin{aligned} \psi _A(\lambda ,n,\mathbf{t}) \; = \; \phi _A(\lambda ,n,\mathbf{t}) \, \lambda ^n, \qquad \psi _B(\lambda ,n,\mathbf{t}) \; = \; \phi _B(\lambda ,n,\mathbf{t}) \, e^{-\sigma (n,\mathbf{t})} \, \lambda ^{-n}.\quad \end{aligned}$$
(90)

Theorem 1 can then be alternatively written in terms of \(\phi _A,\phi _B\) by the following corollary.

Corollary 1

The following formula holds true for \(k\ge 2\):

$$\begin{aligned} \sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}(n,\mathbf{t})}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}}&= (-1)^{k-1} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k B(\lambda _{\pi (j)},\lambda _{\pi (j+1)},n,\mathbf{t})- \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2} , \end{aligned}$$
(91)

where \(B(\lambda ,\mu ,n,\mathbf{t})\) is defined by

$$\begin{aligned} B(\lambda ,\mu ,n,\mathbf{t}) := \frac{\phi _A(\lambda ,n,\mathbf{t}) \, \phi _B(\mu ,n-1,\mathbf{t}) \,-\, w(n,\mathbf{t})\, \phi _A(\lambda ,n-1,\mathbf{t}) \,\phi _B(\mu ,n,\mathbf{t})}{\lambda -\mu }.\nonumber \\ \end{aligned}$$
(92)

In particular, let \(\phi _A(\lambda ,n):=e^{(\Lambda -1)^{-1}(y(\lambda ,n))}\), \(\phi _B(\lambda ,n):=e^{(\Lambda -1)^{-1}(z(\lambda ,n))} e^{-s(n)}\) (cf. (60)–(61)), and let \(B(\lambda ,\mu ,n):= \frac{\phi _A(\lambda ,n) \, \phi _B(\mu ,n-1) \,-\, g(n)\, \phi _A(\lambda ,n-1) \,\phi _B(\mu ,n)}{\lambda -\mu }\), then we have

$$\begin{aligned} \sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}(n,\mathbf{0})}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}}&= (-1)^{k-1} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k B(\lambda _{\pi (j)},\lambda _{\pi (j+1)},n) \nonumber \\&\quad -\, \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2} . \end{aligned}$$
(93)

For some particular examples related to matrix models, it turns out that the suitable chosen D coincides, possibly up to simple factors, with certain kernel of the matrix model. However, the D is not unique. We now introduce a formal series \(K(\lambda ,\mu )\) such that the generating series of multi-point correlation functions still has an explicit expression, but this time K is local and is therefore unique for the given solution. The series K is defined by

$$\begin{aligned} K(\lambda ,\mu ) \, := \, \frac{(1\; + \;\alpha (\lambda ))(1\; + \;\alpha (\mu )) - w_0 \, \gamma (\lambda ) \, \Lambda \bigl (\gamma (\mu )\bigr )}{\lambda -\mu }, \end{aligned}$$
(94)

where \(1+\alpha (\lambda )\) is the (1,1) entry of the basic matrix resolvent \(R(\lambda )\), and \(\gamma (\lambda )\) is the (2,1) entry. The next theorem expresses the left-hand side of (85) in terms of K.

Theorem 3

For any \(k\ge 2\), the following formula holds true:

$$\begin{aligned} \sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}}&\; = \; (-1)^{k-1} \frac{\sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k K\bigl (\lambda _{\pi (j)},\lambda _{\pi (j+1)}\bigr )}{ \prod _{i=1}^k \bigl (1\; + \;\alpha (\lambda _i)\bigr )} \nonumber \\&\quad \qquad \quad -\, \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2} . \end{aligned}$$
(95)

Proof

The identity (81) gives

$$\begin{aligned}&\psi _B(\lambda ,n-1,\mathbf{t}) \; = \; \frac{(1\; + \;\alpha (\lambda ,n,\mathbf{t})) \, d(\lambda ,n,\mathbf{t})}{\psi _A(\lambda ,n,\mathbf{t})}, \\&\psi _B(\lambda ,n,\mathbf{t}) \; = \; - \, \frac{\beta (\lambda ,n,\mathbf{t}) \, d(\lambda ,n,\mathbf{t})}{\psi _A(\lambda ,n,\mathbf{t})} \; = \; w_n\,\frac{\gamma (\lambda ,n+1,\mathbf{t}) \, d(\lambda ,n,\mathbf{t})}{\psi _A(\lambda ,n,\mathbf{t})}, \\&\psi _A(\lambda ,n-1,\mathbf{t}) \; = \; \psi _A(\lambda ,n,\mathbf{t}) \, \frac{\gamma (\lambda ,n,\mathbf{t})}{1\; + \;\alpha (\lambda ,n,\mathbf{t})}. \end{aligned}$$

Substituting these expressions into (84), we obtain

$$\begin{aligned} D(\lambda ,\mu ,n,\mathbf{t}) \; = \; d(\mu ,n,\mathbf{t}) \frac{\psi _A(\lambda ,n,\mathbf{t})}{\psi _A(\mu ,n,\mathbf{t})} e(\lambda ,\mu ,n,\mathbf{t}), \end{aligned}$$
(96)

where

$$\begin{aligned} e(\lambda ,\mu ,n,\mathbf{t}){:=} \frac{(1\; + \;\alpha (\lambda ,n,\mathbf{t}))(1\; + \;\alpha (\mu ,n,\mathbf{t})) - w_n(\mathbf{t}) \, \gamma (\lambda ,n,\mathbf{t}) \, \gamma (\mu ,n+1,\mathbf{t}) }{(\lambda -\mu ) \, (1\; + \;\alpha (\lambda ,n,\mathbf{t}))}.\nonumber \\ \end{aligned}$$
(97)

Combining with the definition of \(K(\lambda ,\mu ,n,\mathbf{t})\) and Theorem 1, we find

$$\begin{aligned}&\sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}(n,\mathbf{t})}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}} \nonumber \\&\quad \; = \; (-1)^{k-1} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k K \bigl (\lambda _{\pi (j)},\lambda _{\pi (j+1)},n,\mathbf{t}\bigr ) \,-\, \frac{ \delta _{k,2}}{(\lambda _1-\lambda _2)^2}. \end{aligned}$$
(98)

The theorem is proved. \(\square \)

It seems to be an interesting question to study the geometric and algebraic meaning of the kernel K (as well as D). Below we give without proof some of their properties.

Proposition 4

The functions K and D are related to

$$\begin{aligned}&K(\lambda ,\mu ,n,\mathbf{t}) \; = \;\frac{e^{\sigma (n-1,\mathbf{t})} }{\mu } \, \bigl (1+\alpha (\lambda ,n,\mathbf{t})\bigr ) \, \frac{\psi _A(\mu ,n,\mathbf{t})}{\psi _A(\lambda ,n,\mathbf{t})} \, D(\lambda ,\mu ,n,\mathbf{t}) \\&\quad \quad \quad \quad \quad \; = \;\frac{e^{2\sigma (n-1,\mathbf{t})}}{\lambda \,\mu } \, \psi _A(\mu ,n,\mathbf{t}) \, \psi _B(\lambda ,n-1,\mathbf{t})\, D(\lambda ,\mu ,n,\mathbf{t}) \\&\quad \quad \quad \quad \quad \; = \;\frac{e^{\sigma (n-1,\mathbf{t})}}{\lambda } \, \bigl (1+\alpha (\mu ,n,\mathbf{t})\bigr )\,\frac{\psi _B(\lambda ,n-1,\mathbf{t}) }{\psi _B(\mu ,n-1,\mathbf{t})} \, D(\lambda ,\mu ,n,\mathbf{t}). \end{aligned}$$

We observe that the following three formal series

$$\begin{aligned} K(\lambda ,\mu )-\frac{1+\alpha (\lambda )}{\lambda -\mu }, \quad K(\lambda ,\mu )-\frac{1+\alpha (\mu )}{\lambda -\mu }, \quad K(\lambda ,\mu )-\frac{2+\alpha (\lambda )+\alpha (\mu )}{2 \, (\lambda -\mu )} \end{aligned}$$

all belong to \({{\mathcal {A}}}\left[ \left[ \lambda ^{-1},\mu ^{-1}\right] \right] \). Therefore, the following three formal series

$$\begin{aligned}&\displaystyle K(\lambda ,\mu ,n,\mathbf{t})-\frac{1+\alpha (\lambda ,n,\mathbf{t})}{\lambda -\mu }, \quad K(\lambda ,\mu ,n,\mathbf{t})-\frac{1+\alpha (\mu ,n,\mathbf{t})}{\lambda -\mu }, \\&\displaystyle K(\lambda ,\mu ,n,\mathbf{t})-\frac{2+\alpha (\lambda ,n,\mathbf{t})+\alpha (\mu ,n,\mathbf{t})}{2(\lambda -\mu )} \end{aligned}$$

all belong to \(V[[\mathbf{t}]]\left[ \left[ \lambda ^{-1},\mu ^{-1}\right] \right] \). It follows from this observation and Proposition 4 that

$$\begin{aligned} \frac{e^{s(n-1)} }{\mu }\, D(\lambda ,\mu ,n,\mathbf{0}) \, \biggl (\frac{\mu }{\lambda }\biggr )^n \,-\, \frac{1}{\lambda -\mu } \; \in \; {\widetilde{V}}\left[ \left[ \lambda ^{-1},\mu ^{-1}\right] \right] . \end{aligned}$$
(99)

Remark 3

We could loosen both the conditions for wave functions and the pair condition. Let us say \(\psi _A\) and \(\psi _B\) are pre-wave functions of type A and of type B, respectively, if they satisfy the first equations of (67) and (68). Define \(d_{{\mathrm{pre}}}(\lambda ,n,\mathbf{t})\) and \(D_{{\mathrm{pre}}}(\lambda ,\mu ,n,\mathbf{t})\) by (129) and (140). Then, the following formula holds true:

$$\begin{aligned}&\sum _{i_1,\ldots ,i_k\ge 0} \frac{\Omega _{i_1,\ldots ,i_k}(n,\mathbf{t})}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}} \nonumber \\&\quad \qquad \; = \; \frac{(-1)^{k-1}}{\prod _{j=1}^k d_{{\mathrm{pre}}}(\lambda _j,n,\mathbf{t})} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k D_{{\mathrm{pre}}}(\lambda _{\pi (j)},\lambda _{\pi (j+1)},n,\mathbf{t}) \,-\, \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2} . \end{aligned}$$
(100)

Now, \(\psi _A\) and \(\psi _B\) are determined by \((v(n,\mathbf{t}),w(n,\mathbf{t}) )\) up to

$$\begin{aligned} \psi _A (\lambda ,n,\mathbf{t}) \; \mapsto \; G(\lambda ,\mathbf{t}) \, \psi _A(\lambda ,n,\mathbf{t}), \quad \psi _B (\lambda ,n,\mathbf{t}) \; \mapsto \; E(\lambda ,\mathbf{t}) \, \psi _B (\lambda ,n,\mathbf{t}), \end{aligned}$$

where \(G(\lambda ,\mathbf{t})=1+\sum _{k\ge 1} G_k(\mathbf{t}) \lambda ^{-k}\), \(E(\lambda ,\mathbf{t})=1+\sum _{k\ge 1} E_k(\mathbf{t})\lambda ^{-k}\) with \(G_k(\mathbf{t}), E_k(\mathbf{t}) \in {{\mathbb {C}}}[[\mathbf{t}]]\), \(k\ge 1\). This freedom affects \(D_{{\mathrm{pre}}}(\lambda ,\mu ,n,\mathbf{t})\) and \(d_{{\mathrm{pre}}}(\lambda ,n,\mathbf{t})\) into

$$\begin{aligned} D_{{\mathrm{pre}}}(\lambda ,\mu ,n,\mathbf{t})\mapsto & {} G(\lambda ,\mathbf{t}) \, E(\mu ,\mathbf{t}) \, D_{{\mathrm{pre}}}(\lambda ,\mu ,\mathbf{t}) , \quad \nonumber \\ d_{{\mathrm{pre}}}(\lambda ,n,\mathbf{t})\mapsto & {} G(\lambda ,\mathbf{t}) \, E(\lambda ,\mathbf{t}) \, d_{{\mathrm{pre}}}(\lambda ,\mathbf{t}). \end{aligned}$$

Therefore, it gives rise to each summand of (100) the factor

$$\begin{aligned} \frac{\prod _{j=1}^k G(\lambda _{\pi (j)},\mathbf{t}) E(\lambda _{\pi (j+1)},\mathbf{t})}{\prod _{j=1}^k G(\lambda _j,\mathbf{t}) E(\lambda _j,\mathbf{t})}, \end{aligned}$$

which is equal to 1. Hence, the right-hand side of (100) remains unchanged.

5 Applications

Partition functions in some matrix models and enumerative models are particular tau-functions for the Toda lattice hierarchy. Theorem 1 can then be used for computing their logarithmic derivatives. In this section, we do two explicit computations.

5.1 Application I: enumeration of ribbon graphs

The initial data of the GUE solution to the Toda lattice hierarchy are given by \(f(n)=0\) and \(g(n)=n\); see, for example, [10] for the proof. For this case, we can take \(V={{\mathbb {Q}}}[n]\) and \({\widetilde{V}}=V\). Substituting the initial data in (26), we find

$$\begin{aligned} s(n)&\; = \; - \bigl (1-\Lambda ^{-1}\bigr )^{-1} \log g(n) \nonumber \\&\; = \; - \bigl (1-\Lambda ^{-1}\bigr )^{-1} \log n\; = \;-\log \,\Gamma (n+1) \; + \;C, \end{aligned}$$
(101)

where C is a constant. Below, we fix this constant as 0.

Proposition 5

The \(\psi _A,\psi _B\) defined by

$$\begin{aligned}&\psi _A(\lambda ,n)= \sum _{j\ge 0} (-1)^j \frac{(n-2j+1)_{2j}}{2^j\, j!\, \lambda ^{2j}} \lambda ^n, \end{aligned}$$
(102)
$$\begin{aligned}&\psi _B(\lambda ,n)= \Gamma (n+1) \sum _{j\ge 0} \frac{(n+1)_{2j}}{2^j\,j!\, \lambda ^{2j}} \lambda ^{-n} \end{aligned}$$
(103)

form a particular pair of wave functions associated with \((f(n)=0,g(n)=n)\). Here and below \((a)_{i}\) denotes the increasing Pochhammer symbol defined by \((a)_{i}=a(a+1)\ldots (a+i-1)\).

Proof

It is straightforward to verify that both \(\psi _A\) and \(\psi _B\) satisfy the equation

$$\begin{aligned} \psi (\lambda ,n+1) \; + \; n\, \psi (\lambda ,n-1) \; = \; \lambda \, \psi (\lambda ,n). \end{aligned}$$
(104)

Moreover, from definitions (102)–(103), we see that

$$\begin{aligned} \psi _A \in {\widetilde{V}}\bigl ( \bigl (\lambda ^{-1}\bigr ) \bigr )\, \lambda ^n, \quad \psi _B \in {\widetilde{V}}\bigl ( \bigl (\lambda ^{-1}\bigr ) \bigr )e^{-s(n)} \lambda ^{-n}. \end{aligned}$$

We are left to show that

$$\begin{aligned} \Gamma (n)^{-1} \, \Bigl (\psi _A(\lambda ,n) \, \psi _B(\lambda ,n-1) - \psi _B(\lambda ,n) \, \psi _A(\lambda ,n-1) \Bigr ) \; = \; \lambda . \end{aligned}$$
(105)

Clearly, the meaning of this identity is the following: Both sides of (105) are Laurent series of \(\lambda ^{-1}\) with coefficients in \({\widetilde{V}}=V=Q[n]\), and the equality means all the coefficients should be equal. More precisely, the identity (105) can be equivalently written as the following sequence of identities:

$$\begin{aligned}&\frac{n}{j+1}\sum _{j_1=0}^{j+1} \frac{(-1)^{j_1}}{2} \left( {\begin{array}{c}j+1\\ j_1\end{array}}\right) \left( {\begin{array}{c}n+2j_1-1\\ 2j+1\end{array}}\right) \nonumber \\&\quad + \sum _{j_1=0}^j (-1)^{j_1} \left( {\begin{array}{c}j\\ j_1\end{array}}\right) \left( {\begin{array}{c}n+2j_1\\ 2j+1\end{array}}\right) \; = \; 0,\quad j\ge 0. \end{aligned}$$
(106)

From (64), we know that the left-hand side of (106) as a polynomial of n is a constant for any \(j\ge 0\). Note that the value of the left-hand side of (106) at \(n=0\) is obviously 0 for any \(j\ge 0\). The proposition is proved. \(\square \)

It follows from the above proposition an explicit expression for the \(D(\lambda ,\mu ,n,\mathbf{0})\) (cf. Eq. (84)) associated with the pair (102)–(103):

$$\begin{aligned} \frac{e^{s(n-1)}}{\mu }\, D(\lambda ,\mu ,n,\mathbf{0}) \, \biggl (\frac{\mu }{\lambda }\biggr )^n \; = \; \frac{1}{\lambda -\mu } \; + \; A(\lambda ,\mu ,n), \end{aligned}$$
(107)

with \(A(\lambda ,\mu ,n)\) given by

$$\begin{aligned} A(\lambda ,\mu ,n)&= \sum _{k\ge 1} \frac{(2k-1)!!}{(2k)!}\sum _{p=0}^{2k-1} (-1)^{p+[(p+1)/2]} \left( \begin{array}{c} k-1 \\ {[p/2]} \end{array}\right) \nonumber \\&\quad \qquad \cdot \prod _{j=-p}^{2k-1-p} (n+j) \, \lambda ^{-p-1} \mu ^{-(2k-p)} . \end{aligned}$$
(108)

This explicit expression (108) first appeared in [31]. Denote

$$\begin{aligned} {\widehat{A}}(\lambda ,\mu ,n)\; = \;\frac{1}{\lambda -\mu } \; + \; A(\lambda ,\mu ,n). \end{aligned}$$
(109)

As a corollary of Proposition 5, Theorem 1, and the above (107), we have achieved a new proof of the following theorem of Jian Zhou.

Theorem 4

[31] Fix \(k\ge 2\) being an integer. The generating series of k-point connected GUE correlators has the following expression:

$$\begin{aligned} \sum _{i_1,\ldots ,i_k\ge 1} \frac{\bigl \langle {\mathrm{tr}} \, M^{i_1}\ldots {\mathrm{tr}} \, M^{i_k}\bigr \rangle _{{\mathrm{c}}}}{\lambda _1^{i_1+1} \ldots \lambda _k^{i_k+1}}&\; = \; (-1)^{k-1} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{j=1}^k {\widehat{A}} \, \bigl (\lambda _{\pi (j)},\lambda _{\pi (j+1)},n\bigr ) \nonumber \\&\qquad \quad \,-\, \frac{ \delta _{k,2}}{(\lambda _1-\lambda _2)^2}, \end{aligned}$$
(110)

where \({\widehat{A}}\) is defined by (108)–(109). Here, we recall that for any fixed \(i_1,\ldots ,i_k\), the connected GUE correlator \(\langle {\mathrm{tr}} \, M^{i_1}\ldots {\mathrm{tr}} \, M^{i_k}\rangle _{{\mathrm{c}}}\) is a polynomial of n (cf. [4, 10, 17, 21]).

5.2 Application II: Gromov–Witten invariants of \({\mathbb {P}}^1\) in the stationary sector

The initial data for the Gromov–Witten solution to the Toda lattice hierarchy were, for example, derived in [10,11,12]. It has the following explicit expression:

$$\begin{aligned} f(n)\; = \;n\epsilon +\frac{\epsilon }{2}, \quad g(n)\; = \;1. \end{aligned}$$
(111)

We have

$$\begin{aligned} s(n) \; = \; - \bigl (1-\Lambda ^{-1}\bigr )^{-1} \log 1 \; = \; C, \end{aligned}$$

where C is an arbitrary constant. Below, we take \(C=0\).

Proposition 6

The \(\psi _1,\psi _2\) defined by

$$\begin{aligned}&\psi _A(\lambda ,n) \; = \; \epsilon ^{\frac{\lambda }{\epsilon }-\frac{1}{2}} \, \Gamma \Bigl (\frac{\lambda }{\epsilon }+\frac{1}{2}\Bigr )\, J_{\frac{\lambda }{\epsilon }-n-\frac{1}{2}}\Bigl (\frac{2}{\epsilon }\Bigr ), \end{aligned}$$
(112)
$$\begin{aligned}&\psi _B(\lambda ,n) \; = \; (-1)^{n+1} \epsilon ^{-\frac{\lambda }{\epsilon }-\frac{1}{2}} \, \lambda \, \Gamma \Bigl (-\frac{\lambda }{\epsilon }+\frac{1}{2}\Bigr )\, J_{-\frac{\lambda }{\epsilon }+n+\frac{1}{2}}\Bigl (\frac{2}{\epsilon }\Bigr ) \end{aligned}$$
(113)

form a particular pair of wave functions associated with \(f(n)=n\epsilon +\frac{\epsilon }{2},g(n)=1\). Here, \(J_\nu (y)\) denotes the Bessel function, and the right-hand sides of (112)–(113) are understood as the large \(\lambda \) asymptotics of the corresponding analytic functions.

Proof

Firstly, using the properties of Bessel functions, we can verify that \(\psi _A(\lambda ,n)\) and \(\psi _B(\lambda ,n)\) defined from the above asymptotics satisfy

$$\begin{aligned}&\psi _A(\lambda ,n+1)\; + \; \Bigl (n\epsilon +\frac{\epsilon }{2}\Bigr ) \, \psi _A(\lambda ,n) \; + \; \psi _A(\lambda ,n-1) \; = \; \lambda \, \psi _A(\lambda ,n),\\&\psi _B(\lambda ,n+1)\; + \; \Bigl (n\epsilon +\frac{\epsilon }{2}\Bigr ) \, \psi _B(\lambda ,n) \; + \; \psi _B(\lambda ,n-1) \; = \; \lambda \, \psi _B(\lambda ,n). \end{aligned}$$

Secondly, as \(\lambda \) goes to \(\infty \), the following asymptotics hold true:

$$\begin{aligned}&\epsilon ^{\frac{\lambda }{\epsilon }-\frac{1}{2}} \, \Gamma \Bigl (\frac{\lambda }{\epsilon }+\frac{1}{2}\Bigr )\, J_{\frac{\lambda }{\epsilon }-n-\frac{1}{2}}\Bigl (\frac{2}{\epsilon }\Bigr ) \;\sim \; \lambda ^n \Bigl (1\; + \; {\mathrm{O}}\bigl (\lambda ^{-1}\bigr )\Bigr ),\\&(-1)^{n+1} \epsilon ^{-\frac{\lambda }{\epsilon }-\frac{1}{2}} \, \lambda \, \Gamma \Bigl (-\frac{\lambda }{\epsilon }+\frac{1}{2}\Bigr )\, J_{-\frac{\lambda }{\epsilon }+n+\frac{1}{2}}\Bigl (\frac{2}{\epsilon }\Bigr ) \;\sim \; \lambda ^{-n} \Bigl (1\; + \; {\mathrm{O}}\bigl (\lambda ^{-1}\bigr )\Bigr ). \end{aligned}$$

Thirdly, \(\psi _A\) and \(\psi _B\) also satisfy

$$\begin{aligned} \psi _A(\lambda ,n) \, \psi _B(\lambda ,n-1) \,-\, \psi _B(\lambda ,n) \, \psi _A(\lambda ,n-1) \; = \; \lambda . \end{aligned}$$

We have verified all the defining properties for a pair of wave functions associated with \(f(n)=n\epsilon +\frac{\epsilon }{2},g(n)=1\). The proposition is proved. \(\square \)

Note that

$$\begin{aligned}&\psi _A(\lambda ,n-1)\; = \; \epsilon ^{\frac{\lambda }{\epsilon }-\frac{1}{2}} \, \Gamma \Bigl (\frac{\lambda }{\epsilon }+\frac{1}{2}\Bigr )\, J_{\frac{\lambda }{\epsilon }-n+\frac{1}{2}}\Bigl (\frac{2}{\epsilon }\Bigr ), \end{aligned}$$
(114)
$$\begin{aligned}&\psi _B(\lambda ,n-1)\; = \; (-1)^{n} \epsilon ^{-\frac{\lambda }{\epsilon }-\frac{1}{2}} \, \lambda \, \Gamma \Bigl (-\frac{\lambda }{\epsilon }+\frac{1}{2}\Bigr )\, J_{-\frac{\lambda }{\epsilon }+n-\frac{1}{2}}\Bigl (\frac{2}{\epsilon }\Bigr ), \end{aligned}$$
(115)

and denote

$$\begin{aligned} J_\nu (y) \,=:\, \frac{(y/2)^\nu }{\Gamma (\nu +1)} j_{\nu +\frac{1}{2}}(y^2/4). \end{aligned}$$

It follows from (84), (112)–(115)  that the \(D(\lambda ,\mu ,0,\mathbf{0})\) associated with the pair (112)–(113) has the following explicit expression:

$$\begin{aligned} \frac{1}{\mu }D(\lambda ,\mu ,0,\mathbf{0}) \; = \; - \frac{1}{\epsilon }\, \frac{j_{-\frac{\mu }{\epsilon }}\bigl (\frac{1}{\epsilon ^2}\bigr ) \, j_{\frac{\lambda }{\epsilon }}\bigl (\frac{1}{\epsilon ^2}\bigr ) + \frac{\epsilon ^{-2}}{(\frac{1}{2}-\frac{\mu }{\epsilon })(\frac{1}{2}+\frac{\lambda }{\epsilon })} \, j_{1-\frac{\mu }{\epsilon }}\bigl (\frac{1}{\epsilon ^2}\bigr )\, j_{1+\frac{\lambda }{\epsilon }}\bigl (\frac{1}{\epsilon ^2}\bigr )}{\mu /\epsilon -\lambda /\epsilon }. \end{aligned}$$

Then, according to [12], the function \(\frac{1}{\mu }D(\lambda ,\mu ,0,\mathbf{0})\) has the following expressions:

$$\begin{aligned}&\frac{1}{\mu }D(\lambda ,\mu ,0,\mathbf{0}) \nonumber \\&\quad \; = \; -\frac{1}{\epsilon }\sum _{k=0}^\infty \frac{(a-b-2k+1)_{k-1}}{k! \, (-a+\tfrac{1}{2})_k \, (b+\tfrac{1}{2})_k} \epsilon ^{-2k} \end{aligned}$$
(116)
$$\begin{aligned}&\quad \; = \; \frac{-1}{\epsilon (a-b)} ~ {}_2 F_3\Bigl (\frac{b-a}{2}, \frac{b-a+1}{2}; \, \frac{1}{2} -a \,, \, \frac{1}{2} + b, b-a+1; \, -4 \epsilon ^{-2}\Bigr ) \end{aligned}$$
(117)
$$\begin{aligned}&\quad \ \sim \; \frac{-1}{\epsilon (a-b)} \,-\, \sum _{p,q\ge 0} \frac{(-1)^{q+1}}{a^{p+1} b^{q+1}} \sum _{k\ge 1} \frac{\epsilon ^{-2k-1}}{k!} \nonumber \\&\quad \qquad \quad \sum _{1\le i,j\le k} (-1)^{i+j}\frac{(i+j-2k)_{k-1} \bigl (i-\frac{1}{2}\bigr )^p \bigl (j-\frac{1}{2}\bigr )^q}{(i-1)!(j-1)!(k-i)!(k-j)!} \;=:\; {\widehat{A}} \, (\lambda ,\mu ), \end{aligned}$$
(118)

where \(a:=\frac{\mu }{\epsilon }\), \(b:=\frac{\lambda }{\epsilon }\), the \((a-b+1)_{-1}\) of (116) is defined as \(1/(a-b)\), and \(\sim \) in (118) is taken as \(a,b\rightarrow \infty \) away from the half integers. The explicit expression (118) first appeared in [12]. So we have completed a new proof of the following theorem.

Theorem 5

[12] The generating series of k-point (\(k\ge 2\)) Gromov–Witten invariants of \({\mathbb {P}}^1\) in the stationary sector has the following explicit expression:

$$\begin{aligned}&\epsilon ^k \! \sum _{i_1,\ldots ,i_k\ge 0} \frac{(i_1+1)! \ldots (i_k+1)!}{\lambda _1^{i_1+2} \ldots \lambda _k^{i_k+2}} \langle \tau _{i_1}(\omega ) \ldots \tau _{i_k}(\omega ) \rangle (\epsilon ) \nonumber \\&\quad \; = \; (-1)^{k-1} \sum _{\pi \in {\mathcal {S}}_k/C_k} \prod _{i=1}^k {\widehat{A}} \, \bigl (\lambda _{\pi (i)}, \lambda _{\pi (i+1)}\bigr ) \,-\, \frac{\delta _{k,2}}{(\lambda _1-\lambda _2)^2}, \end{aligned}$$
(119)

where \({\widehat{A}} \, (\lambda ,\mu )\) is explicitly defined in (118), and

$$\begin{aligned} \langle \tau _{i_1}(\omega ) \ldots \tau _{i_k}(\omega ) \rangle (\epsilon ) := \sum _{g\ge 0} \epsilon ^{2g-2} \sum _{d\ge 0} \int _{\left[ \overline{{\mathcal {M}}}_{g,k}({\mathbb {P}}^1,d)\right] ^{{\mathrm{virt}}}} {\mathrm{ev}}_1^*(\omega ) \ldots {\mathrm{ev}}_k^*(\omega ) \, \psi _1^{i_1} \ldots \psi _k^{i_k}. \end{aligned}$$
(120)

(See, for example, [12] for the notation about the integral in the right-hand side of (120).)