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

Principal Hierarchies for Frobenius Manifolds with Rational and Trigonometric Superpotentials

Shilin Ma Shilin Ma, School of Mathematics and Statistics, Xi’an Jiaotong University, Xi’an 710049, P. R. China, [email protected]
Abstract.

In this paper, we construct the principal hierarchies for Frobenius manifolds with rational and trigonometric superpotentials, as well as their almost dualities. We demonstrate that in both cases, submanifolds with even superpotentials form natural Frobenius submanifolds, and their principal hierarchies can be obtained as restrictions of the principal hierarchies for the original Frobenius manifolds. Furthermore, we introduce a natural rank-1 extension for each of these Frobenius manifolds, providing solutions to the associated open WDVV equations. The principal hierarchy for each extension is also explicitly constructed.

1. Introduction

The concept of Frobenius manifolds, first introduced by Dubrovin in [1], provides a geometric framework for capturing the associativity equations inherent in two-dimensional topological field theory (2D TFT). This concept is highly relevant across various areas of mathematical physics, including Gromov-Witten theory, singularity theory, and integrable systems, among others. Its significance is further highlighted by applications in works such as [2, 3, 4, 5, 6, 7] and their references.

Every Frobenius manifold is associated with an integrable hierarchy of hydrodynamic type, referred to as the principal hierarchy for the Frobenius manifold. This hierarchy involves unknown functions depending on a single scalar spatial variable and various time variables. In the case of semisimple Frobenius manifold, this hierarchy can be deformed into a dispersive hierarchy, referred to as the Dubrovin-Zhang hierarchy. The tau function, determined by the string equation of the Dubrovin-Zhang hierarchy, provides the partition function for the corresponding 2D TFT. This intricate relationship between Frobenius manifold and 2D TFT has significantly advanced our understanding of their geometric and algebraic structures. For further details, see [4, 8, 9].

Although the abstract theory of Dubrovin and Zhang is well-established, explicitly constructing the principal hierarchy for a given Frobenius manifold presents a certain level of difficulty. For relatively low-dimensional Frobenius manifold, one can directly derive the principal hierarchy by solving the PDE systems that govern its Hamiltonian densities. Examples of this process are provided in [4, 10, 11]. For higher-dimensional cases, although constructions have been provided, as in [12, 13, 14], these rely on the specific structures of the corresponding Frobenius manifolds and lack general applicability.

The main result of this paper is the explicit construction of the principal hierarchies for Frobenius manifolds with rational and trigonometric superpotentials, respectively. Our approach is as follows: we first provide suitable representations for the cotangent spaces and derive the explicit formulas for the Hamiltonian structures associated with the flat metrics for these Frobenius manifolds. Then, we reformulate the PDE systems governing the Hamiltonian densities of the principal hierarchies into algebraic equations involving the superpotentials, which admit straightforward solutions. Using a similar approach, we have also constructed the principal hierarchies for the almost dualities [15] of these Frobenius manifolds.

Another important class of Frobenius manifolds consists of those with even rational and trigonometric superpotentials. We will show that these manifolds are natural Frobenius submanifolds [16, 17] of the two types of Frobenius manifolds discussed above, and that their associated principal hierarchies are direct restrictions of those of the parent manifolds.

The open WDVV equations were first introduced in [18] in the context of open Gromov-Witten theory. P. Rossi observed that a solution to the open WDVV equations is equivalent to a flat F-manifold, which serves as a rank-1 extension of the given Frobenius manifold. This extension was systematically studied in [19]. The descendant potential and Virasoro constraints for a flat F-manifold were constructed in [20] for genus-zero cases and in [21] for higher genera. The dispersive deformation of the principal hierarchy for a flat F-manifold, as a generalization of the DR hierarchy for a Frobenius manifold, was studied in [22]. Examples related to the open Gromov-Witten theory of a point and open r-spin theory were investigated in [23, 24, 25] and [26, 27], respectively. In the present paper, we show that there exist natural rank-1 extensions for Frobenius manifolds with rational and trigonometric superpotentials, and explicitly construct the principal hierarchies for these extensions.

We note that the construction presented in this paper has a certain level of generality. For instance, in [28], we employed a similar approach to construct the principal hierarchy for the infinite-dimensional Frobenius manifold underlying the genus-zero universal Whitham hierarchy.

Let us now state our main results precisely.

Given positive integers n0,,nmn_{0},\cdots,n_{m}, let McKPM^{cKP} denote the space of rational functions of the form:

λ(z)=1n0zn0+a0,n02zn02++a0,0+i=1mj=1niai,j(zai,0)j,\lambda(z)=\frac{1}{n_{0}}z^{n_{0}}+a_{0,n_{0}-2}z^{n_{0}-2}+\cdots+a_{0,0}+\sum_{i=1}^{m}\sum_{j=1}^{n_{i}}a_{i,j}(z-a_{i,0})^{-j}, (1.1)

where the set of coefficients {a0,i}i=0n02{a1,i}i=0n1{am,i}i=0nm\{a_{0,i}\}_{i=0}^{n_{0}-2}\cup\{a_{1,i}\}_{i=0}^{n_{1}}\cup\cdots\cup\{a_{m,i}\}_{i=0}^{n_{m}} serves as local coordinates on the manifold McKPM^{cKP}. The space McKPM^{cKP} can be regarded as a special type of Hurwitz space and is thus equipped with a Frobenius manifold structure constructed by Dubrovin [1]. When m=0m=0, this structure coincides with that on the orbit space of the Coxeter group of type AA [29].

The flat coordinates on McKPM^{cKP}, denoted as 𝐭cKP={t0,j}j=1n01{t1,j}j=0n1{tm,j}j=0nm\mathbf{t}^{cKP}=\{t_{0,j}\}_{j=1}^{n_{0}-1}\cup\{t_{1,j}\}_{j=0}^{n_{1}}\cup\cdots\cup\{t_{m,j}\}_{j=0}^{n_{m}}, are given by the following expansions:

z={ti,0+ti,1wi(z)1+,as zai,0,i=1,,m,w0(z)t0,1w0(z)1t0,2w0(z)2+,as z,i=0,z=\begin{cases}t_{i,0}+t_{i,1}w_{i}(z)^{-1}+\cdots,&\text{as }z\to a_{i,0},\ i=1,\cdots,m,\\ w_{0}(z)-t_{0,1}w_{0}(z)^{-1}-t_{0,2}w_{0}(z)^{-2}+\cdots,&\text{as }z\to\infty,\ i=0,\end{cases}

where

wi(z)={(niλ(z))1ni=wi,1(zai,0)1+,as zai,0,i=1,,m,(n0λ(z))1n0=z+w0,1z1+,as z,i=0.w_{i}(z)=\begin{cases}(n_{i}\lambda(z))^{\frac{1}{n_{i}}}=w_{i,1}(z-a_{i,0})^{-1}+\cdots,&\text{as }z\to a_{i,0},\ i=1,\cdots,m,\\ (n_{0}\lambda(z))^{\frac{1}{n_{0}}}=z+w_{0,1}z^{-1}+\cdots,&\text{as }z\to\infty,\ i=0.\end{cases}
Theorem 1.1.

The Hamiltonian densities of the principal hierarchy for the Frobenius manifold McKPM^{cKP} are given by

θt0,j,p(z)=\displaystyle\theta_{t_{0,j},p}(z)= Resc0,j;pw0(z)(p+1)n0jdz,j=1,,n01,\displaystyle-\mathop{\text{\rm Res}}_{\infty}c_{0,j;p}w_{0}(z)^{(p+1)n_{0}-j}\,dz,\quad j=1,\ldots,n_{0}-1,
θti,j,p(z)=\displaystyle\theta_{t_{i,j},p}(z)= Resai,0ci,j;pwi(z)(p+1)nijdz,i=1,,m,j=0,,ni1,\displaystyle\mathop{\text{\rm Res}}_{a_{i,0}}c_{i,j;p}w_{i}(z)^{(p+1)n_{i}-j}\,dz,\quad i=1,\ldots,m,\ j=0,\ldots,n_{i}-1,

and

θti,ni,p(z)=\displaystyle\theta_{t_{i,n_{i}},p}(z)= Resλ(z)pp!(logw0(z)zai,0cpn0)dz+Resai,0λ(z)pp!(log(zai,0)wi(z)cpni)dz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\frac{\lambda(z)^{p}}{p!}(\log\frac{w_{0}(z)}{z-a_{i,0}}-\frac{c_{p}}{n_{0}})\,dz+\mathop{\text{\rm Res}}_{a_{i,0}}\frac{\lambda(z)^{p}}{p!}(\log(z-a_{i,0})w_{i}(z)-\frac{c_{p}}{n_{i}})\,dz
+siResas,0λ(z)pp!log(zai,0)dz,i=1,,m,\displaystyle+\sum_{s\neq i}\mathop{\text{\rm Res}}_{a_{s,0}}\frac{\lambda(z)^{p}}{p!}\log(z-a_{i,0})\,dz,\quad i=1,\ldots,m,

where the constants {ci,j;p}\{c_{i,j;p}\} and {cp}\{c_{p}\} are defined as

ci,j;p=1nij12nij1(p+1)nij,cp=s=1p1s.c_{i,j;p}=\frac{1}{n_{i}-j}\frac{1}{2n_{i}-j}\cdots\frac{1}{(p+1)n_{i}-j},\quad c_{p}=\sum_{s=1}^{p}\frac{1}{s}.

The corresponding Hamiltonian vector fields T,p=𝒫(dθ,p+1)\frac{\partial}{\partial T^{\bullet,p}}=\mathcal{P}(d\theta_{\bullet,p+1}) are given by

λ(z)Tt0,j,p1=\displaystyle\frac{\partial\lambda(z)}{\partial T^{t_{0,j},p-1}}= {(c0,j;p1w0(z)pn0j),0,λ(z)},j=1,,n01,\displaystyle\{(c_{0,j;p-1}w_{0}(z)^{pn_{0}-j})_{\infty,\geq 0},\lambda(z)\},\quad j=1,\ldots,n_{0}-1,
λ(z)Tti,j,p1=\displaystyle\frac{\partial\lambda(z)}{\partial T^{t_{i,j},p-1}}= {(ci,j;p1wi(z)pnij)ai,0,1,λ(z)},i=1,,m,j=0,,ni1,\displaystyle-\{(c_{i,j;p-1}w_{i}(z)^{pn_{i}-j})_{a_{i,0},\leq-1},\lambda(z)\},\quad i=1,\ldots,m,\ j=0,\ldots,n_{i}-1,

and

λ(z)Tti,ni,p1=\displaystyle\frac{\partial\lambda(z)}{\partial T^{t_{i,n_{i}},p-1}}= {(λ(z)pp!(logw0(z)zai,0cpn0)),0(λ(z)pp!(log(zai,0)wi(z)cpni))ai,0,1,λ(z)}\displaystyle\{(\frac{\lambda(z)^{p}}{p!}(\log\frac{w_{0}(z)}{z-a_{i,0}}-\frac{c_{p}}{n_{0}}))_{\infty,\geq 0}-(\frac{\lambda(z)^{p}}{p!}(\log(z-a_{i,0})w_{i}(z)-\frac{c_{p}}{n_{i}}))_{a_{i,0},\leq-1},\lambda(z)\}
si{(λ(z)pp!log(zai,0))as,0,1,λ(z)}+{λ(z)pp!log(zai,0),λ(z)},\displaystyle-\sum_{s\neq i}\{(\frac{\lambda(z)^{p}}{p!}\log(z-a_{i,0}))_{a_{s,0},\leq-1},\lambda(z)\}+\{\frac{\lambda(z)^{p}}{p!}\log(z-a_{i,0}),\lambda(z)\},

where the Poisson bracket is defined as

{f(z,x),g(z,x)}=f(z,x)zg(z,x)xf(z,x)xg(z,x)z.\{f(z,x),g(z,x)\}=\frac{\partial f(z,x)}{\partial z}\frac{\partial g(z,x)}{\partial x}-\frac{\partial f(z,x)}{\partial x}\frac{\partial g(z,x)}{\partial z}.

When m=1m=1, this principal hierarchy was constructed by Aoyama and Kodama using a different method [12], and serves as an extension of the dispersionless limit of the constrained KP hierarchy [30]. The corresponding Dubrovin-Zhang hierarchy governs the generating function enumerating rooted hypermaps on compact two-dimensional surfaces [31].

A similar approach can be used to construct the principal hierarchy for the almost duality of McKPM^{cKP}.

Theorem 1.2.

Let M^cKP\hat{M}^{cKP} be the almost duality of the Frobenius manifold McKPM^{cKP}, then the Hamiltonian densities of the principal hierarchy for M^cKP\hat{M}^{cKP} are given by

Fγ~,p=12πiγ~(log(λ(z)))p+1(p+1)!𝑑z,F_{\tilde{\gamma},p}=\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}\frac{(\log(\lambda(z)))^{p+1}}{(p+1)!}\,dz,

where γ~\tilde{\gamma} is a simple closed curve in the complex plane \mathbb{C} such that the winding number of λ(z)\lambda(z) along γ~\tilde{\gamma} is zero.

In particular, let q1,,qkq_{1},\ldots,q_{k} and p1,,pkp_{1},\ldots,p_{k} be the zeros and poles of λ(z)\lambda(z), respectively, within the region enclosed by γ~\tilde{\gamma}. Then

Fγ~,0=s=1kpss=1kqs.F_{\tilde{\gamma},0}=\sum_{s=1}^{k}p_{s}-\sum_{s=1}^{k}q_{s}.

In the case of m=0m=0, the Hamiltonian densities of this principal hierarchy were constructed by Dubrovin using period integrals [15].

As a direct consequence of Theorem 1.1, we obtain the explicit form of the principal hierarchy for the Frobenius submanifold [16] of McKPM^{cKP} with even superpotential. Assume there exist positive integers mm^{\prime} and n0,n1,,nmn_{0}^{\prime},n_{1}^{\prime},\cdots,n_{m^{\prime}}^{\prime} such that

m=2m1,n0=2n0,n1=2n1,n2j2=n2j1=nj,2jm.m=2m^{\prime}-1,\quad n_{0}=2n_{0}^{\prime},\quad n_{1}=2n_{1}^{\prime},\quad n_{2j-2}=n_{2j-1}=n_{j}^{\prime},\quad 2\leq j\leq m^{\prime}.

Let MDcKPM^{D-cKP} be a submanifold of McKPM^{cKP} consisting of elements of the form:

λ(z)=12n0z2n0+j=0n01b0,jz2j+j=1n1b1,jz2j+i=2mj=1nibi,j(z2bi,0)j,\lambda(z)=\frac{1}{2n_{0}^{\prime}}z^{2n_{0}^{\prime}}+\sum_{j=0}^{n_{0}^{\prime}-1}b_{0,j}z^{2j}+\sum_{j=1}^{n_{1}^{\prime}}b_{1,j}z^{-2j}+\sum_{i=2}^{m^{\prime}}\sum_{j=1}^{n_{i}^{\prime}}b_{i,j}(z^{2}-b_{i,0})^{-j}, (1.2)

which is characterized by the condition λ(z)=λ(z)\lambda(z)=\lambda(-z).

Theorem 1.3.

MDcKPM^{D-cKP} is a natural Frobenius submanifold of McKPM^{cKP}, characterized by the following conditions imposed on the flat coordinates of McKPM^{cKP}:

t0,2=t0,4==t0,2n02=0;\displaystyle t_{0,2}=t_{0,4}=\cdots=t_{0,2n_{0}^{\prime}-2}=0;
t1,0=t1,2==t1,2n1=0;\displaystyle t_{1,0}=t_{1,2}=\cdots=t_{1,2n_{1}^{\prime}}=0;
t2i2,j=t2i1,j,2im, 0jni.\displaystyle t_{2i-2,j}=-t_{2i-1,j},\quad 2\leq i\leq m^{\prime},\ 0\leq j\leq n_{i}^{\prime}.

Moreover, the subhierarchy Tt,p\frac{\partial}{\partial T^{t,p}} of the principal hierarchy for McKPM^{cKP}, where

ttDcKP={t0,2j1}j=1n0{t1,2j1}j=1n1{t2,j}j=0n2{t4,j}j=0n3{t2m2,j}j=0nm,t\in\textbf{t}^{D-cKP}=\{t_{0,2j-1}\}_{j=1}^{n_{0}^{\prime}}\cup\{t_{1,2j-1}\}_{j=1}^{n_{1}^{\prime}}\cup\{t_{2,j}\}_{j=0}^{n_{2}^{\prime}}\cup\{t_{4,j}\}_{j=0}^{n_{3}^{\prime}}\cup\cdots\cup\{t_{2m^{\prime}-2,j}\}_{j=0}^{n_{m^{\prime}}^{\prime}},

can be directly restricted to the submanifold MDcKPM^{D-cKP}, forming the principal hierarchy for MDcKPM^{D-cKP}.

when m=1m^{\prime}=1 and n1=1n_{1}^{\prime}=1, this Frobenius manifold structure coincides with that on the orbit space of the Coxeter group of type DD [32], and the associated principal hierarchy is the dispersionless limit of the Drinfeld-Sokolov hierarchy of type DD [33].

Owing to Theorem 3 in [19], we obtain a solution to the open WDVV equations associated with McKPM^{cKP}, or equivalently, a flat F-manifold structure on McKP×M^{cKP}\times\mathbb{C} which is a rank-1 extension of McKPM^{cKP}. The multiplication \star on this flat F-manifold is given by

(α,0)(β,0)=(αβ,αβΩs)\displaystyle(\partial_{\alpha},0)\star(\partial_{\beta},0)=(\partial_{\alpha}\circ\partial_{\beta},\partial_{\alpha}\partial_{\beta}\Omega\cdot\partial_{s})
(α,0)(0,s)=(0,αλ(s)s)\displaystyle(\partial_{\alpha},0)\star(0,\partial_{s})=(0,\partial_{\alpha}\lambda(s)\cdot\partial_{s})
(0,s)(0,s)=(0,λ(s)s)\displaystyle(0,\partial_{s})\star(0,\partial_{s})=(0,\lambda^{\prime}(s)\cdot\partial_{s})

where α=tα\partial_{\alpha}=\frac{\partial}{\partial t^{\alpha}} for tαtcKPt^{\alpha}\in\textbf{t}^{cKP}, \circ represents the multiplication on McKPM^{cKP}, ss is the coordinate on \mathbb{C}, and

αβΩ=(αλ(s)βλ(s)λ(s)),0+j=1m(αλ(s)βλ(s)λ(s))φj,1.\partial_{\alpha}\partial_{\beta}\Omega=(\frac{\partial_{\alpha}\lambda(s)\partial_{\beta}\lambda(s)}{\lambda^{\prime}(s)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{\alpha}\lambda(s)\partial_{\beta}\lambda(s)}{\lambda^{\prime}(s)})_{\varphi_{j},\leq-1}.
Corollary 1.4.

The principal hierarchy for the flat F-manifold McKP×M^{cKP}\times\mathbb{C} is given by

λ(z)T~s,p=0,s(x)T~s,p=dx(λ(s)p+1(p+1)!),\frac{\partial\lambda(z)}{\partial\tilde{T}^{s,p}}=0,\quad\frac{\partial s(x)}{\partial\tilde{T}^{s,p}}=d_{x}(\frac{\lambda(s)^{p+1}}{(p+1)!}),

and

λ(z)T~t0,n0j,p=λ(z)Tt0,j,p,s(x)T~t0,n0j,p=dx(dθt0,j,p+1(s))+,\displaystyle\frac{\partial\lambda(z)}{\partial\tilde{T}^{t_{0,n_{0}-j},p}}=\frac{\partial\lambda(z)}{\partial T^{t_{0,j},p}},\quad\frac{\partial s(x)}{\partial\tilde{T}^{t_{0,n_{0}-j},p}}=d_{x}(d\theta_{t_{0,j},p+1}(s))_{+},
λ(z)T~ti,nij,p=λ(z)Tti,j,p,s(x)T~t0,nij,p=dx(dθti,j,p+1(s)),\displaystyle\frac{\partial\lambda(z)}{\partial\tilde{T}^{t_{i,n_{i}-j},p}}=\frac{\partial\lambda(z)}{\partial T^{t_{i,j},p}},\quad\frac{\partial s(x)}{\partial\tilde{T}^{t_{0,n_{i}-j},p}}=-d_{x}(d\theta_{t_{i,j},p+1}(s))_{-},

where

dxλ(s)=xλ(s)+λ(s)s(x)x,d_{x}\lambda(s)=\partial_{x}\lambda(s)+\lambda^{\prime}(s)\frac{\partial s(x)}{\partial x},

dθti,j,p(z)d\theta_{t_{i,j},p}(z) is a certain function on λ(z)\lambda(z), and the operators ()+,()(\ )_{+},(\ )_{-} are defined in Subsection 3.2.

When m=0m=0, this hierarchy coincides with the dispersionless limit of the open Gelfand-Dickey hierarchy, which is conjectured to govern the generating function of the open r-spin intersection numbers [27].

Let us proceed to consider the Frobenius manifold with trigonometric superpotential. Let MTodaM^{Toda} be the space of functions of the form:

λ(φ)=1n0en0φ+a0,n01e(n01)φ++a0,0+i=1mj=1niai,j(eφai,0)j,\lambda(\varphi)=\frac{1}{n_{0}}e^{n_{0}\varphi}+a_{0,n_{0}-1}e^{(n_{0}-1)\varphi}+\cdots+a_{0,0}+\sum_{i=1}^{m}\sum_{j=1}^{n_{i}}a_{i,j}(e^{\varphi}-a_{i,0})^{-j},

with a1,0=0a_{1,0}=0, equipped with the Hurwitz Frobenius manifold structure constructed by Dubrovin [1]. In the case of m=1m=1, this structure coincides with that on the orbit space of the extended affine Weyl group of type AA [34].

The flat coordinate system on MTodaM^{Toda}, denoted as

𝐭Toda={t0,j}j=1n01{t1,j}j=0n1{tm,j}j=0nm,\mathbf{t}^{Toda}=\{t_{0,j}\}_{j=1}^{n_{0}-1}\cup\{t_{1,j}\}_{j=0}^{n_{1}}\cup\cdots\cup\{t_{m,j}\}_{j=0}^{n_{m}},

is given by

φ={ti,0+ti,1wi(φ)1+,eφai,0,i=2,,m,log(w1(φ))+t1,0+t1,1w1(φ)1+,eφ0,log(w0(φ))t0,1w0(φ)1t0,2w0(φ)2,eφ,\varphi=\left\{\begin{aligned} &t_{i,0}+t_{i,1}w_{i}(\varphi)^{-1}+\cdots,\quad&e^{\varphi}&\to a_{i,0},\ i=2,\cdots,m,\\ &-\log(w_{1}(\varphi))+t_{1,0}+t_{1,1}w_{1}(\varphi)^{-1}+\cdots,\quad&e^{\varphi}&\to 0,\\ &\log(w_{0}(\varphi))-t_{0,1}w_{0}(\varphi)^{-1}-t_{0,2}w_{0}(\varphi)^{-2}-\cdots,\quad&e^{\varphi}&\to\infty,\end{aligned}\right.

where

wi(φ)={(niλ(φ))1ni=wi,1(eφai,0)1+,eφai,0,i=1,,m,(n0λ(φ))1n0=eφ+w0,0+w0,1eφ+,eφ,i=0.w_{i}(\varphi)=\left\{\begin{aligned} &(n_{i}\lambda(\varphi))^{\frac{1}{n_{i}}}=w_{i,1}(e^{\varphi}-a_{i,0})^{-1}+\cdots,\quad&e^{\varphi}&\to a_{i,0},\ i=1,\cdots,m,\\ &(n_{0}\lambda(\varphi))^{\frac{1}{n_{0}}}=e^{\varphi}+w_{0,0}+w_{0,1}e^{-\varphi}+\cdots,\quad&e^{\varphi}&\to\infty,\ i=0.\end{aligned}\right.
Theorem 1.5.

The Hamiltonian densities of the principal hierarchy for the Frobenius manifold MTodaM^{Toda} are given by

θt0,j,p(z)=\displaystyle\theta_{t_{0,j},p}(z)= Resc0,j;pw0(z)(p+1)n0jdzz,j=1,,n01;\displaystyle-\mathop{\text{\rm Res}}_{\infty}c_{0,j;p}w_{0}(z)^{(p+1)n_{0}-j}\frac{dz}{z},\quad j=1,\ldots,n_{0}-1;
θti,j,p(z)=\displaystyle\theta_{t_{i,j},p}(z)= Resai,0ci,j;pwi(z)(p+1)nijdzz,i=1,,m,j=0,,ni1,\displaystyle\mathop{\text{\rm Res}}_{a_{i,0}}c_{i,j;p}w_{i}(z)^{(p+1)n_{i}-j}\frac{dz}{z},\quad i=1,\ldots,m,\ j=0,\ldots,n_{i}-1,

and

θti,ni,p(z)=\displaystyle\theta_{t_{i,n_{i}},p}(z)= Resλ(z)pp!(logw0(z)zai,0cpn0)dzz+Resai,0λ(z)pp!(log(zai,0)wi(z)cpni)dzz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\frac{\lambda(z)^{p}}{p!}(\log\frac{w_{0}(z)}{z-a_{i,0}}-\frac{c_{p}}{n_{0}})\frac{dz}{z}+\mathop{\text{\rm Res}}_{a_{i,0}}\frac{\lambda(z)^{p}}{p!}(\log(z-a_{i,0})w_{i}(z)-\frac{c_{p}}{n_{i}})\frac{dz}{z}
+siResas,0λ(z)pp!log(zai,0)dzz,i=1,,m,\displaystyle+\sum_{s\neq i}\mathop{\text{\rm Res}}_{a_{s,0}}\frac{\lambda(z)^{p}}{p!}\log(z-a_{i,0})\frac{dz}{z},\quad i=1,\ldots,m,

where z=eφz=e^{\varphi}. The corresponding Hamiltonian vector fields T,p=𝒫(dθ,p+1)\frac{\partial}{\partial T^{\bullet,p}}=\mathcal{P}(d\theta_{\bullet,p+1}) are

λ(z)Tt0,j,p1=\displaystyle\frac{\partial\lambda(z)}{\partial T^{t_{0,j},p-1}}= {(c0,j;p1w0(z)pn0j),0,λ(z)},j=1,,n01;\displaystyle\{(c_{0,j;p-1}w_{0}(z)^{pn_{0}-j})_{\infty,\geq 0},\lambda(z)\},\quad j=1,\ldots,n_{0}-1;
λ(z)Tti,j,p1=\displaystyle\frac{\partial\lambda(z)}{\partial T^{t_{i,j},p-1}}= {(ci,j;p1wi(z)pnij)ai,0,1,λ(z)},i=1,,m,j=0,,ni1,\displaystyle-\{(c_{i,j;p-1}w_{i}(z)^{pn_{i}-j})_{a_{i,0},\leq-1},\lambda(z)\},\quad i=1,\ldots,m,\ j=0,\ldots,n_{i}-1,

and

λ(z)Tti,ni,p1=\displaystyle\frac{\partial\lambda(z)}{\partial T^{t_{i,n_{i}},p-1}}= {(λ(z)pp!(logw0(z)zai,0cpn0)),0(λ(z)pp!(log(zai,0)wi(z)cpni))ai,0,1,λ(z)}\displaystyle\{(\frac{\lambda(z)^{p}}{p!}(\log\frac{w_{0}(z)}{z-a_{i,0}}-\frac{c_{p}}{n_{0}}))_{\infty,\geq 0}-(\frac{\lambda(z)^{p}}{p!}(\log(z-a_{i,0})w_{i}(z)-\frac{c_{p}}{n_{i}}))_{a_{i,0},\leq-1},\lambda(z)\}
si{(λ(z)pp!log(zai,0))as,0,1,λ(z)}+{λ(z)pp!log(zai,0),λ(z)},\displaystyle-\sum_{s\neq i}\{(\frac{\lambda(z)^{p}}{p!}\log(z-a_{i,0}))_{a_{s,0},\leq-1},\lambda(z)\}+\{\frac{\lambda(z)^{p}}{p!}\log(z-a_{i,0}),\lambda(z)\},

where the Poisson bracket is defined as

{f(z,x),g(z,x)}=zf(z,x)zg(z,x)xzf(z,x)xg(z,x)z.\{f(z,x),g(z,x)\}=z\frac{\partial f(z,x)}{\partial z}\frac{\partial g(z,x)}{\partial x}-z\frac{\partial f(z,x)}{\partial x}\frac{\partial g(z,x)}{\partial z}.

When m=1m=1 and n0=n1=1n_{0}=n_{1}=1, this principal hierarchy coincides with the dispersionless limit of the extended Toda hierarchy [35], which governs the generating function of the Gromov-Witten invariants of 1\mathbb{CP}^{1} [36].

Theorem 1.6.

Let M^Toda\hat{M}^{Toda} be the almost duality of the Frobenius manifold MTodaM^{Toda}, then the Hamiltonian densities of the principal hierarchy for M^Toda\hat{M}^{Toda} are given by

Fγ~,p=12πiγ~(log(λ(z)))p+1(p+1)!dzz,F_{\tilde{\gamma},p}=\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}\frac{(\log(\lambda(z)))^{p+1}}{(p+1)!}\frac{dz}{z}, (1.3)

where γ~\tilde{\gamma} is a simple closed curve in the complex plane \mathbb{C}, such that the winding number of λ(z)\lambda(z) along γ~\tilde{\gamma} is zero.

In particular, let q1,,qkq_{1},\ldots,q_{k} and p1,,pkp_{1},\ldots,p_{k} be the zeros and poles of λ(z)\lambda(z), respectively, within the region enclosed by γ~\tilde{\gamma}. Additionally, let qk+1,,qn,pk+1,,prq_{k+1},\ldots,q_{n},p_{k+1},\ldots,p_{r} be the zeros and poles outside this region. Then we have

Fγ~,0={s=1klog(ps)s=1klog(qs),if 0 is outside the region enclosed by γ~,s=k+1nlog(ps)s=k+1rlog(qs),if 0 is inside the region enclosed by γ~.F_{\tilde{\gamma},0}=\left\{\begin{aligned} &\sum_{s=1}^{k}\log(p_{s})-\sum_{s=1}^{k}\log(q_{s}),\quad&\text{if $0\in\mathbb{C}$ is outside the region enclosed by $\tilde{\gamma}$,}\\ &\sum_{s=k+1}^{n}\log(p_{s})-\sum_{s=k+1}^{r}\log(q_{s}),\quad&\text{if $0\in\mathbb{C}$ is inside the region enclosed by $\tilde{\gamma}$.}\end{aligned}\right.

Let’s consider the submanifold of MTodaM^{Toda} with even superpotential. Assume there exist positive integers m,n0,,nmm^{\prime},n_{0}^{\prime},\ldots,n_{m^{\prime}}^{\prime} such that

n0=n1=n0,n2=2n1,n3=2n2,n2j2=n2j1=nj,m=2m1,n_{0}=n_{1}=n_{0}^{\prime},\quad n_{2}=2n_{1}^{\prime},\quad n_{3}=2n_{2}^{\prime},\quad n_{2j-2}=n_{2j-1}=n_{j}^{\prime},\quad m=2m^{\prime}-1,

where j=3,4,,mj=3,4,\ldots,m^{\prime}. Denote p=c1n0(1n0)1n0zp=c^{-\frac{1}{n_{0}}}(\frac{1}{n_{0}})^{-\frac{1}{n_{0}}}z, where c2n0=a1,n1c^{2}n_{0}=a_{1,n_{1}}, then the elements in MTodaM^{Toda} can be expressed as:

λ(p)=j=0n0b0,jpj+j=1n0b1,jpj+i=1mj=1nibi,j(pbi,0)j,\lambda(p)=\sum_{j=0}^{n_{0}}b_{0,j}p^{j}+\sum_{j=1}^{n_{0}}b_{1,j}p^{-j}+\sum_{i=1}^{m}\sum_{j=1}^{n_{i}}b_{i,j}(p-b_{i,0})^{-j},

where b1,n0=b0,n0b_{1,n_{0}}=b_{0,n_{0}}. Let MCTodaM^{C-Toda} be the space of elements in MTodaM^{Toda} of the following form:

λ(p)=j=0n0b~0,j(p+1p)j+j=1n1b~1,j(p+1p2)j+j=1n2b~2,j(p+1p+2)j+i=3mj=1nib~i,j(p+1pb~i,0)j,\lambda(p)=\sum_{j=0}^{n_{0}^{\prime}}\tilde{b}_{0,j}(p+\frac{1}{p})^{j}+\sum_{j=1}^{n_{1}^{\prime}}\tilde{b}_{1,j}(p+\frac{1}{p}-2)^{-j}+\sum_{j=1}^{n_{2}^{\prime}}\tilde{b}_{2,j}(p+\frac{1}{p}+2)^{-j}+\sum_{i=3}^{m^{\prime}}\sum_{j=1}^{n_{i}^{\prime}}\tilde{b}_{i,j}(p+\frac{1}{p}-\tilde{b}_{i,0})^{-j}, (1.4)

which are characterized by λ(p)=λ(1p)\lambda(p)=\lambda(\frac{1}{p}).

Theorem 1.7.

MCTodaM^{C-Toda} is a natural Frobenius submanifold of MTodaM^{Toda}, characterized by the conditions imposed on the flat coordinates of MTodaM^{Toda} as follows:

t0,j=t1,j,j=1,,n01;\displaystyle t_{0,j}=t_{1,j},\quad j=1,\cdots,n_{0}-1;
t2,2=t2,4==t2,n2=0,t2,012t1,0=log1;\displaystyle t_{2,2}=t_{2,4}=\cdots=t_{2,n_{2}}=0,\quad t_{2,0}-\frac{1}{2}t_{1,0}=-\log 1;
t3,2=t3,4==t3,n3=0,t3,012t1,0=log(1);\displaystyle t_{3,2}=t_{3,4}=\cdots=t_{3,n_{3}}=0,\quad t_{3,0}-\frac{1}{2}t_{1,0}=-\log(-1);
t2i2,j=t2i1,j,3im, 0jni,\displaystyle t_{2i-2,j}=t_{2i-1,j},\quad 3\leq i\leq m^{\prime},\ 0\leq j\leq n_{i}^{\prime},

where the constants log1\log 1 and log(1)\log(-1) depend on the choice of branch of logp\log p. Moreover, the subhierarchy Tt,p\frac{\partial}{\partial T^{t,p}} of the principal hierarchy for MTodaM^{Toda}, where

ttCToda={t0,j}j=1n01{t2,2j1}j=1n1{t3,2j1}j=1n2{t4,j}j=0n3{t2m2,j}j=0nm,t\in\textbf{t}^{C-Toda}=\{t_{0,j}\}_{j=1}^{n_{0}^{\prime}-1}\cup\{t_{2,2j-1}\}_{j=1}^{n_{1}^{\prime}}\cup\{t_{3,2j-1}\}_{j=1}^{n_{2}^{\prime}}\cup\{t_{4,j}\}_{j=0}^{n_{3}^{\prime}}\cup\cdots\cup\{t_{2m^{\prime}-2,j}\}_{j=0}^{n_{m^{\prime}}^{\prime}},

can be directly restricted to the submanifold MCTodaM^{C-Toda}, forming the principal hierarchy for MCTodaM^{C-Toda}.

when m=2m^{\prime}=2 and n1=n2=2n_{1}^{\prime}=n_{2}^{\prime}=2, this Frobenius manifold structure coincides with that on the orbit space of the extended affine Weyl group of type DD [34, 37]. As far as we know, the explicit form of the Dubrovin-Zhang hierarchy associated with this Frobenius manifold remains unknown. However, Minanov and Cheng proposed bilinear-type equations governing the descendant potential of this Frobenius manifold [38], and reformulated these equations in the form of Lax equations [39].

There exists a flat F-manifold structure on MToda×M^{Toda}\times\mathbb{C} which is a rank-1 extension of MTodaM^{Toda}. The multiplication on this flat F-manifold is given by

(α,0)(β,0)=(αβ,αβΩs)\displaystyle(\partial_{\alpha},0)\star(\partial_{\beta},0)=(\partial_{\alpha}\circ\partial_{\beta},\partial_{\alpha}\partial_{\beta}\Omega\cdot\partial_{s})
(α,0)(0,s)=(0,αλ(z)s)\displaystyle(\partial_{\alpha},0)\star(0,\partial_{s})=(0,\partial_{\alpha}\lambda(z)\cdot\partial_{s})
(0,s)(0,s)=(0,zλ(z)s)\displaystyle(0,\partial_{s})\star(0,\partial_{s})=(0,z\lambda^{\prime}(z)\cdot\partial_{s})

where ss is the coordinate on \mathbb{C}, z=esz=e^{s}, α=tα\partial_{\alpha}=\frac{\partial}{\partial t^{\alpha}} for tαtTodat^{\alpha}\in\textbf{t}^{Toda}, and

αβΩ=(αλ(z)βλ(z)zλ(z)),0+j=1m(αλ(z)βλ(z)zλ(z))φj,1.\partial_{\alpha}\partial_{\beta}\Omega=(\frac{\partial_{\alpha}\lambda(z)\partial_{\beta}\lambda(z)}{z\lambda^{\prime}(z)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{\alpha}\lambda(z)\partial_{\beta}\lambda(z)}{z\lambda^{\prime}(z)})_{\varphi_{j},\leq-1}.
Corollary 1.8.

The principal hierarchy for the flat F-manifold MToda×M^{Toda}\times\mathbb{C} is given by

λ(z)T~s,p=0,s(x)T~s,p=dx(λ(z)p+1(p+1)!),\frac{\partial\lambda(z)}{\partial\tilde{T}^{s,p}}=0,\quad\frac{\partial s(x)}{\partial\tilde{T}^{s,p}}=d_{x}(\frac{\lambda(z)^{p+1}}{(p+1)!}),

and

λ(z)T~t0,n0j,p=λ(z)Tt0,j,p,s(x)T~t0,n0j,p=dx(dθt0,j,p+1(z))+,\displaystyle\frac{\partial\lambda(z)}{\partial\tilde{T}^{t_{0,n_{0}-j},p}}=\frac{\partial\lambda(z)}{\partial T^{t_{0,j},p}},\quad\frac{\partial s(x)}{\partial\tilde{T}^{t_{0,n_{0}-j},p}}=d_{x}(d\theta_{t_{0,j},p+1}(z))_{+},
λ(z)T~ti,nij,p=λ(z)Tti,j,p,s(x)T~t0,nij,p=dx(dθti,j,p+1(z)),\displaystyle\frac{\partial\lambda(z)}{\partial\tilde{T}^{t_{i,n_{i}-j},p}}=\frac{\partial\lambda(z)}{\partial T^{t_{i,j},p}},\quad\frac{\partial s(x)}{\partial\tilde{T}^{t_{0,n_{i}-j},p}}=-d_{x}(d\theta_{t_{i,j},p+1}(z))_{-},

where

dxλ(z)=xλ(z)+zλ(z)s(x)x,d_{x}\lambda(z)=\partial_{x}\lambda(z)+z\lambda^{\prime}(z)\frac{\partial s(x)}{\partial x},

dθti,j,p(z)d\theta_{t_{i,j},p}(z) is certain function on λ(z)\lambda(z).

We hope that this result can provide insight into defining an open-type extended Toda hierarchy governing the generating function of the open Gromov-Witten invariants of 1\mathbb{CP}^{1} [40].

This paper is organized as follows: In Section 2, we review the definition of Frobenius manifold, almost duality, flat F-manifold, and associated principal hierarchy. In Section 3, we construct the principal hierarchy for the Frobenius manifold with rational superpotential and its almost duality. We then show that this principal hierarchy can be directly restricted to the Frobenius submanifold with even superpotential. Finally, we provide a natural rank-1 extension of this Frobenius manifold and construct the associated principal hierarchy. In Section 4, we apply the same procedure to the Frobenius manifold with trigonometric superpotential.

2. preliminary on Frobenius manifold

In this section, we will recall the definition of Frobenius manifold and associated principal hierarchy.

2.1. Frobenius manifold and principal hierarchy

A Frobenius manifold of charge dd is an nn-dimensional manifold MM, where each tangent space TvMT_{v}M is equipped with a Frobenius algebra structure (Av=TvM,,e,,)(A_{v}=T_{v}M,\circ,e,\langle\cdot,\cdot\rangle) that varies smoothly with vMv\in M. This structure satisfies the following axioms:

  1. (1)

    The bilinear form ,\langle\cdot,\cdot\rangle provides a flat metric on MM, and the unity vector field ee satisfies e=0\nabla e=0, where \nabla is the Levi-Civita connection for the flat metric.

  2. (2)

    Define a 3-tensor cc by c(X,Y,Z):=XY,Zc(X,Y,Z):=\langle X\circ Y,Z\rangle with X,Y,ZTvMX,Y,Z\in T_{v}M. Then, the 4-tensor (Wc)(X,Y,Z)(\nabla_{W}c)(X,Y,Z) is symmetric in X,Y,Z,WTvMX,Y,Z,W\in T_{v}M.

  3. (3)

    There exists a vector field EE, called the Euler vector field, which satisfies 2E=0\nabla^{2}E=0 and

    [E,XY][E,X]YX[E,Y]=XY,\displaystyle[E,X\circ Y]-[E,X]\circ Y-X\circ[E,Y]=X\circ Y,
    E(X,Y)[E,X],YX,[E,Y]=(2d)X,Y\displaystyle E(\langle X,Y\rangle)-\langle[E,X],Y\rangle-\langle X,[E,Y]\rangle=(2-d)\langle X,Y\rangle

    for any vector fields X,YX,Y on MM.

On an nn-dimensional Frobenius manifold MM, we select a set of flat coordinates t=(t1,,tn)t=(t^{1},\ldots,t^{n}) such that e=t1e=\frac{\partial}{\partial t^{1}}. In this coordinate system, the components of the metric ,\langle\ ,\ \rangle are given by:

ηαβ=tα,tβ,α,β=1,,n,\eta_{\alpha\beta}=\left\langle\frac{\partial}{\partial t^{\alpha}},\frac{\partial}{\partial t^{\beta}}\right\rangle,\quad\alpha,\beta=1,\ldots,n,

where ηαβ\eta_{\alpha\beta} defines a constant and non-degenerate n×nn\times n matrix. The inverse of this matrix is denoted by ηαβ\eta^{\alpha\beta}. The metric and its inverse are utilized for index lowering and raising, respectively, with the Einstein summation convention applied to repeated Greek indices.

Furthermore, we denote the components of the 3-tensor cc by:

cαβγ=c(tα,tβ,tγ),α,β,γ=1,,n,c_{\alpha\beta\gamma}=c\left(\frac{\partial}{\partial t^{\alpha}},\frac{\partial}{\partial t^{\beta}},\frac{\partial}{\partial t^{\gamma}}\right),\quad\alpha,\beta,\gamma=1,\ldots,n,

which allows us to express the multiplication structure of the Frobenius algebra TvMT_{v}M in terms of

tαtβ=cαβγtγ,\frac{\partial}{\partial t^{\alpha}}\circ\frac{\partial}{\partial t^{\beta}}=c_{\alpha\beta}^{\gamma}\frac{\partial}{\partial t^{\gamma}},

where the coefficients cαβγc_{\alpha\beta}^{\gamma} are obtained by contracting the 3-tensor cc with the metric ηαβ\eta^{\alpha\beta}:

cαβγ=ηγϵcϵαβ,c_{\alpha\beta}^{\gamma}=\eta^{\gamma\epsilon}c_{\epsilon\alpha\beta},

which satisfy

c1αβ=δαβ,cαβϵcϵγσ=cαγϵcϵβσ.c_{1\alpha}^{\beta}=\delta_{\alpha}^{\beta},\quad c_{\alpha\beta}^{\epsilon}c_{\epsilon\gamma}^{\sigma}=c_{\alpha\gamma}^{\epsilon}c_{\epsilon\beta}^{\sigma}.

According to the definition of Frobenius manifold, there exists a smooth function F(t)F(t) satisfying the following properties:

cαβγ\displaystyle c_{\alpha\beta\gamma} =3Ftαtβtγ,\displaystyle=\frac{\partial^{3}F}{\partial t^{\alpha}\partial t^{\beta}\partial t^{\gamma}},
LieEF\displaystyle\operatorname{Lie}_{E}F =(3d)F+quadratic terms in t.\displaystyle=(3-d)F+\text{quadratic terms in }t.

Hence, F(t)F(t) is a solution to the WDVV equation

3Ftαtβtγηγϵ3Ftϵtσtμ=3Ftαtσtγηγϵ3Ftϵtβtμ.\frac{\partial^{3}F}{\partial t^{\alpha}\partial t^{\beta}\partial t^{\gamma}}\eta^{\gamma\epsilon}\frac{\partial^{3}F}{\partial t^{\epsilon}\partial t^{\sigma}\partial t^{\mu}}=\frac{\partial^{3}F}{\partial t^{\alpha}\partial t^{\sigma}\partial t^{\gamma}}\eta^{\gamma\epsilon}\frac{\partial^{3}F}{\partial t^{\epsilon}\partial t^{\beta}\partial t^{\mu}}.

The third-order derivatives cαβγc_{\alpha\beta\gamma} of F(t)F(t) are known as the 3-point correlator functions in the context of topological field theory.

For a Frobenius manifold MM, its cotangent space TvMT_{v}^{*}M is endowed with a Frobenius algebra structure as well. This structure encompasses an invariant bilinear form and a product, which are defined by:

dtα,dtβ=ηαβ,dtαdtβ=ηαϵcϵγβ.\left\langle dt^{\alpha},dt^{\beta}\right\rangle=\eta^{\alpha\beta},\quad dt^{\alpha}\circ dt^{\beta}=\eta^{\alpha\epsilon}c_{\epsilon\gamma}^{\beta}.

Let us define

gαβ=iE(dtαdtβ),g^{\alpha\beta}=i_{E}\left(dt^{\alpha}\circ dt^{\beta}\right),

then (dtα,dtβ):=gαβ\left(dt^{\alpha},dt^{\beta}\right):=g^{\alpha\beta} establishes a symmetric bilinear form known as the intersection form on TvMT_{v}^{*}M. The intersection form gαβg^{\alpha\beta} and the invariant bilinear form ηαβ\eta^{\alpha\beta} together form a pencil of flat metrics

gαβ+ϵηαβg^{\alpha\beta}+\epsilon\eta^{\alpha\beta}

parameterized by ϵ\epsilon. As a result, they give rise to a bi-hamiltonian structure of hydrodynamic type on the loop space {S1M}\{S^{1}\rightarrow M\}, expressed as:

{,}2+ϵ{,}1.\{\ ,\ \}_{2}+\epsilon\{\ ,\ \}_{1}.

The deformed flat connection on MM, originally introduced by Dubrovin [1], is defined as:

~XY=XY+zXY,X,YVect(M).\widetilde{\nabla}_{X}Y=\nabla_{X}Y+zX\circ Y,\quad X,Y\in\operatorname{Vect}(M).

This connection can be consistently extended to a flat affine connection on M×M\times\mathbb{C}^{*} such that

~Xddz=0,\displaystyle\tilde{\nabla}_{X}\frac{d}{dz}=0,
~ddzddz=0,\displaystyle\tilde{\nabla}_{\frac{d}{dz}}\frac{d}{dz}=0,
~ddzX=zX+EX1z𝒱(X),\displaystyle\tilde{\nabla}_{\frac{d}{dz}}X=\partial_{z}X+E\circ X-\frac{1}{z}\mathcal{V}(X),

where XX is a vector field on M×M\times\mathbb{C}^{*} that vanishes in the \mathbb{C}^{*} component, and 𝒱(X)\mathcal{V}(X) is defined as

𝒱(X):=2d2XXE.\mathcal{V}(X):=\frac{2-d}{2}X-\nabla_{X}E.

There exists a system of deformed flat coordinates v~1(t,z),,v~n(t,z)\tilde{v}_{1}(t,z),\ldots,\tilde{v}_{n}(t,z) that can be expressed in terms of

(v~1(t,z),,v~n(t,z))=(θ1(t,z),,θn(t,z))zμzR.\left(\tilde{v}_{1}(t,z),\ldots,\tilde{v}_{n}(t,z)\right)=\left(\theta_{1}(t,z),\ldots,\theta_{n}(t,z)\right)z^{\mu}z^{R}.

These coordinates are chosen such that the 1-forms

ξα=v~αtβdtβ,α=1,,n,andξn+1=dz,\xi_{\alpha}=\frac{\partial\tilde{v}_{\alpha}}{\partial t^{\beta}}dt^{\beta},\quad\alpha=1,\ldots,n,\quad\text{and}\quad\xi_{n+1}=dz,

constitute a basis of solutions to the system ~ξ=0\widetilde{\nabla}\xi=0. Here, μ=diag(μ1,,μn)\mu=\operatorname{diag}(\mu_{1},\ldots,\mu_{n}) is a diagonal matrix characterized by

𝒱(tα)=μαtα,α=1,,n,\mathcal{V}\left(\frac{\partial}{\partial t^{\alpha}}\right)=\mu_{\alpha}\frac{\partial}{\partial t^{\alpha}},\quad\alpha=1,\ldots,n,

which is called the spectrum of MM, and R=R1++RmR=R_{1}+\ldots+R_{m} is a constant nilpotent matrix satisfying

(Rs)βα=0 if μαμβs,\displaystyle(R_{s})_{\beta}^{\alpha}=0\text{ if }\mu_{\alpha}-\mu_{\beta}\neq s,
(Rs)αγηγβ=(1)s+1(Rs)βγηγα.\displaystyle(R_{s})_{\alpha}^{\gamma}\eta_{\gamma\beta}=(-1)^{s+1}(R_{s})_{\beta}^{\gamma}\eta_{\gamma\alpha}.

The functions θα(t,z)\theta_{\alpha}(t,z), being analytic near z=0z=0, can be represented by a power series expansion:

θα(t,z)=p0θα,p(t)zp,α=1,,n.\theta_{\alpha}(t,z)=\sum_{p\geq 0}\theta_{\alpha,p}(t)z^{p},\quad\alpha=1,\ldots,n.

The coefficients of this expansion satisfy

2θα,p+1(t)tβtγ=cβγϵ(t)θα,p(t)tϵ,\frac{\partial^{2}\theta_{\alpha,p+1}(t)}{\partial t^{\beta}\partial t^{\gamma}}=c_{\beta\gamma}^{\epsilon}(t)\frac{\partial\theta_{\alpha,p}(t)}{\partial t^{\epsilon}}, (2.1)

and

LieE(θα,p(t)tβ)=(p+μα+μβ)θα,p(t)tβ+tβs=1pθϵ,ps(t)(Rs)αϵ.\operatorname{Lie}_{E}\left(\frac{\partial\theta_{\alpha,p}(t)}{\partial t^{\beta}}\right)=\left(p+\mu_{\alpha}+\mu_{\beta}\right)\frac{\partial\theta_{\alpha,p}(t)}{\partial t^{\beta}}+\frac{\partial}{\partial t^{\beta}}\sum_{s=1}^{p}\theta_{\epsilon,p-s}(t)\left(R_{s}\right)_{\alpha}^{\epsilon}. (2.2)

Moreover, the normalization condition111In the referenced work [1], an additional condition θα(t,z),θβ(t,z)=ηαβ\langle\nabla\theta_{\alpha}(t,z),\nabla\theta_{\beta}(t,-z)\rangle=\eta_{\alpha\beta} was considered. However, as it does not significantly alter the properties of the principal hierarchy, we omit it here for computational simplicity is imposed:

θα,0(t)=ηαβtβ.\theta_{\alpha,0}(t)=\eta_{\alpha\beta}t^{\beta}. (2.3)

Given a system of solutions {θα,p}\left\{\theta_{\alpha,p}\right\} to the equations (2.1)-(2.3), the principal hierarchy associated with MM is defined as the following Hamiltonian system on the loop space {S1M}\left\{S^{1}\rightarrow M\right\} :

tγTα,p={tγ(x),θα,p+1(t)𝑑x}1:=ηγβx(θα,p+1(t)tβ),α,β=1,2,,n,p0.\frac{\partial t^{\gamma}}{\partial T^{\alpha,p}}=\left\{t^{\gamma}(x),\int\theta_{\alpha,p+1}(t)\,dx\right\}_{1}:=\eta^{\gamma\beta}\frac{\partial}{\partial x}\left(\frac{\partial\theta_{\alpha,p+1}(t)}{\partial t^{\beta}}\right),\quad\alpha,\beta=1,2,\ldots,n,\ p\geq 0.

These commuting flows are tau-symmetric, which means that

θα,p(t)Tβ,q=θβ,q(t)Tα,p,α,β=1,2,,n,p,q0.\frac{\partial\theta_{\alpha,p}(t)}{\partial T^{\beta,q}}=\frac{\partial\theta_{\beta,q}(t)}{\partial T^{\alpha,p}},\quad\alpha,\beta=1,2,\ldots,n,\ p,q\geq 0.

Furthermore, these flows can be expressed in a bi-hamiltonian recursion form as

Tα,p1=Tα,p(p+μα+12)+s=1pTϵ,ps(Rs)αϵ,\mathcal{R}\frac{\partial}{\partial T^{\alpha,p-1}}=\frac{\partial}{\partial T^{\alpha,p}}\left(p+\mu_{\alpha}+\frac{1}{2}\right)+\sum_{s=1}^{p}\frac{\partial}{\partial T^{\epsilon,p-s}}\left(R_{s}\right)_{\alpha}^{\epsilon},

where ={,}2{,}11\mathcal{R}=\{\ ,\ \}_{2}\cdot\{\ ,\ \}_{1}^{-1}.

2.2. almost duality

The discriminant of a Frobenius manifold is defined as

Σ={pM|det(gαβ(p))=0}.\Sigma=\{p\in M|\det(g^{\alpha\beta}(p))=0\}.

Let M^=MΣ\hat{M}=M\setminus\Sigma, and for any pM^p\in\hat{M}, define the multiplication on TpM^T_{p}\hat{M} as

uv=E1uv,u,vTpM^.u\star v=E^{-1}\circ u\circ v,\quad u,v\in T_{p}\hat{M}.

Then the data set (M^,gαβ,)(\hat{M},g^{\alpha\beta},\star) satisfies the axiom (2) in the definition of Frobenius manifold. This data set is referred to as the almost duality [15] of the Frobenius manifold MM.

Let t^=(t^1,,t^n)\hat{t}=(\hat{t}^{1},\ldots,\hat{t}^{n}) be the flat coordinates of gαβg^{\alpha\beta}, and let c^αβγ(t)\hat{c}_{\alpha\beta}^{\gamma}(t) be the components of the multiplication structure \star in these flat coordinates. Consider the formal power series

θ^α(t^,z)=p0θ^α,p(t^)zp,α=1,,n,\hat{\theta}_{\alpha}(\hat{t},z)=\sum_{p\geq 0}\hat{\theta}_{\alpha,p}(\hat{t})z^{p},\quad\alpha=1,\ldots,n,

such that the following system of equations holds:

2θ^α,p+1(t^)t^βt^γ=c^βγϵ(t^)θ^α,p(t^)t^ϵ,θ^α,0=gαβt^β,\frac{\partial^{2}\hat{\theta}_{\alpha,p+1}(\hat{t})}{\partial\hat{t}^{\beta}\partial\hat{t}^{\gamma}}=\hat{c}_{\beta\gamma}^{\epsilon}(\hat{t})\frac{\partial\hat{\theta}_{\alpha,p}(\hat{t})}{\partial\hat{t}^{\epsilon}},\quad\hat{\theta}_{\alpha,0}=g_{\alpha\beta}\hat{t}^{\beta},

then dθ^(t^,z)d\hat{\theta}(\hat{t},z) provides a set of fundamental solutions near z=0z=0 for the deformed flat connection:

^udeformv=^uv+zE1uv,\hat{\nabla}^{deform}_{u}v=\hat{\nabla}_{u}v+z\cdot E^{-1}\circ u\circ v,

where ^\hat{\nabla} is the Levi-Civita connection of gαβg^{\alpha\beta}. The Hamiltonian system

t^γT^α,p={t^γ(x),θ^α,p+1(t^)𝑑x}2:=gγβx(θ^α,p+1(t^)t^β),α,β=1,2,,n,p0,\frac{\partial\hat{t}^{\gamma}}{\partial\hat{T}^{\alpha,p}}=\left\{\hat{t}^{\gamma}(x),\int\hat{\theta}_{\alpha,p+1}(\hat{t})dx\right\}_{2}:=g^{\gamma\beta}\frac{\partial}{\partial x}\left(\frac{\partial\hat{\theta}_{\alpha,p+1}(\hat{t})}{\partial\hat{t}^{\beta}}\right),\quad\alpha,\beta=1,2,\ldots,n,\ p\geq 0,

defines an integrable hierarchy on the loop space of M^\hat{M}.

2.3. flat F-manifold

A flat F-manifold (M,,,e)(M,\nabla,\circ,e) consists of an analytic manifold MM, a flat torsionless connection \nabla on TMTM, and a commutative associated algebra structure on each tangent space TpMT_{p}M with a unit vector field ee, satisfying the following conditions:

  1. (1)

    e=0\nabla e=0;

  2. (2)

    There exists a vector field Ψ\Psi on MM, called the vector potential, such that

    XY=[X,[Y,Ψ]]X\circ Y=[X,[Y,\Psi]]

    for any flat vector field XX and YY on MM.

For more details, see [41, 42, 19].

Let t=(t1,,tn)\textbf{t}=(t^{1},\dots,t^{n}) denote the flat coordinates associated with the connection \nabla. The principal hierarchy for the flat F-manifold MM is defined as:

Ti,p=Θi,px,i=1,,n,\frac{\partial}{\partial T^{i,p}}=\Theta_{i,p}\circ\partial_{x},\quad i=1,\cdots,n,

where the vector fields Θi,p\Theta_{i,p} on MM satisfy

Θi,0=ti,i=1,,n,\Theta_{i,0}=\frac{\partial}{\partial t^{i}},\quad i=1,\cdots,n,

and

XΘi,p+1=Θi,pX\nabla_{X}\Theta_{i,p+1}=\Theta_{i,p}\circ X

for any vector field XX on MM.

Let MM be a Frobenius manifold with flat coordinates t=(t1,,tn)\textbf{t}=(t^{1},\cdots,t^{n}) and potential F(t)F(\textbf{t}). There exists a flat F-manifold structure on M×M\times\mathbb{C} with vector potential:

Ψ=F(t)tαηαβtβ+Ω(t,s)s\Psi=\frac{\partial F(\textbf{t})}{\partial t^{\alpha}}\eta^{\alpha\beta}\frac{\partial}{\partial t^{\beta}}+\Omega(\textbf{t},s)\frac{\partial}{\partial s}

if and only if Ω(t,s)\Omega(\textbf{t},s) satisfies the open WDVV equations:

cαβδΩ(t,s)tδtγ+Ω(t,s)tαtβΩ(t,s)tγs=cβγδΩ(t,s)tδtα+Ω(t,s)tβtγΩ(t,s)tαsc_{\alpha\beta}^{\delta}\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\delta}\partial t^{\gamma}}+\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\alpha}\partial t^{\beta}}\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\gamma}\partial s}=c_{\beta\gamma}^{\delta}\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\delta}\partial t^{\alpha}}+\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\beta}\partial t^{\gamma}}\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\alpha}\partial s}

and

cαβδΩ(t,s)tδs+Ω(t,s)tαtβΩ(t,s)ss=Ω(t,s)tαsΩ(t,s)tβs,c_{\alpha\beta}^{\delta}\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\delta}\partial s}+\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\alpha}\partial t^{\beta}}\frac{\partial\Omega(\textbf{t},s)}{\partial s\partial s}=\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\alpha}\partial s}\frac{\partial\Omega(\textbf{t},s)}{\partial t^{\beta}\partial s},

where ss denotes the coordinate on \mathbb{C}.

The following lemma by A. Alcolado is useful for constructing such a function Ω(t,s)\Omega(\textbf{t},s).

Lemma 2.1 ([19]).

Let ω=ω(t,s)\omega=\omega(\textbf{t},s) be a smooth function on M×M\times\mathbb{C} satisfying sω0\partial_{s}\omega\neq 0 and

αβω=s(αωβωcαβδδωsω),α,β=1,,n,\partial_{\alpha}\partial_{\beta}\omega=\partial_{s}(\frac{\partial_{\alpha}\omega\partial_{\beta}\omega-c_{\alpha\beta}^{\delta}\partial_{\delta}\omega}{\partial_{s}\omega}),\quad\alpha,\beta=1,\cdots,n,

where α=tα\partial_{\alpha}=\frac{\partial}{\partial t^{\alpha}}. Then, the function Ω(t,s)\Omega(\textbf{t},s) defined by ω=sΩ\omega=\partial_{s}\Omega provides a solution to the open WDVV equations associated with MM.

3. Frobenius manifold with rational superpotential

3.1. Definition of McKPM^{cKP}

Given positive integers mm and n0,,nmn_{0},\ldots,n_{m}, let McKPM^{cKP} be the space of rational functions

λ(z)=1n0zn0+a0,n02zn02++a0,0+i=1mj=1niai,j(zai,0)j\lambda(z)=\frac{1}{n_{0}}z^{n_{0}}+a_{0,n_{0}-2}z^{n_{0}-2}+\cdots+a_{0,0}+\sum_{i=1}^{m}\sum_{j=1}^{n_{i}}a_{i,j}(z-a_{i,0})^{-j}

where ai,ni0,i=1,,ma_{i,n_{i}}\neq 0,\ i=1,\ldots,m. The parameters {a0,i}i=0n02{a1,i}i=0n1{am,i}i=0nm\{a_{0,i}\}_{i=0}^{n_{0}-2}\cup\{a_{1,i}\}_{i=0}^{n_{1}}\cup\cdots\cup\{a_{m,i}\}_{i=0}^{n_{m}} form a coordinate system on McKPM^{cKP}.

For any ,′′,′′′Tλ(z)McKP\partial^{\prime},\partial^{\prime\prime},\partial^{\prime\prime\prime}\in T_{\lambda(z)}M^{cKP}, define the metric

,′′η:=η(,′′)=|λ|<Resdλ=0(λ(z)dz)′′(λ(z)dz)dλ(z)\langle\partial^{\prime},\partial^{\prime\prime}\rangle_{\eta}:=\eta(\partial^{\prime},\partial^{\prime\prime})=\sum_{|\lambda|<\infty}\mathop{\text{\rm Res}}_{d\lambda=0}\frac{\partial^{\prime}(\lambda(z)dz)\partial^{\prime\prime}(\lambda(z)dz)}{d\lambda(z)}

and the (0,3)(0,3)-type tensor

c(,′′,′′′):=|λ|<Resdλ=0(λ(z)dz)′′(λ(z)dz)′′′(λ(z)dz)dλ(z)dz,c(\partial^{\prime},\partial^{\prime\prime},\partial^{\prime\prime\prime}):=\sum_{|\lambda|<\infty}\mathop{\text{\rm Res}}_{d\lambda=0}\frac{\partial^{\prime}(\lambda(z)dz)\partial^{\prime\prime}(\lambda(z)dz)\partial^{\prime\prime\prime}(\lambda(z)dz)}{d\lambda(z)dz},

then the equality

c(,′′,′′′)=η(′′,′′′)c(\partial^{\prime},\partial^{\prime\prime},\partial^{\prime\prime\prime})=\eta(\partial^{\prime}\circ\partial^{\prime\prime},\partial^{\prime\prime\prime})

defines the multiplication \circ on Tλ(z)McKPT_{\lambda(z)}M^{cKP}. Introduce vector fields ee and EE on McKPM^{cKP} such that

Lieeλ(z)=1,LieEλ(z)=λ(z)zn0λ(z),Lie_{e}\lambda(z)=1,\quad Lie_{E}\lambda(z)=\lambda(z)-\frac{z}{n_{0}}\lambda^{\prime}(z),

then the data set (M,η,,e,E)(M,\eta,\circ,e,E) constitutes a semisimple Frobenius manifold with d=12n0d=1-\frac{2}{n_{0}}.

The flat coordinates of the metric η\eta, denoted as

𝐭={t0,j}j=1n01{t1,j}j=0n1{tm,j}j=0nm,\mathbf{t}=\{t_{0,j}\}_{j=1}^{n_{0}-1}\cup\{t_{1,j}\}_{j=0}^{n_{1}}\cup\cdots\cup\{t_{m,j}\}_{j=0}^{n_{m}},

are given by the coefficients of the following series:

z={ti,0+ti,1wi1+,zai,0,i=1,,m,w0t0,1w01t0,2w02+,z,i=0,z=\left\{\begin{aligned} &t_{i,0}+t_{i,1}w_{i}^{-1}+\cdots,\quad&z&\to a_{i,0},\ i=1,\cdots,m,\\ &w_{0}-t_{0,1}w_{0}^{-1}-t_{0,2}w_{0}^{-2}+\cdots,\quad&z&\to\infty,\ \ i=0,\end{aligned}\right.

where

wi={(niλ)1ni=wi,1(zai,0)1+,zai,0,i=1,,m,(n0λ)1n0=z+w0,1z1+,z,i=0.w_{i}=\left\{\begin{aligned} &(n_{i}\lambda)^{\frac{1}{n_{i}}}=w_{i,1}(z-a_{i,0})^{-1}+\cdots,\quad&z&\to a_{i,0},\ i=1,\cdots,m,\\ &(n_{0}\lambda)^{\frac{1}{n_{0}}}=z+w_{0,1}z^{-1}+\cdots,\quad&z&\to\infty,\ i=0.\end{aligned}\right.

Furthermore, we have

λ(z)ti,j={(wi(z)nij1wi(z))ai,0,1,i=1,,m,j=0,,ni,(wi(z)n0j1w0(z)),0,i=0,j=1,,n01.\frac{\partial\lambda(z)}{\partial t_{i,j}}=\left\{\begin{aligned} &-(w_{i}(z)^{n_{i}-j-1}w_{i}^{\prime}(z))_{a_{i,0},\leq-1},\quad i=1,\cdots,m,\ j=0,\cdots,n_{i},\\ &(w_{i}(z)^{n_{0}-j-1}w_{0}^{\prime}(z))_{\infty,\geq 0},\quad i=0,\ j=1,\cdots,n_{0}-1.\end{aligned}\right.

In the flat coordinate system, the vector fields ee and EE can be expressed as

e=t0,n01,E=j=1n01(1+jn0)t0,jt0,j+i=1mj=0ni(1n0+jni)ti,jti,j,e=\frac{\partial}{\partial t_{0,n_{0}-1}},\quad E=\sum_{j=1}^{n_{0}-1}\left(\frac{1+j}{n_{0}}\right)t_{0,j}\frac{\partial}{\partial t_{0,j}}+\sum_{i=1}^{m}\sum_{j=0}^{n_{i}}\left(\frac{1}{n_{0}}+\frac{j}{n_{i}}\right)t_{i,j}\frac{\partial}{\partial t_{i,j}},

thus the spectrum of McKPM^{cKP} is μti,j=12jni\mu_{t_{i,j}}=\frac{1}{2}-\frac{j}{n_{i}}.

Lemma 3.1.

Let \nabla be the Levi-Civita connection associated with the metric η\eta. Then, for any vector fields 1\partial_{1} and 2\partial_{2} on McKPM^{cKP}, it holds that

(12)λ(z)=12λ(z)(1λ(z)2λ(z)λ(z)),0s=1m(1λ(z)2λ(z)λ(z))as,0,1.(\nabla_{\partial_{1}}\partial_{2})\cdot\lambda(z)=\partial_{1}\partial_{2}\lambda(z)-\left(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}_{\infty,\geq 0}-\sum_{s=1}^{m}\left(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}_{a_{s,0},\leq-1}. (3.1)
Proof.

It suffices to prove that the connection \nabla defined by (3.1) satisfies

(12)λ(z)(21)λ(z)=12λ(z)21λ(z)(\nabla_{\partial_{1}}\partial_{2})\cdot\lambda(z)-(\nabla_{\partial_{2}}\partial_{1})\cdot\lambda(z)=\partial_{1}\partial_{2}\lambda(z)-\partial_{2}\partial_{1}\lambda(z) (3.2)

and

12,3η=12,3η+2,13η.\nabla_{\partial_{1}}\left\langle\partial_{2},\partial_{3}\right\rangle_{\eta}=\left\langle\nabla_{\partial_{1}}\partial_{2},\partial_{3}\right\rangle_{\eta}+\left\langle\partial_{2},\nabla_{\partial_{1}}\partial_{3}\right\rangle_{\eta}. (3.3)

The equality (3.2) can be easily verified. For the left-hand side of the equality (3.3), we have

12,3η\displaystyle\nabla_{\partial_{1}}\left\langle\partial_{2},\partial_{3}\right\rangle_{\eta} (3.4)
=\displaystyle= Res(12λ(z)3λ(z)λ(z)+2λ(z)13λ(z)λ(z)2λ(z)3λ(z)1λ(z)(λ(z))2)dz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\left(\frac{\partial_{1}\partial_{2}\lambda(z)\cdot\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}+\frac{\partial_{2}\lambda(z)\cdot\partial_{1}\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}-\frac{\partial_{2}\lambda(z)\cdot\partial_{3}\lambda(z)\cdot\partial_{1}\lambda^{\prime}(z)}{(\lambda^{\prime}(z))^{2}}\right)dz
s=1mResas,0(12λ(z)3λ(z)λ(z)+2λ(z)13λ(z)λ(z)2λ(z)3λ(z)1λ(z)(λ(z))2)dz.\displaystyle-\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\left(\frac{\partial_{1}\partial_{2}\lambda(z)\cdot\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}+\frac{\partial_{2}\lambda(z)\cdot\partial_{1}\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}-\frac{\partial_{2}\lambda(z)\cdot\partial_{3}\lambda(z)\cdot\partial_{1}\lambda^{\prime}(z)}{(\lambda^{\prime}(z))^{2}}\right)dz.

Because 1,2\partial_{1},\partial_{2} commute with z\partial_{z}, applying integration by parts, we obtain

Res2λ(z)3λ(z)1λ(z)(λ(z))2dz\displaystyle\mathop{\text{\rm Res}}_{\infty}\frac{\partial_{2}\lambda(z)\cdot\partial_{3}\lambda(z)\cdot\partial_{1}\lambda^{\prime}(z)}{(\lambda^{\prime}(z))^{2}}dz (3.5)
=\displaystyle= Res2λ(z)λ(z)(3λ(z)1λ(z)λ(z))dz+Res3λ(z)λ(z)(2λ(z)1λ(z)λ(z))dz.\displaystyle\mathop{\text{\rm Res}}_{\infty}\frac{\partial_{2}\lambda(z)}{\lambda^{\prime}(z)}\left(\frac{\partial_{3}\lambda(z)\cdot\partial_{1}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}dz+\mathop{\text{\rm Res}}_{\infty}\frac{\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}\left(\frac{\partial_{2}\lambda(z)\cdot\partial_{1}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}dz.

Thus,

12,3η\displaystyle\nabla_{\partial_{1}}\left\langle\partial_{2},\partial_{3}\right\rangle_{\eta} (3.6)
=\displaystyle= Res(12λ(z)3λ(z)λ(z)+2λ(z)13λ(z)λ(z))dz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\left(\frac{\partial_{1}\partial_{2}\lambda(z)\cdot\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}+\frac{\partial_{2}\lambda(z)\cdot\partial_{1}\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}\right)dz
+Res(2λ(z)λ(z)(3λ(z)1λ(z)λ(z))+3λ(z)λ(z)(2λ(z)1λ(z)λ(z)))dz\displaystyle+\mathop{\text{\rm Res}}_{\infty}\left(\frac{\partial_{2}\lambda(z)}{\lambda^{\prime}(z)}\left(\frac{\partial_{3}\lambda(z)\cdot\partial_{1}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}+\frac{\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}\left(\frac{\partial_{2}\lambda(z)\cdot\partial_{1}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}\right)dz
s=1mResas,0(12λ(z)3λ(z)λ(z)+2λ(z)13λ(z)λ(z))dz\displaystyle-\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\left(\frac{\partial_{1}\partial_{2}\lambda(z)\cdot\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}+\frac{\partial_{2}\lambda(z)\cdot\partial_{1}\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}\right)dz
+s=1mResas,0(2λ(z)λ(z)(3λ(z)1λ(z)λ(z))+3λ(z)λ(z)(2λ(z)1λ(z)λ(z)))dz\displaystyle+\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\left(\frac{\partial_{2}\lambda(z)}{\lambda^{\prime}(z)}\left(\frac{\partial_{3}\lambda(z)\cdot\partial_{1}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}+\frac{\partial_{3}\lambda(z)}{\lambda^{\prime}(z)}\left(\frac{\partial_{2}\lambda(z)\cdot\partial_{1}\lambda(z)}{\lambda^{\prime}(z)}\right)^{\prime}\right)dz
=\displaystyle= 12,3η+2,13η.\displaystyle\left\langle\nabla_{\partial_{1}}\partial_{2},\partial_{3}\right\rangle_{\eta}+\left\langle\partial_{2},\nabla_{\partial_{1}}\partial_{3}\right\rangle_{\eta}.

The lemma is proved. ∎

3.2. cotangent space

For any point λ(z)\lambda(z) in McKPM^{cKP}, a tangent vector Tλ(z)McKP\partial\in T_{\lambda(z)}M^{cKP} can be expressed as a rational function:

(λ(z))=b0,n02zn02++b0,0+i=1mj=1ni+1bi,j(zai,0)j.\partial(\lambda(z))=b_{0,n_{0}-2}z^{n_{0}-2}+\cdots+b_{0,0}+\sum_{i=1}^{m}\sum_{j=1}^{n_{i}+1}b_{i,j}(z-a_{i,0})^{-j}.

To describe a cotangent vector at λ(z)\lambda(z), we consider a collection of disjoint disks D1,,DmD_{1},\ldots,D_{m} on the complex plane \mathbb{C}, such that ai,0Dia_{i,0}\in D_{i} for i=1,,mi=1,\ldots,m. Denote γi=Di\gamma_{i}=\partial D_{i}, D=s=1mDs\textbf{D}=\cup_{s=1}^{m}D_{s}, and Dc=1D\textbf{D}^{c}=\mathbb{P}^{1}\setminus\textbf{D}. Let \mathcal{H} be the space of germs of holomorphic functions on the curves s=1mγs\cup_{s=1}^{m}\gamma_{s}. For any f(z)f(z)\in\mathcal{H}, define

f(z)+:=12πis=1mγsf(p)pz𝑑p,z𝐃̊,f(z):=12πis=1mγsf(p)pz𝑑p,z𝐃c,f(z)_{+}:=\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\frac{f(p)}{p-z}\,dp,\quad z\in\mathring{\mathbf{D}},\quad f(z)_{-}:=-\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\frac{f(p)}{p-z}\,dp,\quad z\in\mathbf{D}^{c},

where f(z)+f(z)_{+} and f(z)f(z)_{-} are holomorphic on D and Dc\textbf{D}^{c}, respectively, and can be analytically continued to some neighborhood of s=1mγs\cup_{s=1}^{m}\gamma_{s}. Thus, f(z)+,f(z)f(z)_{+},f(z)_{-}\in\mathcal{H}, and f(z)=f(z)++f(z)f(z)=f(z)_{+}+f(z)_{-}. Conversely, if there exists a decomposition f(z)=f1(z)+f2(z)f(z)=f_{1}(z)+f_{2}(z), where f1(z)f_{1}(z) and f2(z)f_{2}(z) are holomorphic on D and Dc\textbf{D}^{c}, respectively, and f2()=0f_{2}(\infty)=0, then it follows that f1(z)=f(z)+f_{1}(z)=f(z)_{+} and f2(z)=f(z)f_{2}(z)=f(z)_{-}.

Define the pairing:

ω(z),ξ(z):=12πis=1mγsω(z)ξ(z)𝑑z,ω(z),ξ(z)Tλ(z)McKP.\langle\omega(z),\xi(z)\rangle:=\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\omega(z)\xi(z)\,dz,\quad\omega(z)\in\mathcal{H},\quad\xi(z)\in T_{\lambda(z)}M^{cKP}.

This pairing induces a surjective map from \mathcal{H} to Tλ(z)McKPT^{\ast}_{\lambda(z)}M^{cKP}, so that an element of \mathcal{H} can be regarded as a cotangent vector at λ(z)\lambda(z).

Lemma 3.2.

For any ξ(z)Tλ(z)McKP\xi(z)\in T_{\lambda(z)}M^{cKP}, define linear maps from \mathcal{H} to Tλ(z)McKPT_{\lambda(z)}M^{cKP} as follows:

η(ω(z))=(ω(z))+λ(z)+(ω(z)λ(z))+\eta^{\ast}(\omega(z))=-(\omega(z))_{+}\lambda^{\prime}(z)+(\omega(z)\lambda^{\prime}(z))_{+} (3.7)

and

Cξ(z)(ω(z))=(ω(z)ξ(z))+λ(z)+(ω(z)λ(z))+ξ(z).C_{\xi(z)}(\omega(z))=-(\omega(z)\xi(z))_{+}\lambda^{\prime}(z)+(\omega(z)\lambda^{\prime}(z))_{+}\xi(z). (3.8)

Then we have

η(ω(z)),ξ(z)η=ω(z),ξ(z)\langle\eta^{\ast}(\omega(z)),\xi(z)\rangle_{\eta}=\langle\omega(z),\xi(z)\rangle

and

c(ξ1(z),ξ2(z),η(ω(z)))=Cξ1(z)(ω(z)),ξ2(z)η.c(\xi_{1}(z),\xi_{2}(z),\eta^{\ast}(\omega(z)))=\langle C_{\xi_{1}(z)}(\omega(z)),\xi_{2}(z)\rangle_{\eta}.
Proof.

From direct calculation, we derive

η(ω),ξη=\displaystyle\langle\eta^{\ast}(\omega),\xi\rangle_{\eta}= Res(ωλξλ(ωλ)ξλ)dzs=1mResas,0(ω+λξλ+(ωλ)+ξλ)dz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\left(\frac{\omega_{-}\lambda^{\prime}\xi}{\lambda^{\prime}}-\frac{(\omega\lambda^{\prime})_{-}\xi}{\lambda^{\prime}}\right)dz-\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\left(-\frac{\omega_{+}\lambda^{\prime}\xi}{\lambda^{\prime}}+\frac{(\omega\lambda^{\prime})_{+}\xi}{\lambda^{\prime}}\right)dz
=\displaystyle= Resωξdz+s=1mResas,0ω+ξdz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\omega_{-}\xi\,dz+\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\omega_{+}\xi\,dz
=\displaystyle= ω,ξ.\displaystyle\langle\omega,\xi\rangle.

In a similar manner, we obtain

Cξ1(z)(ω),ξ2η=\displaystyle\langle C_{\xi_{1}(z)}(\omega),\xi_{2}\rangle_{\eta}= Res((ωξ1)λξ2λ(ωλ)ξ1ξ2λ)dz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\left(\frac{(\omega\xi_{1})_{-}\lambda^{\prime}\xi_{2}}{\lambda^{\prime}}-\frac{(\omega\lambda^{\prime})_{-}\xi_{1}\xi_{2}}{\lambda^{\prime}}\right)dz
s=1mResas,0((ωξ1)+λξ2λ+(ωλ)+ξ1ξ2λ)dz\displaystyle-\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\left(-\frac{(\omega\xi_{1})_{+}\lambda^{\prime}\xi_{2}}{\lambda^{\prime}}+\frac{(\omega\lambda^{\prime})_{+}\xi_{1}\xi_{2}}{\lambda^{\prime}}\right)dz
=\displaystyle= Resωξ1ξ2dz+Res(ωλ)ξ1ξ2λdzs=1mResas,0(ωλ)+ξ1ξ2λdz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\omega\xi_{1}\xi_{2}\,dz+\mathop{\text{\rm Res}}_{\infty}\frac{(\omega\lambda^{\prime})_{-}\xi_{1}\xi_{2}}{\lambda^{\prime}}dz-\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\frac{(\omega\lambda^{\prime})_{+}\xi_{1}\xi_{2}}{\lambda^{\prime}}dz
=\displaystyle= Res(ωλξ1ξ2λ(ωλ)ξ1ξ2λ)dz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\left(\frac{\omega_{-}\lambda^{\prime}\xi_{1}\xi_{2}}{\lambda^{\prime}}-\frac{(\omega\lambda^{\prime})_{-}\xi_{1}\xi_{2}}{\lambda^{\prime}}\right)dz
s=1mResas,0(ω+λξ1ξ2λ+(ωλ)+ξ1ξ2λ)dz\displaystyle-\sum_{s=1}^{m}\mathop{\text{\rm Res}}_{a_{s,0}}\left(-\frac{\omega_{+}\lambda^{\prime}\xi_{1}\xi_{2}}{\lambda^{\prime}}+\frac{(\omega\lambda^{\prime})_{+}\xi_{1}\xi_{2}}{\lambda^{\prime}}\right)dz
=\displaystyle= c(ξ1,ξ2,η(ω)).\displaystyle c(\xi_{1},\xi_{2},\eta^{\ast}(\omega)).

The lemma is proved. ∎

3.3. Hamiltonian structure

Let LMcKPLM^{cKP} denote the loop space of McKPM^{cKP}. According to the Dubrovin-Novikov theorem, the Hamiltonian structure 𝒫\mathcal{P} on LMcKPLM^{cKP} corresponding to the flat metric ,η\langle\ ,\ \rangle_{\eta} is given by

𝒫(ω)=ηxω.\mathcal{P}(\omega)=\eta^{\ast}\cdot\nabla_{\partial_{x}}\omega.
Lemma 3.3.

The Hamiltonian operator 𝒫\mathcal{P} has the explicit form

𝒫(ω(z))={ω(z)+,λ(z)}{ω(z),λ(z)}+,\mathcal{P}(\omega(z))=\{\omega(z)_{+},\lambda(z)\}-\{\omega(z),\lambda(z)\}_{+}, (3.9)

where

{f(z,x),g(z,x)}=f(z,x)zg(z,x)xf(z,x)xg(z,x)z.\{f(z,x),g(z,x)\}=\frac{\partial f(z,x)}{\partial z}\frac{\partial g(z,x)}{\partial x}-\frac{\partial f(z,x)}{\partial x}\frac{\partial g(z,x)}{\partial z}.
Proof.

We now deduce the explict form of the operator ηω\eta^{\ast}\cdot\nabla_{\partial}\omega from the equality

ηω,ξη+ω,ξ=ω,ξ.\langle\eta^{\ast}\cdot\nabla_{\partial}\omega,\xi\rangle_{\eta}+\langle\omega,\nabla_{\partial}\xi\rangle=\partial\langle\omega,\xi\rangle.

By direct computation, we have

ω,ξ=12πis=1mγs(ωξ+ωξ)𝑑z,\partial\langle\omega,\xi\rangle=\frac{1}{2\pi\mathrm{i}}\displaystyle\sum_{s=1}^{m}\int_{\gamma_{s}}(\partial\omega\xi+\omega\partial\xi)dz,

and

ω,ξ=\displaystyle\langle\omega,\nabla_{\partial}\xi\rangle= 12πis=1mγs(ω(ξ(ξλλ),0s=1m(ξλλ)as,0,0)dz\displaystyle\frac{1}{2\pi\mathrm{i}}\displaystyle\sum_{s=1}^{m}\int_{\gamma_{s}}(\omega(\partial\xi-(\frac{\xi\partial\lambda}{\lambda^{\prime}})^{\prime}_{\infty,\geq 0}-\displaystyle\sum_{s^{\prime}=1}^{m}(\frac{\xi\partial\lambda}{\lambda^{\prime}})^{\prime}_{a_{s^{\prime},0},\leq 0})dz
=\displaystyle= 12πis=1mγs(ωξω(ξλλ),0ω+s=1m(ξλλ)as,0,0)𝑑z\displaystyle\frac{1}{2\pi\mathrm{i}}\displaystyle\sum_{s=1}^{m}\int_{\gamma_{s}}(\omega\partial\xi-\omega_{-}(\frac{\xi\partial\lambda}{\lambda^{\prime}})^{\prime}_{\infty,\geq 0}-\omega_{+}\displaystyle\sum_{s^{\prime}=1}^{m}(\frac{\xi\partial\lambda}{\lambda^{\prime}})^{\prime}_{a_{s^{\prime},0},\leq 0})dz
=\displaystyle= 12πis=1mγsωξdzResωξλλdz+s=1mResas,0ω+ξλλdz.\displaystyle\frac{1}{2\pi\mathrm{i}}\displaystyle\sum_{s=1}^{m}\int_{\gamma_{s}}\omega\partial\xi dz-\mathop{\text{\rm Res}}_{\infty}\frac{\omega^{\prime}_{-}\xi\partial\lambda}{\lambda^{\prime}}dz+\displaystyle\sum_{s^{\prime}=1}^{m}\mathop{\text{\rm Res}}_{a_{s^{\prime},0}}\frac{\omega^{\prime}_{+}\xi\partial\lambda}{\lambda^{\prime}}dz.

Hence

ηω=\displaystyle\eta^{\ast}\cdot\nabla_{\partial}\omega= ω+λω+λ(ωλωλ)+\displaystyle\omega^{\prime}_{+}\partial\lambda-\partial\omega_{+}\lambda^{\prime}-(\omega^{\prime}\partial\lambda-\partial\omega\lambda^{\prime})_{+}
=\displaystyle= ωλ+ωλ+(ωλωλ).\displaystyle-\omega^{\prime}_{-}\partial\lambda+\partial\omega_{-}\lambda^{\prime}+(\omega^{\prime}\partial\lambda-\partial\omega\lambda^{\prime})_{-}.

Setting =x\partial=\partial_{x}, we obtain the equality (3.9). ∎

3.4. principal hierarchy for McKPM^{cKP}

To prove Theorem 1.1, we first need to establish the following lemma.

Lemma 3.4.

Let Qp(λ),pQ_{p}(\lambda),p\in\mathbb{N}, be analytic functions in λ\lambda that satisfy

Qp(λ)λ=Qp1(λ).\frac{\partial Q_{p}(\lambda)}{\partial\lambda}=Q_{p-1}(\lambda).

Define

Fi,p=12πiγiQp+1(λ(z))𝑑z,F_{i,p}=\frac{1}{2\pi i}\int_{\gamma_{i}}Q_{p+1}(\lambda(z))dz, (3.10)

then we have

ηdFi,p=C(dFi,p1),\eta^{\ast}\cdot\nabla_{\partial}dF_{i,p}=C_{\partial}(dF_{i,p-1}), (3.11)

where \partial is any vector field on McKPM^{cKP}, and the operators η\eta^{\ast}\cdot\nabla_{\partial} and CC_{\partial} are given by equalities (3.7) and (3.8), respectively.

Proof.

The differential of Fi,pF_{i,p} at λ(z)McKP\lambda(z)\in M^{cKP} is given by

dFi,p=Qp+1(λ)λ1γi,dF_{i,p}=\frac{\partial Q_{p+1}(\lambda)}{\partial\lambda}\textbf{1}_{\gamma_{i}}\in\mathcal{H},

where the functions {𝟏γi}i=1m\{\mathbf{1}_{\gamma_{i}}\}_{i=1}^{m} belonging to the space \mathcal{H} are characterized by

𝟏γi|γj=δij,i,j=1,,m.\mathbf{1}_{\gamma_{i}}|_{\gamma_{j}}=\delta_{ij},\quad i,j=1,\ldots,m.

By the identity

{Qp(λ(z)),λ(z)}=0,\{Q_{p}(\lambda(z)),\lambda(z)\}=0,

where {f,g}=fggf\{f,g\}=f^{\prime}\partial g-g^{\prime}\partial f, we obtain

ηdFi,p={(Qp(λ(z))1γi)+,λ(z)}.\eta^{\ast}\cdot\nabla_{\partial}dF_{i,p}=\{(Q_{p}(\lambda(z))\textbf{1}_{\gamma_{i}})_{+},\lambda(z)\}.

For the right-hand side of equality (3.11), we have

C(dFi,p1)=\displaystyle C_{\partial}(dF_{i,p-1})= (Qp(λ)λλ(z)1γi)+λ(z)+(Qp(λ)λλ(z)1γi)+λ(z)\displaystyle-(\frac{\partial Q_{p}(\lambda)}{\partial\lambda}\partial\lambda(z)\textbf{1}_{\gamma_{i}})_{+}\lambda^{\prime}(z)+(\frac{\partial Q_{p}(\lambda)}{\partial\lambda}\lambda^{\prime}(z)\textbf{1}_{\gamma_{i}})_{+}\partial\lambda(z)
=\displaystyle= {(Qp(λ(z))1γi)+,λ(z)})\displaystyle\{(Q_{p}(\lambda(z))\textbf{1}_{\gamma_{i}})_{+},\lambda(z)\})
=\displaystyle= ηdFi,p.\displaystyle\eta^{\ast}\cdot\nabla_{\partial}dF_{i,p}.

Thus, the lemma is proved. ∎

Proof of Theorem 1.1.

For the Hamiltonian density θt0,j,p\theta_{t_{0,j},p}, let MM^{\prime} be a subset of McKPM^{cKP} such that w0=λ1n0w_{0}=\lambda^{\frac{1}{n_{0}}} can be analytically continued onto s=1mγs\cup_{s=1}^{m}\gamma_{s}, that is, w0w_{0}\in\mathcal{H}. Then on MM^{\prime}, θt0,j,p\theta_{t_{0,j},p} can be expressed as a smooth function of the form (3.10), and thus satisfies equation (2.1). The uniqueness of analytic function implies that θt0,j,p\theta_{t_{0,j},p} satisfies equation (2.1) on McKPM^{cKP}.

By a similar method, it can be shown that the remaining Hamiltonian densities θti,j,p\theta_{t_{i,j},p} also satisfy equation (2.1). In particular, for θti,ni,p\theta_{t_{i,n_{i}},p}, consider a subset MM^{\prime} of McKPM^{cKP} such that w0=λ1n0w_{0}=\lambda^{\frac{1}{n_{0}}} can be analytically continued onto s=1mγs\cup_{s=1}^{m}\gamma_{s}, with the winding number 11 along γi\gamma_{i} and 0 along γj\gamma_{j} for jij\neq i, and wi=λ1niw_{i}=\lambda^{\frac{1}{n_{i}}} can be analytically continued onto γi\gamma_{i} with the winding number 1-1. Then on MM^{\prime}, we have

θti,ni,p=\displaystyle\theta_{t_{i,n_{i}},p}= Rescpn0λpp!dz+12πiγiλpp!(logw0wicpni)𝑑z+12πisiγsλpp!logw0dz\displaystyle\mathop{\text{\rm Res}}_{\infty}\frac{c_{p}}{n_{0}}\frac{\lambda^{p}}{p!}dz+\frac{1}{2\pi\mathrm{i}}\int_{\gamma_{i}}\frac{\lambda^{p}}{p!}(\log w_{0}w_{i}-\frac{c_{p}}{n_{i}})dz+\frac{1}{2\pi\mathrm{i}}\sum_{s\neq i}\int_{\gamma_{s}}\frac{\lambda^{p}}{p!}\log w_{0}dz
=12πiγiλpp!(logw0wicpnicpn0)𝑑z+12πisiγsλpp!(logw0cpn0)𝑑z,\displaystyle=\frac{1}{2\pi\mathrm{i}}\int_{\gamma_{i}}\frac{\lambda^{p}}{p!}(\log w_{0}w_{i}-\frac{c_{p}}{n_{i}}-\frac{c_{p}}{n_{0}})dz+\frac{1}{2\pi\mathrm{i}}\sum_{s\neq i}\int_{\gamma_{s}}\frac{\lambda^{p}}{p!}(\log w_{0}-\frac{c_{p}}{n_{0}})dz,

which thus satisfies equation (2.1).

Next, we verify that θti,j,p\theta_{t_{i,j},p} satisfies equation (2.2). Introduce the operator =E+1n0zz\mathcal{E}=E+\frac{1}{n_{0}}z\frac{\partial}{\partial z}, then we have

Lieλ(z)=λ(z)\operatorname{Lie}_{\mathcal{E}}\lambda(z)=\lambda(z)

and

Resai,0Lief(λ(z))dz=LieEResai,0f(λ(z))dz1n0Resai,0f(λ(z))dz,\mathop{\text{\rm Res}}_{a_{i,0}}\operatorname{Lie}_{\mathcal{E}}f(\lambda(z))\,dz=\operatorname{Lie}_{E}\mathop{\text{\rm Res}}_{a_{i,0}}f(\lambda(z))\,dz-\frac{1}{n_{0}}\mathop{\text{\rm Res}}_{a_{i,0}}f(\lambda(z))\,dz,

which implies that

LieEθu,p(t)={(p+1jn0+1n0)θu,p(t),u=t0,j;(p+1jni+1n0)θu,p(t),u=ti,j,jni;(p+1n0)θu,p(t)+s=1m1n0θts,0,p1+1niθti,0,p1,u=ti,ni.\operatorname{Lie}_{E}\theta_{u,p}(t)=\begin{cases}\left(p+1-\frac{j}{n_{0}}+\frac{1}{n_{0}}\right)\theta_{u,p}(t),&u=t_{0,j};\\ \left(p+1-\frac{j}{n_{i}}+\frac{1}{n_{0}}\right)\theta_{u,p}(t),&u=t_{i,j},\ j\neq n_{i};\\ \left(p+\frac{1}{n_{0}}\right)\theta_{u,p}(t)+\sum_{s=1}^{m}\frac{1}{n_{0}}\theta_{t_{s,0},p-1}+\frac{1}{n_{i}}\theta_{t_{i,0},p-1},&u=t_{i,n_{i}}.\end{cases}

Hence, equation (2.2) is satisfied.

We now deduce the Hamiltonian vector fields corresponding to the densities {θti,j,p}\{\theta_{t_{i,j},p}\}. For the density θt0,j,p\theta_{t_{0,j},p}, we have

𝒫(dθt0,j,p)=\displaystyle\mathcal{P}(d\theta_{t_{0,j},p})= {(c0,j;p1w0pn0j)+,λ}\displaystyle\{(c_{0,j;p-1}w_{0}^{pn_{0}-j})_{+},\lambda\}
=\displaystyle= {(c0,j;p1w0pn0j),0,λ}.\displaystyle\{(c_{0,j;p-1}w_{0}^{pn_{0}-j})_{\infty,\geq 0},\lambda\}.

For the case of θti,j,p\theta_{t_{i,j},p} where jnij\neq n_{i}, we have

𝒫(dθti,j,p)=\displaystyle\mathcal{P}(d\theta_{t_{i,j},p})= {(ci,j;p1wipnij𝟏γi),λ}\displaystyle-\{(c_{i,j;p-1}w_{i}^{pn_{i}-j}\mathbf{1}_{\gamma_{i}})_{-},\lambda\}
=\displaystyle= {(ci,j;p1wipnij)ai,0,1,λ}.\displaystyle-\{(c_{i,j;p-1}w_{i}^{pn_{i}-j})_{a_{i,0},\leq-1},\lambda\}.

For the specific density θti,ni,p\theta_{t_{i,n_{i}},p}, we have

𝒫(dθti,ni,p+1)=\displaystyle\mathcal{P}(d\theta_{t_{i,n_{i}},p+1})= {(λpp!(logw0wicpnicpn0)1γi)+,λ}+si{(λpp!(logw0cpn0)1γs)+,λ}\displaystyle\{(\frac{\lambda^{p}}{p!}(\log w_{0}w_{i}-\frac{c_{p}}{n_{i}}-\frac{c_{p}}{n_{0}})\textbf{1}_{\gamma_{i}})_{+},\lambda\}+\displaystyle\sum_{s\neq i}\{(\frac{\lambda^{p}}{p!}(\log w_{0}-\frac{c_{p}}{n_{0}})\textbf{1}_{\gamma_{s}})_{+},\lambda\}
=\displaystyle= {(λpp!(logw0zai,0+logwi(zai,0)cpnicpn0)1γi)+,λ}\displaystyle\{(\frac{\lambda^{p}}{p!}(\log\frac{w_{0}}{z-a_{i,0}}+\log w_{i}(z-a_{i,0})-\frac{c_{p}}{n_{i}}-\frac{c_{p}}{n_{0}})\textbf{1}_{\gamma_{i}})_{+},\lambda\}
+si{(λpp!(logw0zai,0+log(zai,0)cpn0)1γs)+,λ}\displaystyle+\displaystyle\sum_{s\neq i}\{(\frac{\lambda^{p}}{p!}(\log\frac{w_{0}}{z-a_{i,0}}+\log(z-a_{i,0})-\frac{c_{p}}{n_{0}})\textbf{1}_{\gamma_{s}})_{+},\lambda\}
=\displaystyle= {(λpp!(logw0zai,0cpn0)),0,λ}{(λpp!(logwi(zai,0)cpni))ai,0,1,λ}\displaystyle\{(\frac{\lambda^{p}}{p!}(\log\frac{w_{0}}{z-a_{i,0}}-\frac{c_{p}}{n_{0}}))_{\infty,\geq 0},\lambda\}-\{(\frac{\lambda^{p}}{p!}(\log w_{i}(z-a_{i,0})-\frac{c_{p}}{n_{i}}))_{a_{i,0},\leq-1},\lambda\}
+{λpp!(logwi(zai,0)cpni)1γi,λ}+si{(log(zai,0)1γs)+,λ}\displaystyle+\{\frac{\lambda^{p}}{p!}(\log w_{i}(z-a_{i,0})-\frac{c_{p}}{n_{i}})\textbf{1}_{\gamma_{i}},\lambda\}+\displaystyle\sum_{s\neq i}\{(\log(z-a_{i,0})\textbf{1}_{\gamma_{s}})_{+},\lambda\}
=\displaystyle= {λpp!((logw0zai,0cpn0)),0,λ}{(λpp!(log(zai,0)wicpni))ai,0,1,λ}\displaystyle\{\frac{\lambda^{p}}{p!}((\log\frac{w_{0}}{z-a_{i,0}}-\frac{c_{p}}{n_{0}}))_{\infty,\geq 0},\lambda\}-\{(\frac{\lambda^{p}}{p!}(\log(z-a_{i,0})w_{i}-\frac{c_{p}}{n_{i}}))_{a_{i,0},\leq-1},\lambda\}
si{(λpp!log(zai,0))as,0,1,λ}+{λpp!log(zai,0),λ}.\displaystyle-\sum_{s\neq i}\{(\frac{\lambda^{p}}{p!}\log(z-a_{i,0}))_{a_{s,0},\leq-1},\lambda\}+\{\frac{\lambda^{p}}{p!}\log(z-a_{i,0}),\lambda\}.

Thus, the theorem is proved. ∎

3.5. principal hierarchy for MDcKPM^{D-cKP}

Let MDcKPM^{D-cKP} be the submanifold of McKPM^{cKP} consisting of elements of the form (1.2). Let us first show that MDcKPM^{D-cKP} is a natural Frobenius submanifold of McKPM^{cKP}, as defined by Strachan [16].

For λ(z)MDcKP\lambda(z)\in M^{D-cKP}, we have

w0(z)=w0(z),z;\displaystyle w_{0}(-z)=-w_{0}(z),\quad z\to\infty;
w1(z)=w1(z),z0;\displaystyle w_{1}(-z)=-w_{1}(z),\quad z\to 0;
w2i2(z)=w2i1(z),pb2i1,0,i=2,3,,m,\displaystyle w_{2i-2}(-z)=w_{2i-1}(z),\quad p\to b_{2i-1,0},\ i=2,3,\cdots,m^{\prime},

where b2i1,0=b2i2,0b_{2i-1,0}=-b_{2i-2,0}. Hence, for pp\to\infty, we have

z=\displaystyle-z= w0(z)t0,1w01(z)t0,2w02(z)\displaystyle w_{0}(-z)-t_{0,1}w_{0}^{-1}(-z)-t_{0,2}w_{0}^{-2}(-z)-\cdots
=\displaystyle= w0(z)+t0,1w01(z)t0,2w02(z)\displaystyle-w_{0}(z)+t_{0,1}w_{0}^{-1}(z)-t_{0,2}w_{0}^{-2}(z)-\cdots
=\displaystyle= w0(z)+t0,1w01(z)+t0,2w02(z)+.\displaystyle-w_{0}(z)+t_{0,1}w_{0}^{-1}(z)+t_{0,2}w_{0}^{-2}(z)+\cdots.

For p0p\to 0, we have

z=\displaystyle-z= t1,0+t1,1w11(z)+t1,2w12(z)+\displaystyle t_{1,0}+t_{1,1}w_{1}^{-1}(-z)+t_{1,2}w_{1}^{-2}(-z)+\cdots
=\displaystyle= t1,1w11(z)+t1,2w12(z)+\displaystyle t_{1,1}w_{1}^{-1}(-z)+t_{1,2}w_{1}^{-2}(-z)+\cdots
=\displaystyle= t1,1w11(z)+t1,2w12(z)\displaystyle-t_{1,1}w_{1}^{-1}(z)+t_{1,2}w_{1}^{-2}(z)-\cdots
=\displaystyle= t1,1w11(z)t1,2w12(z)+.\displaystyle-t_{1,1}w_{1}^{-1}(z)-t_{1,2}w_{1}^{-2}(z)+\cdots.

For pb2i1,0,i=2,,m,p\to b_{2i-1,0},\ i=2,\cdots,m^{\prime}, we have

z=\displaystyle-z= t2i1,0+t2i1,1w2i11(z)+t2i1,2w2i12(z)+\displaystyle t_{2i-1,0}+t_{2i-1,1}w_{2i-1}^{-1}(-z)+t_{2i-1,2}w_{2i-1}^{-2}(-z)+\cdots
=\displaystyle= t2i1,0+t2i1,1w2i21(z)+t2i1,2w2i22(z)+\displaystyle t_{2i-1,0}+t_{2i-1,1}w_{2i-2}^{-1}(z)+t_{2i-1,2}w_{2i-2}^{-2}(z)+\cdots
=\displaystyle= t2i2,0t2i2,1w2i21(z)t2i2,2w2i22(z).\displaystyle-t_{2i-2,0}-t_{2i-2,1}w_{2i-2}^{-1}(z)-t_{2i-2,2}w_{2i-2}^{-2}(z)-\cdots.

Thus, we obtain the following restrictions for the flat coordinates:

t0,2=t0,4==t0,2n02=0;\displaystyle t_{0,2}=t_{0,4}=\cdots=t_{0,2n_{0}^{\prime}-2}=0;
t1,0=t1,2==t1,2n1=0;\displaystyle t_{1,0}=t_{1,2}=\cdots=t_{1,2n_{1}^{\prime}}=0;
t2i2,j=t2i1,j,i=2,,m,j=0,,ni.\displaystyle t_{2i-2,j}=-t_{2i-1,j},\quad i=2,\cdots,m^{\prime},\ j=0,\cdots,n_{i}^{\prime}.

Hence, MDcKPM^{D-cKP} is a flat submanifold of McKPM^{cKP} with flat coordinates

tDcKP={t0,2j1}j=1n0{t1,2j1}j=1n1{t2,j}j=0n2{t4,j}j=0n3{t2m2,j}j=0nm.\textbf{t}^{D-cKP}=\{t_{0,2j-1}\}_{j=1}^{n_{0}^{\prime}}\cup\{t_{1,2j-1}\}_{j=1}^{n_{1}^{\prime}}\cup\{t_{2,j}\}_{j=0}^{n_{2}^{\prime}}\cup\{t_{4,j}\}_{j=0}^{n_{3}^{\prime}}\cup\cdots\cup\{t_{2m^{\prime}-2,j}\}_{j=0}^{n_{m^{\prime}}^{\prime}}.

On the other hand, let p1,,prp_{1},\cdots,p_{r} be the simple critical points of λ(z)McKP\lambda(z)\in M^{cKP}, where r=dim(McKP)r=dim(M^{cKP}). . For λ(z)MDcKP\lambda(z)\in M^{D-cKP}, we can choose

p2j=p2j1,j=1,2,,r2,p_{2j}=-p_{2j-1},\quad j=1,2,\cdots,\frac{r}{2},

which implies

u2j=u2j1,j=1,2,,r2.u_{2j}=u_{2j-1},\quad j=1,2,\cdots,\frac{r}{2}.

Thus, MDcKPM^{D-cKP} is a caustic submanifold of McKPM^{cKP} (for the definition, see reference [17]). According to Corollary 3.7 in [17], MDcKPM^{D-cKP} is a natural Frobenius submanifold of McKPM^{cKP}.

Remark 3.5.

Observe that the above conclusion also holds for the case n1=0n_{1}^{\prime}=0. In this case, we have

r=dim(McKP)=2dim(MDcKP)1.r=\dim(M^{cKP})=2\dim(M^{D-cKP})-1.

Choose simple critical points p1,,prp_{1},\ldots,p_{r} such that

p1=0,p2j=p2j+1,j=1,2,,r12,p_{1}=0,\quad p_{2j}=-p_{2j+1},\quad j=1,2,\ldots,\frac{r-1}{2},

then we have

u2j=u2j+1,j=1,2,,r12.u_{2j}=u_{2j+1},\quad j=1,2,\ldots,\frac{r-1}{2}.

We now consider the principal hierarchy for MDcKPM^{D-cKP}. Let odd\mathcal{H}^{odd} be the space consisting of elements in \mathcal{H} that satisfy the condition ω(z)=ω(z)\omega(-z)=-\omega(z). For ω(z)odd\omega(z)\in\mathcal{H}^{odd}, we have

ω(z)={ω(z),if z,zDc,ω+(z),if zDc,zD,\omega_{-}(-z)=\begin{cases}-\omega_{-}(z),&\text{if }z,-z\in D^{c},\\ \omega_{+}(z),&\text{if }-z\in D^{c},z\in D,\end{cases}

and

ω+(z)={ω+(z),if z,zD,ω(z),if zD,zDc.\omega_{+}(-z)=\begin{cases}\omega_{+}(z),&\text{if }z,-z\in D,\\ -\omega_{-}(z),&\text{if }-z\in D,z\in D^{c}.\end{cases}

Thus, for λ(z)MDcKP\lambda(z)\in M^{D-cKP}, we have η(ω(z))Tλ(z)MDcKP\eta^{\ast}(\omega(z))\in T_{\lambda(z)}M^{D-cKP}, and

ηωTλ(z)MDcKP\eta^{\ast}\cdot\nabla_{\partial}\omega\in T_{\lambda(z)}M^{D-cKP}

for any Tλ(z)MDcKP\partial\in T_{\lambda(z)}M^{D-cKP}. We observe that the differentials of the Hamiltonian densities

θt0,2j1,p,j=1,2,,n0;\displaystyle\theta_{t_{0,2j-1},p},\quad j=1,2,\cdots,n_{0}^{\prime};
θt1,2j1,p,j=1,2,,n1;\displaystyle\theta_{t_{1,2j-1},p},\quad j=1,2,\cdots,n_{1}^{\prime};

and

θt2i2,j,pθt2i1,j,p,i=2,,m,j=0,,ni\theta_{t_{2i-2,j},p}-\theta_{t_{2i-1,j},p},\quad i=2,\cdots,m^{\prime},\ j=0,\cdots,n_{i}^{\prime}

belong to odd\mathcal{H}^{odd}, hence the corresponding Hamiltonian vector fields form the principal hierarchy for MDcKPM^{D-cKP}.

3.6. principal hierarchy for M^cKP\hat{M}^{cKP}

Let M^cKP\hat{M}^{cKP} be the almost duality of the Frobenius manifold McKPM^{cKP}. We now construct the principal hierarchy for M^cKP\hat{M}^{cKP}.

Suppose IDI\in\textbf{D}, where II is the set of all zeros and poles of λ(z)\lambda(z) in \mathbb{C}. Define the linear map g(ω(z))=CE(ω(z))g^{\ast}(\omega(z))=C_{E}(\omega(z)) from \mathcal{H} to Tλ(z)MT_{\lambda(z)}M as:

g(ω(z))=(λ(z)ω(z))λ(z)λ(z)(λ(z)ω(z))+λ(z)n012πis=1mγsλ(z)ω(z)𝑑z.g^{\ast}(\omega(z))=(\lambda(z)\omega(z))_{-}\lambda^{\prime}(z)-\lambda(z)(\lambda^{\prime}(z)\omega(z))_{-}+\frac{\lambda^{\prime}(z)}{n_{0}}\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\lambda^{\prime}(z)\omega(z)\,dz.

We have

ω(z),ξ(z)=(g(ω(z)),ξ(z))g,\langle\omega(z),\xi(z)\rangle=(g^{\ast}(\omega(z)),\xi(z))_{g},

where

(1,2)g=Resdλ=01logλ(z)2logλ(z)(logλ(z))dz(\partial_{1},\partial_{2})_{g}=\sum\mathop{\text{\rm Res}}_{d\lambda=0}\frac{\partial_{1}\log\lambda(z)\partial_{2}\log\lambda(z)}{(\log\lambda(z))^{\prime}}\,dz

is the intersection form on McKPM^{cKP}. By the following lemma, we derive the explicit form of the Hamiltonian structure 𝒫^\hat{\mathcal{P}} for the flat metric gg.

Lemma 3.6.

Let ^\hat{\nabla} be the Levi-Civita connection associated with gg. Then,

^12λ(z)=12λ(z)1λ(z)2λ(z)λ(z)λ(z)qI(1λ(z)2λ(z)λ(z)λ(z))q,1.\hat{\nabla}_{\partial_{1}}\partial_{2}\cdot\lambda(z)=\partial_{1}\partial_{2}\lambda(z)-\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda(z)}-\lambda(z)\sum_{q\in I}\left(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda(z)\lambda^{\prime}(z)}\right)_{q,\leq-1}^{\prime}.

Morever, we have

g^ω(z)={ω(z),λ(z)}λ(z){(ω(z)λ(z)),λ(z)}λ(z)n012πis=1mγs{ω(z),λ(z)}𝑑z,g^{\ast}\cdot\hat{\nabla}_{\partial}\omega(z)=\{\omega(z),\lambda(z)\}_{-}\lambda(z)-\{(\omega(z)\lambda(z))_{-},\lambda(z)\}-\frac{\lambda^{\prime}(z)}{n_{0}}\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\{\omega(z),\lambda(z)\}\,dz, (3.12)

where {f,g}=fggf\{f,g\}=f^{\prime}\partial g-g^{\prime}\partial f.

Proof.

The compatibility of ^\hat{\nabla} with gg can be verified as follows:

(^12,3)g+(2,^13)g\displaystyle(\hat{\nabla}_{\partial_{1}}\partial_{2},\partial_{3})_{g}+(\partial_{2},\hat{\nabla}_{\partial_{1}}\partial_{3})_{g}
=\displaystyle= (ResqIResq)(12λ(z)3λ(z)λ(z)λ(z)dz+2λ(z)13λ(z)λ(z)λ(z)dz)+K,\displaystyle(-\mathop{\text{\rm Res}}_{\infty}-\sum_{q\in I}\mathop{\text{\rm Res}}_{q})(\frac{\partial_{1}\partial_{2}\lambda(z)\partial_{3}\lambda(z)}{\lambda^{\prime}(z)\lambda(z)}dz+\frac{\partial_{2}\lambda(z)\partial_{1}\partial_{3}\lambda(z)}{\lambda^{\prime}(z)\lambda(z)}dz)+K,

where

K=\displaystyle K= (ResqIResq)(1λ(z)2λ(z)3λ(z)dzλ(z)λ(z)2(1λ(z)2λ(z)λ(z)λ(z))3λ(z)dzλ(z)+c.p.{2,3})\displaystyle(-\mathop{\text{\rm Res}}_{\infty}-\sum_{q\in I}\mathop{\text{\rm Res}}_{q})(-\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)\partial_{3}\lambda(z)dz}{\lambda^{\prime}(z)\lambda(z)^{2}}-(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda(z)\lambda^{\prime}(z)})^{\prime}\frac{\partial_{3}\lambda(z)dz}{\lambda^{\prime}(z)}+c.p.\{2,3\})
=\displaystyle= (ResqIResq)((1λ(z)2λ(z)3λ(z)dzλ(z)λ(z))(1λ(z))+1λ(z)2λ(z)dzλ(z)λ(z)(3λ(z)λ(z))+c.p.{2,3})\displaystyle(-\mathop{\text{\rm Res}}_{\infty}-\sum_{q\in I}\mathop{\text{\rm Res}}_{q})((\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)\partial_{3}\lambda(z)dz}{\lambda^{\prime}(z)\lambda^{\prime}(z)})(\frac{1}{\lambda(z)})^{\prime}+\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)dz}{\lambda(z)\lambda^{\prime}(z)}(\frac{\partial_{3}\lambda(z)}{\lambda^{\prime}(z)})^{\prime}+c.p.\{2,3\})
=\displaystyle= (ResqIResq)((1λ(z)2λ(z)dzλ(z))(3λ(z)λ(z)λ(z))+c.p.{2,3})\displaystyle(-\mathop{\text{\rm Res}}_{\infty}-\sum_{q\in I}\mathop{\text{\rm Res}}_{q})((\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)dz}{\lambda^{\prime}(z)})(\frac{\partial_{3}\lambda(z)}{\lambda^{\prime}(z)\lambda(z)})^{\prime}+c.p.\{2,3\})
=\displaystyle= (ResqIResq)λ(z)1λ(z)(2λ(z)3λ(z)(λ(z)λ(z))2)dz\displaystyle(-\mathop{\text{\rm Res}}_{\infty}-\sum_{q\in I}\mathop{\text{\rm Res}}_{q})\lambda(z)\partial_{1}\lambda(z)(\frac{\partial_{2}\lambda(z)\partial_{3}\lambda(z)}{(\lambda^{\prime}(z)\lambda(z))^{2}})^{\prime}dz
=\displaystyle= (ResqIResq)2λ(z)3λ(z)1(1λ(z)λ(z))dz.\displaystyle(-\mathop{\text{\rm Res}}_{\infty}-\sum_{q\in I}\mathop{\text{\rm Res}}_{q})\partial_{2}\lambda(z)\partial_{3}\lambda(z)\partial_{1}(\frac{1}{\lambda^{\prime}(z)\lambda(z)})dz.

This implies that

(^12,3)g+(2,^13)g=^1(2,3)g.(\hat{\nabla}_{\partial_{1}}\partial_{2},\partial_{3})_{g}+(\partial_{2},\hat{\nabla}_{\partial_{1}}\partial_{3})_{g}=\hat{\nabla}_{\partial_{1}}(\partial_{2},\partial_{3})_{g}.

On the other hand,

ω,^ξ=\displaystyle\langle\omega,\hat{\nabla}_{\partial}\xi\rangle= s=1mγs(ωξωξλλ+(ωλ)qI(ξλλλ)q,1)𝑑z\displaystyle\sum_{s=1}^{m}\int_{\gamma_{s}}(\omega\partial\xi-\frac{\omega\xi\partial\lambda}{\lambda}+(\omega\lambda)^{\prime}\sum_{q\in I}(\frac{\xi\partial\lambda}{\lambda\lambda^{\prime}})_{q,\leq-1})dz
=s=1mγs(ωξ(ωλ)+ξλ+(ωλ)+qI(ξλλλ)q,1)𝑑z\displaystyle=\sum_{s=1}^{m}\int_{\gamma_{s}}(\omega\partial\xi-\frac{(\omega\partial\lambda)_{+}\xi}{\lambda}+(\omega\lambda)^{\prime}_{+}\sum_{q\in I}(\frac{\xi\partial\lambda}{\lambda\lambda^{\prime}})_{q,\leq-1})dz
=s=1mγs(ωξ(ωλ)+ξλ+(ωλ)+ξλλλ)𝑑z,\displaystyle=\sum_{s=1}^{m}\int_{\gamma_{s}}(\omega\partial\xi-\frac{(\omega\partial\lambda)_{+}\xi}{\lambda}+(\omega\lambda)^{\prime}_{+}\frac{\xi\partial\lambda}{\lambda\lambda^{\prime}})dz,

and

(g^ω,ξ)g=\displaystyle(g^{\ast}\cdot\hat{\nabla}_{\partial}\omega,\xi)_{g}= qIResq((ωλλω)+ξλ(ωλ)+ξλλλ+(λω+ωλ)+ξλ)dz\displaystyle\sum_{q\in I}\mathop{\text{\rm Res}}_{q}(\frac{(\omega^{\prime}\partial\lambda-\lambda^{\prime}\partial\omega)_{+}\xi}{\lambda^{\prime}}-\frac{(\omega\lambda^{\prime})_{+}\xi\partial\lambda}{\lambda\lambda^{\prime}}+\frac{(\lambda\partial\omega+\omega\partial\lambda)_{+}\xi}{\lambda})dz
=\displaystyle= s=1mγs((ωλ)+ξλλλ+(λω+ωλ)+ξλ)𝑑z\displaystyle\sum_{s=1}^{m}\int_{\gamma_{s}}(-\frac{(\omega\lambda^{\prime})_{+}\xi\partial\lambda}{\lambda\lambda^{\prime}}+\frac{(\lambda\partial\omega+\omega\partial\lambda)_{+}\xi}{\lambda})dz
=\displaystyle= s=1mγs((ωλ)+ξλλλ+(ωλ)+ξλ+ξω)𝑑z.\displaystyle\sum_{s=1}^{m}\int_{\gamma_{s}}(-\frac{(\omega\lambda^{\prime})_{+}\xi\partial\lambda}{\lambda\lambda^{\prime}}+\frac{(\omega\partial\lambda)_{+}\xi}{\lambda}+\xi\partial\omega)dz.

Hence

(g^ω,ξ)g+ω,^ξ=ω,ξ.(g^{\ast}\cdot\hat{\nabla}_{\partial}\omega,\xi)_{g}+\langle\omega,\hat{\nabla}_{\partial}\xi\rangle=\partial\langle\omega,\xi\rangle.

The lemma is proved. ∎

Proof of Theorem 1.2.

Let Fγj,p=12πiγjQ~p(λ)𝑑zF_{\gamma_{j},p}=\frac{1}{2\pi\mathrm{i}}\int_{\gamma_{j}}\tilde{Q}_{p}(\lambda)\,dz, where Qp(λ)=Q~p(λ)λQ_{p}(\lambda)=\frac{\partial\tilde{Q}_{p}(\lambda)}{\partial\lambda} satisfies the recurrence relation:

λQp+1λ+Qp+1=Qp.\lambda\frac{\partial Q_{p+1}}{\partial\lambda}+Q_{p+1}=Q_{p}.

Then we have dFp=Qp𝟏γjdF_{p}=Q_{p}\mathbf{1}_{\gamma_{j}}, and

g^dFp+1=\displaystyle g^{\ast}\cdot\hat{\nabla}_{\partial}dF_{p+1}= ((Qp+1λλλ𝟏γj+Qp+1λ𝟏γj)λ)+((Qp+1λλλ𝟏γj+Qp+1λ𝟏γj)λ)\displaystyle-((\frac{\partial Q_{p+1}}{\partial\lambda}\lambda\lambda^{\prime}\mathbf{1}_{\gamma_{j}}+Q_{p+1}\lambda^{\prime}\mathbf{1}_{\gamma_{j}})_{-}\partial\lambda)+((\frac{\partial Q_{p+1}}{\partial\lambda}\lambda\partial\lambda\mathbf{1}_{\gamma_{j}}+Q_{p+1}\partial\lambda\mathbf{1}_{\gamma_{j}})_{-}\lambda^{\prime})
=\displaystyle= (Qpλ𝟏γj)λ(Qpλ𝟏γj)λ\displaystyle(Q_{p}\partial\lambda\mathbf{1}_{\gamma_{j}})_{-}\lambda^{\prime}-(Q_{p}\lambda^{\prime}\mathbf{1}_{\gamma_{j}})_{-}\partial\lambda
=\displaystyle= C(dFp).\displaystyle C_{\partial}(dF_{p}).

Consider a subset M^\hat{M}^{\prime} of M^cKP\hat{M}^{cKP} such that for any λ(z)M^\lambda(z)\in\hat{M}^{\prime}, the winding number of λ(z)\lambda(z) along γj\gamma_{j} is zero. Let Q~(p)=(log(λ(z)))p+1(p+1)!\tilde{Q}(p)=\frac{(\log(\lambda(z)))^{p+1}}{(p+1)!}, then

Fγj,p=12πiγ(log(λ(z)))p+1(p+1)!𝑑zF_{\gamma_{j},p}=\frac{1}{2\pi\mathrm{i}}\int_{\gamma}\frac{(\log(\lambda(z)))^{p+1}}{(p+1)!}\,dz

satisfies equality (2.1) on M^\hat{M}^{\prime}. Furthermore, by the uniqueness of analytic function, for any γ~[γj]\tilde{\gamma}\in[\gamma_{j}], Fγ~,pF_{\tilde{\gamma},p} satisfies equality (2.1) on M^cKP\hat{M}^{cKP}, where [γj][\gamma_{j}] denotes the homotopy equivalence class of γj\gamma_{j} in I\mathbb{C}-I.

In particular, let q1,q_{1},\cdots and qk,p1,,pkq_{k},p_{1},\cdots,p_{k} be the zeros and poles of λ(z)\lambda(z) within the region surrounded by γ~\tilde{\gamma}, respectively, then

Fγ~,0=\displaystyle F_{\tilde{\gamma},0}= 12πiγ~log(λ(z))𝑑z\displaystyle\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}\log(\lambda(z))\,dz
=\displaystyle= 12πiγ~log(λ(z)s=1kzpszqs)𝑑z+12πiγ~log(s=1kzqszps)𝑑z\displaystyle\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}\log(\lambda(z)\prod_{s=1}^{k}\frac{z-p_{s}}{z-q_{s}})\,dz+\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}\log(\prod_{s=1}^{k}\frac{z-q_{s}}{z-p_{s}})\,dz
=\displaystyle= Reslog(s=1kzqszps)dz\displaystyle-\mathop{\text{\rm Res}}_{\infty}\log(\prod_{s=1}^{k}\frac{z-q_{s}}{z-p_{s}})\,dz
=\displaystyle= s=1kpss=1kqs.\displaystyle\sum_{s=1}^{k}p_{s}-\sum_{s=1}^{k}q_{s}.

This completes the proof of the theorem. ∎

Corollary 3.7.

The Hamiltonian vector field associated with the Hamiltonian density Fγj,pF_{\gamma_{j},p} for the Hamiltonian structure 𝒫^\hat{\mathcal{P}} takes the form:

λ(z)T^j,p={((log(λ(z)))pp!1γj)+,λ(z)},\frac{\partial\lambda(z)}{\partial\hat{T}^{j,p}}=\{(\frac{(\log(\lambda(z)))^{p}}{p!}\textbf{1}_{\gamma_{j}})_{+},\lambda(z)\},

where {f,g}=fxggxf.\{f,g\}=f^{\prime}\partial_{x}g-g^{\prime}\partial_{x}f.

3.7. rank-1 extension

To construct a rank-1 extension of the Frobenius manifold McKPM^{cKP}, we need the following lemma

Lemma 3.8.

For any vector fields 1,2\partial_{1},\partial_{2} on McKPM^{cKP}, we have

1λ(z)2λ(z)1λ(z)2λ(z)λ=(1λ(z)2λ(z)λ(z)),0+j=1m(1λ(z)2λ(z)λ(z))φj,1.\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)-\partial_{1}\lambda(z)\circ\partial_{2}\lambda(z)}{\lambda^{\prime}}=(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda^{\prime}(z)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda^{\prime}(z)})_{\varphi_{j},\leq-1}.
Proof.

Assume without loss of generality that 1=ui,2=uj\partial_{1}=\partial_{u_{i}},\partial_{2}=\partial_{u_{j}}, where {ui}i=1n\{u_{i}\}_{i=1}^{n} are the canonical coordinates on McKPM^{cKP}. Using the property

uiλ(z)|pj=δij,i,j=1,,n,\partial_{u_{i}}\lambda(z)|_{p_{j}}=\delta_{ij},\quad i,j=1,\cdots,n,

we obtain

1λ(z)2λ(z)1λ(z)2λ(z)λ(z)\displaystyle\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)-\partial_{1}\lambda(z)\circ\partial_{2}\lambda(z)}{\lambda^{\prime}(z)}
=\displaystyle= (1λ(z)2λ(z)1λ(z)2λ(z)λ(z)),0+j=1m(1λ(z)2λ(z)1λ(z)2λ(z)λ(z))φj,1\displaystyle(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)-\partial_{1}\lambda(z)\circ\partial_{2}\lambda(z)}{\lambda^{\prime}(z)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)-\partial_{1}\lambda(z)\circ\partial_{2}\lambda(z)}{\lambda^{\prime}(z)})_{\varphi_{j},\leq-1}
=\displaystyle= (1λ(z)2λ(z)λ(z)),0+j=1m(1λ(z)2λ(z)λ(z))φj,1.\displaystyle(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda^{\prime}(z)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda^{\prime}(z)})_{\varphi_{j},\leq-1}.

Define Ω(t,s)\Omega(\textbf{t},s) such that

sαΩ(t,s)=αλ(s),s2Ω(t,s)=λ(s),\partial_{s}\partial_{\alpha}\Omega(\textbf{t},s)=\partial_{\alpha}\lambda(s),\quad\partial^{2}_{s}\Omega(\textbf{t},s)=\lambda^{\prime}(s),

and

αβΩ(t,s)=(αλ(s)βλ(s)λ(s)),0+j=1m(αλ(s)βλ(s)λ(s))φj,1.\partial_{\alpha}\partial_{\beta}\Omega(\textbf{t},s)=(\frac{\partial_{\alpha}\lambda(s)\partial_{\beta}\lambda(s)}{\lambda^{\prime}(s)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{\alpha}\lambda(s)\partial_{\beta}\lambda(s)}{\lambda^{\prime}(s)})_{\varphi_{j},\leq-1}.

From Lemmas 3.8 and 3.1, we obtain that ω(t,s)=sΩ(t,s)\omega(\textbf{t},s)=\partial_{s}\Omega(\textbf{t},s) satisfies the condition of Lemma 2.1, thus defining a flat F-manifold structure on McKP×M^{cKP}\times\mathbb{C}, with the multiplication of the form:

(α,0)(β,0)=(αβ,αβΩ(t,s)s),\displaystyle(\partial_{\alpha},0)\star(\partial_{\beta},0)=(\partial_{\alpha}\circ\partial_{\beta},\partial_{\alpha}\partial_{\beta}\Omega(\textbf{t},s)\cdot\partial_{s}),
(α,0)(0,s)=(0,αλ(s)s),\displaystyle(\partial_{\alpha},0)\star(0,\partial_{s})=(0,\partial_{\alpha}\lambda(s)\cdot\partial_{s}),
(0,s)(0,s)=(0,λ(s)s).\displaystyle(0,\partial_{s})\star(0,\partial_{s})=(0,\lambda^{\prime}(s)\cdot\partial_{s}).

Here, we denote a vector field on McKP×M^{cKP}\times\mathbb{C} by X=(X¯,X(s)s)X=(\bar{X},X(s)\partial_{s}) , where X¯\bar{X} and X(s)sX(s)\partial_{s} are its components along McKPM^{cKP} and \mathbb{C}, respectively. Let \nabla denote the flat connection on the tangent bundle of McKP×M^{cKP}\times\mathbb{C}, with flat coordinates t{s}\textbf{t}\cup\{s\}. We aim to determine the vector fields

Θ,p=(Θ¯,p,Θ,p(s)s),t{s},\Theta_{\bullet,p}=(\bar{\Theta}_{\bullet,p},\Theta_{\bullet,p}(s)\partial_{s}),\quad\bullet\in\textbf{t}\cup\{s\},

satisfying

XΘ,p+1=XΘ,p\nabla_{X}\Theta_{\bullet,p+1}=X\star\Theta_{\bullet,p} (3.13)

and

Θti,j,p=(ti,j,0),Θs,0=(0,s).\Theta_{t_{i,j},p}=(\frac{\partial}{\partial t^{i,j}},0),\quad\Theta_{s,0}=(0,\partial_{s}). (3.14)

The left-hand side of (3.13) has the form:

(X¯Θ¯,p+1+X(s)sΘ¯,p+1,X¯Θ,p+1(s)s+X(s)sΘ,p+1(s)s).(\nabla_{\bar{X}}\bar{\Theta}_{\bullet,p+1}+X(s)\nabla_{\partial_{s}}\bar{\Theta}_{\bullet,p+1},\partial_{\bar{X}}\Theta_{\bullet,p+1}(s)\partial_{s}+X(s)\partial_{s}\Theta_{\bullet,p+1}(s)\cdot\partial_{s}).

For the right-hand side, we have

(X¯,0)(Θ¯,p,0)=(X¯Θ¯,p,((X¯λ(s)Θ¯,pλ(s)λ(s)),0+j=1m(X¯λ(s)Θ¯,pλ(s)λ(s))φj,1)s),\displaystyle(\bar{X},0)\star(\bar{\Theta}_{\bullet,p},0)=(\bar{X}\circ\bar{\Theta}_{\bullet,p},((\frac{\partial_{\bar{X}}\lambda(s)\partial_{\bar{\Theta}_{\bullet,p}}\lambda(s)}{\lambda^{\prime}(s)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{\bar{X}}\lambda(s)\partial_{\bar{\Theta}_{\bullet,p}}\lambda(s)}{\lambda^{\prime}(s)})_{\varphi_{j},\leq-1})\partial_{s}),
(X¯,0)(0,Θ,p(s)s)=(0,(X¯λ(s))Θ,p(s)s),\displaystyle(\bar{X},0)\star(0,\Theta_{\bullet,p}(s)\partial_{s})=(0,(\partial_{\bar{X}}\lambda(s))\Theta_{\bullet,p}(s)\partial_{s}),
(0,X(s)s)(Θ¯,p,0)=(0,(Θ¯,pλ(s))X(s)s),\displaystyle(0,X(s)\partial_{s})\star(\bar{\Theta}_{\bullet,p},0)=(0,(\partial_{\bar{\Theta}_{\bullet,p}}\lambda(s))X(s)\partial_{s}),
(0,X(s)s)(0,Θ,p(s)s)=(0,λ(s)Θ,pX(s)s).\displaystyle(0,X(s)\partial_{s})\star(0,\Theta_{\bullet,p}(s)\partial_{s})=(0,\lambda^{\prime}(s)\Theta_{\bullet,p}X(s)\partial_{s}).
Theorem 3.9.

The vector fields on McKP×M^{cKP}\times\mathbb{C} of the form

Θs,p=(0,λ(s)pp!s)\displaystyle\Theta_{s,p}=(0,\frac{\lambda(s)^{p}}{p!}\partial_{s})
Θt0,n0j,p=(η(dθt0,j,p),(dθt0,j,p|z=s)+s),\displaystyle\Theta_{t_{0,n_{0}-j},p}=(\eta^{\ast}(d\theta_{t_{0,j},p}),(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}),
Θti,nij,p=(η(dθti,j,p),(dθti,j,p|z=s)s),i=1,,m.\displaystyle\Theta_{t_{i,n_{i}-j},p}=(\eta^{\ast}(d\theta_{t_{i,j},p}),-(d\theta_{t_{i,j},p}|_{z=s})_{-}\partial_{s}),\quad i=1,\cdots,m.

satisfy equations (3.13) and (3.14).

Proof.

Consider the case of Θt0,n0j,p\Theta_{t_{0,n_{0}-j},p} . The left-hand side of equation (3.13) is:

(X¯η(dθt0,j,p+1),(X¯λ(s)dθt0,j,p|z=s)+s+X(s)(λ(s)dθt0,j,p|z=s)+s)\displaystyle(\nabla_{\bar{X}}\eta^{\ast}(d\theta_{t_{0,j},p+1}),(\partial_{\bar{X}}\lambda(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}+X(s)(\lambda^{\prime}(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s})
=\displaystyle= (X¯η(dθt0,j,p),(X¯λ(s)dθt0,j,p|z=s)+s+X(s)(λ(s)dθt0,j,p|z=s)+s).\displaystyle(\bar{X}\circ\eta^{\ast}(d\theta_{t_{0,j},p}),(\partial_{\bar{X}}\lambda(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}+X(s)(\lambda^{\prime}(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}).

The right-hand side of (3.13) is:

(X¯η(dθt0,j,p),),\displaystyle(\bar{X}\circ\eta^{\ast}(d\theta_{t_{0,j},p}),\mathcal{I}),

where

=\displaystyle\mathcal{I}= X¯λ(s)η(dθt0,j,p)|z=sCX¯(dθt0,j,p)|z=sλ(s)s+X¯λ(s)(dθt0,j,p|z=s)+s\displaystyle\frac{\partial_{\bar{X}}\lambda(s)\cdot\eta^{\ast}(d\theta_{t_{0,j},p})|_{z=s}-C_{\bar{X}}(d\theta_{t_{0,j},p})|_{z=s}}{\lambda^{\prime}(s)}\partial_{s}+\partial_{\bar{X}}\lambda(s)\cdot(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}
+X(s)η(dθt0,j,p)|z=ss+X(s)λ(s)(dθt0,j,p|z=s)+s\displaystyle+X(s)\eta^{\ast}(d\theta_{t_{0,j},p})|_{z=s}\partial_{s}+X(s)\lambda^{\prime}(s)\cdot(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}
=\displaystyle= (X¯λ(s)dθt0,j,p|z=s)+sX¯λ(s)(dθt0,j,p|z=s)+s+X¯λ(s)(dθt0,j,p|z=s)+s\displaystyle(\partial_{\bar{X}}\lambda(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}-\partial_{\bar{X}}\lambda(s)\cdot(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}+\partial_{\bar{X}}\lambda(s)\cdot(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}
+X(s)(λ(s)dθt0,j,p|z=s)+s\displaystyle+X(s)(\lambda^{\prime}(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}
=\displaystyle= (X¯λ(s)dθt0,j,p|z=s)+s+X(s)(λ(s)dθt0,j,p|z=s)+s.\displaystyle(\partial_{\bar{X}}\lambda(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}+X(s)(\lambda^{\prime}(s)\cdot d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}.

The remaining cases follow similarly. ∎

Proof of Theorem 1.4.

The principal hierarchy for the flat F-manifold McKP×M^{cKP}\times\mathbb{C} is defined as:

T~,p=Θ,p~x,\frac{\partial}{\partial\tilde{T}^{\bullet,p}}=\Theta_{\bullet,p}\star\tilde{\partial}_{x},

where

~x=(x,sxs).\tilde{\partial}_{x}=(\partial_{x},\frac{\partial s}{\partial x}\partial_{s}).

Consider the case of T~t0,n0j,p\frac{\partial}{\partial\tilde{T}^{t_{0,n_{0}-j},p}}. Through direct computation, we obtain:

T~t0,n0j,p=(Tt0,j,p,𝒥)\frac{\partial}{\partial\tilde{T}^{t_{0,n_{0}-j},p}}=(\frac{\partial}{\partial T^{t_{0,j},p}},\mathcal{J})

where

𝒥=\displaystyle\mathcal{J}= xλ(s)η(dθt0,j,p)|z=sCx(dθt0,j,p)|z=sλ(s)s+xλ(s)(dθt0,j,p|z=s)+s\displaystyle\frac{\partial_{x}\lambda(s)\cdot\eta^{\ast}(d\theta_{t_{0,j},p})|_{z=s}-C_{\partial_{x}}(d\theta_{t_{0,j},p})|_{z=s}}{\lambda^{\prime}(s)}\partial_{s}+\partial_{x}\lambda(s)\cdot(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}
+sxη(dθt0,j,p)|z=ss+λ(s)sx(dθt0,j,p|z=s)+s\displaystyle+\frac{\partial s}{\partial x}\cdot\eta^{\ast}(d\theta_{t_{0,j},p})|_{z=s}\partial_{s}+\lambda^{\prime}(s)\cdot\frac{\partial s}{\partial x}\cdot(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}
=\displaystyle= dxλ(s)η(dθt0,j,p)|z=sCx(dθt0,j,p)|z=sλ(s)s+dxλ(s)(dθt0,j,p|z=s)+s\displaystyle\frac{d_{x}\lambda(s)\cdot\eta^{\ast}(d\theta_{t_{0,j},p})|_{z=s}-C_{\partial_{x}}(d\theta_{t_{0,j},p})|_{z=s}}{\lambda^{\prime}(s)}\partial_{s}+d_{x}\lambda(s)\cdot(d\theta_{t_{0,j},p}|_{z=s})_{+}\partial_{s}
=\displaystyle= dxλ(s)(dθt0,j,p|z=sλ(s))+λ(s)s+(dθt0,j,p|z=sxλ(s))+sxλ(s)(dθt0,j,p|z=sλ(s))+λ(s)s\displaystyle\frac{d_{x}\lambda(s)\cdot(d\theta_{t_{0,j},p}|_{z=s}\cdot\lambda^{\prime}(s))_{+}}{\lambda^{\prime}(s)}\partial_{s}+(d\theta_{t_{0,j},p}|_{z=s}\cdot\partial_{x}\lambda(s))_{+}\partial_{s}-\frac{\partial_{x}\lambda(s)\cdot(d\theta_{t_{0,j},p}|_{z=s}\cdot\lambda^{\prime}(s))_{+}}{\lambda^{\prime}(s)}\partial_{s}
=\displaystyle= (dθt0,j,p|z=sdxλ(s))+s.\displaystyle(d\theta_{t_{0,j},p}|_{z=s}\cdot d_{x}\lambda(s))_{+}\partial_{s}.

The remaining cases follow similarly. ∎

For m=0m=0, this hierarchy is the dispersionless limit of the open Gelfand-Dickey hierarchy:

Ltp=[(Lpn0),L],stp=xρ((Lpn0)+)(1),p=1,2,\displaystyle\frac{\partial L}{\partial t_{p}}=[(L^{\frac{p}{n_{0}}}),L],\quad\frac{\partial s}{\partial t_{p}}=\partial_{x}\cdot\rho((L^{\frac{p}{n_{0}}})_{+})(1),\quad p=1,2,\cdots

where L=xn0+j=0n02aj(x)xjL=\partial_{x}^{n_{0}}+\sum_{j=0}^{n_{0}-2}a_{j}(x)\partial_{x}^{j}, and

ρ:jaj(x)xjjaj(x)(x+s(x))j\rho:\sum_{j}a_{j}(x)\partial^{j}_{x}\to\sum_{j}a_{j}(x)(\partial_{x}+s(x))^{j}

is an automorphism of the pseudo-differential operator algebra 𝒜\mathcal{A} [43].

The commutativity of the open Gelfand-Dickey hierarchy can be shown as follows. Denote

ResAdx=b1\mathop{\text{\rm Res}}Adx=b_{-1}

for A=jbjj𝒜A=\sum_{j}b_{j}\partial^{j}\in\mathcal{A}, then we have

Resρ(A)dx=ResAdx.\mathop{\text{\rm Res}}\rho(A)dx=\mathop{\text{\rm Res}}Adx.

From the equality

ρ((Lpn0)+)(1)=\displaystyle\rho((L^{\frac{p}{n_{0}}})_{+})(1)= Resρ((Lpn0)+)x1dx\displaystyle\mathop{\text{\rm Res}}\rho((L^{\frac{p}{n_{0}}})_{+})\cdot\partial_{x}^{-1}dx
=\displaystyle= Res(Lp2)+(xs(x))1dx\displaystyle\mathop{\text{\rm Res}}(L^{\frac{p}{2}})_{+}\cdot(\partial_{x}-s(x))^{-1}dx

and the zero curvature equation:

(Lpn0)+tq(Lqn0)+tp=[(Lqn0)+,(Lpn0)+],\frac{\partial(L^{\frac{p}{n_{0}}})_{+}}{\partial t_{q}}-\frac{\partial(L^{\frac{q}{n_{0}}})_{+}}{\partial t_{p}}=[(L^{\frac{q}{n_{0}}})_{+},(L^{\frac{p}{n_{0}}})_{+}],

we obtain

2stptq2stqtp\displaystyle\frac{\partial^{2}s}{\partial t_{p}\partial t_{q}}-\frac{\partial^{2}s}{\partial t_{q}\partial t_{p}} =xρ([(Lpn0)+,(Lqn0)+])(1)x(Lpn0)+(Lqn0)+(1)+x(Lqn0)+(Lpn0)+(1)\displaystyle=\partial_{x}\cdot\rho([(L^{\frac{p}{n_{0}}})_{+},(L^{\frac{q}{n_{0}}})_{+}])(1)-\partial_{x}\cdot(L^{\frac{p}{n_{0}}})_{+}\cdot(L^{\frac{q}{n_{0}}})_{+}(1)+\partial_{x}\cdot(L^{\frac{q}{n_{0}}})_{+}\cdot(L^{\frac{p}{n_{0}}})_{+}(1)
=0.\displaystyle=0.

Let s(x)=v(x)xs(x)=\frac{\partial v(x)}{\partial x}, and by applying the identity:

ev(x)xev(x)=x+s(x),e^{-v(x)}\cdot\partial_{x}\cdot e^{v(x)}=\partial_{x}+s(x),

we obtain

stp=xev(x)(L+pn0)(ev(x)).\displaystyle\frac{\partial s}{\partial t_{p}}=\partial_{x}\cdot e^{-v(x)}\cdot(L^{\frac{p}{n_{0}}}_{+})(e^{v(x)}).

Thus,

ev(x)tp=(L+pn0)(ev(x)).\displaystyle\frac{\partial e^{v(x)}}{\partial t_{p}}=(L^{\frac{p}{n_{0}}}_{+})(e^{v(x)}).

This version of the open Gelfand-Dickey hierarchy appeared in [27].

4. Frobenius manifold with trigonometric superpotential

4.1. Definition of MTodaM^{Toda}

Given positive integers n0,,nmn_{0},\ldots,n_{m}, let MTodaM^{Toda} be the space of functions

λ(φ)=1n0en0φ+a0,n01e(n01)φ++a0,0+i=1mj=1niai,j(eφai,0)j,\lambda(\varphi)=\frac{1}{n_{0}}e^{n_{0}\varphi}+a_{0,n_{0}-1}e^{(n_{0}-1)\varphi}+\cdots+a_{0,0}+\sum_{i=1}^{m}\sum_{j=1}^{n_{i}}a_{i,j}(e^{\varphi}-a_{i,0})^{-j},

where a1,0=0a_{1,0}=0. For any ,′′,′′′Tλ(z)MToda\partial^{\prime},\partial^{\prime\prime},\partial^{\prime\prime\prime}\in T_{\lambda(z)}M^{Toda}, the flat metric on MTodaM^{Toda} is defined as:

,′′η:=η(,′′)=|λ|<Resdλ=0(λ(z)dz)′′(λ(z)dz)z2dλ(z),\langle\partial^{\prime},\partial^{\prime\prime}\rangle_{\eta}:=\eta(\partial^{\prime},\partial^{\prime\prime})=\sum_{|\lambda|<\infty}\mathop{\text{\rm Res}}_{d\lambda=0}\frac{\partial^{\prime}(\lambda(z)dz)\partial^{\prime\prime}(\lambda(z)dz)}{z^{2}d\lambda(z)},

and the (0,3)(0,3)-type tensor is given by:

c(,′′,′′′):=|λ|<Resdλ=0(λ(z)dz)′′(λ(z)dz)′′′(λ(z)dz)z2dλ(z)dz,c(\partial^{\prime},\partial^{\prime\prime},\partial^{\prime\prime\prime}):=\sum_{|\lambda|<\infty}\mathop{\text{\rm Res}}_{d\lambda=0}\frac{\partial^{\prime}(\lambda(z)dz)\partial^{\prime\prime}(\lambda(z)dz)\partial^{\prime\prime\prime}(\lambda(z)dz)}{z^{2}d\lambda(z)dz},

where z=eφz=e^{\varphi}. The equality

c(,′′,′′′)=η(′′,′′′)c(\partial^{\prime},\partial^{\prime\prime},\partial^{\prime\prime\prime})=\eta(\partial^{\prime}\circ\partial^{\prime\prime},\partial^{\prime\prime\prime})

defines the multiplication structure on Tλ(z)MTodaT_{\lambda(z)}M^{Toda}. Introduce the vector fields ee and EE on MTodaM^{Toda}, such that

Lieeλ(z)=1,LieEλ(z)=λ(z)zn0λ(z).Lie_{e}\lambda(z)=1,\quad Lie_{E}\lambda(z)=\lambda(z)-\frac{z}{n_{0}}\lambda^{\prime}(z).

The data set (MToda,η,,e,E)(M^{Toda},\eta,\circ,e,E) forms a semisimple Frobenius manifold with charge d=1d=1.

The flat coordinate system for the metric η\eta, denoted as

𝐭={t0,j}j=1n01{t1,j}j=0n1{tm,j}j=0nm,\mathbf{t}=\{t_{0,j}\}_{j=1}^{n_{0}-1}\cup\{t_{1,j}\}_{j=0}^{n_{1}}\cup\cdots\cup\{t_{m,j}\}_{j=0}^{n_{m}},

is determined by the following expansion for φ\varphi:

φ={ti,0+ti,1wi1+,eφai,0,i=2,,m,log(w1)+t1,0+t1,1w11+,eφ0,log(w0)t0,1w01t0,2w02,eφ,\varphi=\left\{\begin{aligned} &t_{i,0}+t_{i,1}w_{i}^{-1}+\cdots,\quad&e^{\varphi}&\to a_{i,0},\ i=2,\cdots,m,\\ &-\log(w_{1})+t_{1,0}+t_{1,1}w_{1}^{-1}+\cdots,\quad&e^{\varphi}&\to 0,\\ &\log(w_{0})-t_{0,1}w_{0}^{-1}-t_{0,2}w_{0}^{-2}-\cdots,\quad&e^{\varphi}&\to\infty,\end{aligned}\right. (4.1)

where

wi={(niλ)1ni=wi,1(zai,0)1+,zai,0,i=1,,m,(n0λ)1n0=z+w0,0+w0,1z1+,z,i=0.w_{i}=\left\{\begin{aligned} &(n_{i}\lambda)^{\frac{1}{n_{i}}}=w_{i,1}(z-a_{i,0})^{-1}+\cdots,\quad&z&\to a_{i,0},\ i=1,\cdots,m,\\ &(n_{0}\lambda)^{\frac{1}{n_{0}}}=z+w_{0,0}+w_{0,1}z^{-1}+\cdots,\quad&z&\to\infty,\ i=0.\\ \end{aligned}\right. (4.2)

Furthermore, we have

z1ti,jλ(z)={(wi(z)nij1wi(z))ai,0,1,i=1,,m,j=0,,ni,(wi(z)n0j1w0(z)),0,i=0,j=1,,n01.z^{-1}\partial_{t_{i,j}}\lambda(z)=\left\{\begin{aligned} &-(w_{i}(z)^{n_{i}-j-1}w_{i}^{\prime}(z))_{a_{i,0},\leq-1},\quad i=1,\cdots,m,\ j=0,\cdots,n_{i},\\ &(w_{i}(z)^{n_{0}-j-1}w_{0}^{\prime}(z))_{\infty,\geq 0},\quad i=0,\ j=1,\cdots,n_{0}-1.\end{aligned}\right.

In this flat coordinate system, the vector fields ee and EE can be expressed as

e=t1,n1e=\frac{\partial}{\partial t_{1,n_{1}}}

and

E=j=1n01jn0t0,jt0,j+(1n0+1n1)t1,0+i=1mj=1n1jniti,jti,j+i=2m1n0ti,0.E=\sum_{j=1}^{n_{0}-1}\frac{j}{n_{0}}t_{0,j}\frac{\partial}{\partial t_{0,j}}+(\frac{1}{n_{0}}+\frac{1}{n_{1}})\frac{\partial}{\partial t_{1,0}}+\sum_{i=1}^{m}\sum_{j=1}^{n_{1}}\frac{j}{n_{i}}t_{i,j}\frac{\partial}{\partial t_{i,j}}+\sum_{i=2}^{m}\frac{1}{n_{0}}\frac{\partial}{\partial t_{i,0}}.
Lemma 4.1.

Let \nabla be the Levi-Civita connection associated with the metric η\eta. Then, for any vector fields 1\partial_{1} and 2\partial_{2} on MTodaM^{Toda}, we have

(12)λ(z)=12λ(z)z(1λ(z)2λ(z)zλ(z)),0s=1mz(1λ(z)2λ(z)zλ(z))as,0,1.(\nabla_{\partial_{1}}\partial_{2})\cdot\lambda(z)=\partial_{1}\partial_{2}\lambda(z)-z(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{z\lambda^{\prime}(z)})^{\prime}_{\infty,\geq 0}-\sum_{s=1}^{m}z(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{z\lambda^{\prime}(z)})^{\prime}_{a_{s,0},\leq-1}. (4.3)
Proof.

The proof follows the approach of Lemma 3.1. ∎

4.2. cotangent space and Hamiltionian structure

For any point λ(z)\lambda(z) in MTodaM^{Toda}, a tangent vector Tλ(z)MToda\partial\in T_{\lambda(z)}M^{Toda} can be represented as ξ(z)=λ(z)\xi(z)=\partial\lambda(z), where

ξ(z)=b0,n01zn01++b0,0+j=1n1b1,jzj+i=2mj=1ni+1bi,j(zai,0)j.\xi(z)=b_{0,n_{0}-1}z^{n_{0}-1}+\cdots+b_{0,0}+\sum_{j=1}^{n_{1}}b_{1,j}z^{-j}+\sum_{i=2}^{m}\sum_{j=1}^{n_{i}+1}b_{i,j}(z-a_{i,0})^{-j}.

To describe a cotangent vector at λ(z)\lambda(z), we follow the procedure in Section 3.2. Consider disjoint disks D1,,DmD_{1},\ldots,D_{m} in the complex plane with ai,0Dia_{i,0}\in D_{i}, and let γi=Di\gamma_{i}=\partial D_{i}. Define the space \mathcal{H} of analytic function germs on s=1mγs\cup_{s=1}^{m}\gamma_{s}, and introduce the following pairing:

ω(z),ξ(z):=12πis=1mγsω(z)ξ(z)dzz,ω(z),ξ(z)Tλ(z)MToda.\langle\omega(z),\xi(z)\rangle:=\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\omega(z)\xi(z)\frac{dz}{z},\quad\omega(z)\in\mathcal{H},\quad\xi(z)\in T_{\lambda(z)}M^{Toda}.

This pairing induces a surjective map from \mathcal{H} to Tλ(z)MTodaT^{\ast}_{\lambda(z)}M^{Toda}, allowing elements of \mathcal{H} to be regarded as cotangent vectors at λ(z)\lambda(z).

Lemma 4.2.

For any ξ(z)Tλ(z)MToda\xi(z)\in T_{\lambda(z)}M^{Toda}, define the linear maps from \mathcal{H} to Tλ(z)MTodaT_{\lambda(z)}M^{Toda} as

η(ω(z))=(ω(z))+λ(z)z+(ω(z)λ(z))+z,\eta^{\ast}(\omega(z))=-(\omega(z))_{+}\lambda^{\prime}(z)z+(\omega(z)\lambda^{\prime}(z))_{+}z, (4.4)

and

Cξ(z)(ω(z))=(ω(z)ξ(z))+λ(z)z+(ω(z)λ(z))+zξ(z).C_{\xi(z)}(\omega(z))=-(\omega(z)\xi(z))_{+}\lambda^{\prime}(z)z+(\omega(z)\lambda^{\prime}(z))_{+}z\xi(z). (4.5)

We then obtain

η(ω(z)),ξ(z)η=ω(z),ξ(z),\langle\eta^{\ast}(\omega(z)),\xi(z)\rangle_{\eta}=\langle\omega(z),\xi(z)\rangle,

and

c(ξ1(z),ξ2(z),η(ω(z)))=Cξ1(ω(z)),ξ2(z)η.c(\xi_{1}(z),\xi_{2}(z),\eta^{\ast}(\omega(z)))=\langle C_{\xi_{1}}(\omega(z)),\xi_{2}(z)\rangle_{\eta}.
Proof.

The proof is analogous to the argument presented in Lemma 3.2. ∎

Using the equality

𝒫(ω)=ηxω,\mathcal{P}(\omega)=\eta^{\ast}\cdot\nabla_{\partial_{x}}\omega,

we can derive the explicit form of the Hamiltonian structure 𝒫\mathcal{P} associated with the flat metric ,η\langle\ ,\ \rangle_{\eta}.

Corollary 4.3.

The dispersionless Hamiltonian operator 𝒫\mathcal{P} associated with the metric ,η\langle\ ,\ \rangle_{\eta} has the form

𝒫(ω(z))={ω(z)+,λ(z)}z{ω(z),λ(z)}+z,\mathcal{P}(\omega(z))=\{\omega(z)_{+},\lambda(z)\}z-\{\omega(z),\lambda(z)\}_{+}z, (4.6)

where

{f(z,x),g(z,x)}=f(z,x)zg(z,x)xf(z,x)xg(z,x)z.\{f(z,x),g(z,x)\}=\frac{\partial f(z,x)}{\partial z}\frac{\partial g(z,x)}{\partial x}-\frac{\partial f(z,x)}{\partial x}\frac{\partial g(z,x)}{\partial z}.
Proof.

The proof follows the same approach as in the proof of Lemma 3.3. ∎

4.3. principal hierarchy for MTodaM^{Toda}

To prove Theorem 1.1, we follow a similar approach to that in Section 3.4, requiring the following lemma:

Lemma 4.4.

For any pp\in\mathbb{N}, let Qp(λ)Q_{p}(\lambda) be analytic functions in λ\lambda that satisfy

Qp(λ)λ=Qp1(λ).\frac{\partial Q_{p}(\lambda)}{\partial\lambda}=Q_{p-1}(\lambda).

Define

Fi,p=12πiγiQp+1(λ(z))dzz,F_{i,p}=\frac{1}{2\pi i}\int_{\gamma_{i}}Q_{p+1}(\lambda(z))\,\frac{dz}{z}, (4.7)

then

ηdFi,p=C(dFi,p1),\eta^{\ast}\cdot\nabla_{\partial}dF_{i,p}=C_{\partial}(dF_{i,p-1}), (4.8)

where the operators η\eta^{\ast}\cdot\nabla_{\partial} and CC_{\partial} are defined by (4.4) and (4.5), respectively.

Proof of Theorem 1.5.

The proof follows a similar approach to that of Theorem 1.1. As an instance, for the density θti,ni,p\theta_{t_{i,n_{i}},p}, let MMTodaM^{\prime}\subset M^{Toda} be such that w0=λ1n0w_{0}=\lambda^{\frac{1}{n_{0}}} can be analytically continued onto s=1mγs\cup_{s=1}^{m}\gamma_{s}, with the winding number 11 along γi\gamma_{i} and 0 along γj\gamma_{j} for jij\neq i. Similarly, wi=λ1niw_{i}=\lambda^{\frac{1}{n_{i}}} can be analytically continued onto γi\gamma_{i} with the winding number 1-1. Then, on MM^{\prime}, we have

θti,ni,p=\displaystyle\theta_{t_{i,n_{i}},p}= Rescpn0λpp!dzz+12πiγiλpp!(logw0wicpni)dzz+12πisiγsλpp!logw0dzz,\displaystyle\mathop{\text{\rm Res}}_{\infty}\frac{c_{p}}{n_{0}}\frac{\lambda^{p}}{p!}\frac{dz}{z}+\frac{1}{2\pi\mathrm{i}}\int_{\gamma_{i}}\frac{\lambda^{p}}{p!}(\log w_{0}w_{i}-\frac{c_{p}}{n_{i}})\frac{dz}{z}+\frac{1}{2\pi\mathrm{i}}\sum_{s\neq i}\int_{\gamma_{s}}\frac{\lambda^{p}}{p!}\log w_{0}\frac{dz}{z},
=\displaystyle= 12πiγiλpp!(logw0wicpnicpn0)dzz+12πisiγsλpp!(logw0cpn0)dzz,\displaystyle\frac{1}{2\pi\mathrm{i}}\int_{\gamma_{i}}\frac{\lambda^{p}}{p!}(\log w_{0}w_{i}-\frac{c_{p}}{n_{i}}-\frac{c_{p}}{n_{0}})\frac{dz}{z}+\frac{1}{2\pi\mathrm{i}}\sum_{s\neq i}\int_{\gamma_{s}}\frac{\lambda^{p}}{p!}(\log w_{0}-\frac{c_{p}}{n_{0}})\frac{dz}{z},

which satisfies equality (2.1).

Introduce the operator =E+1n0zz\mathcal{E}=E+\frac{1}{n_{0}}z\frac{\partial}{\partial z}, then

Lieλ(z)=λ(z),\operatorname{Lie}_{\mathcal{E}}\lambda(z)=\lambda(z),

and

Resai,0Lief(λ(z))dzz=LieEResai,0f(λ(z))dzz.\mathop{\text{\rm Res}}_{a_{i,0}}\operatorname{Lie}_{\mathcal{E}}f(\lambda(z))\frac{dz}{z}=\operatorname{Lie}_{E}\mathop{\text{\rm Res}}_{a_{i,0}}f(\lambda(z))\frac{dz}{z}. (4.9)

Thus, we obtain

LieEθu,p(t)={(p+1jn0)θu,p(t),u=t0,j;(p+1jni)θu,p(t),u=ti,j,jni;pθu,p(t)+s=1m1n0θts,0,p1+1niθti,0,p1,u=ti,ni.\operatorname{Lie}_{E}\theta_{u,p}(t)=\begin{cases}\left(p+1-\frac{j}{n_{0}}\right)\theta_{u,p}(t),&u=t_{0,j};\\ \left(p+1-\frac{j}{n_{i}}\right)\theta_{u,p}(t),&u=t_{i,j},\ j\neq n_{i};\\ p\theta_{u,p}(t)+\sum_{s=1}^{m}\frac{1}{n_{0}}\theta_{t_{s,0},p-1}+\frac{1}{n_{i}}\theta_{t_{i,0},p-1},&u=t_{i,n_{i}}.\end{cases}

Therefore, equality (2.2) holds.

Finally, using formula (4.6), we deduce the Hamiltonian vector fields corresponding to the densities θα,p\theta_{\alpha,p}.

This completes the proof of the theorem. ∎

4.4. principal hierarchy for MCTodaM^{C-Toda}

Let MCTodaM^{C-Toda} be the submanifold of MTodaM^{Toda} consisting of functions of the form (1.4). We will first show that MCTodaM^{C-Toda} is a natural Frobenius submanifold of MTodaM^{Toda}.

For any λ(z)MCToda\lambda(z)\in M^{C-Toda}, we have:

w0(1p)=w1(p),p0;\displaystyle w_{0}(\frac{1}{p})=w_{1}(p),\quad p\to 0;
w2(1p)=w2(p),p1;\displaystyle w_{2}(\frac{1}{p})=-w_{2}(p),\quad p\to 1;
w3(1p)=w3(p),p1;\displaystyle w_{3}(\frac{1}{p})=-w_{3}(p),\quad p\to-1;
w2i2(1p)=w2i1(p),pb2i1,0,i=3,4,,m,\displaystyle w_{2i-2}(\frac{1}{p})=w_{2i-1}(p),\quad p\to b_{2i-1,0},\ i=3,4,\cdots,m^{\prime},

where b2i1,0+1b2i1,0=b~i,0b_{2i-1,0}+\frac{1}{b_{2i-1,0}}=\tilde{b}_{i,0}. For p0p\to 0, we have

φ~=\displaystyle-\tilde{\varphi}= logw0(1p)12t1,0t0,1w01(1p)\displaystyle\log w_{0}(\frac{1}{p})-\frac{1}{2}t_{1,0}-t_{0,1}w_{0}^{-1}(\frac{1}{p})-\cdots
=\displaystyle= logw1(p)12t1,0t0,1w11(p)\displaystyle\log w_{1}(p)-\frac{1}{2}t_{1,0}-t_{0,1}w_{1}^{-1}(p)-\cdots
=\displaystyle= logw1(p)12t1,0t1,1w11(p).\displaystyle\log w_{1}(p)-\frac{1}{2}t_{1,0}-t_{1,1}w_{1}^{-1}(p)-\cdots.

For p1p\to 1:

φ~=\displaystyle-\tilde{\varphi}= t2,012t1,0+w31(1p)+\displaystyle t_{2,0}-\frac{1}{2}t_{1,0}+w_{3}^{-1}(\frac{1}{p})+\cdots
=\displaystyle= log1+t2,1w21(1p)+\displaystyle-\log 1+t_{2,1}w_{2}^{-1}(\frac{1}{p})+\cdots
=\displaystyle= log1t2,1w21(p)+t2,2w22(p)\displaystyle-\log 1-t_{2,1}w_{2}^{-1}(p)+t_{2,2}w_{2}^{-2}(p)-\cdots
=\displaystyle= log1t2,1w21(p)t2,2w22(p).\displaystyle-\log 1-t_{2,1}w_{2}^{-1}(p)-t_{2,2}w_{2}^{-2}(p)-\cdots.

For p1p\to-1:

φ~=\displaystyle-\tilde{\varphi}= t3,012t1,0+w31(1p)+\displaystyle t_{3,0}-\frac{1}{2}t_{1,0}+w_{3}^{-1}(\frac{1}{p})+\cdots
=\displaystyle= log(1)+t3,1w31(1p)+\displaystyle-\log(-1)+t_{3,1}w_{3}^{-1}(\frac{1}{p})+\cdots
=\displaystyle= log(1)t3,1w31(p)+t3,2w32(p)\displaystyle-\log(-1)-t_{3,1}w_{3}^{-1}(p)+t_{3,2}w_{3}^{-2}(p)-\cdots
=\displaystyle= log(1)t3,1w31(p)t3,2w32(p).\displaystyle-\log(-1)-t_{3,1}w_{3}^{-1}(p)-t_{3,2}w_{3}^{-2}(p)-\cdots.

For pb2i1=1b2i2p\to b_{2i-1}=\frac{1}{b_{2i-2}}:

φ~=\displaystyle-\tilde{\varphi}= t2i2,012t1,0+t2i2,1w2i21(1p)+\displaystyle t_{2i-2,0}-\frac{1}{2}t_{1,0}+t_{2i-2,1}w_{2i-2}^{-1}(\frac{1}{p})+\cdots
=\displaystyle= t2i2,012t1,0+t2i2,1w2i11(p)+\displaystyle t_{2i-2,0}-\frac{1}{2}t_{1,0}+t_{2i-2,1}w_{2i-1}^{-1}(p)+\cdots
=\displaystyle= t2i1,012t1,0+t2i1,1w2i11(p)+.\displaystyle t_{2i-1,0}-\frac{1}{2}t_{1,0}+t_{2i-1,1}w_{2i-1}^{-1}(p)+\cdots.

Thus, the flat coordinates satisfy the following constraints:

t0,j=t1,j,j=1,,n01;\displaystyle t_{0,j}=t_{1,j},\quad j=1,\cdots,n_{0}-1;
t2,2=t2,4==t2,n2=0,t2,012t1,0=log1;\displaystyle t_{2,2}=t_{2,4}=\cdots=t_{2,n_{2}}=0,\quad t_{2,0}-\frac{1}{2}t_{1,0}=-\log 1;
t3,2=t3,4==t3,n3=0,t3,012t1,0=log(1);\displaystyle t_{3,2}=t_{3,4}=\cdots=t_{3,n_{3}}=0,\quad t_{3,0}-\frac{1}{2}t_{1,0}=-\log(-1);
t2i2,j=t2i1,j,i=3,,m,j=0,,n2i2.\displaystyle t_{2i-2,j}=t_{2i-1,j},\quad i=3,\cdots,m^{\prime},\ j=0,\cdots,n_{2i-2}.

Here, the values of log1\log 1 and log(1)\log(-1) depend on the chosen branch of logp\log p. Hence, MCTodaM^{C-Toda} forms a flat submanifold of MTodaM^{Toda} with flat coordinates:

𝐭={t1,j}j=0n0{t2,2j1}j=1n1{t3,2j1}j=1n2{t4,j}j=0n3{t2m2,j}j=0nm.\mathbf{t}=\{t_{1,j}\}_{j=0}^{n_{0}^{\prime}}\cup\{t_{2,2j-1}\}_{j=1}^{n_{1}^{\prime}}\cup\{t_{3,2j-1}\}_{j=1}^{n_{2}^{\prime}}\cup\{t_{4,j}\}_{j=0}^{n_{3}^{\prime}}\cup\cdots\cup\{t_{2m^{\prime}-2,j}\}_{j=0}^{n_{m^{\prime}}^{\prime}}.

On the other hand, let p1,,prp_{1},\cdots,p_{r} be the simple critical points of λ(z)MToda\lambda(z)\in M^{Toda}, where r=dim(MToda)r=dim(M^{Toda}). The critical values uj=λ(pj)u_{j}=\lambda(p_{j}) form the canonical coordinates for MTodaM^{Toda}. For λ(z)MCToda\lambda(z)\in M^{C-Toda}, we can choose

p2j=1p2j1,j=1,2,,r2,p_{2j}=\frac{1}{p_{2j-1}},\quad j=1,2,\cdots,\frac{r}{2},

which implies

u2j=u2j1,j=1,2,,r2,u_{2j}=u_{2j-1},\quad j=1,2,\cdots,\frac{r}{2},

Thus, MCTodaM^{C-Toda} is a caustic submanifold of MTodaM^{Toda}. According to Corollary 3.7 in [17], MCTodaM^{C-Toda} is a natural Frobenius submanifold of MTodaM^{Toda}.

Remark 4.5.

Note that the above conclusion also holds when n1=0n_{1}^{\prime}=0 or n2=0n_{2}^{\prime}=0. In the case n1=0n_{1}^{\prime}=0, we have

r=dim(MToda)=2dim(MCToda)1.r=\dim(M^{Toda})=2\dim(M^{C-Toda})-1.

Choose simple critical points p1,,prp_{1},\ldots,p_{r} such that

p1=1,p2j=1p2j+1,j=1,2,.r12,p_{1}=1,\quad p_{2j}=\frac{1}{p_{2j+1}},\quad j=1,2,\ldots.\frac{r-1}{2},

This implies

u2j=u2j+1,j=1,2,,r12.u_{2j}=u_{2j+1},\quad j=1,2,\ldots,\frac{r-1}{2}.

Let us now consider the principal hierarchy for MCTodaM^{C-Toda}. Define odd\mathcal{H}^{odd} as the subspace of \mathcal{H} consisting of elements satisfying ω(z1)=ω(z)\omega(z^{-1})=-\omega(z). For λ(z)MCToda\lambda(z)\in M^{C-Toda} and ω(z)odd\omega(z)\in\mathcal{H}^{odd}, the following relations hold:

ω(z1)={ω(z),if z,z1Dc,ω+(z),if z1Dc,zD,\omega_{-}(z^{-1})=\begin{cases}-\omega_{-}(z),&\text{if }z,z^{-1}\in D^{c},\\ \omega_{+}(z),&\text{if }z^{-1}\in D^{c},z\in D,\end{cases}

and

ω+(z1)={ω+(z),if z,z1D,ω(z),if z1D,zDc.\omega_{+}(z^{-1})=\begin{cases}\omega_{+}(z),&\text{if }z,z^{-1}\in D,\\ -\omega_{-}(z),&\text{if }z^{-1}\in D,z\in D^{c}.\end{cases}

Therefore, η(ω(z))Tλ(z)MCToda\eta^{\ast}(\omega(z))\in T_{\lambda(z)}M^{C-Toda}, and for any Tλ(z)MCToda\partial\in T_{\lambda(z)}M^{C-Toda}, it follows that:

ηωTλ(z)MCToda.\eta^{\ast}\cdot\nabla_{\partial}\omega\in T_{\lambda(z)}M^{C-Toda}.

Since the differentials of the Hamiltonian densities

θt0,j,p+θt1,j,p,j=1,2,,n0;\displaystyle\theta_{t_{0,j},p}+\theta_{t_{1,j},p},\quad j=1,2,\dots,n_{0}^{\prime};
θt2,2j1,p,j=1,2,,n1;\displaystyle\theta_{t_{2,2j-1},p},\quad j=1,2,\dots,n_{1}^{\prime};
θt3,2j1,p,j=1,2,,n2;\displaystyle\theta_{t_{3,2j-1},p},\quad j=1,2,\dots,n_{2}^{\prime};
θt2i2,j,p+θt2i1,j,p,i=3,,m,j=0,,ni\displaystyle\theta_{t_{2i-2,j},p}+\theta_{t_{2i-1,j},p},\quad i=3,\dots,m^{\prime},\ j=0,\dots,n_{i}^{\prime}

belong to odd\mathcal{H}^{odd}, the corresponding Hamiltonian vector fields can be restricted to the loop space of MCTodaM^{C-Toda}, thereby forming the principal hierarchy for MCTodaM^{C-Toda}.

4.5. principal hierarchy for M^Toda\hat{M}^{Toda}

Let M^Toda\hat{M}^{Toda} denote the almost duality of the Frobenius manifold MTodaM^{Toda}. In this subsection, we will construct the principal hierarchy for M^Toda\hat{M}^{Toda}.

Suppose IDI\in\textbf{D}, where II is the set of all zeros and poles of λ(z)\lambda(z) in \mathbb{C}. Define the linear map from \mathcal{H} to Tλ(z)MTodaT_{\lambda(z)}M^{Toda} as g(ω(z))=CE(ω(z))g^{\ast}(\omega(z))=C_{E}(\omega(z)), that is,

g(ω(z))=(λ(z)ω(z))λ(z)zzλ(z)(λ(z)ω(z))+zλ(z)n012πis=1mγsλ(z)ω(z)𝑑z,g^{\ast}(\omega(z))=(\lambda(z)\omega(z))_{-}\lambda^{\prime}(z)z-z\lambda(z)(\lambda^{\prime}(z)\omega(z))_{-}+\frac{z\lambda^{\prime}(z)}{n_{0}}\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\lambda^{\prime}(z)\omega(z)\,dz,

then we have

ω(z),ξ(z)=(g(ω(z)),ξ(z))g,\langle\omega(z),\xi(z)\rangle=(g^{\ast}(\omega(z)),\xi(z))_{g},

where

(1,2)g=Resdλ=01logλ(z)2logλ(z)(logλ(z))z2dz.(\partial_{1},\partial_{2})_{g}=\sum\mathop{\text{\rm Res}}_{d\lambda=0}\frac{\partial_{1}\log\lambda(z)\partial_{2}\log\lambda(z)}{(\log\lambda(z))^{\prime}z^{2}}\,dz.

Let ^\hat{\nabla} be the Levi-Civita connection associated with the intersection form gg, we have

^12λ(z)=12λ(z)1λ(z)2λ(z)λ(z)zλ(z)qI(1λ(z)2λ(z)zλ(z)λ(z))q,1.\hat{\nabla}_{\partial_{1}}\partial_{2}\cdot\lambda(z)=\partial_{1}\partial_{2}\lambda(z)-\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{\lambda(z)}-z\lambda(z)\sum_{q\in I}\left(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{z\lambda(z)\lambda^{\prime}(z)}\right)_{q,\leq-1}^{\prime}.

Furthermore,

g^ω(z)={ω(z),λ(z)}λ(z)z{(ω(z)λ(z)),λ(z)}zzλ(z)n012πis=1mγs{ω(z),λ(z)}𝑑z,g^{\ast}\cdot\hat{\nabla}_{\partial}\omega(z)=\{\omega(z),\lambda(z)\}_{-}\lambda(z)z-\{(\omega(z)\lambda(z))_{-},\lambda(z)\}z-\frac{z\lambda^{\prime}(z)}{n_{0}}\frac{1}{2\pi\mathrm{i}}\sum_{s=1}^{m}\int_{\gamma_{s}}\{\omega(z),\lambda(z)\}\,dz,

where {f,g}=fggf\{f,g\}=f^{\prime}\partial g-g^{\prime}\partial f. Setting =x\partial=\partial_{x}, we obtain the explicit form of the Hamiltonian structure 𝒫^\hat{\mathcal{P}} for the flat metric gg.

Proof of Theorem 1.6.

Consider functions Fp=12πiγjQ~p(λ)dzzF_{p}=\frac{1}{2\pi\mathrm{i}}\int_{\gamma_{j}}\tilde{Q}_{p}(\lambda)\frac{dz}{z} on MTodaM^{Toda}, where Qp(λ)=Q~p(λ)λQ_{p}(\lambda)=\frac{\partial\tilde{Q}_{p}(\lambda)}{\partial\lambda} satisfy the recurrence relation:

λQp+1λ+Qp+1=Qp.\lambda\frac{\partial Q_{p+1}}{\partial\lambda}+Q_{p+1}=Q_{p}. (4.10)

We have dFp=Qp𝟏γjdF_{p}=Q_{p}\mathbf{1}_{\gamma_{j}} and

g^dFp+1=\displaystyle g^{\ast}\cdot\hat{\nabla}_{\partial}dF_{p+1}= ((Qp+1λλλ𝟏γj+Qp+1λ𝟏γj)λz)\displaystyle-\left(\left(\frac{\partial Q_{p+1}}{\partial\lambda}\lambda\lambda^{\prime}\mathbf{1}_{\gamma_{j}}+Q_{p+1}\lambda^{\prime}\mathbf{1}_{\gamma_{j}}\right)_{-}\partial\lambda z\right)
+((Qp+1λλλ𝟏γj+Qp+1λ𝟏γj)λz)\displaystyle+\left(\left(\frac{\partial Q_{p+1}}{\partial\lambda}\lambda\partial\lambda\mathbf{1}_{\gamma_{j}}+Q_{p+1}\partial\lambda\mathbf{1}_{\gamma_{j}}\right)_{-}\lambda^{\prime}z\right)
=\displaystyle= ((Qpλ𝟏γj)λz)((Qpλ𝟏γj)λz)\displaystyle\left((Q_{p}\partial\lambda\mathbf{1}_{\gamma_{j}})_{-}\lambda^{\prime}z\right)-\left((Q_{p}\lambda^{\prime}\mathbf{1}_{\gamma_{j}})_{-}\partial\lambda z\right)
=\displaystyle= C(dFp).\displaystyle C_{\partial}(dF_{p}).

Let MM^{\prime} be a subspace of MTodaM^{Toda} such that for any λ(z)M\lambda(z)\in M^{\prime}, the winding number of λ(z)\lambda(z) along γj\gamma_{j} is zero. Define Q~(p)=(log(λ(z)))p+1(p+1)!\tilde{Q}(p)=\frac{(\log(\lambda(z)))^{p+1}}{(p+1)!}, then Fγj,pF_{\gamma_{j},p} satisfies equality (4.10) on MM^{\prime}. Furthermore, by the uniqueness of analytic function, for any γ~[γj]\tilde{\gamma}\in[\gamma_{j}], Fγ~,pF_{\tilde{\gamma},p} satisfies equality (2.1), where [γj][\gamma_{j}] denotes the homotopy equivalence class of γj\gamma_{j} in I\mathbb{C}-I.

In particular, let q1,,qk,p1,,pkq_{1},\ldots,q_{k},p_{1},\ldots,p_{k} be the zeros and poles of λ(z)\lambda(z) within the region surrounded by γ~\tilde{\gamma}. Additionally, let qk+1,,qn,pk+1,,prq_{k+1},\ldots,q_{n},p_{k+1},\ldots,p_{r} be the zeros and poles outside this region. If 00\in\mathbb{C} is outside the region surrounded by γ~\tilde{\gamma}, we have

Fγ~,0=\displaystyle F_{\tilde{\gamma},0}= 12πiγ~log(λ(z))dzz\displaystyle\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}\log(\lambda(z))\frac{dz}{z}
=\displaystyle= 12πiγ~(log(λ(z)s=1kzpszqs)+log(s=1kzqszps))dzz\displaystyle\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}(\log(\lambda(z)\prod_{s=1}^{k}\frac{z-p_{s}}{z-q_{s}})+\log(\prod_{s=1}^{k}\frac{z-q_{s}}{z-p_{s}}))\frac{dz}{z}
=\displaystyle= Res0log(s=1kzqszps)dzz\displaystyle-\mathop{\text{\rm Res}}_{0}\log(\prod_{s=1}^{k}\frac{z-q_{s}}{z-p_{s}})\,\frac{dz}{z}
=\displaystyle= s=1klog(ps)s=1klog(qs).\displaystyle\sum_{s=1}^{k}\log(p_{s})-\sum_{s=1}^{k}\log(q_{s}).

Otherwise

Fγ~,0=\displaystyle F_{\tilde{\gamma},0}= 12πiγ~log(λ(z))dzz\displaystyle\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}\log(\lambda(z))\frac{dz}{z}
=\displaystyle= 12πiγ~(log(λ(z)s=1kzpszqs)+log(s=1kzqszps))dzz\displaystyle\frac{1}{2\pi\mathrm{i}}\int_{\tilde{\gamma}}(\log(\lambda(z)\prod_{s=1}^{k}\frac{z-p_{s}}{z-q_{s}})+\log(\prod_{s=1}^{k}\frac{z-q_{s}}{z-p_{s}}))\frac{dz}{z}
=\displaystyle= Res0log(λ(z)s=1kzpszqs)dzz\displaystyle\mathop{\text{\rm Res}}_{0}\log(\lambda(z)\prod_{s=1}^{k}\frac{z-p_{s}}{z-q_{s}})\frac{dz}{z}
=\displaystyle= s=k+1rlog(ps)s=k+1nlog(qs).\displaystyle\sum_{s=k+1}^{r}\log(p_{s})-\sum_{s=k+1}^{n}\log(q_{s}).

The theorem is proved. ∎

Corollary 4.6.

The Hamiltonian vector fields T^γi,p=𝒫^(dFγi,p+1)\frac{\partial}{\partial\hat{T}^{\gamma_{i},p}}=\hat{\mathcal{P}}(dF_{\gamma_{i},p+1}) take the form:

λ(z)T^γi,p1={((log(λ(z)))pp!𝟏γi)+,λ(z)},\frac{\partial\lambda(z)}{\partial\hat{T}^{\gamma_{i},p-1}}=\{(\frac{(\log(\lambda(z)))^{p}}{p!}\mathbf{1}_{\gamma_{i}})_{+},\lambda(z)\},

where

{f(z,x),g(z,x)}=z(f(z,x)zg(z,x)xf(z,x)xg(z,x)z).\{f(z,x),g(z,x)\}=z\left(\frac{\partial f(z,x)}{\partial z}\frac{\partial g(z,x)}{\partial x}-\frac{\partial f(z,x)}{\partial x}\frac{\partial g(z,x)}{\partial z}\right). (4.11)

4.6. rank-1 extension

By using a similar approach as in subsection 3.7, we can construct a rank-1 extension of MTodaM^{Toda} using the following lemma.

Lemma 4.7.

For any vector field 1,2\partial_{1},\partial_{2} on MTodaM^{Toda}, we have

1λ(z)2λ(z)1λ(z)2λ(z)zλ=(1λ(z)2λ(z)zλ(z)),0+j=1m(1λ(z)2λ(z)zλ(z))φj,1.\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)-\partial_{1}\lambda(z)\circ\partial_{2}\lambda(z)}{z\lambda^{\prime}}=(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{z\lambda^{\prime}(z)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{1}\lambda(z)\partial_{2}\lambda(z)}{z\lambda^{\prime}(z)})_{\varphi_{j},\leq-1}.
Corollary 4.8.

Let z=esz=e^{s} and define Ω(t,s)\Omega(\textbf{t},s) such that

sαΩ=αλ(z),s2Ω=zλ(z),αβΩ=(αλ(z)βλ(z)zλ(z)),0+j=1m(αλ(z)βλ(z)zλ(z))φj,1.\partial_{s}\partial_{\alpha}\Omega=\partial_{\alpha}\lambda(z),\quad\partial^{2}_{s}\Omega=z\lambda^{\prime}(z),\quad\partial_{\alpha}\partial_{\beta}\Omega=(\frac{\partial_{\alpha}\lambda(z)\partial_{\beta}\lambda(z)}{z\lambda^{\prime}(z)})_{\infty,\geq 0}+\sum_{j=1}^{m}(\frac{\partial_{\alpha}\lambda(z)\partial_{\beta}\lambda(z)}{z\lambda^{\prime}(z)})_{\varphi_{j},\leq-1}.

Then ω=sΩ\omega=\partial_{s}\Omega satisfies the condition of Lemma 2.1, thus defining a flat F-manifold structure on MToda×M^{Toda}\times\mathbb{C}, with multiplication given by the following expressions:

(α,0)(β,0)=(αβ,αβΩs),\displaystyle(\partial_{\alpha},0)\star(\partial_{\beta},0)=(\partial_{\alpha}\circ\partial_{\beta},\partial_{\alpha}\partial_{\beta}\Omega\cdot\partial_{s}),
(α,0)(0,s)=(0,αλ(z)s),\displaystyle(\partial_{\alpha},0)\star(0,\partial_{s})=(0,\partial_{\alpha}\lambda(z)\cdot\partial_{s}),
(0,s)(0,s)=(0,zλ(z)s).\displaystyle(0,\partial_{s})\star(0,\partial_{s})=(0,z\lambda^{\prime}(z)\cdot\partial_{s}).

The principal hierarchy for the flat F-manifold MToda×M^{Toda}\times\mathbb{C} is given by

T~,p=Θ,p~x,\frac{\partial}{\partial\tilde{T}^{\bullet,p}}=\Theta_{\bullet,p}\star\tilde{\partial}_{x},

where

~x=(x,sxs),\tilde{\partial}_{x}=(\partial_{x},\frac{\partial s}{\partial x}\partial_{s}),

and the vector fields Θ,p\Theta_{\bullet,p} are given by the following theorem.

Theorem 4.9.

The vector fields on MToda×M^{Toda}\times\mathbb{C} of the form

Θs,p=(0,λ(z)pp!s)\displaystyle\Theta_{s,p}=(0,\frac{\lambda(z)^{p}}{p!}\partial_{s})
Θt0,n0j,p=(η(dθt0,j,p),(dθt0,j,p|z=es)+s),\displaystyle\Theta_{t_{0,n_{0}-j},p}=(\eta^{\ast}(d\theta_{t_{0,j},p}),(d\theta_{t_{0,j},p}|_{z=e^{s}})_{+}\partial_{s}),
Θti,nij,p=(η(dθti,j,p),(dθti,j,p|z=es)s),i=1,,m,\displaystyle\Theta_{t_{i,n_{i}-j},p}=(\eta^{\ast}(d\theta_{t_{i,j},p}),-(d\theta_{t_{i,j},p}|_{z=e^{s}})_{-}\partial_{s}),\quad i=1,\cdots,m,

satisfy equations of the form (3.13) and (3.14).

By applying a similar approach to the proof of Corollary 1.4, we derive the explicit form of the principal hierarchy for the flat F-manifold MToda×M^{Toda}\times\mathbb{C} as follows:

T~,p=(0,dxλ(z)λ(z)pp!);\displaystyle\frac{\partial}{\partial\tilde{T}^{\bullet,p}}=(0,d_{x}\lambda(z)\cdot\frac{\lambda(z)^{p}}{p!});
T~t0,n0j,p=(Tt0,j,p,(dθt0,j,p|z=esdxλ(z))+s);\displaystyle\frac{\partial}{\partial\tilde{T}^{t_{0,n_{0}-j},p}}=(\frac{\partial}{\partial T^{t_{0,j},p}},(d\theta_{t_{0,j},p}|_{z=e^{s}}\cdot d_{x}\lambda(z))_{+}\partial_{s});
T~ti,nij,p=(Tti,j,p,(dθti,j,p|z=esdxλ(z))s),i=1,,m,\displaystyle\frac{\partial}{\partial\tilde{T}^{t_{i,n_{i}-j},p}}=(\frac{\partial}{\partial T^{t_{i,j},p}},-(d\theta_{t_{i,j},p}|_{z=e^{s}}\cdot d_{x}\lambda(z))_{-}\partial_{s}),\quad i=1,\cdots,m,

where

dxλ(z)=xλ(z)+esλ(z)sx.d_{x}\lambda(z)=\partial_{x}\lambda(z)+e^{s}\lambda^{\prime}(z)\frac{\partial s}{\partial x}.

References

  • [1] Dubrovin B. Geometry of 2d topological field theories. In Integrable Systems and Quantum Groups, pages 120–348. Springer, 1998.
  • [2] Manin I. Frobenius manifolds, quantum cohomology, and moduli spaces, volume 47. American Mathematical Soc., 1999.
  • [3] Hertling C. Frobenius manifolds and moduli spaces for singularities, volume 151. Cambridge University Press, 2002.
  • [4] Dubrovin B. and Zhang Y. Normal forms of hierarchies of integrable PDEs, Frobenius manifolds and Gromov-Witten invariants. arXiv preprint math/0108160, 2001.
  • [5] Givental A. and Milanov E. Simple singularities and integrable hierarchies. In The Breadth of Symplectic and Poisson Geometry: Festschrift in Honor of Alan Weinstein, pages 173–201. Springer, 2005.
  • [6] Liu S.-Q., Ruan Y., and Zhang Y. BCFG Drinfeld–Sokolov hierarchies and FJRW-theory. Inventiones mathematicae, 201:711–772, 2015.
  • [7] Dubrovin B., Liu S.-Q., Yang D., and Zhang Y. Hodge integrals and tau-symmetric integrable hierarchies of Hamiltonian evolutionary PDEs. Advances in Mathematics, 293:382–435, 2016.
  • [8] Liu S.-Q., Wang Z., and Zhang Y. Variational bihamiltonian cohomologies and integrable hierarchies II: Virasoro symmetries. Communications in Mathematical Physics, 395(1):459–519, 2022.
  • [9] Liu S.-Q., Wang Z., and Zhang Y. Variational bihamiltonian cohomologies and integrable hierarchies I: foundations. Communications in Mathematical Physics, 401(1):985–1031, 2023.
  • [10] Chen Y.-T. and Tu M.-H. A note on the dispersionless Dym hierarchy. Letters in Mathematical Physics, 63:125–139, 2003.
  • [11] Tu M.-H. On dispersionless Lax representations for two-primary models. Journal of Physics A: Mathematical and Theoretical, 40(31):9427, 2007.
  • [12] Aoyama S. and Kodama Y. Topological Landau-Ginzburg theory with a rational potential and the dispersionless KP hierarchy. Communications in mathematical physics, 182:185–219, 1996.
  • [13] Raimondo A. Frobenius manifold for the dispersionless Kadomtsev-Petviashvili equation. Communications in Mathematical Physics, 311:557–594, 2012.
  • [14] Carlet G. and Mertens L. P. Principal hierarchies of infinite-dimensional Frobenius manifolds: The extended 2d Toda lattice. Advances in Mathematics, 278:137–181, 2015.
  • [15] Dubrovin B. On almost duality for Frobenius manifolds. American Mathematical Society Translations, 212:75–132, 2004.
  • [16] Strachan I.A.B. Frobenius submanifolds. Journal of Geometry and Physics, 38(3):285–307, 2001.
  • [17] Strachan I.A.B. Frobenius manifolds: natural submanifolds and induced bi-hamiltonian structures. Differential Geometry and its Applications, 20(1):67–99, 2004.
  • [18] Horev A. and Solomon J. P. The open Gromov-Witten-Welschinger theory of blowups of the projective plane. arXiv preprint arXiv:1210.4034, 2012.
  • [19] Alcolado A. Extended Frobenius manifolds and the open WDVV equations. McGill University (Canada), 2017.
  • [20] Basalaev A. and Buryak A. Open WDVV equations and Virasoro constraints. Arnold Mathematical Journal, 5(2):145–186, 2019.
  • [21] Arsie A., Buryak A., Lorenzoni P., and Rossi P. Semisimple flat F-manifolds in higher genus. Communications in Mathematical Physics, 397(1):141–197, 2023.
  • [22] Arsie A., Buryak A., Lorenzoni P., and Rossi P. Flat F-manifolds, F-CohFTs, and integrable hierarchies. Communications in Mathematical Physics, 388(1):291–328, 2021.
  • [23] Pandharipande R., Solomon J. P., and Tessler R. J. Intersection theory on moduli of disks, open KdV and Virasoro. arXiv preprint arXiv:1409.2191, 2014.
  • [24] Buryak A. Equivalence of the open KdV and the open Virasoro equations for the moduli space of Riemann surfaces with boundary. Letters in Mathematical Physics, 105:1427–1448, 2015.
  • [25] Buryak A. Open intersection numbers and the wave function of the KdV hierarchy. Moscow Mathematical Journal, 16(1):27–44, 2016.
  • [26] Buryak A., Clader E., and Tessler R. J. Open r-spin theory and the Gelfand-Dickey wave function. J. Geom. Phys., 137:132, 2019.
  • [27] Buryak A., Clader E., and Tessler R. J. Open rr-spin theory II: the analogue of Witten’s conjecture for rr-spin disks. Journal of Differential Geometry, 128(1):1–75, 2024.
  • [28] Ma S. Infinite-dimensional Frobenius manifolds underlying the genus-zero universal Whitham hierarchy. arXiv preprint arXiv:2406.08239, 2024.
  • [29] Dubrovin B. Differential geometry of the space of orbits of a Coxeter group. Surveys in Differential Geometry, 4(1):181–211, 1998.
  • [30] Liu S.-Q., Zhang Y., and Zhou X. Central invariants of the constrained KP hierarchies. Journal of Geometry and Physics, 97:177–189, 2015.
  • [31] Carlet G., van de Leur J., Posthuma H., and Shadrin S. Enumeration of hypermaps and Hirota equations for extended rationally constrained KP. Communications in Number Theory and Physics, 17(3):643–708, 2023.
  • [32] Zuo D. Frobenius manifolds associated to BlB_{l} and DlD_{l}, revisited. International Mathematics Research Notices, 2007:rnm020, 2007.
  • [33] Liu S.-Q., Wu C.-Z., and Zhang Y. On the Drinfeld–Sokolov hierarchies of D type. International Mathematics Research Notices, 2011(8):1952–1996, 2011.
  • [34] Dubrovin B. and Zhang Y. Extended affine Weyl groups and Frobenius manifolds. Compositio Mathematica, 111(2):167–219, 1998.
  • [35] Carlet G., Dubrovin B., and Zhang Y. The extended Toda hierarchy. Moscow Mathematical Journal, 4(2):313–332, 2004.
  • [36] Zhang Y. On the CP1CP^{1} topological sigma model and the Toda lattice hierarchy. Journal of Geometry and Physics, 40(3-4):215–232, 2002.
  • [37] Dubrovin B., Strachan I.A.B, Zhang Y., and Zuo D. Extended affine Weyl groups of BCD-type: their Frobenius manifolds and Landau–Ginzburg superpotentials. Advances in Mathematics, 351:897–946, 2019.
  • [38] Cheng J. and Milanov T. Hirota quadratic equations for the Gromov–Witten invariants of n2,,2.21.\mathbb{P}_{n-2,,2.2}^{1}. Advances in Mathematics, 388:107860, 2021.
  • [39] Cheng J. and Milanov T. The extended D-toda hierarchy. Selecta Mathematica, 27:1–85, 2021.
  • [40] Buryak A., Zernik A. N., Pandharipande R., and Tessler R. J. Open CP1CP^{1} descendent theory I: The stationary sector. Advances in Mathematics, 401:108249, 2022.
  • [41] Manin Y. I. F-manifolds with flat structure and Dubrovin’s duality. Advances in Mathematics, 198(1):5–26, 2005.
  • [42] Arsie A. and Lorenzoni P. Flat F-manifolds, Miura invariants, and integrable systems of conservation laws. Journal of Integrable Systems, 3(1):xyy004, 2018.
  • [43] Wu C.-Z. and Zhou X. An extension of the Kadomtsev–Petviashvili hierarchy and its Hamiltonian structures. Journal of Geometry and Physics, 106:327–341, 2016.