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

Solutions to SU(n+1){\rm SU}(n+1) Toda system with cone singularities via toric curves on compact Riemann surfaces

Jingyu Mu School of Mathematical Sciences, University of Science and Technology of China, Hefei 230026 China jingyu@mail.ustc.edu.cn Yiqian Shi School of Mathematical Sciences and CAS Wu Wen-Tsun Key Laboratory of Mathematics, University of Science and Technology of China, Hefei 230026 China yqshi@ustc.edu.cn  and  Bin Xu School of Mathematical Sciences and CAS Wu Wen-Tsun Key Laboratory of Mathematics, University of Science and Technology of China, Hefei 230026 China bxu@ustc.edu.cn
Abstract.

On a compact Riemann surface XX with finite punctures P1,,PkP_{1},\ldots,P_{k}, we define toric curves as multi-valued, totally unramified holomorphic maps to n\mathbb{P}^{n} with monodromy in a maximal torus of PSU(n+1){\rm PSU}(n+1). Toric solutions for the SU(n+1){\rm SU}(n+1) system on X{P1,,Pk}X\setminus\{P_{1},\ldots,P_{k}\} are recognized by their associated toric curves in n\mathbb{P}^{n}. We introduce a character n-ensemble as an nn-tuple of meromorphic one-forms with simple poles and purely imaginary periods, generating toric curves on XX minus finitely many points. We establish on XX a correspondence between character nn-ensembles and toric solutions to the SU(n+1){\rm SU}(n+1) system with finitely many cone singularities. Our approach not only broadens seminal solutions for up to two cone singularities on the Riemann sphere, as classified by Jost-Wang (Int. Math. Res. Not., (6):277-290, 2002) and Lin-Wei-Ye (Invent. Math., 190(1):169-207, 2012), but also advances beyond the limits of Lin-Yang-Zhong’s existence theorems (J. Differential Geom., 114(2):337-391, 2020) by introducing a new solution class.

Key words and phrases:
SU(n+1){\rm SU}(n+1) Toda system, regular singularity, unitary curve, Toric solution, character ensemble
2020 Mathematics Subject Classification:
Primary 37K10; Secondary 35J47
Y.S. is supported in part by the National Natural Science Foundation of China (Grant No. 11931009) and Anhui Initiative in Quantum Information Technologies (Grant No. AHY150200). B.X. is supported in part by the Project of Stable Support for Youth Team in Basic Research Field, CAS (Grant No. YSBR-001) and the National Natural Science Foundation of China (Grant Nos. 12271495, 11971450 and 12071449).
B.X. is the corresponding author.

1. Introduction

Gervais-Matsuo [16, Section 2.2.] were the first to demonstrate that totally unramified holomorphic curves [19, p. 270, Proposition] in n{\mathbb{P}}^{n} yield local solutions to SU(n+1){\rm SU}(n+1) Toda system, interpreting these systems as the infinitesimal Plücker formula [19, p. 269] for these curves. Adam Doliwa [9] expanded upon this foundation, extending their findings to Toda systems associated with non-exceptional simple Lie algebras. Over the ensuing years, Luca Battaglia, Chang-shou Lin, and their collaborators [1, 2, 3, 22, 25] have devoted significant research efforts to classifying and empirically substantiating existence theorems for solutions to Toda systems. These systems are associated with any complex simple Lie algebra and feature cone singularities on compact Riemann surfaces (Definition 1). In the next three subsections, we aim to highlight the principal advancements that are pertinent to our thematic focus within this field. To lay the foundation for our discussion, we initially explore the basic correspondence between solutions to the SU(n+1)\mathrm{SU}(n+1) Toda system with cone singularities and their associated unitary curves in n\mathbb{P}^{n} with regular singularities (Definition 2) in the first two subsections.

1.1. A differential-geometric framework for solutions to Toda systems

Building on the seminal contributions of Givental [17] and Positselski [29], this manuscript establishes on Riemann surfaces a new conceptual framework for solutions to SU(n+1)\mathrm{SU}(n+1) Toda system with cone singularities (Definition 1). As elucidated in [28, Theorem 1.2. (ii)], our analytical approach scrutinizes the solutions as nn-tuples of Kähler metrics with cone singularities on Riemann surfaces. Importantly, our methodology not only aligns with but also translates into the ones employed by a spectrum of researchers including Jost-Wang [21], Lin-Wei-Ye [24], Battaglia [1, 2, 3], Lin-Yang-Zhong [25], Chen-Lin [7, 8], and Chen [6], thereby establishing a conceptual equivalence across these varied approaches. This framework is versatile, allowing for an extension to Toda systems associated with any complex simple Lie algebra on Kähler manifolds.

Consider a Riemann surface 𝔛\mathfrak{X}, not necessarily compact, alongside a closed discrete subset 𝔖\mathfrak{S} within 𝔛\mathfrak{X}, where notably, 𝔖\mathfrak{S} is at most countable and possibly empty. We consider a vector \vvω\vv{\omega} of Kähler metrics on the punctured Riemann surface 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}, defined as

(1.1) \vvω:=(ω1=i2eu1dzdz¯,,ωn=i2eundzdz¯),\vv{\omega}:=\left(\omega_{1}=\frac{{\rm i}}{2}e^{u_{1}}\mathrm{d}z\wedge\mathrm{d}\bar{z},\ldots,\omega_{n}=\frac{{\rm i}}{2}e^{u_{n}}\mathrm{d}z\wedge\mathrm{d}\bar{z}\right),

together with the corresponding vector Ric(\vvω)\mathrm{Ric}\big{(}\vv{\omega}\big{)} of Ricci (1,1)(1,1)-forms derived from the metric components of \vvω\vv{\omega},

Ric(\vvω):=(Ric(ω1)=i¯u1,,Ric(ωn)=i¯un).\mathrm{Ric}\big{(}\vv{\omega}\big{)}:=\left(\mathrm{Ric}(\omega_{1})=-{\rm i}\partial\bar{\partial}u_{1},\ldots,\mathrm{Ric}(\omega_{n})=-{\rm i}\partial\bar{\partial}u_{n}\right).

We define \vvω\vv{\omega} as a solution to the SU(n+1)\mathrm{SU}(n+1) Toda system on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S} if it satisfies the equation

(1.2) Ric(\vvω)=\vvω(2aij)n×n\mathrm{Ric}(\vv{\omega})=\vv{\omega}\cdot(2a_{ij})_{n\times n}

on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}, where (aij)(a_{ij}) is the Cartan matrix for 𝔰𝔲(n+1)\mathfrak{su}(n+1), represented by

(aij)n×n=(210012100012100012100012)n×n.(a_{ij})_{n\times n}=\begin{pmatrix}2&-1&0&\ldots&\ldots&0\\ -1&2&-1&0&\ldots&0\\ 0&-1&2&-1&\ldots&0\\ \vdots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&\ldots&0&-1&2&-1\\ 0&\ldots&0&0&-1&2\end{pmatrix}_{n\times n}.

This system, \vvω=Ric(\vvω)(2aij)n×n\vv{\omega}=\mathrm{Ric}\big{(}\vv{\omega}\big{)}\cdot(2a_{ij})_{n\times n}, is identified as the SU(n+1)\text{SU}(n+1) Toda system on the Riemann surface, introducing a factor of 22 for consistency with the notation in [28]. It is important to note that a solution \vvω\vv{\omega} to (1.2) can exhibit wild behavior in the vicinity of punctures within 𝔖\mathfrak{S}. For instance, the integral of ωj\omega_{j} in a neighborhood around a point in 𝔖\mathfrak{S} may diverge. Since the work of Jost-Lin-Wang [20, Proposition 3.1.], there has been an ongoing investigation into solutions of Toda systems with cone singularities. We will soon provide a precise definition for these solutions. To specify, we assign to each point P𝔖P\in\mathfrak{S} a vector \vvγP=(γP,1,,γP,n)\vv{\gamma_{P}}=(\gamma_{P,1},\ldots,\gamma_{P,n}) of real numbers which are greater than 1-1 and do not vanish simultaneously, encapsulated into the \mathbb{R}-divisor vector \vv𝔇:=(𝔇1,,𝔇n)\vv{\mathfrak{D}}:=\big{(}{\mathfrak{D}}_{1},\ldots,{\mathfrak{D}}_{n}\big{)} with 𝔇1=P𝔖γP,1[P],,𝔇n=P𝔖γP,n[P]\mathfrak{D}_{1}=\sum_{P\in\mathfrak{S}}\gamma_{P,1}[P],\ldots,\mathfrak{D}_{n}=\sum_{P\in\mathfrak{S}}\gamma_{P,n}[P]. Denoting by δ\vv𝔇\delta_{\vv{\mathfrak{D}}} the following vector of (1,1)(1,1)-currents [19, Chapter 3],

(1.3) δ\vv𝔇:=2π(P𝔖γP,1δP,,P𝔖γP,nδP),\delta_{\vv{\mathfrak{D}}}:=2\pi\left(\sum_{P\in\mathfrak{S}}\gamma_{P,1}\delta_{P},\ldots,\sum_{P\in\mathfrak{S}}\gamma_{P,n}\delta_{P}\right),

we give the following:

Definition 1.

We consider \vvω\vv{\omega} a solution to the SU(n+1)\mathrm{SU}(n+1) Toda system on 𝔛\mathfrak{X} with cone singularities \vv𝔇\vv{\mathfrak{D}} (i.e., representing \vv𝔇\vv{\mathfrak{D}}) if it has finite area over each compact subset KK of 𝔛\mathfrak{X}, i.e. Kωj<\int_{K}\,\omega_{j}<\infty for all 1jn1\leq j\leq n, and satisfies the system

(1.4) Ric(\vvω)=δ\vv𝔇+\vvω(2aij)n×n\mathrm{Ric}\big{(}\vv{\omega}\big{)}=-\delta_{\vv{\mathfrak{D}}}+\vv{\omega}\cdot(2a_{ij})_{n\times n}

in the sense of (1,1)(1,1)-current on XX. In particular, this system, when restricted to a sufficiently small chart (U,z)(U,z) around each P𝔖P\in\mathfrak{S}, takes the following form:

i2(¯ui+j=1naijeujdzdz¯)=πγP,iδP,1in.\frac{\rm i}{2}\Big{(}\partial\bar{\partial}u_{i}+\sum\limits_{j=1}^{n}a_{ij}e^{u_{j}}\mathrm{d}z\wedge\mathrm{d}\bar{z}\Big{)}=\pi\gamma_{P,i}\delta_{P},\quad 1\leq i\leq n.

By [28, Theorem 1.2. (ii)], for each 1in1\leq i\leq n, the component ωi=i2euidzdz¯\omega_{i}=\frac{{\rm i}}{2}e^{u_{i}}\mathrm{d}z\wedge\mathrm{d}\bar{z} of a solution \vvω\vv{\omega} to (1.4) forms a cone Kähler metric on XX with cone angle 2π(1+γP,i)2\pi(1+\gamma_{P,i}) at P𝔖P\in\mathfrak{S}. In this context, we say that \vvω\vv{\omega} is a solution with cone singularities \vv𝔇\vv{\mathfrak{D}} (i.e., representing \vv𝔇\vv{\mathfrak{D}}).

It is noteworthy that when n=1n=1, the term \vvω=ω1\vv{\omega}=\omega_{1} denotes a cone spherical metric that represents the divisor 𝔇1\mathfrak{D}_{1} on 𝔛\mathfrak{X}. For further insights into cone spherical metrics, see [26] and the references contained therein. Unlike the conventional definition of the Toda system with singularities, our intrinsic approach does not rely on a predetermined background Kähler metric and situates the problem within the context of Kähler Geometry. Specifically, our ongoing project is dedicated to investigating vectors of Kähler metrics with cone singularities along divisors [10] as potential solutions to the Toda systems associated with any complex simple Lie algebra on Kähler manifolds. This approach naturally extends both the Toda system (1.4) on Riemann surfaces and the Monge-Ampère equation concerning cone Kähler-Einstein metrics with positive scalar curvatures on Kähler manifolds of dimension 2\geq 2 ([10]).

1.2. A basic correspondence

Drawing upon the work of Gervais-Matsuo [16, Section 2.2.], Jost-Wang [21, Section 3], and [28, Section 2], in this subsection, we shall present a concise overview of the basic correspondence between solutions to the SU(n+1){\rm SU}(n+1) Toda systems with cone singularities \vv𝔇\vv{\mathfrak{D}} on 𝔛\mathfrak{X}, and totally unramified unitary curves on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S} and with regular singularities \vv𝔇\vv{\mathfrak{D}}. This overview lays a robust foundation for articulating the main results of our manuscript and situating them within the context of classical findings such as [16, 21, 24, 23].

A totally unramified unitary curve ff on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S} is a multi-valued holomorphic map f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n}, characterized by the following properties:

  • The monodromy of ff resides within PSU(n+1){\rm PSU}(n+1), the group of holomorphic isometries preserving the Fubini-Study metric ωFS\omega_{\rm FS} on n\mathbb{P}^{n}.

  • At each point z𝔛𝔖z\in\mathfrak{X}\setminus\mathfrak{S}, any germ 𝔣\mathfrak{f} of ff at zz is totally unramified and, notably, non-degenerate, near zz.

For a more detailed exposition of this concept, please refer to [28, Definition 2.1.].

Definition 2.

We define a totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n} to have regular singularities

\vv𝔇=(P𝔖γP,1[P],,P𝔖γP,n[P])\vv{\mathfrak{D}}=\left(\sum_{P\in\mathfrak{S}}\,\gamma_{P,1}[P],\ldots,\sum_{P\in\mathfrak{S}}\,\gamma_{P,n}[P]\right)

i.e., representing \vv𝔇\vv{\mathfrak{D}}, if, for each point P𝔖P\in\mathfrak{S}, there exists an element φPSU(n+1)\varphi\in{\rm PSU}(n+1) such that the composition φf\varphi\circ f when restricted to a punctured disk {0<|z|<1}𝔛𝔖\{0<|z|<1\}\subset\mathfrak{X}\setminus\mathfrak{S} around PP can be expressed as

(1.5) φf(z)=[zβP,0g0(z):zβP,1g1(z)::zβP,ngn(z)],\varphi\circ f(z)=\left[z^{\beta_{P,0}}g_{0}(z):z^{\beta_{P,1}}g_{1}(z):\ldots:z^{\beta_{P,n}}g_{n}(z)\right],

where the functions g0(z),,gn(z)g_{0}(z),\ldots,g_{n}(z) are holomorphic and nonvanishing at z=0z=0, i.e., PP. The exponents βP,0,,βP,n\beta_{P,0},\ldots,\beta_{P,n} are real numbers satisfying

(1.6) βP,jβP,j1=γP,j+1for all1jn.\beta_{P,j}-\beta_{P,j-1}=\gamma_{P,j}+1\quad\text{for all}\quad 1\leq j\leq n.

In this context, given a point PP in 𝔖\mathfrak{S}, if all γP,j\gamma_{P,j}’s are non-negative integers and do not vanish simultaneously, then P𝔖P\in\mathfrak{S} is referred to as a ramification point of ff ([19, pp. 266-268]); if at least one of {γP,j}j=1n\{\gamma_{P,j}\}_{j=1}^{n} is non-integer, then PP is called a branch point of ff. Both ramification points and branch points are called regular singularities of ff. It is important to note that ff is totally unramified at a point PXP\in X if and only if γP,j=0\gamma_{P,j}=0 for all 1jn1\leq j\leq n.

From a totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n}, one can derive a solution to the SU(n+1){\rm SU}(n+1) Toda system (1.2) on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}. In fact, for 0jn10\leq j\leq n-1, the jj-th associated curve fj:𝔛𝔖G(j+1,n+1)(Λj+1n+1)f_{j}:\mathfrak{X}\setminus\mathfrak{S}\to G(j+1,n+1)\subset\mathbb{P}\left(\Lambda^{j+1}\mathbb{C}^{n+1}\right) of ff is also a unitary curve ([19, pp. 263-264] and [28, Definition 2.1.]). For simplicity, we uniformly adopt the symbol ωFS\omega_{\rm FS} to represent the Fubini-Study metrics on (Λj+1n+1)\mathbb{P}\left(\Lambda^{j+1}\mathbb{C}^{n+1}\right) for all 0jn10\leq j\leq n-1. By employing the infinitesimal Plücker formula, f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n} produces a vector of Kähler metrics

(1.7) \vvω=(f0ωFS,,fn1ωFS),\vv{\omega}=\left(f_{0}^{*}\omega_{\rm FS},\ldots,f_{n-1}^{*}\omega_{\rm FS}\right),

which serves as a solution to (1.2) on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}, as delineated in [28, Lemma 2.2]. Conversely, each solution \vvω\vv{\omega} to the SU(n+1){\rm SU}(n+1) Toda system (1.2) on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S} gives rise to a series of totally unramified unitary curves 𝔛𝔖n\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n}, where any two are distinguishable by a post-composition of an element in PSU(n+1){\rm PSU}(n+1). Furthermore, these curves reconstruct the solution ω\omega as defined in (1.7), and they are termed curves associated with ω\omega. The intricacies will be expounded in Lemma 3. Therefore, we have finalized the exposition detailing the correspondence between solutions to the SU(n+1){\rm SU}(n+1) Toda system, as described in (1.2), and totally unramified unitary curves on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}. In Theorem 4, we further elucidate a refined correspondence between solutions to the SU(n+1){\rm SU}(n+1) Toda system (1.4) with cone singularities \vv𝔇\vv{\mathfrak{D}} on 𝔛\mathfrak{X}, and totally unramified unitary curves on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S} with regular singularities \vv𝔇\vv{\mathfrak{D}}.

1.3. Exploring the concepts of toric curves and toric solutions

We introduce the following innovative concept associated with the SU(n+1)\mathrm{SU}(n+1) Toda system, inspired by its n=1n=1 scenario — specifically, the reducible cone spherical metric [32, 5, 27, 12].

Definition 3.

A toric curve f:𝔛𝔖f:\mathfrak{X}\setminus\mathfrak{S} is defined as a totally unramified unitary curve whose monodromy resides within a maximal torus of PSU(n+1)\mathrm{PSU}(n+1). A solution ω\omega to the SU(n+1)\mathrm{SU}(n+1) Toda system on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S} is termed toric if it yields toric associated curves. This leads to the concepts of a toric curve with regular singularities \vv𝔇\vv{\mathfrak{D}} and a toric solution with cone singularities \vv𝔇\vv{\mathfrak{D}}, respectively.

Having established the foundation for our methodology in addressing Toda systems, now is an opportune moment to review the classifications of solutions and the significant findings regarding their existence. The pioneering effort by Jost-Wang [21] in 2002 revealed that curves associated with finite-area solutions to the SU(n+1){\rm SU}(n+1) Toda system on the complex plane \mathbb{C} extend to rational normal curves from 1\mathbb{P}^{1} to n\mathbb{P}^{n}. By this, they obtained a thorough classification of solutions on 1\mathbb{P}^{1}. In 2007, Eremenko [11, Theorem 2] further classified curves linked to solutions expanding polynomially in area at \infty of \mathbb{C}. Continuing this trajectory, Lin-Wei-Ye [24], five years on, managed to classify all solutions to the SU(n+1){\rm SU}(n+1) system with two cone singularities on 1\mathbb{P}^{1}. Simultaneously, they also characterized the corresponding cone singularities. Building on this, quite recently, Karmakar-Lin-Nie-Wei [22] broadened the scope to include any Toda system tied to a complex simple Lie algebra. The preceding solutions are all toric since the fundamental groups of the underlying surfaces are either trivial or isomorphic to {\mathbb{Z}}. In 2018, Lin-Nie-Wei [23] obtained some existence results about solutions to the SU(n+1){\rm SU}(n+1) system with three cone singularities on the Riemann sphere. Between 2015 and 2016, Battaglia [1, 2, 3] laid down extensive theorems on the existence and absence of solutions for the SU(3){\rm SU}(3) Toda system with cone singularities on compact Riemann surfaces. Following suit, Lin-Yang-Zhong [25] made significant strides by proving existence theorems for Toda systems related to Lie algebras of types AnA_{n}, BnB_{n}, CnC_{n}, and G2G_{2}, featuring cone singularities on compact Riemann surfaces with positive genera. More recently, Chen-Lin [7, 8] have unveiled numerous findings regarding the SU(3){\rm SU}(3) Toda system with cone singularities on tori. In their most recent collaboration [28], Sun and the authors of this manuscript meticulously categorized the solutions to the SU(n+1){\rm SU}(n+1) Toda system, delineated on the disk |z|<1{|z|<1}. These solutions are characterized by a cone singularity at z=0z=0, possess finite area, and inherently exhibit the toric property.

In the left part of this introductory section, we shall focus on toric solutions to the SU(n+1){\rm SU}(n+1) Toda system with cone singularities

\vvD=(D1=i=1kγi,1[Pi],,Dn=i=1kγi,n[Pi]),\vv{D}=\left(D_{1}=\sum_{i=1}^{k}\,\gamma_{i,1}[P_{i}],\ldots,D_{n}=\sum_{i=1}^{k}\,\gamma_{i,n}[P_{i}]\right),

where P1,,PkP_{1},\ldots,P_{k} are distinct points on a compact Riemann surface XX, and (γi,j)1ik1jn\Big{(}\gamma_{i,j}\Big{)}_{\begin{subarray}{c}1\leq i\leq k\\ 1\leq j\leq n\end{subarray}} is a k×nk\times n matrix of real numbers greater than 1-1 such that for each 1ik1\leq i\leq k, at least one of {γij}1jn\{\gamma_{ij}\}_{1\leq j\leq n} does not vanish. We call {P1,,Pk}\{P_{1},\cdots,P_{k}\} the support, and (γi,j)1ik1jn\Big{(}\gamma_{i,j}\Big{)}_{\begin{subarray}{c}1\leq i\leq k\\ 1\leq j\leq n\end{subarray}} the coefficient matrix of \vvD\vv{D}. Chen, Wang, Wu, and the last author [5, Theorems 1.4. and 1.5.] established on XX a correspondence between toric solutions to the SU(2)\mathrm{SU}(2) Toda system with cone singularities, namely reducible cone spherical metrics, and meromorphic one-forms with simple poles and purely imaginary periods on XX. This correspondence is particularly noteworthy as it allows for the explicit characterization of the singularity information of a reducible metric in terms of the corresponding one-form, which was referred to as the character one-form of the metric therein. Moreover, meromorphic one-forms with simple poles and purely imaginary periods are plentifully available on Riemann surfaces, cf. [33, §\S15], [30, §\S8-1] and [14, §\SII.4-5] for precise statements and their proof. In particular, the explicit formulation of such one-forms was meticulously documented in [5, Example 4.7.] on the Riemann sphere. To generalize this correspondence for toric solutions to SU(n+1){\rm SU}(n+1) Toda system with cone singularities on XX, we introduce the following:

Definition 4.

A character nn-ensemble on XX is defined as a vector \vvΩ=(Ω1,,Ωn)\vv{\Omega}=(\Omega_{1},\ldots,\Omega_{n}), composed of nn meromorphic one-forms with simple poles and purely imaginary periods on XX such that there exist finitely many points P1,,PkP_{1},\ldots,P_{k} on XX and

(1.8) f\vvΩ,\vvρ(z):=[1:ρ1exp(zΩ1)::ρnexp(zΩn)]f_{\vv{\Omega},\vv{\rho}}(z):=\left[1:\rho_{1}\cdot\exp\left(\int^{z}\Omega_{1}\right):\ldots:\rho_{n}\cdot\exp\left(\int^{z}\Omega_{n}\right)\right]

generates a family of totally unramified unitary curves mapping from X{P1,,Pk}X\setminus\{P_{1},\ldots,P_{k}\} to n\mathbb{P}^{n}, where \vvρ=(ρ1,,ρn)\vv{\rho}=(\rho_{1},\ldots,\rho_{n})’s vary over all vectors of nn positive real numbers. Importantly, these curves exhibit monodromy within the diagonal maximal torus 𝕋n\mathbb{T}^{n} of PSU(n+1){\rm PSU}(n+1), specified by

𝕋n:={φ([z0::zn])=[eiθ0z0::eiθnzn]|(θ0,,θn)n+1}.\mathbb{T}^{n}:=\left\{\varphi([z_{0}:\ldots:z_{n}])=\left[e^{\mathrm{i}\theta_{0}}z_{0}:\ldots:e^{\mathrm{i}\theta_{n}}z_{n}\right]\,\middle|\,(\theta_{0},\ldots,\theta_{n})\in\mathbb{R}^{n+1}\right\}.

Hence, they are inherently toric curves from X{P1,,Pk}X\setminus\{P_{1},\ldots,P_{k}\} to n\mathbb{P}^{n}. Within the given context, it can be demonstrated that the preceding curves described inherently possess the same regular singularities, denoted by \vvD\vv{D}. These singularities are characterized by a support {P1,,Pk}\{P_{1},\cdots,P_{k}\} and their coefficient matrix (γi,j)1ik1jn\Big{(}\gamma_{i,j}\Big{)}_{\begin{subarray}{c}1\leq i\leq k\\ 1\leq j\leq n\end{subarray}}, both of which are implicitly determined by the nn-ensemble \vvΩ\vv{\Omega} (Proposition 1 and Lemma 5).

1.4. Main results

We establish on XX the following general correspondence for the SU(n+1){\rm SU}(n+1) Toda system which puts the one in [5] as its n=1n=1 scenario.

Theorem 1.

There exists a correspondence between character nn-ensembles and toric solutions ω\omega to the PSU(n+1)\mathrm{PSU}(n+1) Toda system with cone singularities on a compact Riemann surface XX. Specifically, the following two statements hold:

  1. (1)

    Given a character nn-ensemble \vvΩ\vv{\Omega} on XX, it induces a family of toric curves via (1.8) with regular singularities \vvD\vv{D} on XX, which in turn defines a corresponding family of toric solutions representing \vvD\vv{D}, via (1.7).

  2. (2)

    Given a toric solution \vvω\vv{\omega} with cone singularities \vvD\vv{D} on XX, it has an associated toric curve f=[f0,,fn]f=[f_{0},\ldots,f_{n}] with monodromy in 𝕋n\mathbb{T}^{n} and with regular singularities \vvD\vv{D} on XX such that

    Ω:=(d[logf1f0],,d[logfnf0])\Omega:=\left(\mathrm{d}\left[\log\frac{f_{1}}{f_{0}}\right],\ldots,\mathrm{d}\left[\log\frac{f_{n}}{f_{0}}\right]\right)

    forms a character nn-ensemble on XX.

A natural question arises regarding the characterization of regular singularities on curves generated by a character ensemble, as defined in (1.8). The following theorem offers a partial response to this inquiry.

Theorem 2.

Let \vvΩ\vv{\Omega} be a character nn-ensemble and ff one in the family of curves generated by \vvΩ\vv{\Omega} in terms of (1.8) on a compact Riemann surface XX. Denote by 𝒫\mathcal{P} the set of poles of all Ωj\Omega_{j}’s. There holds the following statements:

  • (1)

    A point PP on XX is a branch point of ff if and only if it is a pole of some component Ωj\Omega_{j} of \vvΩ\vv{\Omega} such that the residue of Ωj\Omega_{j} at PP is noninteger.

  • (2)

    A zero of \vvΩ\vv{\Omega}, at which all components of \vvΩ\vv{\Omega} vanish, is a ramification point of ff.

  • (3)

    An algorithm is presented that identifies the regular singularities \vvD\vv{D} of ff in finitely many steps. Its details will be provided within the proof.

Below, we present two examples of toric curves with regular singularities that yield novel toric solutions to the SU(n+1){\rm SU}(n+1) Toda systems with cone singularities. These examples extend beyond the scope of quite recent existence results presented in [25, Theorems 1.8-9.].

Example 1.

Consider a compact Riemann surface XX and let Ω\Omega be a nontrivial meromorphic one-form on XX that has simple poles and purely imaginary periods. Define the vector of scaled one-forms

\vvΩ:=(λ1Ω,,λnΩ),\vv{\Omega}:=(\lambda_{1}\Omega,\ldots,\lambda_{n}\Omega),

where λ1,,λn\lambda_{1},\ldots,\lambda_{n} are distinct nonzero real numbers. This vector \vvΩ\vv{\Omega} constitutes a character nn-ensemble on XX and generates a family of toric solutions that are parametrized by (>0)n(\mathbb{R}_{>0})^{n} as per Equation (1.8).

Furthermore, the singularities of these toric solutions can be fully characterized by \vvΩ\vv{\Omega}. In particular, all the cone singularities of the solutions are confined to the set of zeroes and poles of Ω\Omega. Specifically, at a zero 𝔮\mathfrak{q} of Ω\Omega, each metric component in the toric solution exhibits a cone angle of 2π(1+ord𝔮Ω).2\pi\left(1+\operatorname{ord}_{\mathfrak{q}}\Omega\right).

Example 2.

Identify 1\mathbb{P}^{1} with the Riemann sphere \mathbb{C}\cup\infty. Consider the vector (γ1,,γn)(1,)nn(\gamma_{1},\dots,\gamma_{n})\in(-1,\infty)^{n}\setminus\mathbb{Z}^{n} and a positive integer mm. Given mm distinct points z1,,zmz_{1},\dots,z_{m} on {0}\mathbb{C}\setminus\{0\} and a matrix of non-negative integers (γi,j)1im1jn\Big{(}\gamma_{i,j}\Big{)}_{\begin{subarray}{c}1\leq i\leq m\\ 1\leq j\leq n\end{subarray}}, there exists a family of toric curves from 1{0,,z1,,zm}\mathbb{P}^{1}\setminus\{0,\infty,z_{1},\ldots,z_{m}\} to n\mathbb{P}^{n} parametrized by (>0)n\big{(}\mathbb{R}_{>0}\big{)}^{n} satisfying the following properties:

  • All the curves have the same regular singularities. In particular, they have exactly two branch points at 0 and \infty, with γ[0],j=γj\gamma_{[0],j}=\gamma_{j} for all 1jn1\leq j\leq n;

  • The points z1,,zmz_{1},\dots,z_{m} are all ramification points of them, where γ[zi],j=γi,j\gamma_{[z_{i}],j}=\gamma_{i,j} for all 1im1\leq i\leq m and 1jn1\leq j\leq n.

Moreover, the vector (γ[],j)j=1n\left(\gamma_{[\infty],j}\right)_{j=1}^{n} has finitely many choices, bounded from above by

(i=1m(nγi,1+(n1)γi,2++γi,n)+n+1n).\left(\genfrac{}{}{0.0pt}{}{\sum_{i=1}^{m}\big{(}n\gamma_{i,1}+(n-1)\gamma_{i,2}+\cdots+\gamma_{i,n}\big{)}+n+1}{n}\right).

Outline. The introduction concludes with an outline of the structure for the remainder of this manuscript. Section 2 focuses on establishing the relationship between curves with regular singularities and solutions with cone singularities, as detailed in Theorem 4. The proof of Theorem 2 is presented in Section 3, followed by a detailed substantiation of Theorem 1 in Section 4. Additionally, Section 3 introduces a new concept of a non-degenerate nn-tuple of one-forms. This concept is simpler yet equivalent to the character nn-ensemble, as demonstrated in Proposition 1. Section 5 provides a comprehensive discussion of the two preceding examples. In the final section, we propose three open questions to guide further investigation.

2. Correspondence between curves and solutions

In this section, we shall prove a lemma and a theorem that sequentially facilitate the establishment of the basic correspondence between solutions to the SU(n+1){\rm SU}(n+1) Toda systems with cone singularities and unitary curves with regular singularities. This correspondence was initially outlined in Subsection 1.2. Throughout, we will consistently use the notations previously introduced.

Lemma 3.
  • (1)

    From Curve to Solution: A totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\rightarrow\mathbb{P}^{n} induces a solution \vvω:=(f0ωFS,,fn1ωFS)\vv{\omega}:=\big{(}f_{0}^{*}\omega_{\rm FS},\ldots,f_{n-1}^{*}\omega_{\rm FS}\big{)} to the SU(n+1){\rm SU}(n+1) Toda system on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}. Moreover, any curve φf\varphi\circ f, with φ\varphi in PSU(n+1){\rm PSU}(n+1), produces the same solution as ff.

  • (2)

    From Solution to Curve: Every solution \vvω\vv{\omega} to the SU(n+1){\rm SU}(n+1) Toda system on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S} corresponds to at least one totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\rightarrow\mathbb{P}^{n} that generates \vvω\vv{\omega}. Furthermore, any other curve that corresponds to \vvω\vv{\omega} will be in the form of φf\varphi\circ f for some φ\varphi in PSU(n+1){\rm PSU}(n+1).

Proof.

(1) Consider a totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n}. For the definition of the jj-th associated function fjf_{j} where 0jn0\leq j\leq n, we refer to [19, pp. 263-264] and [28, Definition 2.1]. The proof of the first statement closely mirrors that of Lemma 2.2 in [28], which specifically addresses the plane domain scenario. The second statement follows from the fact that each element φ\varphi of PSU(n+1){\rm PSU}(n+1) preserves the Fubini-Study metric on n\mathbb{P}^{n}.

(2) Consider a solution \vvω=(ω1,,ωn)\vv{\omega}=(\omega_{1},\ldots,\omega_{n}) to the SU(n+1){\rm SU}(n+1) Toda system on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}. Fix a point 𝔭\mathfrak{p} on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}. Restricting \vvω\vv{\omega} to a disc chart (D,|z|<1)(D,|z|<1) around 𝔭\mathfrak{p}, represented as \vvω|D\vv{\omega}|_{D}, results in a solution over DD. Employing [28, Lemma 2.3.], this restriction leads to the formation of a totally unramified holomorphic curve 𝔣𝔭:Dn\mathfrak{f}_{\mathfrak{p}}:D\to\mathbb{P}^{n}, which specifically induces \vvω|D\vv{\omega}|_{D}. Notably, the Fubini-Study form pullback via 𝔣𝔭\mathfrak{f}_{\mathfrak{p}}, denoted 𝔣𝔭ωFS\mathfrak{f}_{\mathfrak{p}}^{*}\omega_{\rm FS}, coincides with the restricted ω1\omega_{1} on DD. According to the local rigidity theorem posited by Eugenio Calabi (refer to [4, Theorem 9] and [18, (4.12)]), the curve 𝔣𝔭\mathfrak{f}_{\mathfrak{p}} is uniquely determined by \vvω\vv{\omega}, up to a transformation in PSU(n+1){\rm PSU}(n+1). Further, leveraging the developing map concept outlined in [31, §\S3.4], a unique, totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n} can be realized through analytic continuations of 𝔣𝔭\mathfrak{f}_{\mathfrak{p}} along curves on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}. Its monodromy results in a group homomorphism

f:π1(𝔛𝔖,𝔭)PSU(n+1).\mathcal{M}_{f}:\pi_{1}\big{(}\mathfrak{X}\setminus\mathfrak{S},\,\mathfrak{p}\bigr{)}\to{\rm PSU}(n+1).

This curve ff not only induces \vvω\vv{\omega} as initially described in (1) but is also uniquely characterized by the same local rigidity theorem within the confines of PSU(n+1){\rm PSU}(n+1). ∎

The lemma above can be generalized to include cases with cone (regular) singularities as follows:

Theorem 4.

We use the notations from Definitions 1 and 2.

  • (1)

    From curve with regular singularities to solution with cone singularities : A totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n} with regular singularities \vv𝔇\vv{\mathfrak{D}} induces a solution \vvω=(f0ωFS,,fn1ωFS)\vv{\omega}=(f_{0}^{*}\omega_{\textnormal{FS}},\ldots,f_{n-1}^{*}\omega_{\textnormal{FS}}) to the SU(n+1)\textnormal{SU}(n+1) Toda system on 𝔛\mathfrak{X} with cone singularities \vv𝔇\vv{\mathfrak{D}}. Moreover, any curve φf\varphi\circ f, where φ\varphi is in PSU(n+1)\textnormal{PSU}(n+1), produces the same solution as ff.

  • (2)

    From solution with cone singularities to curve with regular singularities : Every solution \vvω\vv{\omega} to the SU(n+1)\textnormal{SU}(n+1) Toda system on 𝔛\mathfrak{X} with cone singularities \vv𝔇\vv{\mathfrak{D}} corresponds to at least one totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n} with regular singularities \vv𝔇\vv{\mathfrak{D}} that generates \vvω\vv{\omega}. We call ff a curve associated with \vvω\vv{\omega}. Furthermore, any other curve that corresponds to \vvω\vv{\omega} will be in the form of φf\varphi\circ f for some φ\varphi in PSU(n+1)\textnormal{PSU}(n+1).

Proof.

By Lemma 3, it is sufficient to verify the properties concerning the singularities \vvD\vv{D}.

  • (1)

    Consider a totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n} with regular singularities \vv𝔇\vv{\mathfrak{D}}. According to Lemma 3(1), this curve induces a solution \vvω\vv{\omega} on 𝔛𝔖\mathfrak{X}\setminus\mathfrak{S}. Furthermore, based on [28, Formula 3.] and the infinitesimal Plücker formula, \vvω\vv{\omega} exhibits cone singularities \vv𝔇\vv{\mathfrak{D}}.

  • (2)

    Consider a solution \vvω=(ω1,,ωn)\vv{\omega}=(\omega_{1},\ldots,\omega_{n}) on 𝔛\mathfrak{X} with cone singularities \vvD\vv{D}. Lemma 3(2) states that this solution corresponds to a totally unramified unitary curve f:𝔛𝔖nf:\mathfrak{X}\setminus\mathfrak{S}\to\mathbb{P}^{n}. Moreover, according to [28, Theorem 1.2.(i)], curve ff has regular singularities \vvD\vv{D}.

3. Character ensembles and toric curves with regular singularities

In this section, we shall prove Theorem 2. To this end, we demonstrate in Proposition 1 that character nn-ensembles correspond to nn-tuples of meromorphic one-forms with simple poles and purely imaginary periods. These nn-tuples are non-degenerate in the sense that they define non-degenerate curves under equation (1.8), as established in Definition 5. This explanation renders the concept of a character ensemble much more accessible and relatable.

3.1. A non-degenerate nn-tuple of one-forms is a character nn-ensemble

Definition 5.

Let \vvΩ=(Ω1,,Ωn)\vv{\Omega}=(\Omega_{1},\ldots,\Omega_{n}) be an nn-tuple of meromorphic one-forms on a compact Riemann surface XX, characterized by having simple poles and purely imaginary periods. We define \vvΩ\vv{\Omega} to be non-degenerate if there exists a point 𝔭\mathfrak{p} in XX such that:

  • All components of \vvΩ\vv{\Omega} are holomorphic at 𝔭\mathfrak{p}.

  • In a disc chart (D,|z|<1)(D,|z|<1) around 𝔭\mathfrak{p}, i.e., z=0z=0, the map

    (3.1) 𝔣𝔭(z):=[1:exp(0zΩ1)::exp(0zΩn)]\mathfrak{f_{p}}(z):=\left[1:\exp\left(\int_{0}^{z}\Omega_{1}\right):\ldots:\exp\left(\int_{0}^{z}\Omega_{n}\right)\right]

    defines a holomorphic curve from DD to n\mathbb{P}^{n} that is non-degenerate, meaning that its image is not contained within any hyperplane of n\mathbb{P}^{n}.

It is noteworthy that in this context, the curve 𝔣𝔭(z)\mathfrak{f_{p}}(z) remains non-degenerate near every point 𝔭\mathfrak{p}, where all components of \vvΩ\vv{\Omega} are holomorphic. Denote by 𝒫\mathcal{P} the set of poles of all Ωj\Omega_{j}’s. Similar to Equation (1.8), \vvΩ\vv{\Omega} induces a family of non-degenerate, multi-valued holomorphic curves

(3.2) {f\vvΩ,ρ:\vvρ(>0)n}\left\{f_{\vv{\Omega},\vec{\rho}}:\vv{\rho}\in\big{(}\mathbb{R}_{>0}\big{)}^{n}\right\}

mapping from X𝒫X\setminus\mathcal{P} to n\mathbb{P}^{n}. These curves share identical monodromy in 𝕋n\mathbb{T}^{n} and possess the same regular singularities on XX.

Lemma 5.

Adopting the context of Definition 5 and considering a fixed vector ρ\vec{\rho} consisting of nn positive numbers, we have the following properties for the curve fΩ,ρf_{\vec{\Omega},\vec{\rho}}:

  1. (1)

    A point 𝔭\mathfrak{p} on XX is identified as a branch point of the curve if and only if 𝔭\mathfrak{p} is a pole of some Ωj\Omega_{j} where the residue Res𝔭Ωj{\rm Res}_{\mathfrak{p}}\Omega_{j} is not an integer.

  2. (2)

    The curve is totally unramified on XX, except at a finite number of regular singularities.

Proof.

(1) Take a disc chart (D,|z|<1)(D,|z|<1) around 𝔭\mathfrak{p} and denote by \vv𝔯=(𝔯1,,𝔯𝔫)n\vv{\mathfrak{r}}=\mathfrak{(r_{1},\ldots,r_{n})}\in\mathbb{R}^{n} the vector of residues of \vvΩ=(Ω1,,Ωn)\vv{\Omega}=\left(\Omega_{1},\ldots,\Omega_{n}\right) at 𝔭\mathfrak{p}. We choose a germ 𝔣𝔭\mathfrak{f_{p}} of fΩ,ρf_{\vec{\Omega},\vec{\rho}} in D:={0<|z|<1}D^{*}:=\{0<|z|<1\} with form

[1:z𝔯1h1(z)::z𝔯𝔫hn(z)]=[z0:z𝔯1h1(z)::z𝔯𝔫hn(z)],\left[1:z^{\mathfrak{r_{1}}}h_{1}(z):\ldots:z^{\mathfrak{r_{n}}}h_{n}(z)\right]=\left[z^{0}:z^{\mathfrak{r_{1}}}h_{1}(z):\ldots:z^{\mathfrak{r_{n}}}h_{n}(z)\right],

where each hjh_{j} is holomorphic in DD and does not vanish at z=0z=0.

Assume that the residues 𝔯j\mathfrak{r}_{j} are all integers. In this case, the curve fΩ,ρf_{\vec{\Omega},\vec{\rho}} can be extended holomorphically to the point 𝔭\mathfrak{p}. Therefore, at 𝔭\mathfrak{p}, the curve is either totally unramified or it is a ramification point. Specifically, 𝔭\mathfrak{p} is not a branch point. Furthermore, the ramification indices γ𝔭,j\gamma_{\mathfrak{p},j} for 1jn1\leq j\leq n can be calculated using the algorithm described on pages 266-268 in [19].

Suppose that at least one of the residues 𝔯j\mathfrak{r}_{j} is not an integer. Consequently, the set {0,𝔯1,,𝔯n}\{0,\mathfrak{r}_{1},\cdots,\mathfrak{r}_{n}\} is partitioned into at least two equivalence classes under the modulo \mathbb{Z} equivalence relation. According to the argument used in the proof in [28, Theorem 3.1.], there exists a transformation φPSU(n+1)\varphi\in\mathrm{PSU}(n+1) such that φ(𝔣𝔭)\varphi(\mathfrak{f_{p}}) is represented by

(3.3) [zb0:zb1H1(z)::zbnHn(z)],\left[z^{b_{0}}:z^{b_{1}}H_{1}(z):\ldots:z^{b_{n}}H_{n}(z)\right],

where b0<<bnb_{0}<\ldots<b_{n}, and each function Hj(z)H_{j}(z) is holomorphic in DD and non-vanishing at z=0z=0. Additionally, there is at least one index jj (where 0jn10\leq j\leq n-1) for which b0bj(mod)b_{0}\equiv\ldots\equiv b_{j}\pmod{\mathbb{Z}}, and γ𝔭,j+1=bj+1bj1\gamma_{\mathfrak{p},j+1}=b_{j+1}-b_{j}-1\notin\mathbb{Z}. Therefore, 𝔭\mathfrak{p} is identified as a branch point of the curve fΩ,ρf_{\vec{\Omega},\vec{\rho}}. Notably, we could read all γ𝔭,\gamma_{\mathfrak{p},\cdot} indices as

γ𝔭,j=bj+1bj1forall1jn\gamma_{\mathfrak{p},j}=b_{j+1}-b_{j}-1\quad{\rm for\ all}\quad 1\leq j\leq n

from (3.3), called the quasi-canonical form of ff.

(2) We aim to show that the function fΩ,ρf_{\vec{\Omega},\vec{\rho}} has finitely many ramification points on a compact Riemann surface XX. Assuming the contrary, let {qn}\{q_{n}\} be a sequence of distinct ramification points converging to some point 𝔭\mathfrak{p} on XX. We consider two cases based on the location of 𝔭\mathfrak{p}:

  • (i)

    Case 𝔭𝒫\mathfrak{p}\notin\mathcal{P}: Select a disc chart (D,|z|<1)(D,|z|<1) around 𝔭\mathfrak{p}, which includes {qn}\{q_{n}\} and excludes 𝒫\mathcal{P}. Consider a germ 𝔣p\mathfrak{f}_{p} of ff in DD and define the nn-th associated curve Λn(𝔣𝔭,z)=𝔣𝔭(z)𝔣𝔭(z)𝔣𝔭(n)(z)\Lambda_{n}(\mathfrak{f_{p}},z)=\mathfrak{f_{p}}(z)\wedge\mathfrak{f_{p}}^{\prime}(z)\wedge\ldots\wedge\mathfrak{f_{p}}^{(n)}(z), mapping DD to Λn+1(n+1)\Lambda^{n+1}(\mathbb{C}^{n+1}). According to computations in [19], Λn(𝔣𝔭,z)\Lambda_{n}(\mathfrak{f_{p}},z) vanishes at {qn}{𝔭}\{q_{n}\}\cup\{\mathfrak{p}\} and thus identically on DD. This implies that Λn(f)\Lambda_{n}(f), and consequently ff, is degenerate on X𝒫X\setminus\mathcal{P}. This is a contradiction since both \vvΩ\vv{\Omega} and ff are non-degenerate.

  • (ii)

    Case 𝔭𝒫\mathfrak{p}\in\mathcal{P}: Using a similar setup with a disc chart (D,|z|<1)(D,|z|<1) around 𝔭\mathfrak{p} as in (1), we consider the germ 𝔣𝔭=[zb0H0(z)::zbnHn(z)]\mathfrak{f_{p}}=[z^{b_{0}}H_{0}(z):\ldots:z^{b_{n}}H_{n}(z)] where b0<<bnb_{0}<\ldots<b_{n} and each HjH_{j} is holomorphic in DD and non-vanishing at z=0z=0. By Lemma 4.1. in [28], we have

    Λn(𝔣𝔭,z)\displaystyle\Lambda_{n}(\mathfrak{f_{p}},z) =zi=0nbin(n+1)2Gn(b0,,bn;H0(z),,Hn(z);z)\displaystyle=z^{\sum_{i=0}^{n}b_{i}-\frac{n(n+1)}{2}}\cdot G_{n}\big{(}b_{0},\ldots,b_{n};H_{0}(z),\ldots,H_{n}(z);z\big{)}
    e0en,\displaystyle\quad\cdot e_{0}\wedge\ldots\wedge e_{n},

    where GnG_{n} is holomorphic in DD and non-vanishing at z=0z=0. The vanishing of Λn(𝔣𝔭,z)\Lambda_{n}(\mathfrak{f_{p}},z) at each qnq_{n} and thus identically in DD implies that GnG_{n} vanishes identically, another contradiction.

Now we reach the key proposition of this subsection.

Proposition 1.

On a compact Riemann surface XX, the concept of a character nn-ensemble, as outlined in Definition 4, is equivalent to that of a non-degenerate nn-tuple of one-forms, as described in Definition 5.

Proof.

By definition, a character nn-ensemble is inherently non-degenerate when considered as an nn-tuple of one-forms. This non-degeneracy is essential for defining character nn-ensembles. According to the second statement of Lemma 5, any non-degenerate nn-tuple of one-forms automatically constitutes a character nn-ensemble. Therefore, the properties required for an nn-tuple to be a character nn-ensemble are exactly those that prevent degeneracy among the one-forms in the tuple. ∎

In the analysis that follows, we categorize three specified nn-tuples of one-forms on a compact Riemann surface. We identify the nature of each tuple, highlighting whether it is degenerate, i.e., form a character nn-ensemble.

Example 3.

Let =X\mathfrak{C}=\mathfrak{C}_{X} denote the infinite-dimensional real linear space of meromorphic one-forms with simple poles and purely imaginary periods on a compact Riemann surface XX ([33, §\S15], [30, §\S8-1] and [14, §\SII.4-5]). Denote by =X{0}\mathfrak{C}^{*}=\mathfrak{C}_{X}\setminus\{0\} the non-zero elements of this space.

(1) It is possible to find two linear independent one-forms Ω1,Ω2\Omega_{1},\,\Omega_{2} in X\mathfrak{C}_{X}^{*} such that the pair \vvΩ=(Ω1,Ω2)\vv{\Omega}=(\Omega_{1},\Omega_{2}) is degenerate, meaning that \vvΩ\vv{\Omega} does not form a character 22-ensemble. An example of such forms on the Riemann sphere are Ω1=dzz\Omega_{1}=\frac{{\rm d}z}{z} and Ω2=dzz+1\Omega_{2}=\frac{{\rm d}z}{z+1}. By using (3.2) and setting 𝔭=1\mathfrak{p}=1, we can see that (Ω1,Ω2)(\Omega_{1},\Omega_{2}) generates a line in 2\mathbb{P}^{2}.

(2) Consider a one-form Ω\Omega\in\mathfrak{C}^{*} and nn nonzero real numbers λ1,,λn\lambda_{1},\ldots,\lambda_{n}. The tuple \vvΩ=(λ1Ω,,λnΩ)\vv{\Omega}=(\lambda_{1}\Omega,\ldots,\lambda_{n}\Omega) is non-degenerate if and only if all coefficients λj\lambda_{j} are mutually distinct. We can argue as follows. Define y=exp(0zΩ)y=\exp\left(\int_{0}^{z}\Omega\right). In a small disc chart (D,|z|<1)(D,|z|<1), where Ω\Omega is holomorphic and non-vanishing, we consider the nn-th associated curve Λn(f)\Lambda_{n}(f) of the curve

f(z)=[1:exp(0zΩ1)::exp(0zΩn)]=[1:y1λ::ynλ].f(z)=\left[1:\exp\left(\int_{0}^{z}\Omega_{1}\right):\ldots:\exp\left(\int_{0}^{z}\Omega_{n}\right)\right]=\left[1:y^{\lambda}_{1}:\ldots:y^{\lambda}_{n}\right].

This curve is non-degenerate in DD, thereby confirming that \vvΩ\vv{\Omega} is non-degenerate, as shown by the following computation:

(3.4) Λn(f)=fff(n)=ffynfyn(dydz)n(n+1)2=|1yλ1yλn0λ1yλ11λnyλn10λ1(λ11)(λ1n+1)yλ1nλn(λn1)(λnn+1)yλnn|(y)n(n+1)2e0e1en=i=1nλi1j<in(λiλj)yi=0nλin(n+1)2(y)n(n+1)2e0e1en.\begin{split}\Lambda_{n}(f)&=f\wedge f^{\prime}\wedge\ldots\wedge f^{(n)}\\ &=f\wedge\frac{\partial f}{\partial y}\wedge\ldots\wedge\frac{\partial^{n}f}{\partial y^{n}}\cdot\left(\frac{{\rm d}y}{{\rm d}z}\right)^{\frac{n(n+1)}{2}}\\ &=\begin{vmatrix}1&y^{\lambda_{1}}&\ldots&y^{\lambda_{n}}\\ 0&\lambda_{1}y^{\lambda_{1}-1}&\ldots&\lambda_{n}y^{\lambda_{n}-1}\\ \vdots&\vdots&\ddots&\vdots\\ 0&\lambda_{1}(\lambda_{1}-1)\ldots(\lambda_{1}-n+1)y^{\lambda_{1}-n}&\ldots&\lambda_{n}(\lambda_{n}-1)\ldots(\lambda_{n}-n+1)y^{\lambda_{n}-n}\end{vmatrix}\\ &\cdot(y^{\prime})^{\frac{n(n+1)}{2}}e_{0}\wedge e_{1}\wedge\ldots\wedge e_{n}\\ &=\prod_{i=1}^{n}\lambda_{i}\prod_{1\leq j<i\leq n}(\lambda_{i}-\lambda_{j})\cdot y^{\sum_{i=0}^{n}\lambda_{i}-\frac{n(n+1)}{2}}\cdot(y^{\prime})^{\frac{n(n+1)}{2}}e_{0}\wedge e_{1}\wedge\ldots\wedge e_{n}.\end{split}

(3) This example can be generalized as follows: Let Ω1,,Ωn\Omega_{1},\ldots,\Omega_{n} be nn distinct one-forms in \mathfrak{C}^{*}, and assume there is a point 𝔭\mathfrak{p} on XX where the residues of each Ωj\Omega_{j} at 𝔭\mathfrak{p} are distinct nonzero numbers. Under these conditions, \vvΩ=(Ω1,,Ωn)\vv{\Omega}=(\Omega_{1},\ldots,\Omega_{n}) forms a non-degenerate nn-tuple. The proof is similar to (2).

3.2. Proof of Theorem 2

The first statement of the theorem aligns with Lemma 5(1). We proceed to demonstrate the second statement as follows:

Consider a disc chart (D,|z|<1)(D,|z|<1) centered at a zero 𝔭\mathfrak{p} of \vvΩ\vv{\Omega}. Let (k1,,kn)>0n(k_{1},\ldots,k_{n})\in\mathbb{Z}_{>0}^{n} represent the vector of multiplicities of \vvΩ=(Ω1,,Ωn)\vv{\Omega}=(\Omega_{1},\ldots,\Omega_{n}) at 𝔭\mathfrak{p}. We choose a germ 𝔣𝔭\mathfrak{f_{p}} of fΩ,ρf_{\vec{\Omega},\vec{\rho}} in DD as follows:

[1:C1+zk1+1h1(z)::Cn+zkn+1hn(z)],\left[1:C_{1}+z^{k_{1}+1}h_{1}(z):\ldots:C_{n}+z^{k_{n}+1}h_{n}(z)\right],

where each hj(z)h_{j}(z) is a holomorphic function in DD that does not vanish at z=0z=0, and each CjC_{j} is a nonzero complex number. Given that each ki+1k_{i}+1 is an integer greater than 1, and using the calculations presented on pages 266-268 in [19], we determine that the ramification index of f=f0f=f_{0} at 𝔭\mathfrak{p} is min(k1,,kn)>0\min(k_{1},\ldots,k_{n})>0. Consequently, 𝔭\mathfrak{p} is confirmed as a ramification point of ff.

At last, we present an algorithm to identify the regular singularities of the curve f(z)=[1:exp(zΩ1)::exp(zΩn)]f(z)=\left[1:\exp\left(\int^{z}\Omega_{1}\right):\ldots:\exp\left(\int^{z}\Omega_{n}\right)\right] generated by a character nn-ensemble Ω\vec{\Omega}. The procedure is as follows:

  1. (1)

    Cover the manifold XX using a finite collection of disk charts (Di,|zi|<1)(D_{i},|z_{i}|<1) for i=1,,Ni=1,\dots,N. Ensure that each pole in the set 𝒫\mathcal{P} is contained within exactly one chart.

  2. (2)

    Within these charts, and excluding the poles in 𝒫\mathcal{P}, solve the equation Λn(f)=0\Lambda_{n}(f)=0 to identify 𝒵\mathcal{Z}, the set of all ramification points of ff not in 𝒫\mathcal{P}.

  3. (3)

    For each point 𝔭𝒫𝒵\mathfrak{p}\in\mathcal{P}\cup\mathcal{Z}, select the appropriate chart (D,|z|<1)(D,|z|<1) around it. Execute the process outlined in the proof of Lemma 5(1) to determine the quasi-canonical form of ff. This step allows us to ascertain all γ𝔭,\gamma_{\mathfrak{p},\cdot} indices.

4. Correspondence between character ensembles and toric solutions

In this section, we prove Theorem 1, which establishes a correspondence between character ensembles and toric solutions with cone singularities on a compact Riemann surface XX.

Proof.

The first statement of Theorem 1 follows from Proposition 1 and Lemma 5.

(2) Consider a toric solution \vvω\vv{\omega} with cone singularities at \vvD\vv{D} on a compact Riemann surface XX. Let f:X{P1,,Pk}nf:X\setminus\{P_{1},\ldots,P_{k}\}\to\mathbb{CP}^{n} be an associated curve derived from this solution. This curve is characterized as a totally unramified unitary curve, and its monodromy is constrained within a maximal torus of the group PSU(n+1)\mathrm{PSU}(n+1). To specifically align the monodromy of the curve within 𝕋n\mathbb{T}^{n}, a subgroup of the maximal torus, we select an appropriate element φPSU(n+1)\varphi\in\mathrm{PSU}(n+1). Applying φ\varphi, the transformed curve φf:X{P1,,Pk}n\varphi\circ f:X\setminus\{P_{1},\ldots,P_{k}\}\to\mathbb{CP}^{n} indeed has its monodromy contained within 𝕋n\mathbb{T}^{n}. We will proceed under the assumption that this simplification applies to the curve ff.

Choose a sufficiently small disc chart (D,|z|<1)(D,|z|<1) around 𝔭{P1,,Pk}\mathfrak{p}\in\{P_{1},\ldots,P_{k}\}. Using the argument in the proof of [28, Theorem 3.1.], the restriction of f=[f0::fn]f=[f_{0}:\ldots:f_{n}] to DD can be expressed as:

(4.1) f(z)=[zb0ϕ0(z)::zbnϕn(z)]f(z)=\left[z^{b_{0}}\phi_{0}(z):\ldots:z^{b_{n}}\phi_{n}(z)\right]

where bib_{i}\in\mathbb{R} and ϕi\phi_{i} is a holomorphic function that does not vanish on DD for all 1in1\leq i\leq n.

We define an nn-tuple Ω=(Ω1,,Ωn)\vec{\Omega}=(\Omega_{1},\ldots,\Omega_{n}) by:

(4.2) Ωk=d(logfkf0),k=1,,n.\Omega_{k}=\mathrm{d}\left(\log\frac{f_{k}}{f_{0}}\right),\quad k=1,\ldots,n.

By computation, we find:

(4.3) Ωk=((bkb0)1z+ϕk(z)ϕk(z)ϕ0(z)ϕ0(z))dz\Omega_{k}=\left((b_{k}-b_{0})\frac{1}{z}+\frac{\phi^{\prime}_{k}(z)}{\phi_{k}(z)}-\frac{\phi^{\prime}_{0}(z)}{\phi_{0}(z)}\right)\mathrm{d}z

Hence, each Ωj\Omega_{j} has at most simple poles. Additionally, it can be verified that:

(4.4) 2Ωk=d(log|fkf0|2),2\Re\,\Omega_{k}=\mathrm{d}\left(\log\left|\frac{f_{k}}{f_{0}}\right|^{2}\right),

i.e., the real part of each Ωk\Omega_{k} is exact outside of its poles. Consequently, each Ωk\Omega_{k} is a meromorphic one-form with simple poles and purely imaginary periods. Since Ω\vec{\Omega} is non-degenerate (as ff is non-degenerate), it is a character nn-ensemble on XX by Proposition 1.

By Theorem 1.5 in [5], the regular singularities of a toric curve to 1\mathbb{P}^{1}, which is derived from a character one-form, are confined to the union of the zeros and poles of the one-form. However, the following example demonstrates that a toric curve to n\mathbb{P}^{n}, generated from a character nn-ensemble for n>1n>1, can possess ramification points that do not coincide with the zeros or the poles of any ensemble component.

Example 4.

Consider the toric curve defined by f=[1:z:z2::zn1,zn+1]f=\left[1:z:z^{2}:\ldots:z^{n-1},z^{n+1}\right], where n>1n>1. z=0z=0 is a ramification point. The associated character nn-ensemble, \vvΩ=(Ω1,,Ωn)\vv{\Omega}=(\Omega_{1},\ldots,\Omega_{n}), is given by:

Ωk=kzdz,for k=1,2,,n1;Ωn=n+1zdz\Omega_{k}=\frac{k}{z}\,{\rm d}z,\quad\text{for }k=1,2,\ldots,n-1;\quad\Omega_{n}=\frac{n+1}{z}\,{\rm d}z

Here, z=0z=0 is a common pole of all components Ωj\Omega_{j}. However, applying a non-degenerate linear transformation to ff results in:

f~=[1:1+z:1+z+z2::1+z+zn1:1+z+zn+1],\tilde{f}=\left[1:1+z:1+z+z^{2}:\ldots:1+z+z^{n-1}:1+z+z^{n+1}\right],

with the modified character nn-ensemble \vvΩ~=(Ω~1,,Ω~n)\vv{\widetilde{\Omega}}=(\widetilde{\Omega}_{1},\ldots,\widetilde{\Omega}_{n}) represented as:

Ω~1=dz1+z,Ω~2=1+2z1+z+z2dz,,Ω~n1=1+(n1)zn21+z+zn1dz,Ω~n=1+(n+1)zn1+z+zn+1dz\begin{split}&\widetilde{\Omega}_{1}=\frac{{\rm d}z}{1+z},\;\widetilde{\Omega}_{2}=\frac{1+2z}{1+z+z^{2}}\,{\rm d}z,\ldots,\\ &\widetilde{\Omega}_{n-1}=\frac{1+(n-1)z^{n-2}}{1+z+z^{n-1}}\,{\rm d}z,\;\widetilde{\Omega}_{n}=\frac{1+(n+1)z^{n}}{1+z+z^{n+1}}\,{\rm d}z\end{split}

Although z=0z=0 remains a ramification point for f~\tilde{f}, it is notably no longer a zero or pole of any new one-forms.

5. Two examples

In this section, we explore the practical application of the correspondence between character ensembles and toric solutions, as discussed in the previous section. We conduct a thorough analysis of Examples 1 and 2. These cases introduce innovative toric solutions to the SU(n+1){\rm SU}(n+1) Toda system with cone singularities. Notably, these examples expand upon the recent findings detailed in [25, Theorems 1.8-9], extending the known boundaries of this research area.

5.1. Example 1

This subsection details Example 1. Utilizing Example 3 (2) and Proposition 1, we can see that Ω\Omega constitutes a character nn-ensemble on XX. This ensemble generates a family of toric curves parametrized by (>0)n(\mathbb{R}_{>0})^{n} as per Equation (1.8), all sharing the same regular singularities, denoted by \vvD\vv{D}. According to Theorem 4, these curves are associated with a family of toric solutions that exhibit the same cone singularities \vvD\vv{D}. Notations from Example 3 (2) are used herein.

By Equation (3.4), the singularities of the curve

f(z)=[1:exp(0zλ1Ω)::exp(0zλnΩ)]=[1:yλ1,,yλn]f(z)=\left[1:\exp\left(\int_{0}^{z}\lambda_{1}\Omega\right):\ldots:\exp\left(\int_{0}^{z}\lambda_{n}\Omega\right)\right]=\left[1:y^{\lambda_{1}},\ldots,y^{\lambda_{n}}\right]

are limited to the set of zeros and poles of Ω\Omega. Considering a point 𝔭\mathfrak{p} within this set, we categorize the analysis into the following two cases:

  1. (1)

    𝔭\mathfrak{p} is a zero of Ω\Omega with order kk: We establish that 𝔭\mathfrak{p} is a ramification point of the curve ff with all ramification indices equal to kk. Assuming without loss of generality that y(𝔭)=1y(\mathfrak{p})=1, we express y(z)=1+zk+1ϕ(z)y(z)=1+z^{k+1}\phi(z) in a small disc chart (D,|z|<1)(D,|z|<1) around 𝔭\mathfrak{p}, where ϕ(z)\phi(z) is holomorphic and non-vanishing in DD.

    Expanding each yλiy^{\lambda_{i}} around 𝔭\mathfrak{p} into a power series, we obtain:

    yλi=1+λizk+1φ(z)+λi(λi1)2z2k+2φ(z)2+,i=1,2,,n.y^{\lambda_{i}}=1+\lambda_{i}z^{k+1}\varphi(z)+\frac{\lambda_{i}(\lambda_{i}-1)}{2}z^{2k+2}\varphi(z)^{2}+\ldots,\quad i=1,2,\ldots,n.

    Consequently, ff can be expressed as:

    f=(1,yλ1,yλ2,,yλn)=(1,zk+1φ(z),z2k+2φ(z)2,)(1110λ1λn0λ1(λ11)2λn(λn1)2)\begin{split}f&=(1,y^{\lambda_{1}},y^{\lambda_{2}},\ldots,y^{\lambda_{n}})\\ &=\left(1,z^{k+1}\varphi(z),z^{2k+2}\varphi(z)^{2},\ldots\right)\cdot\begin{pmatrix}1&1&\ldots&1\\ 0&\lambda_{1}&\ldots&\lambda_{n}\\ 0&\frac{\lambda_{1}(\lambda_{1}-1)}{2}&\ldots&\frac{\lambda_{n}(\lambda_{n}-1)}{2}\\ \vdots&\vdots&\ddots&\vdots\end{pmatrix}\end{split}

    Right-multiplying ff by an (n+1)×(n+1)(n+1)\times(n+1) invertible matrix corresponds to applying column transformations to the ×(n+1)\infty\times(n+1) infinite matrix on the right-hand side of (1). The first n+1n+1 rows of this matrix:

    (1110λ1λn0λ1(λ11)2λn(λn1)20λ1(λ11)(λ1n+1)n!λn(λn1)(λnn+1)n!)\begin{pmatrix}1&1&\ldots&1\\ 0&\lambda_{1}&\ldots&\lambda_{n}\\ 0&\frac{\lambda_{1}(\lambda_{1}-1)}{2}&\ldots&\frac{\lambda_{n}(\lambda_{n}-1)}{2}\\ \vdots&\vdots&\ddots&\vdots\\ 0&\frac{\lambda_{1}(\lambda_{1}-1)\ldots(\lambda_{1}-n+1)}{n!}&\ldots&\frac{\lambda_{n}(\lambda_{n}-1)\ldots(\lambda_{n}-n+1)}{n!}\end{pmatrix}

    forms an (n+1)×(n+1)(n+1)\times(n+1) invertible matrix, with a determinant given by:

    i=1ni!i=1nλi1j<in(λiλj)0.\prod_{i=1}^{n}i!\prod_{i=1}^{n}\lambda_{i}\prod_{1\leq j<i\leq n}(\lambda_{i}-\lambda_{j})\neq 0.

    Thus, it can be transformed into a lower triangular matrix through a finite number of column transformations, such as the Gaussian elimination process. Correspondingly, ff has been transformed into the quasi-canonical form:

    (1,zk+1φ(z)+,z2k+2φ(z)2+,)\left(1,z^{k+1}\varphi(z)+\ldots,z^{2k+2}\varphi(z)^{2}+\ldots,\ldots\ldots\right)

    under finitely many transforms in PGL(n+1,){\rm PGL}(n+1,\mathbb{C}). By the definition of ramification indices [19, pp. 266-268] for a holomorphic curve, applying such transformations to the curve does not alter its ramification indices. Therefore, all the nn ramification indices of ff at 𝔭\mathfrak{p} coincide with those of the curve (1), each equal to kk.

  2. (2)

    𝔭\mathfrak{p} is a pole of Ω\Omega: Assume Res𝔭(Ω)=a{0}{\rm Res}_{\mathfrak{p}}(\Omega)=a\in\mathbb{R}\setminus\{0\}. In a small disk chart (D,|z|<1)(D,|z|<1) around point 𝔭\mathfrak{p}, the function y=exp(0zΩ)y=\exp\left(\int_{0}^{z}\Omega\right) is represented as zaϕ(z)z^{a}\phi(z), where ϕ(z)\phi(z) is holomorphic and nonvanishing within DD. Accordingly, we define f(z)=[1:zλ1aϕ(z)λ1::zλnaϕ(z)λn]f(z)=\left[1:z^{\lambda_{1}a}\phi(z)^{\lambda_{1}}:\ldots:z^{\lambda_{n}a}\phi(z)^{\lambda_{n}}\right], representing a totally unramified unitary curve in DD^{*} with a potential regular singularity at z=0z=0 by Equation (3.4).

    By Theorem 4, ff induces a solution \vvω\vv{\omega} to the SU(n+1){\rm SU}(n+1) Toda system on DD, characterized by a possible cone singularity at z=0z=0. Theorem 4 assures that the singularities of both types at z=0z=0 are equivalent. Given the distinct real numbers 0,λ1a,λ2a,,λna0,\lambda_{1}a,\lambda_{2}a,\ldots,\lambda_{n}a, we reorder them in ascending order to achieve μ0<μ1<<μn\mu_{0}<\mu_{1}<\ldots<\mu_{n}. Concurrently, the components of ff are rearranged in ascending powers of zz to attain the quasi-canonical form:

    [zμ0ϕ(z)μ0a:zμ1ϕ(z)μ1a::zμnϕ(z)μna].\left[z^{\mu_{0}}\phi(z)^{\frac{\mu_{0}}{a}}:z^{\mu_{1}}\phi(z)^{\frac{\mu_{1}}{a}}:\ldots:z^{\mu_{n}}\phi(z)^{\frac{\mu_{n}}{a}}\right].

    The solution \vvω\vv{\omega}, and consequently its cone singularity at z=0z=0, remain unaltered, as this constitutes merely a transformation within PSU(n+1){\rm PSU}(n+1). It is established that γ𝔭,i=μiμi11\gamma_{\mathfrak{p},i}=\mu_{i}-\mu_{i-1}-1 for all 1in1\leq i\leq n. Depending on the specific λ\lambda values, the point 𝔭\mathfrak{p} may be classified as a branch point, a ramification point, or a totally unramified point of ff.

5.2. Example 1 provides new solutions with cone singularities on a compact Riemann surface

The curve in this example corresponds to a novel class of solutions to the SU(n+1){\rm SU}(n+1) Toda system with cone singularities on a compact Riemann surface. Lin-Yang-Zhong in [25, Theorems 1.8-9] presented various sufficient conditions for the existence of such solutions on a compact Riemann surface XX of genus gX>0g_{X}>0. Specifically, by transforming Equation (1.4), it can be reformulated into the equivalent form:

Δgu~i+j=1naijeu~jK0=4πk=1mγk,iδpki=1,2,,n\Delta_{g}\tilde{u}_{i}+\sum_{j=1}^{n}a_{ij}e^{\tilde{u}_{j}}-K_{0}=4\pi\sum_{k=1}^{m}\gamma_{k,i}\delta_{p_{k}}\quad i=1,2,\ldots,n

where K0K_{0} is the Gaussian curvature function of the background conformal metric gg on XX and Δg\Delta_{g} its Laplacian. The function u~i\tilde{u}_{i} represents a global transformation of uiu_{i} from Equation (1.1), adjusted by the background metric in each coordinate chart. The singularity coefficients {γi,j}\{\gamma_{i,j}\} remain consistent with those described in (1.4). Lin-Yang-Zhong defined ρi\rho_{i} for all 1in1\leq i\leq n as follows:

ρi:=4πk=1mj=1naijγk,j+j=1naijXK0\rho_{i}:=4\pi\sum_{k=1}^{m}\sum_{j=1}^{n}a^{ij}\gamma_{k,j}+\sum_{j=1}^{n}a^{ij}\int_{X}K_{0}

Here, (aij)=(j(n+1i)n+1)(a^{ij})=\left(\frac{j(n+1-i)}{n+1}\right) represents the inverse of the Cartan matrix (aij)(a_{ij}) for 𝔰𝔲(n+1)\mathfrak{su}(n+1). According to Theorem 1.9 in [25], if ρiΓi\rho_{i}\notin\Gamma_{i} for any i=1,2,ni=1,2,\ldots n, a solution to the Toda system is feasible. The definition of Γi\Gamma_{i} (1in)(1\leq i\leq n) is detailed in [25, pp. 340-341], noting that each includes 4π4\pi\mathbb{N}.

In this example, Ω\Omega is a meromorphic one-form on a compact Riemann surface with positive genus, featuring exactly two simple poles (denoted p1p_{1} and p2p_{2}). Assuming 0<λ1<<λn0<\lambda_{1}<\ldots<\lambda_{n}, with Resp1(ω)=a0{\rm Res}_{p_{1}}(\omega)=a\neq 0, the corresponding γ1,j=(λjλj1)a1\gamma_{1,j}=(\lambda_{j}-\lambda_{j-1})a-1 for j=1,2,,nj=1,2,\ldots,n (assuming λ0=0\lambda_{0}=0 for consistency). According to the residue theorem, Resp2(ω)=a{\rm Res}_{p_{2}}(\omega)=-a, and as previously noted, γ2,j=(λnj+1λnj)a1\gamma_{2,j}=(\lambda_{n-j+1}-\lambda_{n-j})a-1. Additionally, the other m2m-2 singular points p3,,pmp_{3},\ldots,p_{m} are identified as ramification points, with γi,jki\gamma_{i,j}\equiv k_{i} for i=3,,mi=3,\ldots,m and j=1,2,,nj=1,2,\ldots,n, where kik_{i} denotes the order of zero of Ω\Omega at pip_{i}. After calculations, we derive:

ρi=4πk=1mj=1naijγk,j+j=1naijXK0=4π(j=1nj(n+1i)n+1((λjλj1)a1)+j=1nj(n+1i)n+1((λnj+1λnj)a1))+4πk=3mj=1nj(n+1i)n+1ki+j=1nj(n+1i)n+12π(22gX)=4π(n+1i)aλn+4π(k3++kn1gX)n(n+1i)\begin{split}\rho_{i}=&4\pi\sum_{k=1}^{m}\sum_{j=1}^{n}a^{ij}\gamma_{k,j}+\sum_{j=1}^{n}a^{ij}\int_{X}K_{0}\\ =&4\pi\Big{(}\sum_{j=1}^{n}\frac{j(n+1-i)}{n+1}((\lambda_{j}-\lambda_{j-1})a-1)+\sum_{j=1}^{n}\frac{j(n+1-i)}{n+1}((\lambda_{n-j+1}-\lambda_{n-j})a-1)\Big{)}\\ &+4\pi\sum_{k=3}^{m}\sum_{j=1}^{n}\frac{j(n+1-i)}{n+1}k_{i}+\sum_{j=1}^{n}\frac{j(n+1-i)}{n+1}\cdot 2\pi(2-2g_{X})\\ =&4\pi(n+1-i)a\lambda_{n}+4\pi(k_{3}+\cdots+k_{n}-1-g_{X})\cdot n(n+1-i)\end{split}

Choosing aλna\lambda_{n} to be a positive integer, by this calculation, we can see that all ρi\rho_{i}’s lie within 4πΓi4\pi\mathbb{N}\subset\Gamma_{i} for all 1in1\leq i\leq n, violating the criteria of Theorem 1.9 in [25]. Furthermore, Theorem 1.8 from Lin-Yang-Zhong also outlines sufficient conditions for the existence of solutions to the SU(n+1){\rm SU}(n+1) Toda system, requiring all {γi,j}\{\gamma_{i,j}\} to be non-negative integers. By choosing appropriate λ1,,λn\lambda_{1},\ldots,\lambda_{n} such that p1p_{1} and p2p_{2} act as branch points, these conditions are not met. Both theorems stipulate that the genus of XX must be positive.

In summary, the solutions of the Toda system associated with the curves in Example 1 on a general compact Riemann surface and Example 2 on the Riemann sphere introduce new possibilities beyond the scope outlined in [25].

5.3. Example 2

We establish the existence of the necessary curves and provide an approximate enumeration of all possible singularity data at \infty for these curves in this example.

Existence. The following lemma is useful for establishing existence.

Lemma 6.

Let α>0\alpha>0. For any polynomial P(z)P(z), there exists a unique polynomial Q(z)Q(z), having the same degree as P(z)P(z), such that

(5.1) ddz(zαQ(z))=zα1P(z).\frac{{\rm d}}{{\rm d}z}\Big{(}z^{\alpha}Q(z)\Big{)}=z^{\alpha-1}P(z).
Proof.

First, consider the monomial Q(z)=zjα+jQ(z)=\frac{z^{j}}{\alpha+j} and observe the derivative

ddz(zαQ(z))=zα1(αQ(z)+zQ(z))=zα1zj.\frac{{\rm d}}{{\rm d}z}\Big{(}z^{\alpha}Q(z)\Big{)}=z^{\alpha-1}\big{(}\alpha Q(z)+zQ^{\prime}(z)\big{)}=z^{\alpha-1}\cdot z^{j}.

This calculation allows us to solve the linear ODE (5.1) for any polynomial P(z)P(z). The uniqueness of the solution is straightforward to establish. ∎

Proposition 2.

We adopt the notations from Example 2 and assume 0=β0<β1<<βn0=\beta_{0}<\beta_{1}<\ldots<\beta_{n} with the increments βiβi1=1+γi\beta_{i}-\beta_{i-1}=1+\gamma_{i} for all 1in1\leq i\leq n. Consequently, there exist nn polynomials φ1(z),φ2(z),,φn(z)\varphi_{1}(z),\varphi_{2}(z),\ldots,\varphi_{n}(z), each non-vanishing at z=0z=0, such that the family of curves

f\vvρ(z)=[1:ρ1zβ1φ1(z)::ρnzβnφn(z)],\vvρ=(ρ1,,ρn)(>0)n,f_{\vv{\rho}}(z)=\left[1:\rho_{1}z^{\beta_{1}}\varphi_{1}(z):\ldots:\rho_{n}z^{\beta_{n}}\varphi_{n}(z)\right],\quad\vv{\rho}=(\rho_{1},\ldots,\rho_{n})\in\left(\mathbb{R}_{>0}\right)^{n},

meets the requirements specified in Example 2.

Proof.

We shall construct these nn polynomials φ1(z),φ2(z),,φn(z)\varphi_{1}(z),\varphi_{2}(z),\ldots,\varphi_{n}(z) by using the iteration argument.

Step 1. Take

ψ1(z):=i=1m(zzi)γi,1,ψk(z):=i=1m(zzi)γi,1φk(1)(z),2kn,\begin{split}&\psi_{1}(z):=\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}},\\ &\psi_{k}(z):=\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}}\cdot\varphi_{k}^{(1)}(z),\quad 2\leq k\leq n,\end{split}

where φ2(1)(z),,φn(1)(z)\varphi_{2}^{(1)}(z),\ldots,\varphi_{n}^{(1)}(z) are (n1)(n-1) polynomials that will be chosen later. Given that 0<β1<<βn0<\beta_{1}<\ldots<\beta_{n}, Lemma 6 ensures the existence of nn polynomials φ1(z),,φn(z)\varphi_{1}(z),\ldots,\varphi_{n}(z) such that the derivative of f^(z):=(1,zβ1φ1(z),zβ2φ2(z),,zβnφn(z))\hat{f}(z):=\left(1,z^{\beta_{1}}\varphi_{1}(z),z^{\beta_{2}}\varphi_{2}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z)\right) is given by

(5.2) df^(z)dz=(0,zβ11ψ1(z),,zβn1ψn(z))=:zβ11i=1m(zzi)γi,1\vvg1(z),\frac{{\rm d}\hat{f}(z)}{{\rm d}z}=\left(0,z^{\beta_{1}-1}\psi_{1}(z),\ldots,z^{\beta_{n}-1}\psi_{n}(z)\right)=:z^{\beta_{1}-1}\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}}\vv{g_{1}}(z),

where \vvg1(z)=(0,1,zβ2β1φ2(1)(z),,zβnβ1φn(1)(z))\vv{g_{1}}(z)=\left(0,1,z^{\beta_{2}-\beta_{1}}\varphi_{2}^{(1)}(z),\ldots,z^{\beta_{n}-\beta_{1}}\varphi_{n}^{(1)}(z)\right). Moreover, we could determine the polynomial φ1(z)\varphi_{1}(z) of degree =1mγ,1\sum_{\ell=1}^{m}\,\gamma_{\ell,1}, which does not vanish at z=0z=0.

Step 2. Differentiating df^(z)dz\frac{{\rm d}\hat{f}(z)}{{\rm d}z} again, we obtain

(5.3) d2f^(z)dz2=ddz(zβ11i=1m(zzi)γi,1)\vvg1(z)+zβ11i=1m(zzi)γi,1d\vvg1(z)dz.\frac{{\rm d}^{2}\hat{f}(z)}{{\rm d}z^{2}}=\frac{{\rm d}}{{\rm d}z}\left(z^{\beta_{1}-1}\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}}\right)\cdot\vv{g_{1}}(z)+z^{\beta_{1}-1}\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}}\cdot\frac{{\rm d}\vv{g_{1}}(z)}{{\rm d}z}.

For any (n2)(n-2) polynomials φ3(2)(z),,φn(2)(z)\varphi_{3}^{(2)}(z),\ldots,\varphi_{n}^{(2)}(z) that will be chosen later, Lemma 6 ensures there exist (n1)(n-1) polynomials φ2(1)(z),,φn(1)(z)\varphi_{2}^{(1)}(z),\ldots,\varphi_{n}^{(1)}(z) such that

(5.4) d\vvg1dz=ddz(0,1,zβ2β1φ2(1)(z),,zβnβ1φn(1)(z))=zβ2β11i=1m(zzi)γi,2(0,0,1,zβ3β2φ3(2)(z),,zβnβ2φn(2)(z)):=zβ2β11i=1m(zzi)γi,2\vvg2(z),\begin{split}\frac{{\rm d}\vv{g_{1}}}{{\rm d}z}&=\frac{\rm d}{{\rm d}z}\left(0,1,z^{\beta_{2}-\beta_{1}}\varphi_{2}^{(1)}(z),\ldots,z^{\beta_{n}-\beta_{1}}\varphi_{n}^{(1)}(z)\right)\\ &=z^{\beta_{2}-\beta_{1}-1}\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,2}}\cdot\Big{(}0,0,1,z^{\beta_{3}-\beta_{2}}\varphi_{3}^{(2)}(z),\ldots,z^{\beta_{n}-\beta_{2}}\varphi_{n}^{(2)}(z)\Big{)}\\ &:=z^{\beta_{2}-\beta_{1}-1}\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,2}}\cdot\vv{g_{2}}(z),\end{split}

where \vvg2(z)=(0,0,1,zβ3β2φ3(2)(z),,zβnβ2φn(2)(z))\vv{g_{2}}(z)=\left(0,0,1,z^{\beta_{3}-\beta_{2}}\varphi_{3}^{(2)}(z),\ldots,z^{\beta_{n}-\beta_{2}}\varphi_{n}^{(2)}(z)\right). Moreover, by Equations (5.4) and (5.2), we could determine these three polynomials φ2(1)(z)\varphi_{2}^{(1)}(z), ψ2(z)\psi_{2}(z) and φ2(z)\varphi_{2}(z) such that

degφ2==1m(γ,1+γ,2)\deg\,\varphi_{2}=\sum_{\ell=1}^{m}\left(\gamma_{\ell,1}+\gamma_{\ell,2}\right)

and all of them do not vanish at z=0z=0. Substituting this equation into (5.3) yields

(5.5) d2f^(z)dz2=(zβ11i=1m(zzi)γi,1)\vvg1(z)+zβ22i=1m(zzi)γi,1+γi,2\vvg2(z).\frac{{\rm d}^{2}\hat{f}(z)}{{\rm d}z^{2}}=\left(z^{\beta_{1}-1}\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}}\right)^{\prime}\cdot\vv{g_{1}}(z)+z^{\beta_{2}-2}\prod_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}+\gamma_{i,2}}\cdot\vv{g_{2}}(z).

Repeating the above process, we continue to differentiate d2f^(z)dz2\frac{{\rm d}^{2}\hat{f}(z)}{{\rm d}z^{2}} and \vvg2\vv{g_{2}} and obtain

g3=(0,0,0,1,zβ4β3φ4(3)(z),,zβnβ3φn(3)(z))\displaystyle\vec{g}_{3}=\Big{(}0,0,0,1,z^{\beta_{4}-\beta_{3}}\varphi_{4}^{(3)}(z),\ldots,z^{\beta_{n}-\beta_{3}}\varphi_{n}^{(3)}(z)\Big{)}
\displaystyle\vdots
gn(0,0,,0,1)\displaystyle\vec{g}_{n}\equiv\Big{(}0,0,\ldots,0,1\Big{)}

In summary, for each increment in the subscript ii of \vvgi\vv{g_{i}}, the vector-valued function \vvgi\vv{g_{i}} exhibits one additional vanishing component in its earlier entries. After carrying out this procedure nn times, we find all the nn polynomials φ1(z),,φn(z)\varphi_{1}(z),\ldots,\varphi_{n}(z) such that

(5.6) degφk=j=1k=1mγ,j,1kn,\deg\,\varphi_{k}=\sum_{j=1}^{k}\sum_{\ell=1}^{m}\,\gamma_{\ell,j},\quad 1\leq k\leq n,

and all of them do not vanish at z=0z=0.

Step 4. In the final step, we show that the curve

f(z):=[f^(z)]=[1:zβ1φ1(z)::zβnφn(z)]f(z):=[\hat{f}(z)]=\left[1:z^{\beta_{1}}\varphi_{1}(z):\ldots:z^{\beta_{n}}\varphi_{n}(z)\right]

satisfies the requirements specified in Example 2. Since f^(z)\hat{f}(z) itself is of quasi-canonical form near z=0z=0, it has the desired singularity information there, i.e., γ[0],j=γj=βjβj11\gamma_{[0],j}=\gamma_{j}=\beta_{j}-\beta_{j-1}-1 for all 1jn1\leq j\leq n. By the induction argument, we obtain

(5.7) f^(k)(z)zβkki=1m(zzi)γi,1++γi,k\vvgk(z)=a0(k)f^(z)+a1(k)f^(z)++ak1(k)f^(k1)(z),1kn,\begin{split}&\hat{f}^{(k)}(z)-z^{\beta_{k}-k}\prod\limits_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}+\ldots+\gamma_{i,k}}\cdot\vv{g_{k}}(z)\\ &=a_{0}^{(k)}\hat{f}(z)+a_{1}^{(k)}\hat{f}^{\prime}(z)+\ldots+a_{k-1}^{(k)}\hat{f}^{(k-1)}(z),\quad 1\leq k\leq n,\end{split}

where all ai(j)a_{i}^{(j)} are multi-valued holomorphic functions. Then we obtain the following equation for each 1kn1\leq k\leq n,

(5.8) Λk(f^)=f^f^f^(k)=f^zβ11i=1m(zzi)γi,1\vvg1zβkki=1m(zzi)γi,1++γi,k\vvgk=zi=1kβik(k+1)2i=1m(zzi)kγi,1+(k1)γi,2++γi,kf^\vvg1\vvgk=zi=1kβik(k+1)2i=1m(zzi)kγi,1+(k1)γi,2++γi,k(1zβ1φ1(z)zβ2φ2(z)zβnφn(z))T(01zβ2β1φ2(1)(z)zβnβ1φn(1)(z))T(001zβk+1βkφk+1(k)(z)zβnβkφn(k)(z))T.\begin{split}\Lambda_{k}(\hat{f})&=\hat{f}\wedge\hat{f}^{\prime}\wedge\ldots\wedge\hat{f}^{(k)}\\ &=\hat{f}\wedge z^{\beta_{1}-1}\prod\limits_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}}\cdot\vv{g_{1}}\wedge\ldots\wedge z^{\beta_{k}-k}\prod\limits_{i=1}^{m}(z-z_{i})^{\gamma_{i,1}+\ldots+\gamma_{i,k}}\cdot\vv{g_{k}}\\ &=z^{\sum\limits_{i=1}^{k}\beta_{i}-\frac{k(k+1)}{2}}\prod\limits_{i=1}^{m}(z-z_{i})^{k\gamma_{i,1}+(k-1)\gamma_{i,2}+\ldots+\gamma_{i,k}}\hat{f}\wedge\vv{g_{1}}\wedge\ldots\wedge\vv{g_{k}}\\ &=z^{\sum\limits_{i=1}^{k}\beta_{i}-\frac{k(k+1)}{2}}\prod\limits_{i=1}^{m}(z-z_{i})^{k\gamma_{i,1}+(k-1)\gamma_{i,2}+\ldots+\gamma_{i,k}}\cdot\\ &\begin{pmatrix}1\\ z^{\beta_{1}}\varphi_{1}(z)\\ z^{\beta_{2}}\varphi_{2}(z)\\ \vdots\\ z^{\beta_{n}}\varphi_{n}(z)\end{pmatrix}^{\rm T}\wedge\begin{pmatrix}0\\ 1\\ z^{\beta_{2}-\beta_{1}}\varphi_{2}^{(1)}(z)\\ \vdots\\ z^{\beta_{n}-\beta_{1}}\varphi_{n}^{(1)}(z)\end{pmatrix}^{\rm T}\wedge\ldots\wedge\begin{pmatrix}0\\ \vdots\\ 0\\ 1\\ z^{\beta_{k+1}-\beta_{k}}\varphi_{k+1}^{(k)}(z)\\ \vdots\\ z^{\beta_{n}-\beta_{k}}\varphi_{n}^{(k)}(z)\end{pmatrix}^{\rm T}.\end{split}

Notably, the first term of f^\vvg1\vvgk\hat{f}\wedge\vv{g_{1}}\wedge\ldots\wedge\vv{g_{k}} equals e0eke_{0}\wedge\ldots\wedge e_{k}. From this relationship, Λn(f^)\Lambda_{n}(\hat{f}) remains non-zero outside the set {0,,z1,,zm}\{0,\infty,z_{1},\ldots,z_{m}\}, where the curve ff is totally unramified. Moreover, for the non-degenerate unitary curve f:{0}nf:\mathbb{C}\setminus\{0\}\to\mathbb{P}^{n}, Λn(f^)\Lambda_{n}(\hat{f}) vanishes at z1,,zmz_{1},\ldots,z_{m}, which constitute all the ramification points of ff.

For each 1m1\leq\ell\leq m, by applying a linear transformation in GL(n+1,)\mathrm{GL}(n+1,\mathbb{C}) to f^\hat{f}, we can achieve the quasi-canonical form of f^\hat{f} near its ramification point zkz_{k}. This form is given by

f^(z)=(1,(zz)γ[z],1+1g1(z),(zz)γ[z],1+γ[z],2+2g2(z),\displaystyle\hat{f}(z)=\biggl{(}1,(z-z_{\ell})^{\gamma_{[z_{\ell}],1}+1}g_{1}(z),(z-z_{\ell})^{\gamma_{[z_{\ell}],1}+\gamma_{[z_{\ell}],2}+2}g_{2}(z),
,(zz)γ[z],1++γ[z],n+ngn(z)),\displaystyle\ldots,(z-z_{\ell})^{\gamma_{[z_{\ell}],1}+\ldots+\gamma_{[z_{\ell}],n}+n}g_{n}(z)\biggr{)},

where g1,,gng_{1},\ldots,g_{n} are multi-valued holomorphic functions that do not vanish at zz_{\ell}. By straightforward computation, for each 1kn1\leq k\leq n, the kk-th associated curve Λk(f^)\Lambda_{k}(\hat{f}) is expressed as

(5.9) Λk(f^)=i=1m(zz)kγ[z],1+(k1)γ[z],2++γ[z],k\vvλk(z)\Lambda_{k}(\hat{f})=\prod_{i=1}^{m}(z-z_{\ell})^{k\gamma_{[z_{\ell}],1}+(k-1)\gamma_{[z_{\ell}],2}+\ldots+\gamma_{[z_{\ell}],k}}\cdot\vv{\lambda_{k}}(z)

where \vvλk(z)\vv{\lambda_{k}}(z) is a multi-valued holomorphic curve valued in Λk+1(n+1)\Lambda^{k+1}(\mathbb{C}^{n+1}) and does not vanish at zz_{\ell}. By using this equation and [28, Lemma 4.1.], we obtain

(5.10) Λn(f^)=Constzi=1nβin(n+1)2=1m(zz)nγ[z],1+(n1)γ[z],2++γ[z],ne0en.\Lambda_{n}(\hat{f})={\rm Const}\cdot z^{\sum_{i=1}^{n}\,\beta_{i}-\frac{n(n+1)}{2}}\prod_{\ell=1}^{m}(z-z_{\ell})^{n\gamma_{[z_{\ell}],1}+(n-1)\gamma_{[z_{\ell}],2}+\ldots+\gamma_{[z_{\ell}],n}}\cdot e_{0}\wedge\ldots\wedge e_{n}.

Comparing this with Equation (5.8) yields the following system of linear equations:

kγ[z],1+(k1)γ[z],2++γ[z],k=kγ,1+(k1)γ,2++γ,k,1kn.k\gamma_{[z_{\ell}],1}+(k-1)\gamma_{[z_{\ell}],2}+\ldots+\gamma_{[z_{\ell}],k}=k\gamma_{\ell,1}+(k-1)\gamma_{\ell,2}+\ldots+\gamma_{\ell,k},\quad 1\leq k\leq n.

Consequently, we find that γ[z],k=γ,k\gamma_{[z_{\ell}],k}=\gamma_{\ell,k} for all 1m1\leq\ell\leq m and 1kn1\leq k\leq n.

Remark 1.

The singularity information at \infty of the curve

f^(z)=(zβ0φ0(z),zβ1φ1(z),,zβnφn(z))withβ0=1andφ0(z)1,\hat{f}(z)=\left(z^{\beta_{0}}\varphi_{0}(z),z^{\beta_{1}}\varphi_{1}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z)\right)\quad\text{with}\quad\beta_{0}=1\quad\text{and}\quad\varphi_{0}(z)\equiv 1,

constructed in the proof is given by

(γ[],i)i=1n=(βnβn11+i=1mγi,n,,β2β11+i=1mγi,2,β11+i=1mγi,1).(\gamma_{[\infty],i})_{i=1}^{n}=\left(\beta_{n}-\beta_{n-1}-1+\sum_{i=1}^{m}\gamma_{i,n},\ldots,\beta_{2}-\beta_{1}-1+\sum_{i=1}^{m}\gamma_{i,2},\beta_{1}-1+\sum_{i=1}^{m}\gamma_{i,1}\right).

Indeed, substituting z=1wz=\frac{1}{w} into f^(z)\hat{f}(z), we can obtain its local expression near \infty as

(5.11) f^(w)=(wβ0degφ0Φ0(w),wβ1degφ1Φ1(w),,wβndegφnΦn(w)),\hat{f}(w)=\left(w^{-\beta_{0}-\deg\varphi_{0}}\Phi_{0}(w),w^{-\beta_{1}-\deg\varphi_{1}}\Phi_{1}(w),\ldots,w^{-\beta_{n}-\deg\varphi_{n}}\Phi_{n}(w)\right),

where Φ0(w),,Φn(w)\Phi_{0}(w),\ldots,\Phi_{n}(w) are polynomials non-vanishing at w=0w=0. Since β0<<βn\beta_{0}<\ldots<\beta_{n} and degφ0<<degφn\deg\varphi_{0}<\ldots<\deg\varphi_{n}, we obtain the quasi-canonical form of f^(w)\hat{f}(w) near w=0w=0 as

(wβndegφnΦn(w),,wβ1degφ1Φ1(w),,wβ0degφ0Φ0(w)).\left(w^{-\beta_{n}-\deg\varphi_{n}}\Phi_{n}(w),\ldots,w^{-\beta_{1}-\deg\varphi_{1}}\Phi_{1}(w),\ldots,w^{-\beta_{0}-\deg\varphi_{0}}\Phi_{0}(w)\right).

The statement follows from Equation (5.6).

Counting. We shall enumerate the possibilities of (γ[],i)i=1(\gamma_{[\infty],i})_{i=1}^{\infty} for the curves in Example 2. To this end, we need the following lemma.

Lemma 7.

Consider β0<<βn\beta_{0}<\ldots<\beta_{n}, (n+1)(n+1) polynomials φ0(z),φ1,,φn(z)\varphi_{0}(z),\varphi_{1},\ldots,\varphi_{n}(z) and the following unitary curve on {0}\mathbb{C}\setminus\{0\}

f^(z)=(zβ0φ0(z),zβ1φ1(z),,zβnφn(z)).\hat{f}(z)=\left(z^{\beta_{0}}\varphi_{0}(z),z^{\beta_{1}}\varphi_{1}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z)\right).

Then by applying a suitable non-degenerate linear transformation to f^(z)\hat{f}(z), we can obtain

(5.12) g^(z)=(zβ0ψ0(z),zβ1ψ1(z),,zβnψn(z))\hat{g}(z)=\big{(}z^{\beta_{0}}\psi_{0}(z),z^{\beta_{1}}\psi_{1}(z),\ldots,z^{\beta_{n}}\psi_{n}(z)\big{)}

where ψ0(z),,ψn(z)\psi_{0}(z),\ldots,\psi_{n}(z) are polynomials such that β0+degψ0,β1+degψ1,,βn+degψn\beta_{0}+\deg\psi_{0},\beta_{1}+\deg\psi_{1},\ldots,\beta_{n}+\deg\psi_{n} are mutually distinct. Moreover, these two curves share the same regular singularities on the Riemann sphere \mathbb{C}\cup\infty.

Proof.

For simplicity, we define the term “quasi-degree” for the expressions β0+degφ0,β1+degφ1,,βn+degφn\beta_{0}+\deg\varphi_{0},\beta_{1}+\deg\varphi_{1},\ldots,\beta_{n}+\deg\varphi_{n}. These quasi-degrees correspond to the quasi-polynomials zβ0φ0(z),zβ1φ1(z),,zβnφn(z)z^{\beta_{0}}\varphi_{0}(z),z^{\beta_{1}}\varphi_{1}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z), where each component can include non-integer powers of zz.

Should any two quasi-degrees be equal, such that βi+degφi=βj+degφj\beta_{i}+\deg\varphi_{i}=\beta_{j}+\deg\varphi_{j} for i<ji<j, we can alter zβiφi(z)z^{\beta_{i}}\varphi_{i}(z) by adding a suitable multiple of zβjφj(z)z^{\beta_{j}}\varphi_{j}(z). This operation transforms zβiφi(z)z^{\beta_{i}}\varphi_{i}(z) into zβiφ~i(z)z^{\beta_{i}}\tilde{\varphi}_{i}(z), where degφ~i(z)<degφi(z)\deg\tilde{\varphi}_{i}(z)<\deg\varphi_{i}(z) and φ~i(0)0\tilde{\varphi}_{i}(0)\neq 0. Importantly, the terms zβiz^{\beta_{i}} and zβjz^{\beta_{j}} remain unchanged.

Consider the scenario where zβmφm(z)z^{\beta_{m}}\varphi_{m}(z) has the highest quasi-degree among all components, and it has the largest index among those with this quasi-degree. Beginning with this component, we use the aforementioned transformation to adjust other components with the same quasi-degree, reducing their degrees. This ensures that βm+degφm\beta_{m}+\deg\varphi_{m} stands out as the unique highest quasi-degree. Subsequently, we identify the next highest quasi-degree among the remaining components and apply similar transformations. After a finite series of these steps, the quasi-degrees of all components become distinct. This results in the transformed curve maintaining its form as g^(z)=(zβ0ψ0(z),zβ1ψ1(z),,zβnψn(z))\hat{g}(z)=\big{(}z^{\beta_{0}}\psi_{0}(z),z^{\beta_{1}}\psi_{1}(z),\ldots,z^{\beta_{n}}\psi_{n}(z)\big{)}.

Since curve g^\hat{g} differs from curve f^\hat{f} only by a non-degenerate linear transformation, they share the same regular singularities on the Riemann sphere.

By Lemma 7, we assume that for each curve

f^(z)=(zβ0φ0(z),zβ1φ1(z),,zβnφn(z))\hat{f}(z)=\left(z^{\beta_{0}}\varphi_{0}(z),z^{\beta_{1}}\varphi_{1}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z)\right)

in Example 2, the (n+1)(n+1) positive numbers β0+degφ0,,βn+degφn\beta_{0}+\deg\varphi_{0},\ldots,\beta_{n}+\deg\varphi_{n} are mutually distinct. Substituting z=1wz=\frac{1}{w} into f^\hat{f}, we obtain its local expression near z=z=\infty, i.e., w=0w=0, as

(5.13) f^(w)=(wβ0degφ0Φ0(w),wβ1degφ1Φ1(w),,wβndegφnΦn(w))\hat{f}(w)=\left(w^{-\beta_{0}-\deg\varphi_{0}}\Phi_{0}(w),w^{-\beta_{1}-\deg\varphi_{1}}\Phi_{1}(w),\ldots,w^{-\beta_{n}-\deg\varphi_{n}}\Phi_{n}(w)\right)

where Φ0,Φ1,,Φn\Phi_{0},\Phi_{1},\ldots,\Phi_{n} are polynomials non-vanishing at w=0w=0. Since β0+degφ0,β1+degφ1,,βn+degφn\beta_{0}+\deg\varphi_{0},\beta_{1}+\deg\varphi_{1},\ldots,\beta_{n}+\deg\varphi_{n} are pairwise distinct, we can rearrange the components in equation (4.41) according to the ascending order of the powers of ww. This arrangement yields the quasi-canonical form of f^(w)\hat{f}(w) near w=0w=0, which determines the sequence (γ[],i)i=1(\gamma_{[\infty],i})_{i=1}^{\infty}. The challenge then reduces to determining how many distinct degree vectors \vvd=(degφ0,degφ1,,degφn)\vv{d}=(\deg\varphi_{0},\deg\varphi_{1},\dots,\deg\varphi_{n}) are possible. The subsequent lemma is critical for this counting.

Lemma 8.

j=0ndegφj=i=1m(nγi,1+(n1)γi,2++γi,n)\displaystyle{\sum_{j=0}^{n}\,\deg\,\varphi_{j}=\sum\limits_{i=1}^{m}\Big{(}n\gamma_{i,1}+(n-1)\gamma_{i,2}+\ldots+\gamma_{i,n}\Big{)}}.

Proof.

Denote by AA the sum i=1m(nγi,1+(n1)γi,2++γi,n)\sum_{i=1}^{m}\left(n\gamma_{i,1}+(n-1)\gamma_{i,2}+\ldots+\gamma_{i,n}\right). Consider Λn(f^(z))\Lambda_{n}(\hat{f}(z)) as the Wronskian Wn+1(zβ0φ0(z),,zβnφn(z))W_{n+1}\left(z^{\beta_{0}}\varphi_{0}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z)\right) of the terms zβ0φ0(z),,zβnφn(z)z^{\beta_{0}}\varphi_{0}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z). This can be expanded by columns into a linear combination of finitely many quasi-monomials of the form Wn+1(zβ0+k0,zβ1+k1,,zβn+kn)W_{n+1}\left(z^{\beta_{0}+k_{0}},z^{\beta_{1}+k_{1}},\dots,z^{\beta_{n}+k_{n}}\right). Among these, the quasi-monomial summand

Wn+1(zβ0+degφ0,zβ1+degφ1,,zβn+degφn)W_{n+1}\left(z^{\beta_{0}+\deg\varphi_{0}},z^{\beta_{1}+\deg\varphi_{1}},\dots,z^{\beta_{n}+\deg\varphi_{n}}\right)

in Wn+1(zβ0φ0(z),,zβnφn(z))W_{n+1}(z^{\beta_{0}}\varphi_{0}(z),\ldots,z^{\beta_{n}}\varphi_{n}(z)) has the highest quasi-degree, calculated as i=0n(βi+degφi)n(n+1)2\sum_{i=0}^{n}(\beta_{i}+\deg\varphi_{i})-\frac{n(n+1)}{2}. The statement follows from equation (5.10). ∎

By this lemma, the degree vector \vvd=(degφ0,degφ1,,degφn)\vv{d}=(\deg\varphi_{0},\deg\varphi_{1},\dots,\deg\varphi_{n}) could achieve a finite number of possible values, not exceeding (A+n+1n)\begin{pmatrix}A+n+1\\ n\end{pmatrix}. Therefore, we complete the counting process.

In summary, we conclude the detailed discussion of Example 2.

6. Three open questions

We conclude this manuscript by posing the following three open questions.

Question 1.

Utilizing the notations from Subsection 1.3, we inquire about the characterization of the coefficient matrices Γ:=(γi,j)1ik1jn\Gamma:=\big{(}\gamma_{i,j}\big{)}_{\begin{subarray}{c}1\leq i\leq k\\ 1\leq j\leq n\end{subarray}} associated with the regular singularities \vvD\vv{D} represented by toric solutions to the SU(n+1){\rm SU}(n+1) Toda system on the Riemann sphere. Remarkably, Lin-Wei-Ye [24] addressed the case when k=2k=2, and Alexandre Eremenko [13] resolved the case when n=1n=1 for this inquiry.

Question 2.

A question akin to the one previously discussed emerges for compact Riemann surfaces of positive genus. In particular, for a specified genus gg, we aim to characterize the k×nk\times n matrices Γ\Gamma that correspond to regular singularities \vvD\vv{D} manifested by toric solutions on compact Riemann surfaces of genus gg. It is noteworthy that Gendron-Tahar [15] addressed this issue for the scenario when n=1n=1.

Question 3.

Consider a k×nk\times n matrix Γ\Gamma and a non-negative integer gg. The task is to determine the dimension of the moduli space of toric solutions on compact Riemann surfaces of genus gg, where the regular singularities \vvD\vv{D} are characterized by the coefficient matrix Γ\Gamma. Notably, Sicheng Lu and the last author explored this scenario for the case n=1n=1 in their study [26].

Acknowledgements: B.X. expresses sincere gratitude to Professor Guofang Wang at the University of Freiburg for introducing him to the field of Toda systems during the summer of 2018 and providing invaluable suggestions in the spring of 2024. Our heartfelt appreciation also goes to Professor Zhaohu Nie at the University of Utah, who kindly addressed several questions from B.X. related to Toda systems.

References

  • [1] Luca Battaglia. Existence and multiplicity result for the singular Toda system. J. Math. Anal. Appl., 424(1):49–85, 2015.
  • [2] Luca Battaglia, Aleks Jevnikar, Andrea Malchiodi, and David Ruiz. A general existence result for the Toda system on compact surfaces. Adv. Math., 285:937–979, 2015.
  • [3] Luca Battaglia and Andrea Malchiodi. Existence and non-existence results for the SU(3)SU(3) singular Toda system on compact surfaces. J. Funct. Anal., 270(10):3750–3807, 2016.
  • [4] Eugenio Calabi. Isometric imbedding of complex manifolds. Ann. of Math. (2), 58:1–23, 1953.
  • [5] Qing Chen, Wei Wang, Yingyi Wu, and Bin Xu. Conformal metrics with constant curvature one and finitely many conical singularities on compact Riemann surfaces. Pacific J. Math., 273(1):75–100, 2015.
  • [6] Zhijie Chen. Singular SU(3){\rm SU}(3) Toda system on the flat tori. Sci Sin Math, 53(8):1067–1084, 2023. In Chinese.
  • [7] Zhijie Chen and Chang-Shou Lin. Sharp results for SU(3)SU(3) toda system with critical parameters and monodromy of a third order linear ode. Preprint.
  • [8] Zhijie Chen and Chang-Shou Lin. On number and evenness of solutions of the SU(3)SU(3) Toda system on flat tori with non-critical parameters. J. Differential Geom., 125(1):85–120, 2023.
  • [9] Adam Doliwa. Holomorphic curves and Toda systems. Lett. Math. Phys., 39(1):21–32, 1997.
  • [10] S. K. Donaldson. Kähler metrics with cone singularities along a divisor. In Essays in mathematics and its applications, pages 49–79. Springer, Heidelberg, 2012.
  • [11] A. Eremenko. A Toda lattice in dimension 2 and Nevanlinna theory. Zh. Mat. Fiz. Anal. Geom., 3(1):39–46, 129, 2007.
  • [12] Alexandre Eremenko. Co-axial monodromy. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 20(2):619–634, 2020.
  • [13] Alexandre Eremenko. Co-axial monodromy. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 20(2):619–634, 2020.
  • [14] Hershel M. Farkas and Irwin Kra. Riemann surfaces, volume 71 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1992.
  • [15] Quentin Gendron and Guillaume Tahar. Dihedral monodromy of cone spherical metrics. Illinois J. Math., 67(3):457–483, 2023.
  • [16] Jean-Loup Gervais and Yutaka Matsuo. Classical AnA_{n}-WW-geometry. Comm. Math. Phys., 152(2):317–368, 1993.
  • [17] A. B. Givental’. Plücker formulas and Cartan matrices. Uspekhi Mat. Nauk, 44(3(267)):155–156, 1989.
  • [18] P. Griffiths. On Cartan’s method of Lie groups and moving frames as applied to uniqueness and existence questions in differential geometry. Duke Math. J., 41:775–814, 1974.
  • [19] Phillip Griffiths and Joseph Harris. Principles of algebraic geometry. Wiley Classics Library. John Wiley & Sons, Inc., New York, 1994. Reprint of the 1978 original.
  • [20] Jürgen Jost, Changshou Lin, and Guofang Wang. Analytic aspects of the Toda system. II. Bubbling behavior and existence of solutions. Comm. Pure Appl. Math., 59(4):526–558, 2006.
  • [21] Jürgen Jost and Guofang Wang. Classification of solutions of a Toda system in 2{\mathbb{R}}^{2}. Int. Math. Res. Not., (6):277–290, 2002.
  • [22] D. Karmakar, C.-S. Lin, Z. Nie, and J. Wei. Total masses of solutions to general toda systems with singular sources. To appear in Advances in Mathematics, 2024.
  • [23] Chang-Shou Lin, Zhaohu Nie, and Juncheng Wei. Toda systems and hypergeometric equations. Trans. Amer. Math. Soc., 370(11):7605–7626, 2018.
  • [24] Chang-Shou Lin, Juncheng Wei, and Dong Ye. Classification and nondegeneracy of SU(n+1)SU(n+1) Toda system with singular sources. Invent. Math., 190(1):169–207, 2012.
  • [25] Chang-Shou Lin, Wen Yang, and Xuexiu Zhong. A priori estimates of Toda systems, I: the Lie algebras of 𝐀n{\bf A}_{n}, 𝐁n{\bf B}_{n}, 𝐂n{\bf C}_{n} and 𝐆2{\bf G}_{2}. J. Differential Geom., 114(2):337–391, 2020.
  • [26] Sicheng Lu and Bin Xu. Moduli space of dihedral spherical surfaces and measured foliations. https://arxiv.org/abs/2312.01807, 2024.
  • [27] Gabriele Mondello and Dmitri Panov. Spherical metrics with conical singularities on a 2-sphere: angle constraints. International Mathematics Research Notices, 2016(16):4937–4995, 10 2015.
  • [28] Jingyu Mu, Yiqian Shi, Tianyang Sun, and Bin Xu. Classifying solutions of SU(n + 1) toda system around a singular source. Proc. Amer. Math. Soc., 152(6):2585–2600, June 2024.
  • [29] L. E. Positselskii. Local Plücker formulas for a semisimple Lie group. Funktsional. Anal. i Prilozhen., 25(4):74–76, 1991.
  • [30] George Springer. Introduction to Riemann surfaces. Chelsea Publishing Co., New York, 1981.
  • [31] William P. Thurston. Three-dimensional geometry and topology. Vol. 1, volume 35 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1997.
  • [32] Masaaki Umehara and Kotaro Yamada. Metrics of constant curvature 1 with three conical singularities on the 22-sphere. Illinois Journal of Mathematics, 44(1):72 – 94, 2000.
  • [33] Hermann Weyl. The concept of a Riemann surface. ADIWES International Series in Mathematics. Addison-Wesley Publishing Co., Inc., Reading, Mass.-London, german edition, 1964.