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

The Convergence of Prescribed Combinatorial Ricci Flows for Total Geodesic Curvatures in Spherical Background Geometry

Guangming Hu, Ziping Lei,Yu Sun and Puchun Zhou
Abstract

In this paper, we study the existence and rigidity of (degenerated) circle pattern metric with prescribed total geodesic curvatures in spherical background geometry. To find the (degenerated) circle pattern metric with prescribed total geodesic curvatures, we define some prescribed combinatorial Ricci flows and study the convergence of flows for (degenerated) circle pattern metrics. We solve the prescribed total geodesic curvature problem and provide two methods to find the degenerated circle pattern metric with prescribed total geodesic curvatures. As far as we know, this is the first degenerated result for total geodesic curvatures in spherical background geometry.

Mathematics Subject Classification (2020): 52C25, 52C26, 53A70.

1 Introduction

1.1 Background

In differential geometry, Hamilton [5] introduced the Ricci flow defined by some equations dgijdt=2Kgij\frac{dg_{ij}}{dt}=-2Kg_{ij}, where gijg_{ij} is the Riemannian metric and KK is the Gaussian curvature. This is an important tool to solve the Poincaré conjecture [7, 8, 9] and Thurston’s geometrization conjecture [1]. Moreover, the Ricci flow is used to prove the uniformization theorem [2].

In discrete geometry, Chow and Luo [3] constructed the combinatorial Ricci flow which is an analogy of the Ricci flow in smooth case. Given a closed surface SS with a triangulation 𝒯=(V,E,F)\mathscr{T}=(V,E,F) and a weight Φ:E[0,π2]\Phi:E\to[0,\frac{\pi}{2}], then we can define a map 𝐫=(r1,,r|V|):V(0,+)\mathbf{r}=(r_{1},\cdots,r_{|V|}):V\to(0,+\infty), the 𝐫\mathbf{r} is called a metric on the weighted triangulation (𝒯,Φ)(\mathscr{T},\Phi). Using the metric 𝐫\mathbf{r}, for any edge eijEe_{ij}\in E joining viv_{i} and vjv_{j}, we can define the length lijl_{ij} of eije_{ij} in spherical (Euclidean, hyperbolic) background geometry, i.e.

lij=arccos(cosricosrjsinrisinrjcos(Φ(eij)))(𝕊2),l_{ij}=\arccos(\cos r_{i}\cos r_{j}-\sin r_{i}\sin r_{j}\cos(\Phi(e_{ij})))\leavevmode\nobreak\ (\mathbb{S}^{2}),
lij=ri2+rj2+2rirjcos(Φ(eij))(2)l_{ij}=\sqrt{r_{i}^{2}+r_{j}^{2}+2r_{i}r_{j}\cos(\Phi(e_{ij}))}\leavevmode\nobreak\ (\mathbb{R}^{2})

and

lij=arccosh(coshricoshrj+sinhrisinhrjcos(Φ(eij)))(2).l_{ij}=\operatorname{arccosh}(\cosh r_{i}\cosh r_{j}+\sinh r_{i}\sinh r_{j}\cos(\Phi(e_{ij})))\leavevmode\nobreak\ (\mathbb{H}^{2}).

Then there exists a local spherical (Euclidean, hyperbolic) metric on each triangle ijkF\triangle_{ijk}\in F whose vertices are viv_{i}, vjv_{j} and vkv_{k}. Gluing these spherical (Euclidean, hyperbolic) triangles along their sides together, then we obtain a spherical (Euclidean, hyperbolic) conical metric with singularities at the vertices on the closed surface SS. Denote ϑijk\vartheta_{i}^{jk} as the angle at the vertex viv_{i} in the spherical (Euclidean, hyperbolic) triangle ijkF\triangle_{ijk}\in F. The combinatorial Gaussian curvature KiK_{i} at the vertex viv_{i} is defined as

Ki=2πijkFϑijk,K_{i}=2\pi-\sum_{\triangle_{ijk}\in F}\vartheta_{i}^{jk},

where the sum is taken over all triangles with the vertex viv_{i} in FF. Then Chow and Luo [3] defined the combinatorial Ricci flow in spherical (Euclidean, hyperbolic) background geometry, i.e.

(1.1) dridt=Kisinri(𝕊2),\frac{dr_{i}}{dt}=-K_{i}\sin r_{i}\leavevmode\nobreak\ (\mathbb{S}^{2}),
(1.2) dridt=Kiri(2)\frac{dr_{i}}{dt}=-K_{i}r_{i}\leavevmode\nobreak\ (\mathbb{R}^{2})

and

(1.3) dridt=Kisinhri(2).\frac{dr_{i}}{dt}=-K_{i}\sinh r_{i}\leavevmode\nobreak\ (\mathbb{H}^{2}).

Chow and Luo [3] studied the combinatorial Ricci flow and obtained the long time existence of the solution to the combinatorial Ricci flow. Moreover, the solution to the combinatorial Ricci flow will converge exponentially fast to the metric with constant curvature in some cases. They obtained the following two theorems:

Theorem 1.1.

Suppose (𝒯,Φ)(\mathscr{T},\Phi) is a weighted triangulation of a closed connected surface SS. Given any initial metric based on the weighted triangulation, the solution to the combinatorial Ricci flow (1.2) in the Euclidean background geometry with the given initial value exists for all time and converges if and only if for any proper subset IVI\subset V,

(1.4) 2π|I|χ(S)/|V|>(e,v)Lk(I)(πΦ(e))+2πχ(FI),2\pi|I|\chi(S)/|V|>-\sum_{(e,v)\in Lk(I)}(\pi-\Phi(e))+2\pi\chi(F_{I}),

where FIF_{I} is the subcomplex of 𝒯\mathscr{T} consisting of all simplex whose vertices are contained in II, (e,v)Lk(I)(e,v)\in\operatorname{Lk}(I) is a triangle in FF with a vertex vv in (e,v)(e,v) belongs to II and neither of the endpoints of the edge ee in (e,v)(e,v) opposite vv belongs to II and Lk(I)\operatorname{Lk}(I) is the set of all such triangles.

Furthermore, if the solution converges, then it converges exponentially fast to the metric with constant curvature.

Theorem 1.2.

Suppose (𝒯,Φ)(\mathscr{T},\Phi) is a weighted triangulation of a closed connected surface SS of negative Euler characteristic. Given any initial metric, the solution to (1.3) in the hyperbolic background geometry with the given initial value exists for all time and converges if and only if the following two conditions hold.

  1. 1.

    For any three edges e1,e2,e3e_{1},e_{2},e_{3} forming a null homotopic loop in SS, if i=13Φ(ei)π\sum_{i=1}^{3}\Phi\left(e_{i}\right)\geq\pi, then these three edges form the boundary of a triangle in FF.

  2. 2.

    For any four edges e1,e2,e3,e4e_{1},e_{2},e_{3},e_{4} forming a null homotopic loop in SS, if i=14Φ(ei)\sum_{i=1}^{4}\Phi\left(e_{i}\right)\geq 2π2\pi, then these eie_{i} ’s form the boundary of the union of two adjacent triangles in FF.

Furthermore, if it converges, then it converges exponentially fast to a hyperbolic metric on SS so that all vertex angles are 2π2\pi.

Besides, Chow and Luo [3] posed some questions and one of them is whether the limit limt+ri(t)\lim_{t\to+\infty}r_{i}(t) of the solution ri(t)r_{i}(t) to combinatorial Ricci flow always exists in the extended set [0,+][0,+\infty] when the combinatorial conditions (1.4), 1 and 2 are not valid.

Takatsu [10] studied the question and solved it, based on an infinitesimal description of degenerated circle pattern metrics. She obtained the following theorem:

Theorem 1.3.

Let (𝒯,Φ)(\mathscr{T},\Phi) be a weighted triangulation of a surface SS of nonpositive Euler characteristic such that ϕ(I)0\phi(I)\leq 0 holds for any subset IVI\subset V and ZT:={zVZ_{T}:=\{z\in V\mid there exists a proper subset ZZ of VV such that zZz\in Z and ϕ(Z)=0}\phi(Z)=0\} is nonempty, where

ϕ(I):=(e,v)Lk(I)(πΦ(e))+2πχ(FI).\phi(I):=-\sum_{(e,v)\in Lk(I)}(\pi-\Phi(e))+2\pi\chi\left(F_{I}\right).

Then for any metric 𝐫\mathbf{r} on (𝒯,Φ)(\mathscr{T},\Phi), the solution {𝐫(t)}t0\{\mathbf{r}(t)\}_{t\geq 0} to combinatorial Ricci flow with initial data 𝐫\mathbf{r} does not converge on >0|V|\mathbb{R}_{>0}^{|V|} at infinity. However,

limt+Ki(𝐫(t))=0\lim_{t\rightarrow+\infty}K_{i}(\mathbf{r}(t))=0

holds for any vertex viv_{i}.

On the one hand, for χ(S)=0\chi(S)=0, the solution {𝐫(t)}t0\{\mathbf{r}(t)\}_{t\geq 0} to combinatorial Ricci flow does not converge on 0|V|\mathbb{R}_{\geq 0}^{|V|} at infinity. However, if we fix an arbitrary vV\ZTv\in V\backslash Z_{T}, then the limit

ρu:=limt+ru(t)rv(t)\rho_{u}:=\lim_{t\rightarrow+\infty}\frac{r_{u}(t)}{r_{v}(t)}

exists for any uVu\in V, where ZT={zVρz=0}Z_{T}=\left\{z\in V\mid\rho_{z}=0\right\} holds and (ρu)uV\ZT\left(\rho_{u}\right)_{u\in V\backslash Z_{T}} is a unique circle pattern metric with normalization ρv=1\rho_{v}=1 on a certain weighted triangulation with vertices V\ZTV\backslash Z_{T}.

On the other hand, for χ(S)<0\chi(S)<0, the solution {𝐫(t)}t0\{\mathbf{r}(t)\}_{t\geq 0} to combinatorial Ricci flow converges on 0|V|\mathbb{R}_{\geq 0}^{|V|} at infinity, where we have ZT={zVlimtrz(t)=0}Z_{T}=\left\{z\in V\mid\lim_{t\rightarrow\infty}r_{z}(t)=0\right\} holds and the limit of (rv(t))vV\ZT\left(r_{v}(t)\right)_{v\in V\backslash Z_{T}} at infinity is a unique circle pattern metric on a certain weighted triangulation with vertices V\ZTV\backslash Z_{T}.

Recently, Nie [6] defined a new combinatorial scalar curvature, i.e. the total geodesic curvature and constructed a convex functional with total geodesic curvature. He obtained the rigidity of circle patterns in spherical background geometry. Followed his work, the last author of this paper and his collaborators [4] defined a combinatorial curvature flow which is an analogy of the prescribed combinatorial Ricci flow. They obtained an algorithm to find the desired ideal circle pattern.

Motivated by [3, 10], we obtain some results for total geodesic curvature in spherical background geometry. In this paper, we will define the prescribed combinatorial Ricci flow for total geodesic curvature in spherical background geometry and study the convergence of the solution to prescribed combinatorial Ricci flow.

1.2 Main results

1.2.1 Spherical conical metrics on surfaces

Given a closed topological surface SS and a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F) of SS. Denote V,EV,E and FF as set of 0-cells, 11-cells and 22-cells, respectively. By |E||E| we denote the number of 1-cells and by |F||F| we denote the number of 2-cells. We can define a function 𝐫\mathbf{r} on VV and a function Φ\Phi on EE, i.e. 𝐫=(r1,,r|F|):V(0,π2]\mathbf{r}=(r_{1},\cdots,r_{|F|}):V\to(0,\frac{\pi}{2}] and Φ:E(0,π2)\Phi:E\to(0,\frac{\pi}{2}). Besides, the function Φ\Phi is called a weight.

For any 2-cell fFf\in F, we choose a auxiliary point pffp_{f}\in f and add an edge between each vertex on f\partial f and pp. Then we obtain a triangulation of SS as shown in Figure 1 .

Refer to caption
Figure 1: The cellular decomposition and triangulation of SS

For any 1-cell eEe\in E, there exists a quadrilateral Q(e)Q(e), where the vertices of Q(e)Q(e) are the end points of ee and auxiliary points of the 2-cells on the two sides of ee. Denote the two auxiliary points as p,pp,{p}^{\prime} and f(p),f(p)Ff(p),f({p}^{\prime})\in F denotes the 2-cells containing the points p,pp,{p}^{\prime}, respectively.

Given the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|}, for any radii 𝐫(0,π2]|F|\mathbf{r}\in(0,\frac{\pi}{2}]^{|F|}, by Lemma 3.1, there exists a local spherical metric on the quadrilateral Q(e)Q(e) such that Q(e)Q(e) is the corresponding spherical quadrilateral of two spherical disks, where the radii of the disks are rf(p)r_{f(p)} and rf(p)r_{f({p}^{\prime})} and their intersection angle is Φ(e)\Phi(e) as shown in Figure 2.

Refer to caption
Figure 2: The spherical quadrilateral Q(e)Q(e)

Gluing these spherical quadrilaterals along their sides together, then we obtain a spherical conical metric on the closed topological surface SS. Denote the spherical conical metric as μ\mu, we obtain a spherical conical surface (S,μ)(S,\mu).

By definition, these spherical disks form a circle pattern 𝒫\mathcal{P} on (S,μ)(S,\mu) which realizes the weight Φ\Phi. By (μ,𝒫)(\mu,\mathcal{P}) we denote the spherical conical metric μ\mu with circle pattern 𝒫\mathcal{P}. Besides, if the radii 𝐫(0,π2)|F|\mathbf{r}\in(0,\frac{\pi}{2})^{|F|}, then the spherical conical metric μ\mu is called a circle pattern metric. If there exists some radii equal to π2\frac{\pi}{2}, then the spherical conical metric μ\mu is called a degenerated circle pattern metric.

Remark 1.4.

Given a closed topological surface SS with a cellular decomposition Σ\Sigma, by the above argument, for the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the radii 𝐫(0,π2]|E|\mathbf{r}\in(0,\frac{\pi}{2}]^{|E|}, there exists the corresponding circle pattern 𝒫\mathcal{P} which realizes the weight Φ\Phi and the radii 𝐫\mathbf{r}.

For a disk D𝕊2D\subset\mathbb{S}^{2}, suppose that the radius of DD is r(0,π2]r\in(0,\frac{\pi}{2}]. Then the D\partial D has constant geodesic curvature k(D)=cotrk(\partial D)=\cot r. Hence we can define a map ι\iota, i.e.

ι:(0,π2]|F|0|F|𝐫=(r1,,r|F|)Tk=(cotr1,,cotr|F|)T\begin{array}[]{cccc}\iota:(0,\frac{\pi}{2}]^{|F|}&\longrightarrow&\mathbb{R}^{|F|}_{\geq 0}\\ \mathbf{r}=(r_{1},\cdots,r_{|F|})^{T}&\longmapsto&k=(\cot r_{1},\cdots,\cot r_{|F|})^{T}\\ \end{array}

The map ι\iota is bijective.

Besides, for simplicity, in this paper, if the geodesic curvatures k0|F|k\in\mathbb{R}_{\geq 0}^{|F|}, then kk is also called a spherical conical metric. If the geodesic curvatures k>0|F|k\in\mathbb{R}_{>0}^{|F|}, then kk is also called a circle pattern metric. If there exists some geodesic curvatures equal to 0, then kk is also called a degenerated circle pattern metric.

For any prescribed total geodesic curvature L^\hat{L}, we want to know whether there exists some (degenerated) circle pattern metrics with total geodesic curvature L^\hat{L}. In other words, we consider the following problem:

Prescribed Total Geodesic Curvature Problem: Is there a (degenerated) circle pattern metric with the prescribed total geodesic curvature L^\hat{L}? If it exists, how to find it?

Our first results solve the first part of prescribed total geodesic curvature problem:

Theorem 1.5.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F) and the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|}. For the prescribed total geodesic curvature L^=(L^1,,L^|F|)T\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{|F|})^{T} on the face set FF, there exists a circle pattern metric on SS with total geodesic curvature L^\hat{L} if and only if the prescribed total geodesic curvature L^1\hat{L}\in\mathcal{L}_{1}, where

1={(L1,,L|F|)T>0|F||fFLf<2eEFΦ(e),FF}.\mathcal{L}_{1}=\{(L_{1},\cdots,L_{|F|})^{T}\in\mathbb{R}^{|F|}_{>0}\leavevmode\nobreak\ |\leavevmode\nobreak\ \sum_{f\in F^{\prime}}L_{f}<2\sum_{e\in E_{F^{\prime}}}\Phi(e),\forall F^{\prime}\subset F\}.

Moreover, the circle pattern metric is unique up to isometry if it exists.

Remark 1.6.

Theorem 1.5 is first proved in [6] and we provide a concrete proof in this paper.

Theorem 1.7.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F) and the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|}. For the prescribed total geodesic curvature L^=(L^1,,L^|F|)T\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{|F|})^{T} on the face set FF, there exists a degenerated circle pattern metric on SS with total geodesic curvature L^\hat{L} if and only if the prescribed total geodesic curvature L^¯\hat{L}\in\bar{\partial}\mathcal{L}, where

¯=1m|F|1,1i1<<im|F|i1im{0},\bar{\partial}\mathcal{L}=\bigcup_{1\leq m\leq|F|-1,1\leq i_{1}<\cdots<i_{m}\leq|F|}\mathcal{L}_{i_{1}\cdots i_{m}}\cup\{0\},
i1im={(0,,Li1,,0,,Lim,0,,0)T|F|fFi1imLf<2eEFi1imΦ(e),Fi1imFi1im;Lij>0,1jm}.\mathcal{L}_{i_{1}\cdots i_{m}}=\left\{\begin{array}[]{l|l}(0,\cdots,L_{i_{1}},\cdots,0,\cdots,L_{i_{m}},0,\cdots,0)^{T}\in\mathbb{R}^{|F|}&\begin{array}[]{l}\sum_{f\in F_{i_{1}\cdots i_{m}}^{\prime}}L_{f}<2\sum_{e\in E_{F_{i_{1}\cdots i_{m}}^{\prime}}}\Phi(e),\\ \forall F_{i_{1}\cdots i_{m}}^{\prime}\subset F_{i_{1}\cdots i_{m}};L_{i_{j}}>0,1\leq j\leq m\end{array}\end{array}\right\}.

Moreover, the degenerated circle pattern metric is unique up to isometry if it exists.

Theorem 1.8.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F) and the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|}. For the prescribed total geodesic curvature L^=(L^1,,L^|F|)T\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{|F|})^{T} on the face set FF, there exists a spherical conical metric on SS with total geodesic curvature L^\hat{L} if and only if the prescribed total geodesic curvature L^\hat{L}\in\mathcal{L}, where =1¯\mathcal{L}=\mathcal{L}_{1}\cup\bar{\partial}\mathcal{L}. Moreover, the spherical conical metric is unique up to isometry if it exists.

1.2.2 Prescribed combinatorial Ricci flows for total geodesic curvatures

Given L^=(L^1,,L^|F|)T1\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{|F|})^{T}\in\mathcal{L}_{1}, then we can define the prescribed combinatorial Ricci flows as follows, i.e.

(1.5) dkidt=(LiL^i)ki,iF.\frac{dk_{i}}{dt}=-(L_{i}-\hat{L}_{i})k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F.

For any ki1im|F|k\in\mathbb{R}_{i_{1}\cdots i_{m}}^{|F|}, where 1m|F|11\leq m\leq|F|-1, 1i1<<im|F|1\leq i_{1}<\cdots<i_{m}\leq|F| and

i1im|F|={(0,,ki1,0,,kim,,0)|F||ki1,,kim>0},\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}}=\{(0,\cdots,k_{i_{1}},0,\cdots,k_{i_{m}},\cdots,0)\in\mathbb{R}^{|F|}\leavevmode\nobreak\ |\leavevmode\nobreak\ k_{i_{1}},\cdots,k_{i_{m}}>0\},

then k~=(ki1,,kim)>0m\tilde{k}=(k_{i_{1}},\cdots,k_{i_{m}})\in\mathbb{R}_{>0}^{m}. Given L^i1im\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}}, we construct the prescribed combinatorial Ricci flows as follows, i.e.

(1.6) dk~idt=(Li(k~)L^i)k~i,iFi1im,\frac{d\tilde{k}_{i}}{dt}=-(L_{i}(\tilde{k})-\hat{L}_{i})\tilde{k}_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F_{i_{1}\cdots i_{m}},

where Fi1im={i1,,im}F_{i_{1}\cdots i_{m}}=\{i_{1},\cdots,i_{m}\}.

Given the prescribed total geodesic curvature L^i1im\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}}, we can define the prescribed combinatorial Ricci flows as follows, i.e.

(1.7) dkidt=(LiL^i)ki,iFi1im,dkidt=Liki,iFFi1im.\frac{dk_{i}}{dt}=-(L_{i}-\hat{L}_{i})k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F_{i_{1}\cdots i_{m}},\leavevmode\nobreak\ \frac{dk_{i}}{dt}=-L_{i}k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F\setminus F_{i_{1}\cdots i_{m}}.

The following results solve the second part of prescribed total geodesic curvature problem:

Theorem 1.9.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^=(L^1,,L^|F|)T1\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{|F|})^{T}\in\mathcal{L}_{1} on the face set FF. For any initial geodesic curvature k(0)>0|F|k(0)\in\mathbb{R}_{>0}^{|F|}, the solution of the prescribed combinatorial Ricci flows (1.5) converges to the unique circle pattern metric with the total geodesic curvature L^\hat{L} up to isometry.

Remark 1.10.

Theorem 1.9 is first proved in [4] and we provide a new proof in this paper.

Theorem 1.11.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^i1im(1m|F|1\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}}(1\leq m\leq|F|-1, 1i1<<im|F|)1\leq i_{1}<\cdots<i_{m}\leq|F|) on the face set FF. For any initial geodesic curvature k~(0)>0m\tilde{k}(0)\in\mathbb{R}_{>0}^{m} on the face set Fi1imF_{i_{1}\cdots i_{m}} and 0 on the face set FFi1imF\setminus F_{i_{1}\cdots i_{m}}, the solution of the prescribed combinatorial Ricci flows (1.6) converges to the unique degenerated circle pattern metric with the total geodesic curvature L~=(L^i1,,L^im)T\tilde{L}=(\hat{L}_{i_{1}},\cdots,\hat{L}_{i_{m}})^{T} on Fi1imF_{i_{1}\cdots i_{m}} and 0 on FFi1imF\setminus F_{i_{1}\cdots i_{m}} up to isometry. Moreover, if the solution converges to the degenerated circle pattern metric k^\hat{k}, then k^i1im|F|\hat{k}\in\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}}.

Theorem 1.12.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^i1im(1m|F|1\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}}(1\leq m\leq|F|-1, 1i1<<im|F|)1\leq i_{1}<\cdots<i_{m}\leq|F|) on the face set FF. For any initial geodesic curvature k(0)>0|F|k(0)\in\mathbb{R}_{>0}^{|F|}, the solution of the prescribed combinatorial Ricci flows (1.7) converges to the unique degenerated circle pattern metric with the total geodesic curvature L^\hat{L} up to isometry. Moreover, if the solution converges to the degenerated circle pattern metric k^\hat{k}, then k^i1im|F|\hat{k}\in\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}}.

Remark 1.13.

By Theorem 1.11 and Theorem 1.12, for finding the degenerated circle pattern metric with the total geodesic curvature L^i1im(1m|F|1,1i1<<im|F|)\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}}\leavevmode\nobreak\ (1\leq m\leq|F|-1,1\leq i_{1}<\cdots<i_{m}\leq|F|), we have two methods. The first one is by constructing prescribed combinatorial Ricci flows (1.6) and the second one is by constructing prescribed combinatorial Ricci flows (1.7). Then the solutions of the two flows will converge to the same degenerated circle pattern metric with the total geodesic curvature L^\hat{L} in different dimensions (see figure 3).

Refer to caption
Figure 3: The solutions of the two flows converge to k^\hat{k}

Organization This paper is organized as follows. In section 2, we define the circle patterns on spherical surfaces. In section 3, we study the moduli space of spherical conical metrics. In section 4, we construct the potential functions on bigons and define the moduli space of spherical bigons with given angle. In section 5, we prove the existence and rigidity of circle pattern metrics for prescribed total geodesic curvatures. In section 6, we prove the existence and rigidity of degenerated circle pattern metrics for prescribed total geodesic curvatures. In section 7, we prove the existence and rigidity of spherical conical metrics for prescribed total geodesic curvatures. In section 8, we study the convergence of prescribed combinatorial Ricci flows for circle pattern metrics. In section 9, we study the convergence of prescribed combinatorial Ricci flows for degenerated circle pattern metrics.

2 Circle patterns on spherical surfaces

Let SS be a topological surface, then SS is called a spherical surface if every point pp in SS has a neighborhood UU such that there is a isometry from UU onto an open subset of 𝕊α2\mathbb{S}_{\alpha}^{2}, where 𝕊α2\mathbb{S}_{\alpha}^{2} is a unit sphere with a cone angle α>0\alpha>0. We also often call the spherical surface SS a surface with spherical conical metric.

By Dα(o,r)D_{\alpha}(o,r) we denote the open disk of radius rr in 𝕊α2\mathbb{S}_{\alpha}^{2} centered at the cone point o𝕊α2o\in\mathbb{S}_{\alpha}^{2}. In this paper, we only consider the disk Dα(o,r)D_{\alpha}(o,r) with radius r(0,π2]r\in(0,\frac{\pi}{2}]. By a bigon in a spherical surface SS we denote an open set isometric to the intersection of two disks of radii r1,r2(0,π2]r_{1},r_{2}\in(0,\frac{\pi}{2}] in 𝕊α2\mathbb{S}_{\alpha}^{2} not containing each other. Besides, a bigon does not contain any cone point. The angle of a bigon is the interior angle ϕ\phi formed by its two sides.

Refer to caption
Figure 4: The circle pattern 𝒫\mathcal{P} and cellular decomposition Σ\Sigma on SS

Using similar analysis in proof of Proposition 2 in [6], we have the following lemma:

Lemma 2.1.

Given some disks Di=Dαi(o,ri)(1in)D_{i}=D_{\alpha_{i}}\left(o,r_{i}\right)\leavevmode\nobreak\ (1\leq i\leq n) as above with αi>0,ri(0,π2]\alpha_{i}>0,r_{i}\in(0,\frac{\pi}{2}]. Let η\eta be a local isometry from the disjoint union D1DnD_{1}\sqcup\cdots\sqcup D_{n} to a closed spherical surface Γ\Gamma such that

  1. 1.

    the set Γη(D1Dn)\Gamma\setminus\eta(D_{1}\sqcup\cdots\sqcup D_{n}) only has finitely many points;

  2. 2.

    η\eta is at most 2 to 1;

  3. 3.

    the η(Di)(1in)\eta\left(D_{i}\right)\leavevmode\nobreak\ (1\leq i\leq n) do not contain each other.

By \mathscr{B} we denote the set :={pΓη1(p)has two points}\mathscr{B}:=\{p\in\Gamma\mid\eta^{-1}(p)\leavevmode\nobreak\ \text{has two points}\}. Then η1()\eta^{-1}(\mathscr{B}) is a disjoint union of bigons which fills up the boundary of the Di(1in)D_{i}\leavevmode\nobreak\ (1\leq i\leq n).

For more details, we refer [6] to the readers. Then we can define the circle patterns on spherical surfaces in this paper.

Definition 2.2.

A local isometry η:D1DnΓ\eta:D_{1}\sqcup\cdots\sqcup D_{n}\rightarrow\Gamma as in Lemma 2.1 is called a circle pattern on Γ\Gamma and each connected component of the set Γ\mathscr{B}\subset\Gamma is called a bigon of the circle pattern. For simplicity, we also call Di(1in)D_{i}\leavevmode\nobreak\ (1\leq i\leq n) an immersed disk in Γ\Gamma and call the set of immersed disks 𝒫={D1,,Dn}\mathcal{P}=\left\{D_{1},\cdots,D_{n}\right\} a circle pattern (so that each bigon is either an intersection component of two such disks Di,DjD_{i},D_{j} or a "self-intersection component" of a single disk DiD_{i}).

Let (G𝒫,Φ)(G_{\mathcal{P}},\Phi) be the weighted graph on Γ\Gamma with the circle pattern 𝒫\mathcal{P} which satisfies

  1. 1.

    the vertex set V𝒫V_{\mathcal{P}} is the complement of D1DnD_{1}\cup\cdots\cup D_{n} in Γ\Gamma: V𝒫=Γ\D1DnV_{\mathcal{P}}=\Gamma\backslash D_{1}\cup\cdots\cup D_{n};

  2. 2.

    the edges in edge set E𝒫E_{\mathcal{P}} are 1 to 1 correspondence with the bigons: given a bigon BB, then the corresponding edge eE𝒫e\in E_{\mathcal{P}} is a curve in BB joining the two endpoints of the bigon BB;

  3. 3.

    each edge eE𝒫e\in E_{\mathcal{P}} is weighted by the angle Φ(e)(0,π)\Phi(e)\in(0,\pi) of the corresponding bigon;

  4. 4.

    each face fF𝒫f\in F_{\mathcal{P}} corresponds to a disk Di𝒫D_{i}\in\mathcal{P}: ff is contained in D¯i\bar{D}_{i} and each vertex on f\partial f is in Di\partial D_{i}.

Remark 2.3.

Given a closed spherical surface SS with a circle pattern 𝒫\mathcal{P}, then the weighted graph (G𝒫,Φ)(G_{\mathcal{P}},\Phi) gives a cellular decomposition Σ\Sigma of the surface SS (see figure 4).

Refer to caption
Figure 5: The bigon of angle ϕ\phi on 𝕊2\mathbb{S}^{2}

3 The moduli space of spherical conical metrics

Lemma 3.1.

For any r1,r2(0,π2]r_{1},r_{2}\in(0,\frac{\pi}{2}] and any ϕ(0,π2)\phi\in(0,\frac{\pi}{2}), there exists a bigon of angle ϕ\phi which is formed by the disk Di𝕊2D_{i}\subset\mathbb{S}^{2}, where the DiD_{i} has radius rir_{i}, i=1,2i=1,2. Moreover, the bigon is unique up to isometry.

Proof.

First consider the case r1r2r_{1}\leq r_{2}. Let Di𝕊2D_{i}\subset\mathbb{S}^{2} be the disk which has radius rir_{i}, i=1,2i=1,2. We suppose that the distance between the centers of D1D_{1} and D2D_{2} is r3r_{3}, the intersection angle of D1D_{1} and D2D_{2} is ϕ\phi. By the law of cosines in spherical geometry, we have the following identity, i.e.

cosr3=cosr1cosr2+sinr1sinr2cos(πϕ)=cosr1cosr2sinr1sinr2cosϕ.\cos r_{3}=\cos r_{1}\cos r_{2}+\sin r_{1}\sin r_{2}\cos(\pi-\phi)=\cos r_{1}\cos r_{2}-\sin r_{1}\sin r_{2}\cos\phi.

For r1,r2(0,π2]r_{1},r_{2}\in(0,\frac{\pi}{2}], ϕ(0,π2)\phi\in(0,\frac{\pi}{2}), we have

cos(r1+r2)<cosr3=cosr1cosr2sinr1sinr2cosϕ<cos(r1+r2).\cos(r_{1}+r_{2})<\cos r_{3}=\cos r_{1}\cos r_{2}-\sin r_{1}\sin r_{2}\cos\phi<\cos(r_{1}+r_{2}).

Since 0r1+r2,r2r1π0\leq r_{1}+r_{2},r_{2}-r_{1}\leq\pi, 0<r3<r1+r2π0<r_{3}<r_{1}+r_{2}\leq\pi, we know that the r3r_{3} is unique and r3(r2r1,r1+r2)r_{3}\in(r_{2}-r_{1},r_{1}+r_{2}). The D1D2D_{1}\cap D_{2} is the bigon of angle ϕ\phi which is formed by the disks D1,D2𝕊2D_{1},D_{2}\subset\mathbb{S}^{2} as shown in Figure 5. Since r3r_{3} is unique, the bigon is unique up to isometry.

For the case r1>r2r_{1}>r_{2}, it is similar to the analysis above. This completes the proof. ∎

Remark 3.2.

For the disks D1,D2𝕊2D_{1},D_{2}\subset\mathbb{S}^{2} with radii r1,r2(0,π2]r_{1},r_{2}\in(0,\frac{\pi}{2}] and the intersection angle ϕ(0,π2)\phi\in(0,\frac{\pi}{2}), D¯1\bar{D}_{1} do not contain the center of D2D_{2} and D¯2\bar{D}_{2} do not contain the center of D1D_{1}.

Given a cellular decomposition Σ\Sigma with the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} on the closed topological surface SS, then we can define the moduli space 𝒯\mathcal{T} of all spherical conical metrics with circle patterns on SS which realize the weighted graph (Σ,Φ)(\Sigma,\Phi), i.e.

𝒯:={(μ,𝒫)μ is a spherical conical metric on S;𝒫 is a circle pattern on (S,μ) such that G𝒫 is isotopic toΣand realizes the weightΦ}/,\mathcal{T}:=\left\{\begin{array}[]{l|l}(\mu,\mathcal{P})&\begin{array}[]{l}\mu\text{ is a spherical conical metric on }S;\\ \mathcal{P}\text{ is a circle pattern on }(S,\mu)\text{ such that }G_{\mathcal{P}}\\ \text{ is isotopic to}\leavevmode\nobreak\ \Sigma\leavevmode\nobreak\ \text{and realizes the weight}\leavevmode\nobreak\ \Phi\end{array}\end{array}\right\}/\sim,

where (μ1,𝒫1)(μ2,𝒫2)(\mu_{1},\mathcal{P}_{1})\sim(\mu_{2},\mathcal{P}_{2}) if and only if there exists an isometry f:SSf:S\to S such that fμ2=μ1,f𝒫2=𝒫1f^{*}\mu_{2}=\mu_{1},f^{*}\mathcal{P}_{2}=\mathcal{P}_{1} and ff is isotopic to the identity map. Then we can define a map κ\kappa from (0,π2]|F|(0,\frac{\pi}{2}]^{|F|} to 𝒯\mathcal{T}. By Lemma 3.1, the map κ\kappa is well-defined. Besides, we have that (0,π2]|F|𝒯(0,\frac{\pi}{2}]^{|F|}\cong\mathcal{T}, i.e.

Proposition 3.3.

Given the weighted graph (Σ,Φ)(\Sigma,\Phi) on SS, then the map κ\kappa

κ:(0,π2]|F|𝒯𝐫=(r1,,r|F|)T[(μ,𝒫)]\begin{array}[]{cccc}\kappa:(0,\frac{\pi}{2}]^{|F|}&\longrightarrow&\mathcal{T}\\ \mathbf{r}=(r_{1},\cdots,r_{|F|})^{T}&\longmapsto&[(\mu,\mathcal{P})]\\ \end{array}

is bijective.

Proof.

If κ(𝐫1)=κ(𝐫2)\kappa(\mathbf{r}_{1})=\kappa(\mathbf{r}_{2}), let κ(𝐫i)=[(μi,𝒫i)],i=1,2\kappa(\mathbf{r}_{i})=[(\mu_{i},\mathcal{P}_{i})],i=1,2. Then we know that (μ1,𝒫1)(μ2,𝒫2)(\mu_{1},\mathcal{P}_{1})\sim(\mu_{2},\mathcal{P}_{2}). By definition, the radii of the corresponding disks in circle patterns 𝒫1\mathcal{P}_{1} and 𝒫2\mathcal{P}_{2} are equal, i.e. 𝐫1=𝐫2\mathbf{r}_{1}=\mathbf{r}_{2}. The map κ\kappa is injective.

For any [(μ,𝒫)]𝒯[(\mu,\mathcal{P})]\in\mathcal{T}, there exists a corresponding radii 𝐫(0,π2]|F|\mathbf{r}\in(0,\frac{\pi}{2}]^{|F|}. Given the weight ϕ\phi, by Lemma 3.1, the local spherical metric generated by 𝐫\mathbf{r} is isometric to the corresponding local metric of (μ,𝒫)(\mu,\mathcal{P}). Hence the spherical metric generated by 𝐫\mathbf{r} is isometric to the metric of (μ,𝒫)(\mu,\mathcal{P}), i.e. κ(𝐫)=[(μ,𝒫)]\kappa(\mathbf{r})=[(\mu,\mathcal{P})]. The map κ\kappa is surjective. ∎

By Proposition 3.3, we know that (0,π2]|F|𝒯0|F|(0,\frac{\pi}{2}]^{|F|}\cong\mathcal{T}\cong\mathbb{R}^{|F|}_{\geq 0}.

4 Potential functions on bigons

Given ϕ(0,π2)\phi\in(0,\frac{\pi}{2}), we consider two disks D1,D2𝕊2D_{1},D_{2}\subset\mathbb{S}^{2}. Suppose that the radius of disk DiD_{i} is ri(0,π2]r_{i}\in(0,\frac{\pi}{2}], i=1,2i=1,2. Besides, let the intersection angle of D1D_{1} and D2D_{2} be ϕ\phi. By BB we denote the bigon which is formed by the intersection of disks D1D_{1} and D2D_{2}. Let i\ell_{i} denote the length of the sides of bigon BB, i=1,2i=1,2. By kik_{i} we denote the geodesic curvature of Di\partial D_{i}, i.e. ki=cotrik_{i}=\cot r_{i}, i=1,2i=1,2. By LiL_{i} we denote the total geodesic curvature of the sides of bigon BB, i.e. Li=ikiL_{i}=\ell_{i}k_{i}, i=1,2i=1,2 (see Figure 6).

Refer to caption
Figure 6: The disk D1,D2𝕊2D_{1},D_{2}\subset\mathbb{S}^{2}
Lemma 4.1.

Given ϕ(0,π2)\phi\in(0,\frac{\pi}{2}), then the 1-form ωϕ=1dk1+2dk2\omega_{\phi}=\ell_{1}dk_{1}+\ell_{2}dk_{2} is closed on >02\mathbb{R}^{2}_{>0}, >0×{k2=0}\mathbb{R}_{>0}\times\{k_{2}=0\} and {k1=0}×>0\{k_{1}=0\}\times\mathbb{R}_{>0}.

Proof.

By Lemma 8 in [6], the 1-form ωϕ\omega_{\phi} is closed on >02\mathbb{R}^{2}_{>0}. On >0×{k2=0}\mathbb{R}_{>0}\times\{k_{2}=0\}, ωϕ=1dk1\omega_{\phi}=\ell_{1}dk_{1}. Since 1\ell_{1} is a smooth function with respect to k1k_{1}, dωϕ=0d\omega_{\phi}=0, ωϕ\omega_{\phi} is closed. Similarly, we know that ωϕ\omega_{\phi} is also closed on {k1=0}×>0\{k_{1}=0\}\times\mathbb{R}_{>0}. ∎

Using change of variables Ki=lnkiK_{i}=\ln k_{i}, i=1,2i=1,2, by Lemma 4.1 the 1-form ωϕ=L1dK1+L2dK2\omega_{\phi}=L_{1}dK_{1}+L_{2}dK_{2} is closed on 2\mathbb{R}^{2}, ωϕ=L1dK1\omega_{\phi}=L_{1}dK_{1} is closed on ×{}\mathbb{R}\times\{-\infty\} and ωϕ=L2dK2\omega_{\phi}=L_{2}dK_{2} is closed on {}×\{-\infty\}\times\mathbb{R}. On 2\mathbb{R}^{2}, we can define a potential function

Λϕ(K1,K2):=0(K1,K2)ωϕ.\Lambda_{\phi}(K_{1},K_{2}):=\int_{0}^{\left(K_{1},K_{2}\right)}\omega_{\phi}.

On ×{}\mathbb{R}\times\{-\infty\}, we define a potential function

Λϕ(K1):=0K1ωϕ.\Lambda_{\phi}(K_{1}):=\int_{0}^{K_{1}}\omega_{\phi}.

On {}×\{-\infty\}\times\mathbb{R}, we define a potential function

Λϕ(K2):=0K2ωϕ.\Lambda_{\phi}(K_{2}):=\int_{0}^{K_{2}}\omega_{\phi}.

These three functions are well-defined. Besides, we have that Λϕ(K1,K2)=(L1,L2)T\nabla\Lambda_{\phi}(K_{1},K_{2})=(L_{1},L_{2})^{T}, Λϕ(K1)=L1\Lambda^{{}^{\prime}}_{\phi}(K_{1})=L_{1} and Λϕ(K2)=L2\Lambda^{{}^{\prime}}_{\phi}(K_{2})=L_{2}.

Lemma 4.2.

We have that L1K1>0\frac{\partial L_{1}}{\partial K_{1}}>0 on ×{}\mathbb{R}\times\{-\infty\}, L2K2>0\frac{\partial L_{2}}{\partial K_{2}}>0 on {}×\{-\infty\}\times\mathbb{R} and the matrix

(L1K1L1K2L2K1L2K2)\left(\begin{array}[]{ll}\frac{\partial L_{1}}{\partial K_{1}}&\frac{\partial L_{1}}{\partial K_{2}}\\ \frac{\partial L_{2}}{\partial K_{1}}&\frac{\partial L_{2}}{\partial K_{2}}\end{array}\right)

is positive definite on 2\mathbb{R}^{2}.

Proof.

By Lemma 9 in [6], we know that on 2\mathbb{R}^{2}, the matrix

(L1K1L1K2L2K1L2K2)\left(\begin{array}[]{ll}\frac{\partial L_{1}}{\partial K_{1}}&\frac{\partial L_{1}}{\partial K_{2}}\\ \frac{\partial L_{2}}{\partial K_{1}}&\frac{\partial L_{2}}{\partial K_{2}}\end{array}\right)

is positive definite.

On ×{}\mathbb{R}\times\{-\infty\}, since k2=0k_{2}=0, L2=0L_{2}=0. By Gauss-Bonnet formula, we have

Area(B)=2ϕL1.\text{Area}(B)=2\phi-L_{1}.

By Lemma 9 in [6], we know that Area(B)K1<0\frac{\partial\text{Area}(B)}{\partial K_{1}}<0, then L1K1>0\frac{\partial L_{1}}{\partial K_{1}}>0. Similarly, we know that L2K2>0\frac{\partial L_{2}}{\partial K_{2}}>0 on {}×\{-\infty\}\times\mathbb{R}. ∎

By Lemma 4.2, Λϕ(K1,K2)\Lambda_{\phi}(K_{1},K_{2}) is a strictly convex function on 2\mathbb{R}^{2}, Λϕ(K1)\Lambda_{\phi}(K_{1}) is a strictly convex function on ×{}\mathbb{R}\times\{-\infty\} and Λϕ(K2)\Lambda_{\phi}(K_{2}) is a strictly convex function on {}×\{-\infty\}\times\mathbb{R}.

We can define the moduli space ϕ\mathcal{B}_{\phi} of spherical bigons with given angle ϕ(0,π2)\phi\in(0,\frac{\pi}{2}), i.e.

ϕ:={BB is a spherical bigon which is formed bythe intersection of two disksD1,D2𝕊2with given intersection angleϕ(0,π2),wherethe radiir1,r2 of the two disks are in (0,π2]}/,\mathcal{B}_{\phi}:=\left\{\begin{array}[]{l|l}B&\begin{array}[]{l}B\text{ is a spherical bigon which is formed by}\\ \text{the intersection of two disks}\leavevmode\nobreak\ D_{1},D_{2}\subset\mathbb{S}^{2}\leavevmode\nobreak\ \text{with}\\ \text{ given intersection angle}\leavevmode\nobreak\ \phi\in(0,\frac{\pi}{2}),\leavevmode\nobreak\ \text{where}\\ \text{the radii}\leavevmode\nobreak\ r_{1},r_{2}\leavevmode\nobreak\ \text{ of the two disks are in }\leavevmode\nobreak\ (0,\frac{\pi}{2}]\\ \end{array}\end{array}\right\}/\sim,

where B1B2B_{1}\sim B_{2} if and only if B1B_{1} and B2B_{2} are isomorphic.

Then we can construct a map γ\gamma from ϕ\mathcal{B}_{\phi} to (0,π2]2(0,\frac{\pi}{2}]^{2}, i.e.

γ:ϕ(0,π2]2[B](r1,r2),\begin{aligned} \gamma:\mathcal{B}_{\phi}&\longrightarrow(0,\frac{\pi}{2}]^{2}\\ {[B]}&\longmapsto\left(r_{1},r_{2}\right)\end{aligned},

The map γ\gamma is well-defined. Besides, by Lemma 3.1, γ\gamma is a bijective map, ϕ(0,π2]2\mathcal{B}_{\phi}\cong(0,\frac{\pi}{2}]^{2}.

Then we define some subspaces of the moduli space ϕ\mathcal{B}_{\phi}, i.e.

ϕ0:={BB is a spherical bigon which is formed bythe intersection of two disksD1,D2𝕊2with given intersection angleϕ(0,π2),wherethe radiir1,r2 of the two disks areπ2}/,\mathcal{B}^{0}_{\phi}:=\left\{\begin{array}[]{l|l}B&\begin{array}[]{l}B\text{ is a spherical bigon which is formed by}\\ \text{the intersection of two disks}\leavevmode\nobreak\ D_{1},D_{2}\subset\mathbb{S}^{2}\leavevmode\nobreak\ \text{with}\\ \text{ given intersection angle}\leavevmode\nobreak\ \phi\in(0,\frac{\pi}{2}),\leavevmode\nobreak\ \text{where}\\ \text{the radii}\leavevmode\nobreak\ r_{1},r_{2}\leavevmode\nobreak\ \text{ of the two disks are}\leavevmode\nobreak\ \frac{\pi}{2}\\ \end{array}\end{array}\right\}/\sim,
ϕ1:={BB is a spherical bigon which is formed bythe intersection of two disks D1,D2𝕊2with given intersection angleϕ(0,π2),wherethe radiir1,r2 of the two disks satisfyr1(0,π2),r2=π2}/,\mathcal{B}^{1}_{\phi}:=\left\{\begin{array}[]{l|l}B&\begin{array}[]{l}B\text{ is a spherical bigon which is formed by}\\ \text{the intersection of two disks }\leavevmode\nobreak\ D_{1},D_{2}\subset\mathbb{S}^{2}\leavevmode\nobreak\ \text{with}\\ \text{ given intersection angle}\leavevmode\nobreak\ \phi\in(0,\frac{\pi}{2}),\leavevmode\nobreak\ \text{where}\\ \text{the radii}\leavevmode\nobreak\ r_{1},r_{2}\leavevmode\nobreak\ \text{ of the two disks satisfy}\\ r_{1}\in(0,\frac{\pi}{2}),r_{2}=\frac{\pi}{2}\\ \end{array}\end{array}\right\}/\sim,
ϕ2:={BB is a spherical bigon which is formed bythe intersection of two disks D1,D2𝕊2with given intersection angleϕ(0,π2),wherethe radiir1,r2 of the two disks satisfyr1=π2,r2(0,π2)}/,\mathcal{B}^{2}_{\phi}:=\left\{\begin{array}[]{l|l}B&\begin{array}[]{l}B\text{ is a spherical bigon which is formed by}\\ \text{the intersection of two disks }\leavevmode\nobreak\ D_{1},D_{2}\subset\mathbb{S}^{2}\leavevmode\nobreak\ \text{with}\\ \text{ given intersection angle}\leavevmode\nobreak\ \phi\in(0,\frac{\pi}{2}),\leavevmode\nobreak\ \text{where}\\ \text{the radii}\leavevmode\nobreak\ r_{1},r_{2}\leavevmode\nobreak\ \text{ of the two disks satisfy}\\ r_{1}=\frac{\pi}{2},r_{2}\in(0,\frac{\pi}{2})\\ \end{array}\end{array}\right\}/\sim,
ϕ3:={BB is a spherical bigon which is formed bythe intersection of two disks D1,D2𝕊2with given intersection angleϕ(0,π2),wherethe radiir1,r2 of the two disks are in (0,π2)}/,\mathcal{B}^{3}_{\phi}:=\left\{\begin{array}[]{l|l}B&\begin{array}[]{l}B\text{ is a spherical bigon which is formed by}\\ \text{the intersection of two disks }\leavevmode\nobreak\ D_{1},D_{2}\subset\mathbb{S}^{2}\leavevmode\nobreak\ \text{with}\\ \text{ given intersection angle}\leavevmode\nobreak\ \phi\in(0,\frac{\pi}{2}),\leavevmode\nobreak\ \text{where}\\ \text{the radii}\leavevmode\nobreak\ r_{1},r_{2}\leavevmode\nobreak\ \text{ of the two disks are in }\leavevmode\nobreak\ (0,\frac{\pi}{2})\\ \end{array}\end{array}\right\}/\sim,

It is easy to know that the subspace ϕ0,ϕ1,ϕ2\mathcal{B}^{0}_{\phi},\mathcal{B}^{1}_{\phi},\mathcal{B}^{2}_{\phi} and ϕ3\mathcal{B}^{3}_{\phi} are mutually disjoint. Besides, we have that ϕ=ϕ0ϕ1ϕ2ϕ3.\mathcal{B}_{\phi}=\mathcal{B}^{0}_{\phi}\sqcup\mathcal{B}^{1}_{\phi}\sqcup\mathcal{B}^{2}_{\phi}\sqcup\mathcal{B}^{3}_{\phi}. Similarly, we have that ϕ0{(π2,π2)},ϕ1(0,π2)×{π2},ϕ2{π2}×(0,π2)\mathcal{B}^{0}_{\phi}\cong\{(\frac{\pi}{2},\frac{\pi}{2})\},\mathcal{B}^{1}_{\phi}\cong(0,\frac{\pi}{2})\times\{\frac{\pi}{2}\},\mathcal{B}^{2}_{\phi}\cong\{\frac{\pi}{2}\}\times(0,\frac{\pi}{2}) and ϕ3(0,π2)2\mathcal{B}^{3}_{\phi}\cong(0,\frac{\pi}{2})^{2}.

Lemma 4.3.

We have that ×{}Δ1,{}×Δ2\mathbb{R}\times\{-\infty\}\cong\Delta_{1},\{-\infty\}\times\mathbb{R}\cong\Delta_{2}, where Δ1={(x,0)| 0<x<2ϕ},Δ2={(0,y)| 0<y<2ϕ}\Delta_{1}=\{(x,0)\leavevmode\nobreak\ |\leavevmode\nobreak\ 0<x<2\phi\},\Delta_{2}=\{(0,y)\leavevmode\nobreak\ |\leavevmode\nobreak\ 0<y<2\phi\}.

Proof.

On ×{}\mathbb{R}\times\{-\infty\}, by Lemma 4.2, it is easy to know that Λϕ(K1)\Lambda^{{}^{\prime}}_{\phi}(K_{1}) is a embedding map. For the map Λϕ(K1)\Lambda^{{}^{\prime}}_{\phi}(K_{1}), we have

Λϕ(K1):\displaystyle\Lambda^{{}^{\prime}}_{\phi}(K_{1}):\mathbb{R} >0\displaystyle\longrightarrow\mathbb{R}_{>0}
K1\displaystyle K_{1} L1(K1).\displaystyle\longmapsto L_{1}(K_{1}).

On ×{}\mathbb{R}\times\{-\infty\}, L2=0L_{2}=0. By Gauss-Bonnet formula, we have that Area(B)=2ϕL1>0\text{Area}(B)=2\phi-L_{1}>0, i.e. 0<L1<2ϕ0<L_{1}<2\phi. Hence the image set of Λϕ(K1)\Lambda^{{}^{\prime}}_{\phi}(K_{1}) is contained in Δ~1\tilde{\Delta}_{1}, where Δ~1={x| 0<x<2ϕ}\tilde{\Delta}_{1}=\{x\leavevmode\nobreak\ |\leavevmode\nobreak\ 0<x<2\phi\}. Then we will show that 0 and 2ϕ2\phi are the limit points of the map Λϕ(K1)\Lambda^{{}^{\prime}}_{\phi}(K_{1}).

We can choose a sequence {n}n+\{n\}_{n\in\mathbb{N}_{+}}, then k1(n)=en+(n+),r1(n)=arccotk1(n)0(n+)k_{1}(n)=e^{n}\to+\infty\leavevmode\nobreak\ (n\to+\infty),r_{1}(n)=\operatorname{arccot}k_{1}(n)\to 0\leavevmode\nobreak\ (n\to+\infty). Hence we have that Area(B(n))0(n+)\text{Area}(B(n))\to 0\leavevmode\nobreak\ (n\to+\infty). On ×{}\mathbb{R}\times\{-\infty\}, L2(n)=0L_{2}(n)=0. By Gauss-Bonnet formula, we have that Area(B(n))=2ϕL1(n)0(n+)\text{Area}(B(n))=2\phi-L_{1}(n)\to 0\leavevmode\nobreak\ (n\to+\infty), i.e. L1(n)2ϕ(n+)L_{1}(n)\to 2\phi\leavevmode\nobreak\ (n\to+\infty).

We choose another sequence {n}n+\{-n\}_{n\in\mathbb{N}_{+}}, then k1(n)=enk_{1}(-n)=e^{-n}. Since 1(n)π\ell_{1}(-n)\leq\pi, then 0L1(n)=1(n)k1(n)πen0(n+)0\leq L_{1}(-n)=\ell_{1}(-n)k_{1}(-n)\leq\pi e^{-n}\to 0\leavevmode\nobreak\ (n\to+\infty). Hence we have that L1(n)0(n+)L_{1}(-n)\to 0\leavevmode\nobreak\ (n\to+\infty).

Since 0 and 2ϕ2\phi are the limit points of the map Λϕ(K1)\Lambda^{{}^{\prime}}_{\phi}(K_{1}) and Λϕ(K1)\Lambda^{{}^{\prime}}_{\phi}(K_{1}) is a embedding map, it is easy to know that Λϕ(K1)\Lambda^{{}^{\prime}}_{\phi}(K_{1}) is homeomorphism from \mathbb{R} to Δ~1\tilde{\Delta}_{1}, i.e. ×{}Δ1\mathbb{R}\times\{-\infty\}\cong\Delta_{1}.

Using the above same argument, we know that {}×Δ2\{-\infty\}\times\mathbb{R}\cong\Delta_{2}. ∎

By Lemma 4.3, we have that ϕ1(0,π2)×{π2}×{}Δ1\mathcal{B}^{1}_{\phi}\cong(0,\frac{\pi}{2})\times\{\frac{\pi}{2}\}\cong\mathbb{R}\times\{-\infty\}\cong\Delta_{1} and ϕ2{π2}×(0,π2){}×Δ2\mathcal{B}^{2}_{\phi}\cong\{\frac{\pi}{2}\}\times(0,\frac{\pi}{2})\cong\{-\infty\}\times\mathbb{R}\cong\Delta_{2}. By Lemma 11 in [6], ϕ3(0,π2)22Δ3\mathcal{B}^{3}_{\phi}\cong(0,\frac{\pi}{2})^{2}\cong\mathbb{R}^{2}\cong\Delta_{3}, where Δ3={(x,y)>02|x+y<2ϕ}\Delta_{3}=\{(x,y)\in\mathbb{R}^{2}_{>0}\leavevmode\nobreak\ |\leavevmode\nobreak\ x+y<2\phi\}. Besides, we have ϕ0{(π2,π2)}{(0,0)}\mathcal{B}^{0}_{\phi}\cong\{(\frac{\pi}{2},\frac{\pi}{2})\}\cong\{(0,0)\}.

5 Existence and rigidity of circle pattern metrics for prescribed total geodesic curvatures

Given a closed topological surface SS and a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F) of SS. For simplicity, we can suppose that F={1,,|F|}F=\{1,\cdots,|F|\}. By Lf,eL_{f,e} we denote the total geodesic curvature of an arc Cf,eC_{f,e}, where the arc Cf,eC_{f,e} is the side of the bigon corresponding to 2-cell fFf\in F and 1-cell eEe\in E. By BeB_{e} we denote the bigon corresponding to ee (see Figure 7).

Refer to caption
Figure 7: The position of f,e,Bef,e,B_{e} and Cf,eC_{f,e}

Then we can define the total geodesic curvature LfL_{f} of a 2-cell ff, i.e. Lf:=eEfLf,eL_{f}:=\sum_{e\in E_{f}}L_{f,e}, where EfE_{f} is the set of 1-cells contained in f\partial f.

Given the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|}, for 𝐫=(r1,,r|F|)(0,π2)|F|\mathbf{r}=(r_{1},\cdots,r_{|F|})\in(0,\frac{\pi}{2})^{|F|}, using change of variables Ki=lncotriK_{i}=\ln\cot r_{i}, i=1,,|F|i=1,\cdots,|F|, we consider the new variable K=(K1,,K|F|)|F|K=(K_{1},\cdots,K_{|F|})\in\mathbb{R}^{|F|}.

We can define a potential function

Λ(K):=eEΛΦ(e)(Kf1(e),Kf2(e))\Lambda(K):=\sum_{e\in E}\Lambda_{\Phi(e)}(K_{f_{1}(e)},K_{f_{2}(e)})

on |F|\mathbb{R}^{|F|}, where f1(e)f_{1}(e) and f2(e)f_{2}(e) are the 2-cells on two sides of 1-cell ee. Besides, we have

(5.1) Λ(K)Ki=eEiΛΦ(e)(Kf1(e),Kf2(e))Ki=eEiLi,e=Li>0, 1i|F|.\displaystyle\frac{\partial\Lambda(K)}{\partial K_{i}}=\sum_{e\in E_{i}}\frac{\partial\Lambda_{\Phi(e)}(K_{f_{1}(e)},K_{f_{2}(e)})}{\partial K_{i}}=\sum_{e\in E_{i}}L_{i,e}=L_{i}>0,\leavevmode\nobreak\ 1\leq i\leq|F|.

Hence the Hessian of Λ\Lambda is a Jacobi matrix, i.e.

HessΛ=M=(L1K1L1K|F|L|F|K1L|F|K|F|).\text{Hess}\leavevmode\nobreak\ \Lambda=M=\begin{pmatrix}\frac{\partial L_{1}}{\partial K_{1}}&\cdots&\frac{\partial L_{1}}{\partial K_{|F|}}\\ \vdots&\ddots&\vdots\\ \frac{\partial L_{|F|}}{\partial K_{1}}&\cdots&\frac{\partial L_{|F|}}{\partial K_{|F|}}\\ \end{pmatrix}.
Proposition 5.1.

The Jacobi matrix MM is positive definite.

Proof.

By Lemma 9 in [6], we have that

LiKi=(eEiLi,e)Ki=eEiLi,eKi>0,i=1,,|F|,\frac{\partial L_{i}}{\partial K_{i}}=\frac{\partial(\sum_{e\in E_{i}}L_{i,e})}{\partial K_{i}}=\sum_{e\in E_{i}}\frac{\partial L_{i,e}}{\partial K_{i}}>0,\leavevmode\nobreak\ i=1,\cdots,|F|,

and

LiKj=(eEiLi,e)Kj=eEiLi,eKj0, 1ij|F|.\frac{\partial L_{i}}{\partial K_{j}}=\frac{\partial(\sum_{e\in E_{i}}L_{i,e})}{\partial K_{j}}=\sum_{e\in E_{i}}\frac{\partial L_{i,e}}{\partial K_{j}}\leq 0,\leavevmode\nobreak\ 1\leq i\neq j\leq|F|.

Hence we obtain

|LiKi|ji|LjKi|=LiKi+jiLjKi=(fFLf)Ki.\left|\frac{\partial L_{i}}{\partial K_{i}}\right|-\sum_{j\neq i}\left|\frac{\partial L_{j}}{\partial K_{i}}\right|=\frac{\partial L_{i}}{\partial K_{i}}+\sum_{j\neq i}\frac{\partial L_{j}}{\partial K_{i}}=\frac{\partial(\sum_{f\in F}L_{f})}{\partial K_{i}}.

Besides, we have

fFLf=fFeEfLf,e=eEfF{e}Lf,e,\sum_{f\in F}L_{f}=\sum_{f\in F}\sum_{e\in E_{f}}L_{f,e}=\sum_{e\in E}\sum_{f\in F_{\{}e\}}L_{f,e},

where F{e}F_{\{e\}} is the set of 2-cells on two sides of 1-cell ee. Then by Gauss-Bonnet formula, we have

fF{e}Lf,e=2Φ(e)Area(Be).\sum_{f\in F_{\{}e\}}L_{f,e}=2\Phi(e)-\text{Area}(B_{e}).

Hence we obtain

fFLf=eE2Φ(e)Area(Be).\sum_{f\in F}L_{f}=\sum_{e\in E}2\Phi(e)-\text{Area}(B_{e}).

By Lemma 9 in [6], we have

(fFLf)Ki=eE(Area(Be))Ki>0,\frac{\partial(\sum_{f\in F}L_{f})}{\partial K_{i}}=-\sum_{e\in E}\frac{\partial(\text{Area}(B_{e}))}{\partial K_{i}}>0,

Then we obtain

|LiKi|ji|LjKi|=(fFLf)Ki>0.\left|\frac{\partial L_{i}}{\partial K_{i}}\right|-\sum_{j\neq i}\left|\frac{\partial L_{j}}{\partial K_{i}}\right|=\frac{\partial(\sum_{f\in F}L_{f})}{\partial K_{i}}>0.

Hence MM is a strictly diagonally dominant matrix with positive diagonal entries, i.e. MM is positive definite. ∎

Corollary 5.2.

The potential function Λ\Lambda is strictly convex on |F|\mathbb{R}^{|F|}.

Proof.

By Proposition 5.1, HessΛ\text{Hess}\leavevmode\nobreak\ \Lambda is positive definite, then Λ\Lambda is strictly convex on |F|\mathbb{R}^{|F|}. ∎

For a subset FFF^{\prime}\subset F, by EFE_{F^{\prime}} we denote the set

EF={eE|fF,s.t.eEf}.E_{F^{\prime}}=\{e\in E\leavevmode\nobreak\ |\leavevmode\nobreak\ \exists f\in F^{\prime},s.t.\leavevmode\nobreak\ e\in E_{f}\}.

Then we have the following theorem, i.e.

Theorem 5.3.

Λ\nabla\Lambda is a homeomorphism from |F|\mathbb{R}^{|F|} to 1\mathcal{L}_{1}, where

1={(L1,,L|F|)T>0|F||fFLf<2eEFΦ(e),FF}.\mathcal{L}_{1}=\{(L_{1},\cdots,L_{|F|})^{T}\in\mathbb{R}^{|F|}_{>0}\leavevmode\nobreak\ |\leavevmode\nobreak\ \sum_{f\in F^{\prime}}L_{f}<2\sum_{e\in E_{F^{\prime}}}\Phi(e),\forall F^{\prime}\subset F\}.
Proof.

By (5.1), we have

Λ:|F|\displaystyle\nabla\Lambda:\mathbb{R}^{|F|} >0|F|\displaystyle\longrightarrow\mathbb{R}^{|F|}_{>0}
K=(K1,,K|F|)T\displaystyle K=(K_{1},\cdots,K_{|F|})^{T} L=(L1,,L|F|)T.\displaystyle\mapsto L=(L_{1},\cdots,L_{|F|})^{T}.

For any subset FFF^{\prime}\subset F, we obtain

fFLf=fFeEfLf,e=eEFfF{e}FLf,e.\sum_{f\in F^{\prime}}L_{f}=\sum_{f\in F^{\prime}}\sum_{e\in E_{f}}L_{f,e}=\sum_{e\in E_{F^{\prime}}}\sum_{f\in F_{\{e\}}\cap F^{\prime}}L_{f,e}.

Then by Gauss-Bonnet formula, we have

Area(Be)=2Φ(e)fF{e}Lf,e>0,\text{Area}(B_{e})=2\Phi(e)-\sum_{f\in F_{\{e\}}}L_{f,e}>0,

i.e. fF{e}Lf,e<2Φ(e)\sum_{f\in F_{\{e\}}}L_{f,e}<2\Phi(e). Hence we obtain

fFLf=eEFfF{e}FLf,eeEFfF{e}Lf,e<2eEFΦ(e),\sum_{f\in F^{\prime}}L_{f}=\sum_{e\in E_{F^{\prime}}}\sum_{f\in F_{\{e\}}\cap F^{\prime}}L_{f,e}\leq\sum_{e\in E_{F^{\prime}}}\sum_{f\in F_{\{e\}}}L_{f,e}<2\sum_{e\in E_{F^{\prime}}}\Phi(e),

hence (L1,,L|F|)T1(L_{1},\cdots,L_{|F|})^{T}\in\mathcal{L}_{1}, Λ\nabla\Lambda is a map from |F|\mathbb{R}^{|F|} to 1\mathcal{L}_{1}.

By Proposition 5.1 and Corollary 5.2, Λ\nabla\Lambda is a embedding map. Hence we only need to show that the image set of Λ\nabla\Lambda is 1\mathcal{L}_{1}. Now we need to analysis the boundary of its image.

Choose a point a=(a1,,a|F|)T|F|a=(a_{1},\cdots,a_{|F|})^{T}\in\partial\mathbb{R}^{|F|}, we define two subset W1,W2FW_{1},W_{2}\subset F, i.e.

W1={iF|ai=+},W2={iF|ai=},W_{1}=\{i\in F\leavevmode\nobreak\ |\leavevmode\nobreak\ a_{i}=+\infty\},\leavevmode\nobreak\ W_{2}=\{i\in F\leavevmode\nobreak\ |\leavevmode\nobreak\ a_{i}=-\infty\},

we know that W1W_{1}\neq\emptyset or W2W_{2}\neq\emptyset.

We choose a sequence {an}n=1+|F|\{a^{n}\}_{n=1}^{+\infty}\subset\mathbb{R}^{|F|} such that limn+an=a\lim_{n\to+\infty}a^{n}=a, where an=(a1n,,a|F|n)Ta^{n}=(a_{1}^{n},\cdots,a_{|F|}^{n})^{T}. Then we have

Λ(an)=L(an)=ΔLn,kin=Δexp(ain),n1.\nabla\Lambda(a^{n})=L(a^{n})\overset{\Delta}{=}L^{n},\leavevmode\nobreak\ k_{i}^{n}\overset{\Delta}{=}\exp(a_{i}^{n}),\leavevmode\nobreak\ \forall n\geq 1.

Now we need to show that {Ln}n=1+\{L^{n}\}_{n=1}^{+\infty} converges to the boundary of 1\mathcal{L}_{1}.

If W1W_{1}\neq\emptyset, we know that for each iW1i\in W_{1}, kin=exp(ain)+(n+)k_{i}^{n}=\exp(a_{i}^{n})\to+\infty\leavevmode\nobreak\ (n\to+\infty), and rin=arccotkin0(n+)r_{i}^{n}=\operatorname{arccot}k_{i}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty). Besides, note that

(5.2) limn+fW1Lfn=limn+fW1eEfLf,en=limn+eEW1fW1F{e}Lf,en.\displaystyle\lim_{n\to+\infty}\sum_{f\in W_{1}}L_{f}^{n}=\lim_{n\to+\infty}\sum_{f\in W_{1}}\sum_{e\in E_{f}}L_{f,e}^{n}=\lim_{n\to+\infty}\sum_{e\in E_{W_{1}}}\sum_{f\in W_{1}\cap F_{\{e\}}}L_{f,e}^{n}.

Now we need to show

(5.3) limn+fW1F{e}Lf,en=2Φ(e).\displaystyle\lim_{n\to+\infty}\sum_{f\in W_{1}\cap F_{\{e\}}}L_{f,e}^{n}=2\Phi(e).

Since W1W_{1}\neq\emptyset and eEW1e\in E_{W_{1}}, the set W1F{e}W_{1}\cap F_{\{e\}} has 1 or 2 elements.

If W1F{e}W_{1}\cap F_{\{e\}} has 1 elements, we can suppose that W1F{e}={h}W_{1}\cap F_{\{e\}}=\{h\} and F{e}={h,j}F_{\{e\}}=\{h,j\}. By Gauss-Bonnet formula, we have

(5.4) Lh,en+Lj,en=2Φ(e)Area(Ben).\displaystyle L_{h,e}^{n}+L_{j,e}^{n}=2\Phi(e)-\text{Area}(B_{e}^{n}).

Since jW1j\notin W_{1}, we know that ajn,kjn+,rjn0(n+)a_{j}^{n},k_{j}^{n}\nrightarrow+\infty,r_{j}^{n}\nrightarrow 0\leavevmode\nobreak\ (n\to+\infty). Since hW1h\in W_{1}, we know that rhn0(n+)r_{h}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty). Then we have that j,en,Area(Ben)0(n+)\ell_{j,e}^{n},\text{Area}(B_{e}^{n})\to 0\leavevmode\nobreak\ (n\to+\infty). Hence we obtain Lj,en=j,enkjn0(n+)L_{j,e}^{n}=\ell_{j,e}^{n}k_{j}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty).

Using (5.4), we have

limn+Lh,en=2Φ(e)limn+Area(Ben)=2Φ(e),\lim_{n\to+\infty}L_{h,e}^{n}=2\Phi(e)-\lim_{n\to+\infty}\text{Area}(B_{e}^{n})=2\Phi(e),

i.e. limn+fW1F{e}Lf,en=2Φ(e)\lim_{n\to+\infty}\sum_{f\in W_{1}\cap F_{\{e\}}}L_{f,e}^{n}=2\Phi(e).

If W1F{e}W_{1}\cap F_{\{e\}} has 2 elements, we can suppose that W1F{e}={m,s}W_{1}\cap F_{\{e\}}=\{m,s\}, then F{e}={m,s}F_{\{e\}}=\{m,s\}. By Gauss-Bonnet formula, we have

(5.5) Lm,en+Ls,en=2Φ(e)Area(Ben).\displaystyle L_{m,e}^{n}+L_{s,e}^{n}=2\Phi(e)-\text{Area}(B_{e}^{n}).

Since m,sW1m,s\in W_{1}, we know that kmn,ksn+(n+)k_{m}^{n},k_{s}^{n}\to+\infty\leavevmode\nobreak\ (n\to+\infty), rmn,rsn0(n+)r_{m}^{n},r_{s}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty). Then we have Area(Ben)0(n+)\text{Area}(B_{e}^{n})\to 0\leavevmode\nobreak\ (n\to+\infty). Using (5.5), we obtain

limn+Lm,en+Ls,en=2Φ(e)limn+Area(Ben)=2Φ(e),\lim_{n\to+\infty}L_{m,e}^{n}+L_{s,e}^{n}=2\Phi(e)-\lim_{n\to+\infty}\text{Area}(B_{e}^{n})=2\Phi(e),

i.e. limn+fW1F{e}Lf,en=2Φ(e)\lim_{n\to+\infty}\sum_{f\in W_{1}\cap F_{\{e\}}}L_{f,e}^{n}=2\Phi(e).

Hence if W1W_{1}\neq\emptyset, by (5.2) and (5.3), we have

limn+fW1Lfn=limn+eEW1fW1F{e}Lf,en=eEW1limn+fW1F{e}Lf,en=2eEW1Φ(e),\lim_{n\to+\infty}\sum_{f\in W_{1}}L_{f}^{n}=\lim_{n\to+\infty}\sum_{e\in E_{W_{1}}}\sum_{f\in W_{1}\cap F_{\{e\}}}L_{f,e}^{n}=\sum_{e\in E_{W_{1}}}\lim_{n\to+\infty}\sum_{f\in W_{1}\cap F_{\{e\}}}L_{f,e}^{n}=2\sum_{e\in E_{W_{1}}}\Phi(e),

i.e. {Ln}n=1+\{L^{n}\}_{n=1}^{+\infty} converges to the boundary of 1\mathcal{L}_{1}.

If W2W_{2}\neq\emptyset, then for each iW2i\in W_{2}, kin=exp(ain)0(n+)k_{i}^{n}=\exp(a_{i}^{n})\to 0\leavevmode\nobreak\ (n\to+\infty). For any 1-cell eEie\in E_{i}, since 0i,en2π0\leq\ell_{i,e}^{n}\leq 2\pi and kin0(n+)k_{i}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty), then limn+Li,en=limn+i,enkin=0\lim_{n\to+\infty}L_{i,e}^{n}=\lim_{n\to+\infty}\ell_{i,e}^{n}k_{i}^{n}=0. Hence we obtain

limn+Lin=limn+eEiLi,en=0,\lim_{n\to+\infty}L_{i}^{n}=\lim_{n\to+\infty}\sum_{e\in E_{i}}L_{i,e}^{n}=0,

i.e. {Ln}n=1+\{L^{n}\}_{n=1}^{+\infty} converges to the boundary of 1\mathcal{L}_{1}.

By Brouwer’s Theorem on the Invariance of Domain and the above analysis, we know that the image set of Λ\nabla\Lambda is 1\mathcal{L}_{1}. Hence Λ\nabla\Lambda is a homeomorphism from |F|\mathbb{R}^{|F|} to 1\mathcal{L}_{1}. ∎

We can construct a map ς\varsigma from >0|F|\mathbb{R}^{|F|}_{>0} to |F|\mathbb{R}^{|F|}, i.e.

ς:>0|F||F|k=(k1,,k|F|)TK=(lnk1,,lnk|F|)T,\begin{array}[]{cccc}\varsigma:\mathbb{R}^{|F|}_{>0}&\longrightarrow&\mathbb{R}^{|F|}\\ k=(k_{1},\cdots,k_{|F|})^{T}&\longmapsto&K=(\ln k_{1},\cdots,\ln k_{|F|})^{T},\\ \end{array}

the map ς\varsigma is a homeomorphism. By Theorem 5.3, we know that the map 1:=Λς\mathcal{E}_{1}:=\nabla\Lambda\circ\varsigma, i.e.

1:>0|F|1k1(k)=Λ(K)=L(K)\begin{array}[]{cccc}\mathcal{E}_{1}:\mathbb{R}^{|F|}_{>0}&\longrightarrow&\mathcal{L}_{1}\\ k&\longmapsto&\mathcal{E}_{1}(k)=\nabla\Lambda(K)=L(K)\\ \end{array}

is a homeomorphism from >0|F|\mathbb{R}^{|F|}_{>0} to 1\mathcal{L}_{1}. Then we completed the proof of Theorem 1.5.

6 Existence and rigidity of degenerated circle pattern metrics for prescribed total geodesic curvatures

By i1im|F|\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}} we denote the set

i1im|F|={(0,,ki1,0,,kim,,0)|F||ki1,,kim>0}.\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}}=\{(0,\cdots,k_{i_{1}},0,\cdots,k_{i_{m}},\cdots,0)\in\mathbb{R}^{|F|}\leavevmode\nobreak\ |\leavevmode\nobreak\ k_{i_{1}},\cdots,k_{i_{m}}>0\}.

By i0|F|\partial_{i}\mathbb{R}^{|F|}_{\geq 0} we denote the set

i0|F|={(k1,,k|F|)|F||icomponents>0,|F|icomponents=0}.\partial_{i}\mathbb{R}^{|F|}_{\geq 0}=\{(k_{1},\cdots,k_{|F|})\in\mathbb{R}^{|F|}\leavevmode\nobreak\ |\leavevmode\nobreak\ i\leavevmode\nobreak\ \text{components}>0,\leavevmode\nobreak\ |F|-i\leavevmode\nobreak\ \text{components}=0\}.

Then we have

10|F|=1i|F|{(0,,ki,,0)|F||ki>0}=1i|F|i|F|,\partial_{1}\mathbb{R}^{|F|}_{\geq 0}=\bigcup_{1\leq i\leq|F|}\{(0,\cdots,k_{i},\cdots,0)\in\mathbb{R}^{|F|}\leavevmode\nobreak\ |\leavevmode\nobreak\ k_{i}>0\}=\bigcup_{1\leq i\leq|F|}\mathbb{R}^{|F|}_{i},
20|F|=1i<j|F|{(0,,ki,,kj,,0)|F||ki,kj>0}=1i<j|F|ij|F|,\partial_{2}\mathbb{R}^{|F|}_{\geq 0}=\bigcup_{1\leq i<j\leq|F|}\{(0,\cdots,k_{i},\cdots,k_{j},\cdots,0)\in\mathbb{R}^{|F|}\leavevmode\nobreak\ |\leavevmode\nobreak\ k_{i},k_{j}>0\}=\bigcup_{1\leq i<j\leq|F|}\mathbb{R}^{|F|}_{ij},
\cdots
|F|10|F|=1i1<<i|F|1|F|{(ki,,0,,ki|F|1)|F||ki1,,ki|F|1>0}=1i1<<i|F|1|F|i1i|F|1|F|.\partial_{|F|-1}\mathbb{R}^{|F|}_{\geq 0}=\bigcup_{1\leq i_{1}<\cdots<i_{|F|-1}\leq|F|}\{(k_{i},\cdots,0,\cdots,k_{i_{|F|-1}})\in\mathbb{R}^{|F|}\leavevmode\nobreak\ |\leavevmode\nobreak\ k_{i_{1}},\cdots,k_{i_{|F|-1}}>0\}=\bigcup_{1\leq i_{1}<\cdots<i_{|F|-1}\leq|F|}\mathbb{R}^{|F|}_{i_{1}\cdots i_{|F|-1}}.

For 1m|F|11\leq m\leq|F|-1, we have that

m0|F|=1i1<<im|F|i1im|F|.\partial_{m}\mathbb{R}^{|F|}_{\geq 0}=\bigcup_{1\leq i_{1}<\cdots<i_{m}\leq|F|}\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}}.

By ¯0|F|\bar{\partial}\mathbb{R}^{|F|}_{\geq 0} we denote the set

¯0|F|=m=1|F|1m0|F|{0}.\bar{\partial}\mathbb{R}^{|F|}_{\geq 0}=\bigcup_{m=1}^{|F|-1}\partial_{m}\mathbb{R}^{|F|}_{\geq 0}\cup\{0\}.

Then we consider the set 1m|F|\mathbb{R}^{|F|}_{1\cdots m}, i.e.

1m|F|={(k1,,km,0,,0)|F||k1,,km>0}.\mathbb{R}^{|F|}_{1\cdots m}=\{(k_{1},\cdots,k_{m},0,\cdots,0)\in\mathbb{R}^{|F|}\leavevmode\nobreak\ |\leavevmode\nobreak\ k_{1},\cdots,k_{m}>0\}.

Given the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|}, for 𝐫=(r1,r2,,rm,π2,,π2)(0,π2)m×{π2}|F|m\mathbf{r}=(r_{1},r_{2},\cdots,r_{m},\frac{\pi}{2},\cdots,\frac{\pi}{2})\in(0,\frac{\pi}{2})^{m}\times\{\frac{\pi}{2}\}^{|F|-m}, using change of variables Ki=lncotriK_{i}=\ln\cot r_{i}, i=1,,mi=1,\cdots,m, we consider the new variable K=(K1,,Km)mK=(K_{1},\cdots,K_{m})\in\mathbb{R}^{m}.

We can define a potential function

Λ1(K):=eE,f1(e),f2(e)FmΛΦ(e)(Kf1(e),Kf2(e))+eE,f1(e)Fm,f2(e)FmΛΦ(e)(Kf1(e))\Lambda_{1}(K):=\sum_{e\in E,f_{1}(e),f_{2}(e)\in F_{m}}\Lambda_{\Phi(e)}(K_{f_{1}(e)},K_{f_{2}(e)})+\sum_{e\in E,f_{1}(e)\in F_{m},f_{2}(e)\notin F_{m}}\Lambda_{\Phi(e)}(K_{f_{1}(e)})

on m\mathbb{R}^{m}, where Fm={1,2,,m}F_{m}=\{1,2,\cdots,m\}, f1(e)f_{1}(e) and f2(e)f_{2}(e) are the 2-cells on two sides of 1-cell ee. Besides, note that at least one of these two sums exists.

For 1im1\leq i\leq m, we have

Λ1(K)Ki\displaystyle\frac{\partial\Lambda_{1}(K)}{\partial K_{i}} =eEi,f1(e),f2(e)FmΛΦ(e)(Kf1(e),Kf2(e))Ki+eEi,f1(e)Fm,f2(e)FmΛΦ(e)(Kf1(e))Ki\displaystyle=\sum_{e\in E_{i},f_{1}(e),f_{2}(e)\in F_{m}}\frac{\partial\Lambda_{\Phi(e)}(K_{f_{1}(e)},K_{f_{2}(e)})}{\partial K_{i}}+\sum_{e\in E_{i},f_{1}(e)\in F_{m},f_{2}(e)\notin F_{m}}\frac{\partial\Lambda_{\Phi(e)}(K_{f_{1}(e)})}{\partial K_{i}}
=eEi,f1(e),f2(e)FmLi,e(Kf1(e),Kf2(e))+eEi,f1(e)Fm,f2(e)FmLi,e(Ki).\displaystyle=\sum_{e\in E_{i},f_{1}(e),f_{2}(e)\in F_{m}}L_{i,e}(K_{f_{1}(e)},K_{f_{2}(e)})+\sum_{e\in E_{i},f_{1}(e)\in F_{m},f_{2}(e)\notin F_{m}}L_{i,e}(K_{i}).

Since iFmi\in F_{m}, then for any 1-cell eEie\in E_{i}, we know that f1(e),f2(e)Fmf_{1}(e),f_{2}(e)\in F_{m} or f1(e)Fm,f2(e)Fmf_{1}(e)\in F_{m},f_{2}(e)\notin F_{m}. Hence we have

(6.1) Li=eEiLi,e=eEi,f1(e),f2(e)FmLi,e(Kf1(e),Kf2(e))+eEi,f1(e)Fm,f2(e)FmLi,e(Ki)=Λ1(K)Ki.\displaystyle L_{i}=\sum_{e\in E_{i}}L_{i,e}=\sum_{e\in E_{i},f_{1}(e),f_{2}(e)\in F_{m}}L_{i,e}(K_{f_{1}(e)},K_{f_{2}(e)})+\sum_{e\in E_{i},f_{1}(e)\in F_{m},f_{2}(e)\notin F_{m}}L_{i,e}(K_{i})=\frac{\partial\Lambda_{1}(K)}{\partial K_{i}}.

Then the Hessian of Λ1\Lambda_{1} is a Jacobi matrix, i.e.

HessΛ1=M1=(L1K1L1KmLmK1LmKm).\text{Hess}\leavevmode\nobreak\ \Lambda_{1}=M_{1}=\begin{pmatrix}\frac{\partial L_{1}}{\partial K_{1}}&\cdots&\frac{\partial L_{1}}{\partial K_{m}}\\ \vdots&\ddots&\vdots\\ \frac{\partial L_{m}}{\partial K_{1}}&\cdots&\frac{\partial L_{m}}{\partial K_{m}}\\ \end{pmatrix}.
Proposition 6.1.

The Jacobi matrix M1M_{1} is positive definite.

Proof.

For any 1im1\leq i\leq m, by (6.1), we have

LiKi=eEi,f1(e),f2(e)FmLi,e(Kf1(e),Kf2(e))Ki+eEi,f1(e)Fm,f2(e)FmLi,e(Ki)Ki.\frac{\partial L_{i}}{\partial K_{i}}=\sum_{e\in E_{i},f_{1}(e),f_{2}(e)\in F_{m}}\frac{\partial L_{i,e}(K_{f_{1}(e)},K_{f_{2}(e)})}{\partial K_{i}}+\sum_{e\in E_{i},f_{1}(e)\in F_{m},f_{2}(e)\notin F_{m}}\frac{\partial L_{i,e}(K_{i})}{\partial K_{i}}.

By Lemma 9 in [6] and Lemma 4.2, we have Li,e(Kf1(e),Kf2(e))Ki>0,Li,e(Ki)Ki>0\frac{\partial L_{i,e}(K_{f_{1}(e)},K_{f_{2}(e)})}{\partial K_{i}}>0,\frac{\partial L_{i,e}(K_{i})}{\partial K_{i}}>0. Besides, at least one of these two sums exists, then we obtain that LiKi>0\frac{\partial L_{i}}{\partial K_{i}}>0 for 1im1\leq i\leq m.

For 1ijm1\leq i\neq j\leq m, by (6.1), we have

LiKj=eEi,f1(e),f2(e)FmLi,e(Kf1(e),Kf2(e))Kj.\frac{\partial L_{i}}{\partial K_{j}}=\sum_{e\in E_{i},f_{1}(e),f_{2}(e)\in F_{m}}\frac{\partial L_{i,e}(K_{f_{1}(e)},K_{f_{2}(e)})}{\partial K_{j}}.

For any 1-cell eEie\in E_{i} such that f1(e),f2(e)Fmf_{1}(e),f_{2}(e)\in F_{m}, we know that iF{e}i\in F_{\{e\}}. If jF{e}j\in F_{\{e\}}, by Lemma 9 in [6], then we have

Li,e(Kf1(e),Kf2(e))Kj=Li,e(Ki,Kj)Kj<0.\frac{\partial L_{i,e}(K_{f_{1}(e)},K_{f_{2}(e)})}{\partial K_{j}}=\frac{\partial L_{i,e}(K_{i},K_{j})}{\partial K_{j}}<0.

If jF{e}j\notin F_{\{e\}}, then we have

Li,e(Kf1(e),Kf2(e))Kj=0.\frac{\partial L_{i,e}(K_{f_{1}(e)},K_{f_{2}(e)})}{\partial K_{j}}=0.

Then we obtain LiKj0\frac{\partial L_{i}}{\partial K_{j}}\leq 0 for 1ijm1\leq i\neq j\leq m.

Hence for 1im1\leq i\leq m, we have

|LiKi|1jim|LjKi|\displaystyle\left|\frac{\partial L_{i}}{\partial K_{i}}\right|-\sum_{1\leq j\neq i\leq m}\left|\frac{\partial L_{j}}{\partial K_{i}}\right| =(fFmLf)Ki=(fFmeEfLf,e)Ki=(eEFmfFmF{e}Lf,e)Ki\displaystyle=\frac{\partial(\sum_{f\in F_{m}}L_{f})}{\partial K_{i}}=\frac{\partial(\sum_{f\in F_{m}}\sum_{e\in E_{f}}L_{f,e})}{\partial K_{i}}=\frac{\partial(\sum_{e\in E_{F_{m}}}\sum_{f\in F_{m}\cap F_{\{e\}}}L_{f,e})}{\partial K_{i}}
=eEFm(fFmF{e}Lf,e)Ki=eEi(fFmF{e}Lf,e)Ki.\displaystyle=\sum_{e\in E_{F_{m}}}\frac{\partial(\sum_{f\in F_{m}\cap F_{\{e\}}}L_{f,e})}{\partial K_{i}}=\sum_{e\in E_{i}}\frac{\partial(\sum_{f\in F_{m}\cap F_{\{e\}}}L_{f,e})}{\partial K_{i}}.

For 1-cell eEie\in E_{i}, we can suppose that F{e}={i,h}F_{\{e\}}=\{i,h\}. Then FmF{e}={i}F_{m}\cap F_{\{e\}}=\{i\} or {i,h}\{i,h\}.

If FmF{e}={i,h}F_{m}\cap F_{\{e\}}=\{i,h\}, by Gauss-Bonnet formula, we have

fFmF{e}Lf,e=Li,e+Lh,e=2Φ(e)Area(Be).\sum_{f\in F_{m}\cap F_{\{e\}}}L_{f,e}=L_{i,e}+L_{h,e}=2\Phi(e)-\text{Area}(B_{e}).

By Lemma 9 in [6], we have

(fFmF{e}Lf,e)Ki=Area(Be)Ki>0.\frac{\partial(\sum_{f\in F_{m}\cap F_{\{e\}}}L_{f,e})}{\partial K_{i}}=-\frac{\partial\text{Area}(B_{e})}{\partial K_{i}}>0.

If FmF{e}={i}F_{m}\cap F_{\{e\}}=\{i\}, then by Lemma 4.2, we have

(fFmF{e}Lf,e)Ki=Li,e(Ki)Ki>0.\frac{\partial(\sum_{f\in F_{m}\cap F_{\{e\}}}L_{f,e})}{\partial K_{i}}=\frac{\partial L_{i,e}(K_{i})}{\partial K_{i}}>0.

Then we obtain

|LiKi|1jim|LjKi|=eEi(fFmF{e}Lf,e)Ki>0.\left|\frac{\partial L_{i}}{\partial K_{i}}\right|-\sum_{1\leq j\neq i\leq m}\left|\frac{\partial L_{j}}{\partial K_{i}}\right|=\sum_{e\in E_{i}}\frac{\partial(\sum_{f\in F_{m}\cap F_{\{e\}}}L_{f,e})}{\partial K_{i}}>0.

Hence M1M_{1} is a strictly diagonally dominant matrix with positive diagonal entries, i.e. M1M_{1} is positive definite. ∎

Corollary 6.2.

The potential function Λ1\Lambda_{1} is strictly convex on m\mathbb{R}^{m}.

Proof.

By Proposition 6.1, HessΛ1\text{Hess}\leavevmode\nobreak\ \Lambda_{1} is positive definite, then Λ1\Lambda_{1} is strictly convex on m\mathbb{R}^{m}. ∎

Theorem 6.3.

Λ1\nabla\Lambda_{1} is a homeomorphism from m\mathbb{R}^{m} to ~1m\tilde{\mathcal{L}}_{1\cdots m}, where

~1m={(L1,,Lm)T>0m|fFmLf<2eEFmΦ(e),FmFm}.\tilde{\mathcal{L}}_{1\cdots m}=\{(L_{1},\cdots,L_{m})^{T}\in\mathbb{R}^{m}_{>0}\leavevmode\nobreak\ |\leavevmode\nobreak\ \sum_{f\in F_{m}^{\prime}}L_{f}<2\sum_{e\in E_{F_{m}^{\prime}}}\Phi(e),\forall F_{m}^{\prime}\subset F_{m}\}.
Proof.

By (6.1), we have

Λ1:m\displaystyle\nabla\Lambda_{1}:\mathbb{R}^{m} >0m\displaystyle\longrightarrow\mathbb{R}^{m}_{>0}
K=(K1,,Km)T\displaystyle K=(K_{1},\cdots,K_{m})^{T} L=(L1,,Lm)T.\displaystyle\mapsto L=(L_{1},\cdots,L_{m})^{T}.

For any subset FmFmF_{m}^{\prime}\subset F_{m}, we obtain

fFmLf=fFmeEfLf,e=eEFmfF{e}FmLf,e.\sum_{f\in F_{m}^{\prime}}L_{f}=\sum_{f\in F_{m}^{\prime}}\sum_{e\in E_{f}}L_{f,e}=\sum_{e\in E_{F_{m}^{\prime}}}\sum_{f\in F_{\{e\}}\cap F_{m}^{\prime}}L_{f,e}.

Then by Gauss-Bonnet formula, we have

Area(Be)=2Φ(e)fF{e}Lf,e>0,\text{Area}(B_{e})=2\Phi(e)-\sum_{f\in F_{\{e\}}}L_{f,e}>0,

i.e. fF{e}Lf,e<2Φ(e)\sum_{f\in F_{\{e\}}}L_{f,e}<2\Phi(e). Hence we obtain

fFmLf=eEFmfF{e}FmLf,eeEFmfF{e}Lf,e<2eEFmΦ(e),\sum_{f\in F_{m}^{\prime}}L_{f}=\sum_{e\in E_{F_{m}^{\prime}}}\sum_{f\in F_{\{e\}}\cap F_{m}^{\prime}}L_{f,e}\leq\sum_{e\in E_{F_{m}^{\prime}}}\sum_{f\in F_{\{e\}}}L_{f,e}<2\sum_{e\in E_{F_{m}^{\prime}}}\Phi(e),

hence (L1,,Lm)T~1m(L_{1},\cdots,L_{m})^{T}\in\tilde{\mathcal{L}}_{1\cdots m}, Λ1\nabla\Lambda_{1} is a map from m\mathbb{R}^{m} to ~1m\tilde{\mathcal{L}}_{1\cdots m}.

By Proposition 6.1 and Corollary 6.2, Λ1\nabla\Lambda_{1} is a embedding map. Using the method in the proof of Theorem 5.3, we have a similar result that the image set of Λ1\nabla\Lambda_{1} is ~1m\tilde{\mathcal{L}}_{1\cdots m}. Hence Λ1\nabla\Lambda_{1} is a homeomorphism from m\mathbb{R}^{m} to ~1m\tilde{\mathcal{L}}_{1\cdots m}. ∎

We can construct a map ς1\varsigma_{1} from >0m\mathbb{R}^{m}_{>0} to m\mathbb{R}^{m}, i.e.

ς1:>0mmk=(k1,,km)TK=(lnk1,,lnkm)T,\begin{array}[]{cccc}\varsigma_{1}:\mathbb{R}^{m}_{>0}&\longrightarrow&\mathbb{R}^{m}\\ k=(k_{1},\cdots,k_{m})^{T}&\longmapsto&K=(\ln k_{1},\cdots,\ln k_{m})^{T},\\ \end{array}

the map ς1\varsigma_{1} is a homeomorphism. By Theorem 6.3, we know that the map ~:=Λ1ς1\tilde{\mathcal{E}}:=\nabla\Lambda_{1}\circ\varsigma_{1}, i.e.

~:>0m~1mk~(k)=Λ1(K)=L(K)\begin{array}[]{cccc}\tilde{\mathcal{E}}:\mathbb{R}^{m}_{>0}&\longrightarrow&\tilde{\mathcal{L}}_{1\cdots m}\\ k&\longmapsto&\tilde{\mathcal{E}}(k)=\nabla\Lambda_{1}(K)=L(K)\\ \end{array}

is a homeomorphism from >0m\mathbb{R}^{m}_{>0} to ~1m\tilde{\mathcal{L}}_{1\cdots m}. In other words, there exists a homeomorphism from 1m|F|\mathbb{R}^{|F|}_{1\cdots m} to 1m\mathcal{L}_{1\cdots m}, where

1m={(L1,,Lm,0,,0)T>0m×{0}|F|mfFmLf<2eEFmΦ(e),FmFm}.\mathcal{L}_{1\cdots m}=\left\{\begin{array}[]{l|l}(L_{1},\cdots,L_{m},0,\cdots,0)^{T}\in\mathbb{R}^{m}_{>0}\times\{0\}^{|F|-m}&\begin{array}[]{l}\sum_{f\in F_{m}^{\prime}}L_{f}<2\sum_{e\in E_{F_{m}^{\prime}}}\Phi(e),\forall F_{m}^{\prime}\subset F_{m}\end{array}\end{array}\right\}.

Similarly, for 1m|F|11\leq m\leq|F|-1, 1i1<<im|F|1\leq i_{1}<\cdots<i_{m}\leq|F|, by Fi1imF_{i_{1}\cdots i_{m}} we denote the set Fi1im={i1,,im}F_{i_{1}\cdots i_{m}}=\{i_{1},\cdots,i_{m}\}. Then there exists a homeomorphism from i1im|F|\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}} to i1im\mathcal{L}_{i_{1}\cdots i_{m}}, where

i1im={(0,,Li1,,0,,Lim,0,,0)T|F|fFi1imLf<2eEFi1imΦ(e),Fi1imFi1im;Lij>0,1jm}.\mathcal{L}_{i_{1}\cdots i_{m}}=\left\{\begin{array}[]{l|l}(0,\cdots,L_{i_{1}},\cdots,0,\cdots,L_{i_{m}},0,\cdots,0)^{T}\in\mathbb{R}^{|F|}&\begin{array}[]{l}\sum_{f\in F_{i_{1}\cdots i_{m}}^{\prime}}L_{f}<2\sum_{e\in E_{F_{i_{1}\cdots i_{m}}^{\prime}}}\Phi(e),\\ \forall F_{i_{1}\cdots i_{m}}^{\prime}\subset F_{i_{1}\cdots i_{m}};L_{i_{j}}>0,1\leq j\leq m\end{array}\end{array}\right\}.

Besides, we define the set ~i1im\tilde{\mathcal{L}}_{i_{1}\cdots i_{m}}, i.e.

~i1im={(Li1,,Lim)T>0mfFi1imLf<2eEFi1imΦ(e),Fi1imFi1im}.\tilde{\mathcal{L}}_{i_{1}\cdots i_{m}}=\left\{\begin{array}[]{l|l}(L_{i_{1}},\cdots,L_{i_{m}})^{T}\in\mathbb{R}_{>0}^{m}&\begin{array}[]{l}\sum_{f\in F_{i_{1}\cdots i_{m}}^{\prime}}L_{f}<2\sum_{e\in E_{F_{i_{1}\cdots i_{m}}^{\prime}}}\Phi(e),\leavevmode\nobreak\ \forall F_{i_{1}\cdots i_{m}}^{\prime}\subset F_{i_{1}\cdots i_{m}}\end{array}\end{array}\right\}.

By ¯\bar{\partial}\mathcal{L} we denote the set

¯=1m|F|1,1i1<<im|F|i1im{0}.\bar{\partial}\mathcal{L}=\bigcup_{1\leq m\leq|F|-1,1\leq i_{1}<\cdots<i_{m}\leq|F|}\mathcal{L}_{i_{1}\cdots i_{m}}\cup\{0\}.

Then by above argument, we completed the proof of Theorem 1.7.

7 Existence and rigidity of spherical conical metrics for prescribed total geodesic curvatures

By i1im\mathcal{E}_{i_{1}\cdots i_{m}} we denote the homeomorphism from i1im|F|\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}} to i1im\mathcal{L}_{i_{1}\cdots i_{m}} and by \mathcal{L} we denote the set =1¯\mathcal{L}=\mathcal{L}_{1}\cup\bar{\partial}\mathcal{L}. Then we can define a map \mathcal{E} from 0|F|\mathbb{R}^{|F|}_{\geq 0} to \mathcal{L}, i.e.

(7.4) (k):={1(k),k>0|F|,i1im(k),ki1im|F|,1m|F|1,1i1<<im|F|,0,k=0.\displaystyle\mathcal{E}(k):=\left\{\begin{array}[]{lc}\mathcal{E}_{1}(k),&k\in\mathbb{R}^{|F|}_{>0},\\ \mathcal{E}_{i_{1}\cdots i_{m}}(k),&k\in\mathbb{R}^{|F|}_{i_{1}\cdots i_{m}},1\leq m\leq|F|-1,1\leq i_{1}<\cdots<i_{m}\leq|F|,\\ 0,&k=0.\end{array}\right.
Theorem 7.1.

\mathcal{E} is a homeomorphism from 0|F|\mathbb{R}^{|F|}_{\geq 0} to \mathcal{L} such that the interior of 0|F|\mathbb{R}^{|F|}_{\geq 0} maps to the interior of \mathcal{L} and ¯0|F|\bar{\partial}\mathbb{R}^{|F|}_{\geq 0} maps to ¯\bar{\partial}\mathcal{L}.

Proof.

Since 1\mathcal{E}_{1}, i1im(1m|F|1,1i1<<im|F|)\mathcal{E}_{i_{1}\cdots i_{m}}(1\leq m\leq|F|-1,1\leq i_{1}<\cdots<i_{m}\leq|F|) are homeomorphisms, by (7.4), then \mathcal{E} is a bijective map from 0|F|\mathbb{R}^{|F|}_{\geq 0} to \mathcal{L}. We only need to show that \mathcal{E} and 1\mathcal{E}^{-1} are continuous from interior to boundary.

We choose a sequence {kn}n=1+>0|F|\{k^{n}\}_{n=1}^{+\infty}\subset\mathbb{R}^{|F|}_{>0} such that limn+kn=0\lim_{n\to+\infty}k^{n}=0, where kn=(k1n,,k|F|n)Tk^{n}=(k_{1}^{n},\cdots,k_{|F|}^{n})^{T}. Then we have 1(kn)=(L1(kn),,L|F|(kn))T\mathcal{E}_{1}(k^{n})=(L_{1}(k^{n}),\cdots,L_{|F|}(k^{n}))^{T}, where Li(kn)=eEii,enkinL_{i}(k^{n})=\sum_{e\in E_{i}}\ell_{i,e}^{n}k_{i}^{n}, 1i|F|1\leq i\leq|F|. Since 0i,enπ,kin0(n+)0\leq\ell_{i,e}^{n}\leq\pi,k_{i}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty), then we have i,enkin0(n+)\ell_{i,e}^{n}k_{i}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty), Li(kn)=eEii,enkin0(n+)L_{i}(k^{n})=\sum_{e\in E_{i}}\ell_{i,e}^{n}k_{i}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty), 1i|F|1\leq i\leq|F|. Hence we obtain 1(kn)0(n+)\mathcal{E}_{1}(k^{n})\to 0\leavevmode\nobreak\ (n\to+\infty).

For any point ki1im|F|k\in\mathbb{R}_{i_{1}\cdots i_{m}}^{|F|}, 1m|F|11\leq m\leq|F|-1, 1i1<<im|F|1\leq i_{1}<\cdots<i_{m}\leq|F|, we choose another sequence {kn}n=1+>0|F|\{k^{n}\}_{n=1}^{+\infty}\subset\mathbb{R}^{|F|}_{>0} such that limn+kn=k\lim_{n\to+\infty}k^{n}=k. Then we need to show that 1(kn)i1im(k)(n+)\mathcal{E}_{1}(k^{n})\to\mathcal{E}_{i_{1}\cdots i_{m}}(k)\leavevmode\nobreak\ (n\to+\infty). For simplicity, we only show that 1(kn)1m(k)=(L1(k),,Lm(k),0,,0)(n+)\mathcal{E}_{1}(k^{n})\to\mathcal{E}_{1\cdots m}(k)=(L_{1}(k),\cdots,L_{m}(k),0,\cdots,0)\leavevmode\nobreak\ (n\to+\infty), the others use the same method.

We can suppose that k=(k1,,km,0,,0)Tk=(k_{1},\cdots,k_{m},0,\cdots,0)^{T}, since kin0,m+1i|F|k_{i}^{n}\to 0,\leavevmode\nobreak\ m+1\leq i\leq|F|, by the analysis above, then we have Li(kn)=eEii,enkin0(n+),m+1i|F|L_{i}(k^{n})=\sum_{e\in E_{i}}\ell_{i,e}^{n}k_{i}^{n}\to 0\leavevmode\nobreak\ (n\to+\infty),\leavevmode\nobreak\ m+1\leq i\leq|F|. Since the total geodesic curvature LiL_{i} is continuous on 0|F|\mathbb{R}^{|F|}_{\geq 0}, then we have Li(kn)Li(k)(n+)L_{i}(k^{n})\to L_{i}(k)\leavevmode\nobreak\ (n\to+\infty), 1im1\leq i\leq m, i.e. 1(kn)1m(k)(n+)\mathcal{E}_{1}(k^{n})\to\mathcal{E}_{1\cdots m}(k)\leavevmode\nobreak\ (n\to+\infty). Hence the map \mathcal{E} is continuous.

For any L¯L\in\bar{\partial}\mathcal{L}, there exists a small enough neighborhood UU of LL such that 1(U)\mathcal{E}^{-1}(U) is bounded in 0|F|\mathbb{R}^{|F|}_{\geq 0}. We choose a sequence {Ln}n=1+U\{L^{n}\}_{n=1}^{+\infty}\subset U such that limn+Ln=L\lim_{n\to+\infty}L^{n}=L. By {kn}n=1+0|F|\{k^{n}\}_{n=1}^{+\infty}\subset\mathbb{R}_{\geq 0}^{|F|} and kk we denote the image of {Ln}n=1+\{L^{n}\}_{n=1}^{+\infty} and LL under the map 1\mathcal{E}^{-1}, respectively. Then we only need to show that knk(n+)k^{n}\to k\leavevmode\nobreak\ (n\to+\infty).

If knk(n+)k^{n}\nrightarrow k\leavevmode\nobreak\ (n\to+\infty), since {kn}n=1+1(U)\{k^{n}\}_{n=1}^{+\infty}\subset\mathcal{E}^{-1}(U) and 1(U)\mathcal{E}^{-1}(U) is bounded, then there exists a subsequence of {kn}n=1+\{k^{n}\}_{n=1}^{+\infty}, which is still denoted by {kn}n=1+\{k^{n}\}_{n=1}^{+\infty}, such that knk(n+)k^{n}\to k^{\prime}\leavevmode\nobreak\ (n\to+\infty), where kkk^{\prime}\neq k. Since \mathcal{E} is continuous, then (kn)=Ln(k)(n+)\mathcal{E}(k^{n})=L^{n}\to\mathcal{E}(k^{\prime})\leavevmode\nobreak\ (n\to+\infty). Besides, we know that LnL=(k)(n+)L^{n}\to L=\mathcal{E}(k)\leavevmode\nobreak\ (n\to+\infty), then we obtain (k)=(k)\mathcal{E}(k)=\mathcal{E}(k^{\prime}). This is a contradiction since the map \mathcal{E} is injective. Hence the map 1\mathcal{E}^{-1} is continuous, \mathcal{E} is a homeomorphism from 0|F|\mathbb{R}^{|F|}_{\geq 0} to \mathcal{L}. ∎

By above argument, we completed the proof of Theorem 1.8.

8 Convergence of prescribed combinatorial Ricci flows for circle pattern metrics

Given L^=(L^1,,L^|F|)T1\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{|F|})^{T}\in\mathcal{L}_{1}, then we can define the prescribed combinatorial Ricci flows as follows, i.e.

(8.1) dkidt=(LiL^i)ki,iF.\frac{dk_{i}}{dt}=-(L_{i}-\hat{L}_{i})k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F.

Using change of variables Ki=lnkiK_{i}=\ln k_{i}, we can rewrite flows (8.1) as the following equivalent prescribed combinatorial Ricci flows

(8.2) dKidt=(LiL^i),iF.\frac{dK_{i}}{dt}=-(L_{i}-\hat{L}_{i}),\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F.
Theorem 8.1.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^=(L^1,,L^|F|)T1\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{|F|})^{T}\in\mathcal{L}_{1} on the face set FF. For any initial geodesic curvature k(0)>0|F|k(0)\in\mathbb{R}_{>0}^{|F|}, the solution of the prescribed combinatorial Ricci flows (8.1) exists for all time t[0,+)t\in[0,+\infty) and is unique.

Proof.

We can only consider the equivalent flows (8.2). Since all (LiL^i)-(L_{i}-\hat{L}_{i}) are smooth functions on >0|F|\mathbb{R}_{>0}^{|F|}, by Peano’s existence theorem in ODE theory, we know that the solution of flows (8.2) exists on [0,ϵ)[0,\epsilon), where ϵ>0\epsilon>0.

By Gauss-Bonnet formula, we have that

Area(Be)=2Φ(e)Li,eLj,e>0,jF{e},\text{Area}(B_{e})=2\Phi(e)-L_{i,e}-L_{j,e}>0,\leavevmode\nobreak\ \leavevmode\nobreak\ j\in F_{\{e\}},

hence we obtain 0<Li,e<2Φ(e)0<L_{i,e}<2\Phi(e). Since Li=eEiLi,eL_{i}=\sum_{e\in E_{i}}L_{i,e}, we have that 0<Li<2eEiΦ(e)0<L_{i}<2\sum_{e\in E_{i}}\Phi(e). Besides, since L^1\hat{L}\in\mathcal{L}_{1}, by definition, we have 0<L^i<fFL^f<2eEFΦ(e)0<\hat{L}_{i}<\sum_{f\in F}\hat{L}_{f}<2\sum_{e\in E_{F}}\Phi(e). Then we obtain that

|LiL^i||Li|+|L^i|2eEiΦ(e)+2eEFΦ(e)<+.|L_{i}-\hat{L}_{i}|\leq|L_{i}|+|\hat{L}_{i}|\leq 2\sum_{e\in E_{i}}\Phi(e)+2\sum_{e\in E_{F}}\Phi(e)<+\infty.

Hence |LiL^i||L_{i}-\hat{L}_{i}| is uniformly bounded by a constant, which depends only on the weight Φ\Phi and cellular decomposition Σ\Sigma. By the extension theorem of solution in ODE theory, the solution of flows (8.2) exists for all time t[0,+)t\in[0,+\infty). By existence and uniqueness theorem of solution in ODE theory, the solution of flows (8.2) is unique. ∎

We need the following result in the theory of negative gradient flows.

Lemma 8.2.

([10], Proposition 2.13) Let h:nh:\mathbb{R}^{n}\rightarrow\mathbb{R} be a smooth convex function and let ξ:[0,+)n\xi:[0,+\infty)\rightarrow\mathbb{R}^{n} be a negative gradient flow of hh. It holds for any τ>0\tau>0 and ξn\xi^{*}\in\mathbb{R}^{n} that

|h(ξ(τ))|2|h(ξ)|2+1τ2|ξξ(0)|2.|\nabla h(\xi(\tau))|^{2}\leq\left|\nabla h\left(\xi^{*}\right)\right|^{2}+\frac{1}{\tau^{2}}\left|\xi^{*}-\xi(0)\right|^{2}.

Then we can prove the Theorem 1.9.

Proof of Theorem 1.9.

We can only consider the equivalent flows (8.2). By Theorem 8.1, we suppose that the solution of flows (8.2) is K(t)K(t), t[0,+)t\in[0,+\infty). We construct a function Λ~=ΛfFKfL^f\tilde{\Lambda}=\Lambda-\sum_{f\in F}K_{f}\hat{L}_{f}, by Corollary 5.2, the function Λ~\tilde{\Lambda} is convex on |F|\mathbb{R}^{|F|}. Besides, we have that Λ~=ΛL^=LL^,\nabla\tilde{\Lambda}=\nabla\Lambda-\hat{L}=L-\hat{L}, then we obtain

dKdt=Λ~(K(t)).\frac{dK}{dt}=-\nabla\tilde{\Lambda}(K(t)).

Since L^1\hat{L}\in\mathcal{L}_{1}, by Theorem 5.3, there exists the unique K^|F|\hat{K}\in\mathbb{R}^{|F|} such that Λ(K^)=L(K^)=L^\nabla\Lambda(\hat{K})=L(\hat{K})=\hat{L}. By Lemma 8.2, for any t>0t>0, we have

|Λ~(K(t))|2|Λ~(K^)|2+1t2|K^K(0)|2,|\nabla\tilde{\Lambda}(K(t))|^{2}\leq|\nabla\tilde{\Lambda}(\hat{K})|^{2}+\frac{1}{t^{2}}|\hat{K}-K(0)|^{2},

i.e. we obtain

|L(K(t))L(K^)|21t2|K^K(0)|20(t+).|L(K(t))-L(\hat{K})|^{2}\leq\frac{1}{t^{2}}|\hat{K}-K(0)|^{2}\to 0\leavevmode\nobreak\ (t\to+\infty).

Hence we know that L(K(t))L(K^)=L^(t+)L(K(t))\to L(\hat{K})=\hat{L}\leavevmode\nobreak\ (t\to+\infty), i.e. Λ(K(t))Λ(K^)(t+)\nabla\Lambda(K(t))\to\nabla\Lambda(\hat{K})\leavevmode\nobreak\ (t\to+\infty). Then by Theorem 5.3, we have that K(t)K^(t+).K(t)\to\hat{K}\leavevmode\nobreak\ (t\to+\infty). This completes the proof. ∎

9 Convergence of prescribed combinatorial Ricci flows for degenerated circle pattern metrics

Given the prescribed total geodesic curvature L^i1im\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}}, where 1m|F|11\leq m\leq|F|-1, 1i1<<im|F|1\leq i_{1}<\cdots<i_{m}\leq|F|, we can define the prescribed combinatorial Ricci flows as follows, i.e.

(9.1) dkidt=(LiL^i)ki,iFi1im,dkidt=Liki,iFFi1im,\frac{dk_{i}}{dt}=-(L_{i}-\hat{L}_{i})k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F_{i_{1}\cdots i_{m}},\leavevmode\nobreak\ \frac{dk_{i}}{dt}=-L_{i}k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F\setminus F_{i_{1}\cdots i_{m}},

where Fi1im={i1,,im}F_{i_{1}\cdots i_{m}}=\{i_{1},\cdots,i_{m}\}. Then we study the existence and convergence of solution of the flows (9.1).

Using change of variables Ki=lnkiK_{i}=\ln k_{i}, we can rewrite flows (9.1) as the following equivalent prescribed combinatorial Ricci flows

(9.2) dKidt=(LiL^i),iFi1im,dKidt=Li,iFFi1im.\frac{dK_{i}}{dt}=-(L_{i}-\hat{L}_{i}),\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F_{i_{1}\cdots i_{m}},\leavevmode\nobreak\ \frac{dK_{i}}{dt}=-L_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F\setminus F_{i_{1}\cdots i_{m}}.

By similar argument in the proof of Theorem 8.1, we have the following theorem.

Theorem 9.1.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^i1im\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}} on the face set FF. For any initial geodesic curvature k(0)>0|F|k(0)\in\mathbb{R}_{>0}^{|F|}, the solution of the prescribed combinatorial Ricci flows (9.1) exists for all time t[0,+)t\in[0,+\infty) and is unique.

For simplicity, we consider the set Fm={1,,m}F_{m}=\{1,\cdots,m\}, where 1m|F|11\leq m\leq|F|-1. Given Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|}, L^=(L^1,,L^m,0,,0)T1m\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{m},0,\cdots,0)^{T}\in\mathcal{L}_{1\cdots m}, then we study the following prescribed combinatorial Ricci flows, i.e.

(9.3) dkidt=(LiL^i)ki, 1im,dkidt=Liki,m+1i|F|.\frac{dk_{i}}{dt}=-(L_{i}-\hat{L}_{i})k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq i\leq m,\leavevmode\nobreak\ \frac{dk_{i}}{dt}=-L_{i}k_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ m+1\leq i\leq|F|.

We can also consider the equivalent prescribed combinatorial Ricci flows, i.e.

(9.4) dKidt=(LiL^i), 1im,dKidt=Li,m+1i|F|.\frac{dK_{i}}{dt}=-(L_{i}-\hat{L}_{i}),\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq i\leq m,\leavevmode\nobreak\ \frac{dK_{i}}{dt}=-L_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ m+1\leq i\leq|F|.
Remark 9.2.

For any K~=(K~1,,K~m)Tm\tilde{K}=(\tilde{K}_{1},\cdots,\tilde{K}_{m})^{T}\in\mathbb{R}^{m}, we have that Λ1(K~)=(L1(K~),,Lm(K~))T\nabla\Lambda_{1}(\tilde{K})=(L_{1}(\tilde{K}),\cdots,L_{m}(\tilde{K}))^{T}. By (6.1), we know that Li(K~)(1im)L_{i}(\tilde{K})(1\leq i\leq m) is actually the total geodesic curvature at the face ii when the radii 𝐫=(arccoteK~1,,arccoteK~m,π2,,π2)T\mathbf{r}=(\operatorname{arccot}e^{\tilde{K}_{1}},\cdots,\operatorname{arccot}e^{\tilde{K}_{m}},\frac{\pi}{2},\cdots,\frac{\pi}{2})^{T} or when the geodesic curvatures k=(eK~1,,eK~m,0,,0)Tk=(e^{\tilde{K}_{1}},\cdots,e^{\tilde{K}_{m}},0,\cdots,0)^{T} or when the K=(K~1,,K~m,,,)TK=(\tilde{K}_{1},\cdots,\tilde{K}_{m},-\infty,\cdots,-\infty)^{T}.
If we use geodesic curvatures k~=(k~1,,k~m)T>0m\tilde{k}=(\tilde{k}_{1},\cdots,\tilde{k}_{m})^{T}\in\mathbb{R}^{m}_{>0} as the variable of LiL_{i}, then the Li(k~)(1im)L_{i}(\tilde{k})(1\leq i\leq m) is actually the total geodesic curvature at the face ii when the geodesic curvatures k=(k~1,,k~m,0,,0)Tk=(\tilde{k}_{1},\cdots,\tilde{k}_{m},0,\cdots,0)^{T}.

For k~=(k~1,,k~m)T>0m\tilde{k}=(\tilde{k}_{1},\cdots,\tilde{k}_{m})^{T}\in\mathbb{R}^{m}_{>0}, we construct the prescribed combinatorial Ricci flows as follows, i.e.

(9.5) dk~idt=(Li(k~)L^i)k~i, 1im.\frac{d\tilde{k}_{i}}{dt}=-(L_{i}(\tilde{k})-\hat{L}_{i})\tilde{k}_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq i\leq m.

Using change of variables K~i=lnk~i\tilde{K}_{i}=\ln\tilde{k}_{i}, we can rewrite flows (9.5) as the following equivalent prescribed combinatorial Ricci flows

(9.6) dK~idt=(Li(K~)L^i), 1im.\frac{d\tilde{K}_{i}}{dt}=-(L_{i}(\tilde{K})-\hat{L}_{i}),\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq i\leq m.

By similar argument in the proof of Theorem 8.1, we have the following theorem.

Theorem 9.3.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^1m\hat{L}\in\mathcal{L}_{1\cdots m} on the face set FF. For any initial geodesic curvature k~(0)>0m\tilde{k}(0)\in\mathbb{R}_{>0}^{m} on the face set FmF_{m} and 0 on the face set FFmF\setminus F_{m}, the solution of the prescribed combinatorial Ricci flows (9.5) exists for all time t[0,+)t\in[0,+\infty) and is unique.

Then we study the convergence of solution to the flows (9.5).

Theorem 9.4.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^=(L^1,,L^m,0,0)T1m\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{m},0\cdots,0)^{T}\in\mathcal{L}_{1\cdots m} on the face set FF. For any initial geodesic curvature k~(0)>0m\tilde{k}(0)\in\mathbb{R}_{>0}^{m} on the face set FmF_{m} and 0 on the face set FFmF\setminus F_{m}, the solution of the prescribed combinatorial Ricci flows (9.5) converges to the unique degenerated circle pattern metric with the total geodesic curvature L~=(L^1,,L^m)T\tilde{L}=(\hat{L}_{1},\cdots,\hat{L}_{m})^{T} on FmF_{m} and 0 on FFmF\setminus F_{m} up to isometry. Moreover, if the solution converges to the degenerated circle pattern metric k^\hat{k}, then k^1m|F|\hat{k}\in\mathbb{R}^{|F|}_{1\cdots m}.

Proof.

We can only consider the equivalent flows (9.6). By Theorem 9.3, we suppose that the solution of flows (9.6) is K~(t)\tilde{K}(t), t[0,+)t\in[0,+\infty). We construct a function Λ~1(K~)=Λ1(K~)fFmK~fL^f\tilde{\Lambda}_{1}(\tilde{K})=\Lambda_{1}(\tilde{K})-\sum_{f\in F_{m}}\tilde{K}_{f}\hat{L}_{f}. By Corollary 6.2, the function Λ~1\tilde{\Lambda}_{1} is convex on m\mathbb{R}^{m}. Besides, we have that

Λ~1(K~)K~i=Λ1(K~)K~iL^i=Li(K~)L^i, 1im,\frac{\partial\tilde{\Lambda}_{1}(\tilde{K})}{\partial\tilde{K}_{i}}=\frac{\partial\Lambda_{1}(\tilde{K})}{\partial\tilde{K}_{i}}-\hat{L}_{i}=L_{i}(\tilde{K})-\hat{L}_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ 1\leq i\leq m,

then we obtain

dK~dt=Λ~1(K~(t)).\frac{d\tilde{K}}{dt}=-\nabla\tilde{\Lambda}_{1}(\tilde{K}(t)).

Since L~~1m\tilde{L}\in\tilde{\mathcal{L}}_{1\cdots m}, by Theorem 6.3, there exists the unique K~=(K^1,,K^m)Tm\tilde{K}=(\hat{K}_{1},\cdots,\hat{K}_{m})^{T}\in\mathbb{R}^{m} such that Λ1(K~)=(L1(K~),,Lm(K~))T=(L^1,,L^m)T=L~\nabla\Lambda_{1}(\tilde{K})=(L_{1}(\tilde{K}),\cdots,L_{m}(\tilde{K}))^{T}=(\hat{L}_{1},\cdots,\hat{L}_{m})^{T}=\tilde{L}. By Lemma 8.2, for any t>0t>0, we have

|Λ~1(K~(t))|2|Λ~1(K~)|2+1t2|K~K~(0)|2,|\nabla\tilde{\Lambda}_{1}(\tilde{K}(t))|^{2}\leq|\nabla\tilde{\Lambda}_{1}(\tilde{K})|^{2}+\frac{1}{t^{2}}|\tilde{K}-\tilde{K}(0)|^{2},

i.e. we obtain

i=1m(Li(K~(t))L^i)21t2|K~K~(0)|20(t+).\sum_{i=1}^{m}(L_{i}(\tilde{K}(t))-\hat{L}_{i})^{2}\leq\frac{1}{t^{2}}|\tilde{K}-\tilde{K}(0)|^{2}\to 0\leavevmode\nobreak\ (t\to+\infty).

Hence for 1im1\leq i\leq m, we know that Li(K~(t))L^i=Li(K~)(t+)L_{i}(\tilde{K}(t))\to\hat{L}_{i}=L_{i}(\tilde{K})\leavevmode\nobreak\ (t\to+\infty) and L(K~(t))L~=L(K~)(t+)L(\tilde{K}(t))\to\tilde{L}=L(\tilde{K})\leavevmode\nobreak\ (t\to+\infty). Hence we have that Λ1(K~(t))Λ1(K~)(t+)\nabla\Lambda_{1}(\tilde{K}(t))\to\nabla\Lambda_{1}(\tilde{K})\leavevmode\nobreak\ (t\to+\infty). Then by Theorem 6.3, we have that K~(t)K~(t+)\tilde{K}(t)\to\tilde{K}\leavevmode\nobreak\ (t\to+\infty). This completes the proof. ∎

Remark 9.5.

By Remark 9.2 and Theorem 9.4, we know that (K~1(t),,K~m(t),,,)T(K^1,,K^m,,,)T(t+)(\tilde{K}_{1}(t),\cdots,\tilde{K}_{m}(t),-\infty,\cdots,-\infty)^{T}\\ \to(\hat{K}_{1},\cdots,\hat{K}_{m},-\infty,\cdots,-\infty)^{T}\leavevmode\nobreak\ (t\to+\infty) and Li(K~1(t),,K~m(t),,,)L^i(t+)L_{i}(\tilde{K}_{1}(t),\cdots,\tilde{K}_{m}(t),-\infty,\cdots,-\infty)\to\hat{L}_{i}\leavevmode\nobreak\ (t\to+\infty), 1im1\leq i\leq m. Besides, we have that L^i=Li(K^1,,K^m,,,)\hat{L}_{i}=L_{i}(\hat{K}_{1},\cdots,\hat{K}_{m},-\infty,\cdots,-\infty).

For any ki1im|F|k\in\mathbb{R}_{i_{1}\cdots i_{m}}^{|F|}, then k~=(ki1,,kim)>0m\tilde{k}=(k_{i_{1}},\cdots,k_{i_{m}})\in\mathbb{R}_{>0}^{m}. Given L^i1im\hat{L}\in\mathcal{L}_{i_{1}\cdots i_{m}}, we construct the prescribed combinatorial Ricci flows as follows, i.e.

(9.7) dk~idt=(Li(k~)L^i)k~i,iFi1im.\frac{d\tilde{k}_{i}}{dt}=-(L_{i}(\tilde{k})-\hat{L}_{i})\tilde{k}_{i},\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F_{i_{1}\cdots i_{m}}.

Using change of variables K~i=lnk~i\tilde{K}_{i}=\ln\tilde{k}_{i}, we can rewrite flows (9.7) as the following equivalent prescribed combinatorial Ricci flows

(9.8) dK~idt=(Li(K~)L^i),iFi1im.\frac{d\tilde{K}_{i}}{dt}=-(L_{i}(\tilde{K})-\hat{L}_{i}),\leavevmode\nobreak\ \leavevmode\nobreak\ \forall i\in F_{i_{1}\cdots i_{m}}.

We can use the same techniques above to obtain the long time existence of the solution to prescribed combinatorial Ricci flows (9.7) and Theorem 1.11.

Then we study the convergence of solution to the flows (9.1).

Theorem 9.6.

Given a closed topological surface SS with a cellular decomposition Σ=(V,E,F)\Sigma=(V,E,F), the weight Φ(0,π2)|E|\Phi\in(0,\frac{\pi}{2})^{|E|} and the prescribed total geodesic curvature L^=(L^1,,L^m,0,0)T1m\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{m},0\cdots,0)^{T}\in\mathcal{L}_{1\cdots m} on the face set FF. For any initial geodesic curvature k(0)>0|F|k(0)\in\mathbb{R}_{>0}^{|F|}, the solution of the prescribed combinatorial Ricci flows (9.3) converges to the unique degenerated circle pattern metric with the total geodesic curvature L^\hat{L} up to isometry. Moreover, if the solution converges to the degenerated circle pattern metric k^\hat{k}, then k^1m|F|\hat{k}\in\mathbb{R}^{|F|}_{1\cdots m}.

Proof.

We consider the equivalent flows (9.4). By Theorem 9.1, we suppose that the solution of flows (9.4) is K(t)K(t), t[0,+)t\in[0,+\infty). We construct a function Λ¯=ΛfFmKfL^f\bar{\Lambda}=\Lambda-\sum_{f\in F_{m}}K_{f}\hat{L}_{f}. By Corollary 5.2, the function Λ¯\bar{\Lambda} is convex on |F|\mathbb{R}^{|F|}. Besides, we have that Λ¯=ΛL^=LL^,\nabla\bar{\Lambda}=\nabla\Lambda-\hat{L}=L-\hat{L}, then we obtain

dKdt=Λ¯(K(t)).\frac{dK}{dt}=-\nabla\bar{\Lambda}(K(t)).

By Lemma 8.2, for any t>0t>0 and K¯(t)=(K~1(t),,K~m(t),lnt,,lnt)T|F|\bar{K}(t)=(\tilde{K}_{1}(t),\cdots,\tilde{K}_{m}(t),-\ln t,\cdots,\ln t)^{T}\in\mathbb{R}^{|F|}, we have

|Λ¯(K(t))|2|Λ¯(K¯(t))|2+1t2|K¯(t)K(0)|2,|\nabla\bar{\Lambda}(K(t))|^{2}\leq|\nabla\bar{\Lambda}(\bar{K}(t))|^{2}+\frac{1}{t^{2}}|\bar{K}(t)-K(0)|^{2},

By Theorem 9.4, we obtain that K~i(t)K^i(t+)\tilde{K}_{i}(t)\to\hat{K}_{i}\leavevmode\nobreak\ (t\to+\infty), 1im1\leq i\leq m. Hence we know that the function K~i(t)(1im)\tilde{K}_{i}(t)\leavevmode\nobreak\ (1\leq i\leq m) is bounded, then we have

1t2|K¯(t)K(0)|20(t+).\frac{1}{t^{2}}|\bar{K}(t)-K(0)|^{2}\to 0\leavevmode\nobreak\ (t\to+\infty).

Besides, we have

|Λ¯(K¯(t))|2=i=1m(Li(K¯(t))L^i)2+i=m+1|F|Li2(K¯(t)).|\nabla\bar{\Lambda}(\bar{K}(t))|^{2}=\sum_{i=1}^{m}(L_{i}(\bar{K}(t))-\hat{L}_{i})^{2}+\sum_{i=m+1}^{|F|}L_{i}^{2}(\bar{K}(t)).

Since K¯(t)=(K~1(t),,K~m(t),lnt,,lnt)T(K^1,,K^m,,,)T(t+)\bar{K}(t)=(\tilde{K}_{1}(t),\cdots,\tilde{K}_{m}(t),-\ln t,\cdots,\ln t)^{T}\to(\hat{K}_{1},\cdots,\hat{K}_{m},-\infty,\cdots,-\infty)^{T}\leavevmode\nobreak\ (t\to+\infty), by Remark 9.5, we have that Li(K¯(t))Li(K^1,,K^m,,,)=L^i(t+)L_{i}(\bar{K}(t))\to L_{i}(\hat{K}_{1},\cdots,\hat{K}_{m},-\infty,\cdots,-\infty)=\hat{L}_{i}\leavevmode\nobreak\ (t\to+\infty), 1im1\leq i\leq m. Besides, the geodesic curvature ki=elnt=1t0(t+)k_{i}=e^{-\ln t}=\frac{1}{t}\to 0\leavevmode\nobreak\ (t\to+\infty), m+1i|F|m+1\leq i\leq|F|, then Li(K¯(t))0(t+)L_{i}(\bar{K}(t))\to 0\leavevmode\nobreak\ (t\to+\infty), m+1i|F|m+1\leq i\leq|F|. Hence we have

|Λ¯(K¯(t))|20(t+).|\nabla\bar{\Lambda}(\bar{K}(t))|^{2}\to 0\leavevmode\nobreak\ (t\to+\infty).

Then we obtain

|Λ¯(K(t))|20(t+),|\nabla\bar{\Lambda}(K(t))|^{2}\to 0\leavevmode\nobreak\ (t\to+\infty),

i.e. L(K(t))L^=(L^1,,L^m,0,0)T(t+)L(K(t))\to\hat{L}=(\hat{L}_{1},\cdots,\hat{L}_{m},0\cdots,0)^{T}\leavevmode\nobreak\ (t\to+\infty).

We define some functions ki(t)=eKi(t)k_{i}(t)=e^{K_{i}(t)}, 1i|F|1\leq i\leq|F| and some constants k^i=eK^i\hat{k}_{i}=e^{\hat{K}_{i}}, 1im1\leq i\leq m. Then it is easy to know that k(t)=(k1(t),,k|F|(t))Tk(t)=(k_{1}(t),\cdots,k_{|F|}(t))^{T} is the solution of the prescribed combinatorial Ricci flows (9.3) and k^=(k^1,,k^m,0,,0)T1m|F|\hat{k}=(\hat{k}_{1},\cdots,\hat{k}_{m},0,\cdots,0)^{T}\in\mathbb{R}_{1\cdots m}^{|F|}.

For simplicity, we can define L(k):=L(lnk1,,lnk|F|)L(k):=L(\ln k_{1},\cdots,\ln k_{|F|}). Then we have L^i=Li(k^), 1im\hat{L}_{i}=L_{i}(\hat{k}),\leavevmode\nobreak\ 1\leq i\leq m and L^=L(k^)\hat{L}=L(\hat{k}). Hence we know that L(k(t))L(k^)=L^(t+)L(k(t))\to L(\hat{k})=\hat{L}\leavevmode\nobreak\ (t\to+\infty). Since {k(t)|t[0,+)}>0|F|\{k(t)\leavevmode\nobreak\ |\leavevmode\nobreak\ t\in[0,+\infty)\}\subset\mathbb{R}_{>0}^{|F|} and k^1m|F|\hat{k}\in\mathbb{R}_{1\cdots m}^{|F|}, by (7.4), we have that (k(t))(k^)(t+)\mathcal{E}(k(t))\to\mathcal{E}(\hat{k})\leavevmode\nobreak\ (t\to+\infty). By Theorem 7.1, 1\mathcal{E}^{-1} is continuous, then we have that k(t)k^(t+)k(t)\to\hat{k}\leavevmode\nobreak\ (t\to+\infty). This completes the proof. ∎

We can use the same techniques above to obtain the Theorem 1.12.

10 Acknowledgments

Guangming Hu is supported by NSF of China (No. 12101275). Ziping Lei is supported by NSF of China (No. 12122119). Puchun Zhou is supported by Shanghai Science and Technology Program [Project No. 22JC1400100].

References

  • [1] H. Cao, X. Zhu, A Complete Proof of the Poincaré and Geometrization Conjectures - application of the Hamilton-Perelman theory of the Ricci flow, Asian J. Math. 10(2), 2006, 165-492.
  • [2] X. X Chen, P. Lu, G. Tian A note on uniformization of Riemann surfaces by Ricci, Proc. Amer. Math. Soc. 134 (2006), 3391-3393
  • [3] B. Chow and F. Luo, Combinatorial Ricci flows on surfaces, J. Differential Geom. 63, 2003, 97–129.
  • [4] H. Ge, B. Hua, and P. Zhou, A combinatorial curvature flow in spherical background geometry. J. Funct. Anal., 286(7), 2024, 110335.
  • [5] R. S. Hamilton, Three-manifolds with positive Ricci curvature, J. Differential Geom., 17(2), 1982, 255–306.
  • [6] X. Nie, On circle patterns and spherical conical metrics, Proc. Amer. Math. Soc., 152 (2024), 843-853.
  • [7] G. Perelman, The entropy formula for the Ricci flow and its geometric applications, arXiv: 0211159.
  • [8] G. Perelman, Ricci flow with surgery on three-manifolds, arXiv: math.DG/0303109.
  • [9] G.Perelman, Finite extinction time for the solutions to the Ricci flow on certain three manifolds, arXiv: math.DG/0307245.
  • [10] A. Takatsu, Convergence of combinatorial ricci flows to degenerate circle patterns, Trans. Amer. Math. Soc., 372(11), 2019, 7597–7617.

Guangming Hu, [email protected]
College of Science, Nanjing University of Posts and Telecommunications, Nanjing, 210003, P.R. China.

Ziping Lei, [email protected]
School of Mathematics, Renmin University of China, Beijing, 100872, P.R. China.

Yu Sun, [email protected]
School of mathematics and physics, Nanjing institute of technology, Nanjing, 211100, P.R. China.

Puchun Zhou, [email protected]
School of Mathematical Sciences, Fudan University, Shanghai, 200433, P.R. China