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

Ollivier Ricci-flow on weighted graphs

Shuliang Bai This author was supported in part by NSFC grant number 12301434.    Yong Lin This author was supported in part by NSFC grant number 12071245.    Linyuan Lu The author was supported in part by NSF grant DMS 2038080.    Zhiyu Wang    Shing-Tung Yau
Abstract

We study the existence of solutions of Ricci flow equations of Ollivier-Lin-Lu-Yau curvature defined on weighted graphs. Our work is motivated by[13] in which the discrete time Ricci flow algorithm has been applied successfully as a discrete geometric approach in detecting communities. Our main result is the existence and uniqueness theorem for solutions to a continuous time normalized Ricci flow. We also display possible solutions to the Ricci flow on path graph and prove the Ricci flow on finite star graph with at least three leaves converges to constant-weighted star.

Keywords: Ricci flow, Discrete Ricci curvature, Uniqueness problem.

1 Introduction

Ricci flow which was introduced by Richard Hamilton [7] in 1980s is a process that deforms the metric of a Riemannian manifold in a way formally analogous to the diffusion of heat. On Riemannian manifolds MM with a smooth Riemannian metric gg, the geometry of (M,g)(M,g) is altered by changing the metric gg via a second-order nonlinear PDE on symmetric (0,2)(0,2)-tensors:

tgij=2Rij,\displaystyle\frac{\partial}{\partial t}g_{ij}=-2R_{ij}, (1)

where RijR_{ij} is the Ricci curvature. A solution to a Ricci flow is a one-parameter family of metrics g(t)g(t) on a smooth manifold MM, defined on a time interval II, and satisfying equation (1). Intuitively, Ricci flow smooths the metric, but can lead to singularities that can be removed. This procedure is known as surgery. Ricci flow(and surgery) were used in an astonishing manner in the landmark work of Perleman [17] for solving the Poincaré conjecture, as well as in the proof of the differentiable sphere theorem by Simon Brendle and Richard Schoen [4].

One might image such powerful method can be applied to discrete geometry, where objects are irregular complex networks. In 2019, [13] claim good community detection on networks using Ricci flow defined on weighted graphs. For community detection, as one of the fundamental problems in network analysis, one can refer to [6, 9, 16]. The algorithm in [13] makes use of discrete Ricci flow based on Ollivier Ricci curvature which was introduced in [15, 14] and an analogous surgery procedure to partition networks which are modeled as weighted graphs. The discrete Ricci flow deforms edge weights as time progresses: edges of large positive Ricci curvature (i.e., sparsely traveled edges) will shrink and edges of very negative Ricci curvature (i.e., heavily traveled edges) will be stretched. By iterating the Ricci flow process, the heavily traveled edges are identified and thus communities can be partitioned. This approach has successfully detected communities for various networks including Zachary’s Karate Club graph, Network of American football games, Facebook Ego Network, etc. The paper is beautiful in applications but lack of solid mathematical results/theorems. There are several fundamental questions needed to be addressed:

  1. 1.

    What are intrinsic metric/curvature in graphs?

  2. 2.

    Is the solution of Ricci-flow equation always exists? At what domain?

  3. 3.

    If the limit object of the Ricci-flow exist? Do they have constant curvature?

The goal of this paper is to give a mathematical framework so that these essential questions can be answered rigorously.

We start with a brief explanation of the idea of discrete Ricci flow in [13] which contains two parts: the geometric meaning of Ollivier Ricci curvature and the role of Ricci flow.

In the setting of graphs, the Ollivier Ricci curvature is based on optimal transport of probability measures associated to a lazy random walk. To generalize the idea behind Ricci curvature on manifolds to discrete space, the spheres are replaced by probability measures μx,μy\mu_{x},\mu_{y} defined on one-step neighborhood of vertices x,yx,y. Measures will be transported by a distance equal to (1κxy)d(x,y)(1-\kappa_{xy})d(x,y), where κxy\kappa_{xy} represents the Ollivier curvature along the geodesic segment xyxy. A natural choice for the distance between measures μx,μy\mu_{x},\mu_{y} is the Wasserstein transportation metric W1W_{1}. Therefore, the Ollivier’s curvature is defined as: κxy=1W1(μx,μy)d(x,y)\kappa_{xy}=1-\frac{W_{1}(\mu_{x},\mu_{y})}{d(x,y)}. By this notion, positive Ollivier Ricci curvature implies that the neighbors of the two centers are close or overlapping, negative Ollivier Ricci curvature implies that the neighbors of two centers are further apart, and zero Ollivier Ricci curvature or near-zero curvature implies that the neighbors are locally embeddable in a flat surface. The Ollivier Ricci curvature can be generalized to weighted graph (V,E,w)(V,E,w) where ww indicates the edge weights. There has been various generalized versions, while the probability measure may vary the essence using optimal transport theory remains unchanged.

The discrete Ricci flow algorithm in [13] on weighted graphs is an evolving process. In each iteration, the Ricci flow process generate a time dependent family of weighted graph (V,E,w(t))(V,E,w(t)) such that the weight wij(t)w_{ij}(t) on edge ijij changes proportional to the Ollivier Ricci curvature κij(t)\kappa_{ij}(t) at edge ijij at time tt. Ollivier[14] suggested to use the following formula for Ricci flow with continuous time parameter tt:

ddtwe(t)=κe(t)we(t),\displaystyle\frac{d}{dt}w_{e}(t)=-\kappa_{e}(t)w_{e}(t), (2)

where eEe\in E, κe\kappa_{e} represents the Ollivier Ricci curvature on ee and wew_{e} indicates the length of edge ee. Then [13] uses the following formula for Ricci flow with discrete time tt:

wij(t+1)=(1ϵκij(t))dt(i,j),\displaystyle w_{ij}(t+1)=(1-\epsilon\kappa_{ij}(t))d^{t}(i,j), (3)

where dt(i,j)d^{t}(i,j) is the associated distance at time tt, i.e the shortest path length between i,ji,j and κij(t)\kappa_{ij}(t) is the Ollivier Ricci curvature on edge (i,j)(i,j) at time tt. Observe such an iteration process, the Ricci flow enlarges the weights on negatively curvatured edges and shrink the weights on positively curvatured edges over time. By iterating the Ricci flow process, edges with high weights are detected so that can be removed by a surgery processes. As a result, the network is naturally partitioned into different communities with relatively large Ricci curvature.

Motivated by work in [13], we propose a theoretic framework for the Ricci flow equations defined on weighted graphs. Since the Ricci flow (2 ) does not preserve the sum of edge length of GG, which would possibly lead to that the graph becomes infinitesimal in the limit if the initial metric satisfies a certain conditions, see an example in Section 4.2. To avoid this, we define the normalized Ricci flow:

ddtwe(t)=κe(t)we(t)+we(t)hE(G)κh(t)wh(t).\displaystyle\frac{d}{dt}w_{e}(t)=-\kappa_{e}(t)w_{e}(t)+w_{e}(t)\displaystyle\sum_{h\in E(G)}\kappa_{h}(t)w_{h}(t). (4)

Here we adopt the Ollivier-Lin-Lu-Yau’s Ricci curvature[10], which is the limit version of Ollivier Ricci curvature. Under this normalized flow, which is equivalent to the unnormalized Ricci flow (2) by scaling the metric in space by a function of tt, the sum of edge length of the solution metric is 11 in time. To see this, let w(t)={we0(t),,wem(t)}\vec{w}(t)=\{w_{e_{0}}(t),\ldots,w_{e_{m}}(t)\} be a solution of the unnormalized equation, let ϕ(t)\phi(t) be a function of time tt and ϕ(t)>0\phi(t)>0. Set w~(t)=ϕ(t)w(t)\vec{\tilde{w}}(t)=\phi(t)\vec{w(t)} and ewe~(t)=1\sum_{e}\tilde{w_{e}}(t)=1, then ϕ(t)=1hwh(t)\phi(t)=\frac{1}{\sum_{h}w_{h}(t)}. Note the edge curvature κ\kappa does not change under a scaling of the metric. Thus κe~=κe\tilde{\kappa_{e}}=\kappa_{e} for all edges eEe\in E. Let t~=t\tilde{t}=t, then

ddt~we~(t)\displaystyle\frac{d}{d\tilde{t}}\tilde{w_{e}}(t) =dϕ(t)we(t)dtdtdt~=(dϕ(t)dtwe(t)+dwe(t)dtϕ(t))×1\displaystyle=\frac{d\phi(t)w_{e}(t)}{dt}\frac{dt}{d\tilde{t}}=\Big{(}\frac{d\phi(t)}{dt}w_{e}(t)+\frac{dw_{e}(t)}{dt}\phi(t)\Big{)}\times 1
=1(hwh(t))2hdwhdtwe(t)κe(t)we(t)ϕ(t)\displaystyle=-\frac{1}{(\sum_{h}w_{h}(t))^{2}}\sum_{h}\frac{dw_{h}}{dt}w_{e}(t)-\kappa_{e}(t)w_{e}(t)\phi(t)
=we~(t)hE(G)κh~(t)wh~(t)κe~(t)we~(t),\displaystyle=\tilde{w_{e}}(t)\displaystyle\sum_{h\in E(G)}\tilde{\kappa_{h}}(t)\tilde{w_{h}}(t)-\tilde{\kappa_{e}}(t)\tilde{w_{e}}(t),

where the last equation is obtained by replacing we(t)w_{e}(t) by 1ϕ(t)we~(t)\frac{1}{\phi(t)}\tilde{w_{e}}(t) for all edges ee.

On the other side, let w~(t),w(t)\tilde{w}(t),w(t) be solutions of the normalized equation and unnormalized equation respectively, we show that for each edge ee, we=we~hwh(t)w_{e}=\tilde{w_{e}}\sum_{h}w_{h}(t). It suffices to show that we~hwh(t)\tilde{w_{e}}\sum_{h}w_{h}(t) satisfies equation (2).

ddtwe~hwh(t)\displaystyle\frac{d}{dt}\tilde{w_{e}}\sum_{h}w_{h}(t) =hwh(t)ddtwe~(t)+we~(t)dhwh(t)dt\displaystyle=\sum_{h}w_{h}(t)\frac{d}{dt}\tilde{w_{e}}(t)+\tilde{w_{e}}(t)\frac{d\sum_{h}w_{h}(t)}{dt}
=hwh(t)(we~(t)hE(G)κh~(t)wh~(t)κe~(t)we~(t))+we~(t)(hE(G)κh(t)wh(t))\displaystyle=\sum_{h}w_{h}(t)\Big{(}\tilde{w_{e}}(t)\displaystyle\sum_{h\in E(G)}\tilde{\kappa_{h}}(t)\tilde{w_{h}}(t)-\tilde{\kappa_{e}}(t)\tilde{w_{e}}(t)\Big{)}+\tilde{w_{e}}(t)(-\displaystyle\sum_{h\in E(G)}\kappa_{h}(t)w_{h}(t))
=κe(t)we(t).\displaystyle=-\kappa_{e}(t)w_{e}(t).

Thus, there is a bijection between solutions of the unnormalized and normalized Ricci flow equations.

In Riemannian manifolds, with the establishment of Ricci flow equations, one of important work is to verify whether this equation always has a unique smooth solution, at least for a short time, on any compact manifold of any dimension for any initial value of the metric. In this paper, we study the problem of the existence and uniqueness of solutions and convergence results to the normalized Ricci flow (4) on connected weighted graphs. The difficulty of the problem lies in that there is no explicit expression for κe(t)\kappa_{e}(t), although κe\kappa_{e} can be written in terms of the infimum of distance-based 1-lipschitz function, there is no common optimal 1-lipschitz function for all edges ee, thus it is not easy to estimate the derivative of the right-hand side of Ricci flow. The main Theorem 3 of this paper proves the long-time existence and uniqueness of solutions for the initial value problem involved Ricci flow equations (4) provided that each edge is always the shortest path connecting its endpoints over time. This theorem also implies a same result for the unnormalized Ricci flow (2). We also prove that several convergence results of Ricci flow on path and star graphs. Our results display different possible solutions of Ricci flow for the path of length 2, a graph minor of any finite path resulted by the Ricci flow accompanied with edge operations (see Theorem 4) and prove that Ricci flow on star graph can deforms any initial metric to a constant-curvatured metric (see Theorem 5).

The paper is organized as follows. In section 2, we introduce the notion of Ollivier-Lin-Lu-Yau Ricci curvature defined on weighted graphs and related lemmas; in section 3, we introduce the Ricci flow equation and prove our main theorem; in section 4.2, we mainly apply the Ricci flow on a tree graph and show different convergence results of normalized Ricci flow.

2 Notations and lemmas

Let G=(V,E,w)G=(V,E,w) be a weighted graph on vertex set VV associated by the edge weight function w:E[0,)w:E\to[0,\infty). For any two vertices x,yx,y, we write xyxy or xyx\sim y to represent an edge e=(x,y)e=(x,y), wxyw_{xy} is always positive if xyx\sim y. For any vertex xVx\in V, denote the neighbors of xx as N(x)N(x) and the degree of xx as dxd_{x}. The length of a path is the sum of edge lengths on the path, for any two vertices x,yx,y, the distance d(x,y)d(x,y) is the length of a minimal weighted path among all paths that connect xx and yy. We call GG a combinatorial graph if wxy=1w_{xy}=1 for xyx\sim y, wxy=0w_{xy}=0 for x≁yx\not\sim y. Next we recall the definition of Ricci curvature defined on weighted graphs.

Definition 1.

A probability distribution over the vertex set V(G)V(G) is a mapping μ:V[0,1]\mu:V\to[0,1] satisfying xVμ(x)=1\sum_{x\in V}\mu(x)=1. Suppose that two probability distributions μ1\mu_{1} and μ2\mu_{2} have finite support. A coupling between μ1\mu_{1} and μ2\mu_{2} is a mapping A:V×V[0,1]A:V\times V\to[0,1] with finite support such that

yVA(x,y)=μ1(x)andxVA(x,y)=μ2(y).\sum\limits_{y\in V}A(x,y)=\mu_{1}(x)\ \text{and}\sum\limits_{x\in V}A(x,y)=\mu_{2}(y).
Definition 2.

The transportation distance between two probability distributions μ1\mu_{1} and μ2\mu_{2} is defined as follows:

W(μ1,μ2)=infAx,yVA(x,y)d(x,y),W(\mu_{1},\mu_{2})=\inf_{A}\sum_{x,y\in V}A(x,y)d(x,y),

where the infimum is taken over all coupling AA between μ1\mu_{1} and μ2\mu_{2}.

By the theory of linear programming, the transportation distance is also equal to the optimal solution of its dual problem. Thus, we also have

W(μ1,μ2)=supfxVf(x)[μ1(x)μ2(x)]W(\mu_{1},\mu_{2})=\sup_{f}\sum_{x\in V}f(x)[\mu_{1}(x)-\mu_{2}(x)]

where ff is 11-Lipschitz function satisfying

|f(x)f(y)|d(x,y)for x,yV(G) .|f(x)-f(y)|\leq d(x,y)\ \textrm{for $\forall x,y\in V(G)$ }.
Definition 3.

[14][10][1] Let G=(V,E,w)G=(V,E,w) be a weighted graph where the distance dd is determined by the weight function ww. For any x,yVx,y\in V and α[0,1]\alpha\in[0,1], the α\alpha-Ricci curvature κα\kappa_{\alpha} is defined to be

κα(x,y)=1W(μxα,μyα)d(x,y),\kappa_{\alpha}(x,y)=1-\frac{W(\mu_{x}^{\alpha},\mu_{y}^{\alpha})}{d(x,y)},

where the probability distribution μxα\mu_{x}^{\alpha} is defined as:

μxα(y)={α,if y=x,(1α)γ(wxy)zxγ(wxz),if yx,0,otherwise,\mu_{x}^{\alpha}(y)=\begin{cases}\alpha,&\text{if $y=x$},\\ \displaystyle(1-\alpha)\frac{\gamma(w_{xy})}{\sum_{z\sim x}\gamma(w_{xz})},&\text{if $y\sim x$},\\ 0,&\text{otherwise},\end{cases}

where γ:++\gamma:\mathbb{R}_{+}\to\mathbb{R}_{+} represents an arbitrary one-to-one function. The Lin-Lu-Yau’s Ollivier Ricci curvature κ(x,y)\kappa(x,y) is defined as

κ(x,y)=limα1κα(x,y)1α.\kappa(x,y)=\lim\limits_{\alpha\to 1}\frac{\kappa_{\alpha}(x,y)}{1-\alpha}.

It is clear that curvature κ\kappa is a continuous function about the weight function ww. On combinatorial graphs, the probability distribution μxα\mu_{x}^{\alpha} is uniform on xx’s neighbors, the above limit expression for Lin-Lu-Yau’s Ollivier curvature is studied in [10, 3] and it turned out that function κα:α\kappa_{\alpha}:\alpha\to\mathbb{R} is a piece-wise linear function with at most three pieces. Therefore one can calculate κ\kappa easily by choosing a large enough value of α\alpha. On weighted graphs, the probability distribution μxα\mu_{x}^{\alpha} is determined by weight ww and function γ\gamma, the distance dd involved is reflected directly by ww. Some authors used a combinatorial distance dd which measures the number of edges in the shorted path instead of the weighted version. For instance, in [12], the authors [12] also simplify the limit expression of κ(x,y)\kappa(x,y) to two different limit-free expressions via graph Laplacian and via transport cost. Although the details are different, the curvature definitions are essentially the same. The limit-free expression of κ(x,y)\kappa(x,y) in [12] is still true under our definition.

To state this limit-free curvature expression, we need to rephrase the notion of Laplacian in order to adapt to Definition 3. Let G=(V,w,μ)G=(V,w,\mu) be a weighted graph, let ff represent a function in {f:V}\{f:V\to\mathbb{R}\}. The gradient of ff is defined by

xyf=f(x)f(y)d(x,y)for xy\nabla_{xy}f=\frac{f(x)-f(y)}{d(x,y)}\ \ \mbox{for $x\neq y$. }

According to Definition 3, the graph Laplacian Δ\Delta is defined via:

Δf(x)=1zxγ(wxz)yxγ(wxy)(f(y)f(x)),\displaystyle\Delta f(x)=\frac{1}{\sum_{z\sim x}\gamma(w_{xz})}\sum_{y\sim x}\gamma(w_{xy})(f(y)-f(x)), (5)

where f{f:V}f\in\{f:V\to\mathbb{R}\}.

The limit-free formulation of the Lin-Lu-Yau Ricci curvature using graph Laplacian and gradient is as follows.

Theorem 1.

[12] (Curvature via the Laplacian) Let G=(V,w,m)G=(V,w,m) be a weighted graph and let xyV(G)x\neq y\in V(G). Then

κ(x,y)=inffLip(1)yxf=1xyΔf,\kappa(x,y)=\inf_{\begin{subarray}{c}f\in Lip(1)\\ \nabla_{yx}f=1\end{subarray}}\nabla_{xy}\Delta f,

where xyf\nabla_{xy}f is the gradient of ff, dd is the combinatorial graph distance.

Remark 1.

Although in Theorem 1, κ(x,y)\kappa(x,y) is defined with the assumption that dd is the usual combinatorial graph distance, however, the proof of Theorem 1 works verbatim when dd is the weighted distance. Please refer to the detailed proofs in [12].

Using the weighted distance in Definition 3, the limit expression is simplified to another limit-free version via a so called \ast-coupling functions[1]. Let μx:=μx0\mu_{x}:=\mu_{x}^{0} be the probability distribution of random walk at xx with idleness equal to zero. For any two vertices uu and vv, a \ast-coupling between μu\mu_{u} and μv\mu_{v} is a mapping B:V×VB:V\times V\to\mathbb{R} with finite support such that

  1. 1.

    0<B(u,v)0<B(u,v), but all other values B(x,y)0B(x,y)\leq 0.

  2. 2.

    x,yVB(x,y)=0\sum\limits_{x,y\in V}B(x,y)=0.

  3. 3.

    yVB(x,y)=μu(x)\sum\limits_{y\in V}B(x,y)=-\mu_{u}(x) for all xx except uu.

  4. 4.

    xVB(x,y)=μv(y)\sum\limits_{x\in V}B(x,y)=-\mu_{v}(y) for all yy except vv.

Because of items (2),(3), and (4), we get

B(u,v)=(x,y)V×V{(u,v)}B(x,y)xμu(x)+yμv(y)2.B(u,v)=\sum_{(x,y)\in V\times V\setminus\{(u,v)\}}-B(x,y)\leq\sum_{x}\mu_{u}(x)+\sum_{y}\mu_{v}(y)\leq 2.
Theorem 2.

(Curvature via Coupling function)[1] Let G=(V,w,m)G=(V,w,m) be a weighted graph and let u,vV(G)u,v\in V(G) and uvu\neq v. Then

κ(u,v)=1d(u,v)supBx,yVB(x,y)d(x,y),\kappa(u,v)=\frac{1}{d(u,v)}\sup\limits_{B}\sum\limits_{x,y\in V}B(x,y)d(x,y),

where the superemum is taken over all weak \ast-coupling BB between μu\mu_{u} and μv\mu_{v}.

Since 0<B(u,v)0<B(u,v) and B(x,y)0B(x,y)\leq 0. then

κ(u,v)1d(u,v)B(u,v)d(u,v)2.\kappa(u,v)\leq\frac{1}{d(u,v)}B(u,v)d(u,v)\leq 2.

A lower bound of κ(u,v)\kappa(u,v) is obtained by using a result of Lemma 3.2 in [1]. We re-organized this result as follows:

Lemma 1.

Let G=(V,E)G=(V,E) be a weighted graph associated by a edge weight function ww where the maximum value of ww is denoted by D(G)D(G). Let uvVu\neq v\in V. Then

κ(u,v)2D(G)d(u,v).\kappa(u,v)\geq-\frac{2D(G)}{d(u,v)}.
Proof.

For any vertex uVu\in V, let Du=xN(u)γ(wux)D_{u}=\sum_{x\in N(u)}\gamma(w_{ux}). Fix an edge uvE(G)uv\in E(G), we define a function B:V×VB:V\times V\to\mathbb{R} for calculating κ(u,v)\kappa(u,v). For any xN(u){v}x\in N(u)\setminus\{v\}, let B(x,y)=γ(wux)DuB(x,y)=-\frac{\gamma(w_{ux})}{D_{u}} if y=vy=v and 0 otherwise. For any yN(v){u}y\in N(v)\setminus\{u\}, let B(x,y)=γ(wvy)DvB(x,y)=-\frac{\gamma(w_{vy})}{D_{v}} if x=ux=u and 0 otherwise. Let B(v,v)=γ(wuv)DuB(v,v)=-\frac{\gamma(w_{uv})}{D_{u}}, B(u,u)=γ(wuv)DvB(u,u)=-\frac{\gamma(w_{uv})}{D_{v}}, and B(u,v)=2B(u,v)=2. The rest of entries are set to 0. It is straightforward to verify the following results:

x,yVB(x,y)=0\sum\limits_{x,y\in V}B(x,y)=0; yVB(x,y)=μu(x)\sum\limits_{y\in V}B(x,y)=-\mu_{u}(x) for all xx except uu; xVB(x,y)=μv(y)\sum\limits_{x\in V}B(x,y)=-\mu_{v}(y) for all yy except vv.

Thus BB is \ast-coupling between μu\mu_{u} and μv\mu_{v}. By Theorem 2, we have

κ(u,v)1d(u,v)x,yVB(x,y)d(x,y)=2xN(u){v}γ(wux)Dud(x,v)d(u,v)yN(v){u}γ(wvy)Dvd(u,y)d(u,v)2xN(u){v}γ(wux)Dud(x,u)+d(u,v)d(u,v)yN(v){u}γ(wvy)Dvd(y,v)+d(u,v)d(u,v)=2xN(u){v}γ(wux)DuyN(v){u}γ(wvy)DvxN(u){v}γ(wux)Dud(x,u)d(u,v)yN(v){u}γ(wvy)Dvd(y,v)d(u,v)=wuvDu+γ(wuv)DvxN(u){v}γ(wux)Dud(x,u)d(u,v)yN(v){u}γ(wvy)Dvd(y,v)d(u,v)=2γ(wuv)Du+2γ(wuv)DvxN(u)γ(wux)Dud(x,u)d(u,v)yN(v)γ(wvy)Dvd(y,v)d(u,v).\displaystyle\begin{split}\kappa(u,v)&\geq\frac{1}{d(u,v)}\sum_{x,y\in V}B(x,y)d(x,y)\\ &=2-\sum_{x\in N(u)\setminus\{v\}}\frac{\gamma(w_{ux})}{D_{u}}\frac{d(x,v)}{d(u,v)}-\sum_{y\in N(v)\setminus\{u\}}\frac{\gamma(w_{vy})}{D_{v}}\frac{d(u,y)}{d(u,v)}\\ &\geq 2-\sum_{x\in N(u)\setminus\{v\}}\frac{\gamma(w_{ux})}{D_{u}}\frac{d(x,u)+d(u,v)}{d(u,v)}-\sum_{y\in N(v)\setminus\{u\}}\frac{\gamma(w_{vy})}{D_{v}}\frac{d(y,v)+d(u,v)}{d(u,v)}\\ &=2-\sum_{x\in N(u)\setminus\{v\}}\frac{\gamma(w_{ux})}{D_{u}}-\sum_{y\in N(v)\setminus\{u\}}\frac{\gamma(w_{vy})}{D_{v}}-\sum_{x\in N(u)\setminus\{v\}}\frac{\gamma(w_{ux})}{D_{u}}\frac{d(x,u)}{d(u,v)}\\ &\quad\quad-\sum_{y\in N(v)\setminus\{u\}}\frac{\gamma(w_{vy})}{D_{v}}\frac{d(y,v)}{d(u,v)}\\ &=\frac{w_{uv}}{D_{u}}+\frac{\gamma(w_{uv})}{D_{v}}-\sum_{x\in N(u)\setminus\{v\}}\frac{\gamma(w_{ux})}{D_{u}}\frac{d(x,u)}{d(u,v)}-\sum_{y\in N(v)\setminus\{u\}}\frac{\gamma(w_{vy})}{D_{v}}\frac{d(y,v)}{d(u,v)}\\ &=\frac{2\gamma(w_{uv})}{D_{u}}+\frac{2\gamma(w_{uv})}{D_{v}}-\sum_{x\in N(u)}\frac{\gamma(w_{ux})}{D_{u}}\frac{d(x,u)}{d(u,v)}-\sum_{y\in N(v)}\frac{\gamma(w_{vy})}{D_{v}}\frac{d(y,v)}{d(u,v)}.\end{split} (6)

Let D(G)D(G) denote the maximum edge length of GG, i.e. d(x,y)D(G)d(x,y)\leq D(G) for all xyx\sim y. Then

κ(u,v)D(G)d(u,v)(xuγ(wux)Du+yvγ(wvy)Dv)=D(G)d(u,v)×2.\displaystyle\begin{split}\kappa(u,v)&\geq-\frac{D(G)}{d(u,v)}\Big{(}\sum_{x\sim u}\frac{\gamma(w_{ux})}{D_{u}}+\sum_{y\sim v}\frac{\gamma(w_{vy})}{D_{v}}\Big{)}=-\frac{D(G)}{d(u,v)}\times 2.\end{split} (7)

3 Continuous Ricci flow process

In this section, we will describe a continuous Ricci flow process on weighted graphs and prove that this Ricci flow has a unique solution that exists for all time.

Let κ:E(G)|E(G)|\kappa:E(G)\to\mathbb{R}^{|E(G)|} be the Ollivier-Lin-Lu-Yau curvature function defined on a weighted graph G=(V,w,μ)G=(V,w,\mu) where ww is the weight function on the edge set of GG and μ={μx:xV(G)}\mu=\{\mu_{x}:x\in V(G)\} be probability distribution function such that for each xV(G)x\in V(G),

μxα(y)={α if x=y,(1α)γ(wxy)zN(x)γ(wxz) if yN(x),0 otherwise,\mu^{\alpha}_{x}(y)=\begin{cases}\alpha&\textrm{ if $x=y$},\\ (1-\alpha)\frac{\gamma(w_{xy})}{\sum_{z\in N(x)}\gamma(w_{xz})}&\textrm{ if $y\in N(x)$},\\ 0&\textrm{ otherwise},\end{cases} (8)

where γ:++\gamma:\mathbb{R}_{+}\to\mathbb{R}_{+} is a Lipschitz function over [δ,1][\delta,1] for all δ>0\delta>0.

Let X(t)=we1(t),we2(t),,wem(t)+mX(t)=\langle w_{e_{1}}(t),w_{e_{2}}(t),\ldots,w_{e_{m}}(t)\rangle\in\mathbb{R}_{+}^{m} where t[0,)t\in[0,\infty) and m=|E(G)|m=|E(G)| denotes the number of edges of GG. Let X0++mX_{0}\in\mathbb{R}_{++}^{m} be an arbitrary vector we1(0),we2(0),,wem(0)\langle w_{e_{1}}(0),w_{e_{2}}(0),\ldots,w_{e_{m}}(0)\rangle with i=1mwei(0)=1\sum_{i=1}^{m}w_{e_{i}}(0)=1. Note X0X_{0} gives the initial metric w(0)\vec{w}(0) for graph GG with all wei(0)w_{e_{i}}(0) strictly greater than zero. We then define a system of ordinary differential equations as follows:

{dwedt=κewe+wehE(G)κhwhX(0)=X0.\begin{cases}\frac{dw_{e}}{dt}=-\kappa_{e}w_{e}+w_{e}\sum_{h\in E(G)}\kappa_{h}w_{h}\\ X(0)=X_{0}.\end{cases} (9)

Now we introduce the continuous Ricci flow process as follows:

Input: An undirected graph GG, merge threshold mt>0mt>0, and termination threshold δ>0\delta>0.
Output: A collection of vertex-disjoint minors of GG, which are the ‘clusters’ if GG is viewed as a network.
1
2Set hierarchy level to be 1.
3Let w\vec{w} be the solution to the system of ODE described in (9) with the exit condition:(I) wuv(t)>d(u,v)(t)w_{uv}(t)>d(u,v)(t) for some uvE(G)uv\in E(G) and some t[0,)t\in[0,\infty); (II) wuv(t)<mtw_{uv}(t)<mt for some uvE(G)uv\in E(G) and some t[0,)t\in[0,\infty);If Condition (I) is met, delete the edge uvuv and restart step 22;If Condition (II) is met, contract the edge uvuv and restart step 22.If w\vec{w} is a non-chaotic solution, go to Step 33. If w\vec{w} is chaotic, slightly perturb the initial conditions.
Let GG^{\prime} be the resulting graph from Step 2. (I) Label each edge of GG^{\prime} with the current hierarchy level.(II) Increase the hierarchy level by 1. Perform Step 22 on GG^{\prime}.
Algorithm 1 Continuous Ricci flow process

We make three observations:

First, there is no edge getting a zero weight at any time tt during the whole Ricci flow process. In the initial weighted graph GG, we(0)0w_{e}(0)\neq 0 for all edges eE(G)e\in E(G). Fix an edge ee, the right-hand side of (9) is bounded below by we(22|E(G)|)w_{e}(-2-2|E(G)|) according to Lemma 1, then we(t)>we(0)e(22|E(G)|)tw_{e}(t)>w_{e}(0)e^{(-2-2|E(G)|)t} which is always positive at finite time.

Second, each edge has a weight assigned to it over time, theoretically, we don’t know if there is an edge meeting the exit condition (I), that is, wuv(t)>d(u,v)(t)w_{uv}(t)>d(u,v)(t) for some uvE(G)uv\in E(G). To fix this possible barrier we choose to delete such edges, notice that the resulting graph is still connected. In addition, the only reason for the reduction in the number of vertices is the exit condition (II). Thus, GG will not degenerate to a point.

Last, since i=1mwei(0)=1\sum_{i=1}^{m}w_{e_{i}}(0)=1, under the assumption that no edges meeting two exit conditions, we claim that the property i=1mwei=1\sum_{i=1}^{m}w_{e_{i}}=1 is always maintained. To see this, let T(t)=hE(Gt)wh(t)T(t)=\sum_{h\in E(G^{t})}w_{h}(t), where GtG^{t} is the resulting graph at time tt. Sum up both sides of equation (9) over all edges of GtG^{t}, we have

dT(t)dt=(T(t)1)hE(Gt)κh(t)wh(t).\frac{dT(t)}{dt}=(T(t)-1)\sum_{h\in E(G^{t})}\kappa_{h}(t)w_{h}(t).

By Theorem 1, the right hand side is a bounded value for all tt, it follows then that T(t)1T(t)-1 has the following form:

T(t)1=ce0t(hE(Gs)κh(s)wh(s))𝑑s,T(t)-1=ce^{\displaystyle\int_{0}^{t}(\sum_{h\in E(G^{s})}\kappa_{h}(s)w_{h}(s))ds},

where cc is a constant depending on T(0)T(0). Since T(0)=1T(0)=1, then c=0c=0 implies T(t)=1T(t)=1 for all t0t\geq 0, done. Therefore, in algorithm 1, for all time tt the sum of weight is at most 11, and for each edge ee, 0<we(t)10<w_{e}(t)\leq 1. In order to remain the sum of weight constant 11, an alternative approach is re-normalize the edge weight after each Ricci flow iteration so that the sum of weight always remains 11, but sum of weight being 11 or not does not affect the validity of the following theorem.

3.1 Existence and uniqueness of the solution

Theorem 3.

For any initial weighted graph GG, by fixing the exit condition (I), there exists a unique solution X(t)X(t), for all time t[0,)t\in[0,\infty), to the system of ordinary differential equations in (9).

Before we prove Theorem 3, we first need some lemmas. By the exit condition (I) stated in above algorithm, once wuv(t)>d(u,v)(t)w_{uv}(t)>d(u,v)(t) for some uvE(G)uv\in E(G) we will delete the edge uvuv, thus wxyw_{xy} always represent the length of edge uvuv. For convenience, we use wxyw_{xy} instead of d(x,y)d(x,y) to represent the distance between any pair of vertices xx and yy.

Lemma 2.

Let G=(V,E,w)G=(V,E,w) be a weighted graph and x,yx,y be two fixed vertices in GG. For any 11-Lipschitz function ff defined on GG and such that 0<f(y)f(x)<wxy0<f(y)-f(x)<w_{xy}, there exists a 11-Lipschitz function ff^{\prime}, such that f(y)f(x)=wxyf^{\prime}(y)-f^{\prime}(x)=w_{xy} and f(z)f(z)wxy(f(y)f(x))f^{\prime}(z)-f(z)\leq w_{xy}-(f(y)-f(x)) for all zVz\in V.

Proof.

Define function gg on GG such that

{g(y)=f(x)+wxy,g(z)=f(z)zV{y}.\begin{cases}g(y)=f(x)+w_{xy},\\ g(z)=f(z)&z\in V\setminus\{y\}.\end{cases}

We have

g(y)g(x)=wxy,g(y)-g(x)=w_{xy},
g(y)f(y)=wxy(f(y)f(x)),g(y)-f(y)=w_{xy}-(f(y)-f(x)),
g(z)f(z)=0wxy(f(y)f(x))zy.g(z)-f(z)=0\leq w_{xy}-(f(y)-f(x))\ \forall z\neq y.

Thus, if gg is 11-Lipschitz on GG, let f=gf^{\prime}=g, we are done.

If gg is not 11-Lipschitz, according to its definition, it then fails at vertex yy and some vertex in V{x,y}V\setminus\{x,y\}, denote such vertices as v1,v2,,vtv_{1},v_{2},\ldots,v_{t} in the order that g(v1)g(v2)g(vt)<g(y)g(v_{1})\leq g(v_{2})\leq\cdots\leq g(v_{t})<g(y). We have |g(vj)g(y)|>wyvj|g(v_{j})-g(y)|>w_{yv_{j}} for each jtj\leq t. That is, either g(vj)g(y)>wyvjg(v_{j})-g(y)>w_{yv_{j}} or g(vj)g(y)<wyvjg(v_{j})-g(y)<-w_{yv_{j}}. Note g(vj)g(y)=g(vj)f(x)wxywxvjwxywyvjg(v_{j})-g(y)=g(v_{j})-f(x)-w_{xy}\leq w_{xv_{j}}-w_{xy}\leq w_{yv_{j}}, thus, it has to be the latter case, that is, g(vj)g(y)<wyvjg(v_{j})-g(y)<-w_{yv_{j}} for all jtj\leq t.

Note at any pair of vertices out of NN, gg is 1-Lipschitz. Further, at vertex xx and vertex vjv_{j}, we have g(vj)g(x)=g(vj)g(y)+wxy<wyvj+wxywxvjg(v_{j})-g(x)=g(v_{j})-g(y)+w_{xy}<-w_{yv_{j}}+w_{xy}\leq w_{xv_{j}}, thus, g(vj)g(x)g(v_{j})-g(x) is strictly less than wxvjw_{xv_{j}}.

Now we create a new function gg^{\prime} from gg so that gg^{\prime} is 1-Lipschitz on GG. Let

{g(vj)=g(vj)+ajjt,g(u)=g(u)otherwise,\begin{cases}g^{\prime}(v_{j})=g(v_{j})+a_{j}&j\leq t,\\ g^{\prime}(u)=g(u)&\text{otherwise},\end{cases}

where values aja_{j} will be chosen from internal

Lj=[g(y)g(vj)wyvj,min{wxy(f(y)f(x)),minzN{wzvj+g(z)g(vj)}}].L_{j}=\Big{[}g(y)-g(v_{j})-w_{yv_{j}},\min\{w_{xy}-(f(y)-f(x)),\ \min_{z\not\in N}\{w_{zv_{j}}+g(z)-g(v_{j})\}\}\Big{]}.

One can check that internal LjL_{j} is non empty and aja_{j} is positive. Further, let aja_{j}s satisfy 0<ajai<g(vi)g(vj)+wvivj0<a_{j}-a_{i}<g(v_{i})-g(v_{j})+w_{v_{i}v_{j}} for all 1j<it1\leq j<i\leq t. Note we are able to achieve this purpose by choosing value of aja_{j} as large as possible in the reverse order (i.e., from ata_{t} to a1a_{1}). Next, we will confirm that gg^{\prime} is 1-Lipschitz on GG.

  • For yy and each vjv_{j}, g(vj)g(y)=g(vj)+ajg(x)wxy<g(vj)+wxyf(y)+f(x)g(x)wxy<g(vj)f(y)<wyvjg^{\prime}(v_{j})-g^{\prime}(y)=g(v_{j})+a_{j}-g(x)-w_{xy}<g(v_{j})+w_{xy}-f(y)+f(x)-g(x)-w_{xy}<g(v_{j})-f(y)<w_{yv_{j}}, thus, gg^{\prime} is 1-Lipschitz at yy and vjv_{j}.

  • For viv_{i} and vjv_{j}, j<ij<i, g(vj)g(vi)=g(vj)g(vi)+ajai<g(vj)g(vi)+g(vi)g(vj)+wvivj<wvivjg^{\prime}(v_{j})-g^{\prime}(v_{i})=g(v_{j})-g(v_{i})+a_{j}-a_{i}<g(v_{j})-g(v_{i})+g(v_{i})-g(v_{j})+w_{v_{i}v_{j}}<w_{v_{i}v_{j}} and g(vj)g(vi)=g(vj)g(vi)+ajai>g(vj)g(vi)>wvjvig^{\prime}(v_{j})-g^{\prime}(v_{i})=g(v_{j})-g(v_{i})+a_{j}-a_{i}>g(v_{j})-g(v_{i})>-w_{v_{j}v_{i}}. Thus, gg^{\prime} is 1-Lipschitz at viv_{i} and vjv_{j}.

  • For zNz\not\in N and each vjv_{j}’s. g(vj)g(z)=g(vj)+ajg(z)g(vj)g(z)+wzvj+g(z)g(vj)wzvjg^{\prime}(v_{j})-g^{\prime}(z)=g(v_{j})+a_{j}-g(z)\leq g(v_{j})-g(z)+w_{zv_{j}}+g(z)-g(v_{j})\leq w_{zv_{j}} and g(vj)g(z)=g(vj)+ajg(z)>g(vj)g(z)>wzvjg^{\prime}(v_{j})-g^{\prime}(z)=g(v_{j})+a_{j}-g(z)>g(v_{j})-g(z)>-w_{zvj}. Thus, function gg^{\prime} is 1-Lipschitz at zz and vjv_{j}’s.

  • For u,vNu,v\notin N, |g(u)g(v)|=|g(u)g(v)|wuv|g^{\prime}(u)-g^{\prime}(v)|=|g(u)-g(v)|\leq w_{uv}. Thus, function gg^{\prime} is 1-Lipschitz out of NN.

To sum up, there exist positive values aia_{i} so that gg^{\prime} is 1-Lip between all pairs of vertices of GG. One can also check that gg^{\prime} satisfies inequalities stated in the lemma:

g(y)g(x)=f(y)+wxyg(x)=wxy,g^{\prime}(y)-g^{\prime}(x)=f(y)+w_{xy}-g(x)=w_{xy},
g(vi)f(vi)=aiwxy(f(y)f(x)),jt,g^{\prime}(v_{i})-f(v_{i})=a_{i}\leq w_{xy}-(f(y)-f(x)),\ j\leq t,
g(z)f(z)=0<wxy(f(y)f(x))zN,g^{\prime}(z)-f(z)=0<w_{xy}-(f(y)-f(x))\ z\notin N,

Let f=gf^{\prime}=g^{\prime}, the proof is complete.

A very similar proof gives the following result:

Lemma 3.

Let G=(V,E,w)G=(V,E,w) be a weighted graph and x,yx,y be two fixed vertices in GG. Let 0<a<wxy0<a<w_{xy}. For any 11-Lipschitz function ff^{\prime} defined on GG such that f(y)f(x)=wxyf^{\prime}(y)-f^{\prime}(x)=w_{xy}, there exists a 11-Lipschitz function ff, such that f(y)f(x)=af(y)-f(x)=a and |f(z)f(z)|wxya|f^{\prime}(z)-f(z)|\leq w_{xy}-a for all zVz\in V.

Proof.

Define function gg on GG such that

{g(y)=f(x)+a,g(z)=f(z)zV{y}.\begin{cases}g(y)=f^{\prime}(x)+a,\\ g(z)=f^{\prime}(z)&z\in V\setminus\{y\}.\end{cases}

We have

g(y)g(x)=a,g(y)-g(x)=a,
g(y)f(y)=wxya,g(y)-f^{\prime}(y)=w_{xy}-a,
g(z)f(z)=0wxyazy.g(z)-f^{\prime}(z)=0\leq w_{xy}-a\ \forall z\neq y.

Thus, if gg is 11-Lipschitz, let f=gf=g, we are done.

If gg is not 11-Lipschitz, then there exists zx,yz\neq x,y so that |g(z)g(y)|>wyz|g(z)-g(y)|>w_{yz} and one can check that it is the case g(z)g(y)>wyzg(z)-g(y)>w_{yz}. Denote such vertices as v1,v2,,vtv_{1},v_{2},\ldots,v_{t} in the order that g(v1)g(v2)g(vt)>g(y)g(v_{1})\geq g(v_{2})\geq\cdots\geq g(v_{t})>g(y). Denote set N={y,v1,v2,,vt}N=\{y,v_{1},v_{2},\ldots,v_{t}\}. Observe that gg is 1-Lip for pair of vertices not in NN and we have |g(vi)g(x)||g(v_{i})-g(x)| not equal to wxviw_{xv_{i}}, as g(vi)g(x)>wyvi+g(y)g(x)=wyvi+a>wxvig(v_{i})-g(x)>w_{yv_{i}}+g(y)-g(x)=w_{yv_{i}}+a>-w_{xv_{i}} and g(vi)g(x)<wyvi+a<wyvi+wxy<wxvig(v_{i})-g(x)<-w_{yv_{i}}+a<-w_{yv_{i}}+w_{xy}<w_{xv_{i}}. Thus adding an appropriate negative value to g(vi)g(v_{i}) will not affect the pair xx and viv_{i}.

Now we create a new function gg^{\prime} from gg so that gg^{\prime} is 1-Lipschitz on GG. Let

{g(vj)=g(vj)ajjt,g(u)=g(u)otherwise,\begin{cases}g^{\prime}(v_{j})=g(v_{j})-a_{j}&j\leq t,\\ g^{\prime}(u)=g(u)&\text{otherwise},\end{cases}

where values aja_{j} will be chosen from internal

Lj=[max{g(vj)g(y)wyvj,g(vi)g(x)wxvi},wxya].L_{j}=\Big{[}\max\{g(v_{j})-g(y)-w_{yv_{j}},g(v_{i})-g(x)-w_{xv_{i}}\},w_{xy}-a\Big{]}.

One can check that internal LjL_{j} is non empty and aja_{j} is positive. Further, let aja_{j}s satisfy 0<ajai<g(vj)g(vi)+wvivj0<a_{j}-a_{i}<g(v_{j})-g(v_{i})+w_{v_{i}v_{j}} for all 1j<it1\leq j<i\leq t, note we are able to achieve this purpose by choosing value of aja_{j} as large as possible in the reverse order (i.e., from vtv_{t} to v1v_{1}). Next, we will confirm that gg^{\prime} is 1-Lipschitz on GG.

  • For yy and each vjv_{j}, g(vj)g(y)=g(vj)ajg(y)<g(vj)g(vj)+g(y)+wyvjg(y)<wyvjg^{\prime}(v_{j})-g^{\prime}(y)=g(v_{j})-a_{j}-g(y)<g(v_{j})-g(v_{j})+g(y)+w_{yv_{j}}-g(y)<w_{yv_{j}}, g(vj)g(y)=g(vj)ajg(y)>f(vj)wxy+af(x)a>wxvjwxy>wyvjg^{\prime}(v_{j})-g^{\prime}(y)=g(v_{j})-a_{j}-g(y)>f^{\prime}(v_{j})-w_{xy}+a-f^{\prime}(x)-a>w_{xv_{j}}-w_{xy}>-w_{yv_{j}}, thus, gg^{\prime} is 1-Lipschitz at yy and vjv_{j}.

  • For viv_{i} and vjv_{j}, j<ij<i, g(vj)g(vi)=g(vj)g(vi)(ajai)<g(vj)g(vi)wvivjg^{\prime}(v_{j})-g^{\prime}(v_{i})=g(v_{j})-g(v_{i})-(a_{j}-a_{i})<g(v_{j})-g(v_{i})\leq w_{v_{i}v_{j}}, and g(vj)g(vi)=g(vj)g(vi)(ajai)>wvjvig^{\prime}(v_{j})-g^{\prime}(v_{i})=g(v_{j})-g(v_{i})-(a_{j}-a_{i})>-w_{v_{j}v_{i}}. Thus, gg^{\prime} is 1-Lipschitz at viv_{i} and vjv_{j}.

  • For zNz\not\in N and each vjv_{j}’s, g(vj)g(z)=g(vj)ajg(z)<g(vj)g(z)wzvjg^{\prime}(v_{j})-g^{\prime}(z)=g(v_{j})-a_{j}-g(z)<g(v_{j})-g(z)\leq w_{zv_{j}} and g(vj)g(z)=g(vj)ajg(z)>g(vj)g(z)g(vj)+g(x)+wxvjwxz+wxvj>wzvjg^{\prime}(v_{j})-g^{\prime}(z)=g(v_{j})-a_{j}-g(z)>g(v_{j})-g(z)-g(v_{j})+g(x)+w_{xv_{j}}\geq-w_{xz}+w_{xv_{j}}>-w_{zvj}. Thus, function gg^{\prime} is still 1-Lipschitz at zz and vjv_{j}’s.

  • For u,vNu,v\notin N, |g(u)g(v)|=|g(u)g(v)|wuv|g^{\prime}(u)-g^{\prime}(v)|=|g(u)-g(v)|\leq w_{uv}. Thus, function gg^{\prime} is 1-Lipschitz out of NN.

To sum up, there exist positive values aia_{i} so that function g(v)g^{\prime}(v) obtained from gg by reducing values {ai}\{a_{i}\} is 1-Lipschitz between all pairs of vertices of GG, and g(y)g(x)=ag^{\prime}(y)-g^{\prime}(x)=a, |g(z)f(z)|<wxya|g^{\prime}(z)-f(z)|<w_{xy}-a for all zVz\in V are satisfied. Let g=fg^{\prime}=f, the proof is complete.

In order to show Theorem 3, we need some classical theorem on the existence and uniqueness of solutions to a system of ordinary differential equations.

Lemma 4.

[5] Suppose that vector-valued function F(t,X)={f1(t,X),,fn(t,X)}F(t,X)=\{f_{1}(t,X),\cdots,f_{n}(t,X)\} is continuous in some n+1n+1 dimensional region:

R={(t,X):|t|a,XX0b},R=\{(t,X):|t|\leq a,\left\lVert X-X_{0}\right\lVert\leq b\},

and is Lipschitz continuous about X=(x1,x2,,xn)X=(x_{1},x_{2},\cdots,x_{n}). Then the the following ODE’s initial value problem

dXdt=F(t,X),X(0)=X0\frac{dX}{dt}=F(t,X),\hskip 14.22636ptX(0)=X_{0}

has a unique solution X=X(t)X=X(t) at region |t|α|t|\leq\alpha, where

α=min{a,bN},N=max(t,X)RF(t,X).\alpha=\min\{a,\frac{b}{N}\},\hskip 14.22636ptN=\max_{(t,X)\in R}\ \left\lVert F(t,X)\right\lVert.

A vector-valued function satisfies a continuous or a Lipschitz condition in a region if and only if its component functions satisfy these conditions in the same region. The following theorem is used to estimate the maximum existence interval of solutions of the following initial value problem:

dydx=f(x,y),y(x0)=y0.\displaystyle\frac{dy}{dx}=f(x,y),\hskip 14.22636pty(x_{0})=y_{0}. (10)

For narrative convenience, we call a function ϕ(x)\phi(x) as right-top solution to (10) if

dϕdx>f(x,y),ϕ(x0)y0.\frac{d\phi}{dx}>f(x,y),\hskip 14.22636pt\phi(x_{0})\geq y_{0}.

And we call a function Φ(x)\Phi(x) as right-bottom solution to (10) if

dΦdx<f(x,y),Φ(x0)y0.\frac{d\Phi}{dx}<f(x,y),\hskip 14.22636pt\Phi(x_{0})\leq y_{0}.
Lemma 5.

[5] Suppose f(x,y)f(x,y) is a continuous function in the region R={(x,y),x0x<b,<y<}R=\{(x,y),x_{0}\leq x<b,-\infty<y<\infty\}, and (x0,y0)R(x_{0},y_{0})\in R. Denote [x0,β1)[x_{0},\beta_{1}) as the maximum existence interval of solution to (10). If (10) has right-top solution ϕ(x)\phi(x) and right-bottom solution Φ(x)\Phi(x) and they have the same interval of solutions [x0,β)[x_{0},\beta), then β1β\beta_{1}\geq\beta.

Now, we are ready to prove Theorem 3.

Proof of Theorem 3.

For a fixed δ>0\delta>0, define

S={w1,w2,,wm:wi>0,i[m]wi1}S=\{\langle w_{1},w_{2},\ldots,w_{m}\rangle:w_{i}>0,\sum_{i\in[m]}w_{i}\leq 1\}

and

Sδ={w1,w2,,wm:wiδ,i[m]wi1}.S_{\delta}=\{\langle w_{1},w_{2},\ldots,w_{m}\rangle:w_{i}\geq\delta,\sum_{i\in[m]}w_{i}\leq 1\}.

We first show that (9) has a unique solution in time interval (0,T)(0,T) for some T>0T>0 in SδS_{\delta} for any positive δ>0\delta>0. Note that

S=δ>0Sδ.S=\displaystyle\bigcup_{\delta>0}S_{\delta}.

It then follows that (9) has a unique solution in SS.

By the existence and uniqueness theorem on systems of ODE, to show (9) has a unique solution in SδS_{\delta}, it suffices to show that κewe\kappa_{e}w_{e} is (uniformly) Lipschitz on SδS_{\delta}.

Let DD be the metric on SδS_{\delta} induced by the \infty-norm, i.e., given w,wSδ\vec{w},\vec{w}^{\prime}\in S_{\delta} with w=w1,,wm\vec{w}=\langle w_{1},\ldots,w_{m}\rangle and w=w1,,wm\vec{w}^{\prime}=\langle w_{1}^{\prime},\ldots,w_{m}^{\prime}\rangle, D(w,w)=maxi[m]|wiwi|D(\vec{w},\vec{w}^{\prime})=\max_{i\in[m]}|w_{i}-w_{i}^{\prime}|. We now show that for a given edge ee, the function μe:wκe(w)we\mu_{e}:\vec{w}\to\kappa_{e}(\vec{w})w_{e} is Lipschitz continuous on SδS_{\delta} equipped with the metric DD.

Fix e=xye=xy. Let w,wSδ\vec{w},\vec{w}^{\prime}\in S_{\delta} be arbitrarily chosen. By Theorem 1,

κ(x,y)=inffLip(1)yxf=1xyΔf.\kappa(x,y)=\displaystyle\inf_{\begin{subarray}{c}f\in Lip(1)\\ \nabla_{yx}f=1\end{subarray}}\nabla_{xy}\Delta f.

Note that |whwh|<ϵ|w^{\prime}_{h}-w_{h}|<\epsilon for any edge hh by our assumption. WLOG that wxy>wxyw^{\prime}_{xy}>w_{xy} and write wxywxy=ϵ0ϵw^{\prime}_{xy}-w_{xy}=\epsilon_{0}\leq\epsilon. Let ff be the function that achieves inf{(Δf(x)Δf(y)):fLip(1),f(y)f(x)=wxy}\inf\{(\Delta f(x)-\Delta f(y)):f\in Lip(1),f(y)-f(x)=w_{xy}\}. Note for these ff, f(y)f(x)=wxy<wxyf(y)-f(x)=w_{xy}<w^{\prime}_{xy}. By Lemma 2, it follows that there exists fLip(1)f^{\prime}\in Lip(1) such that f(y)f(x)=wxyf^{\prime}(y)-f^{\prime}(x)=w^{\prime}_{xy} and |f(z)f(z)|ϵ0|f(z)-f^{\prime}(z)|\leq\epsilon_{0}. Thus,

κewe=inffLip(1)f(y)f(x)=wxy(Δf(x)Δf(y))Δf(x)Δf(y).\kappa_{e}^{\prime}w_{e}^{\prime}=\displaystyle\inf_{\begin{subarray}{c}f\in Lip(1)\\ f(y)-f(x)=w_{xy}^{\prime}\end{subarray}}(\Delta f(x)-\Delta f(y))\leq\Delta f^{\prime}(x)-\Delta f^{\prime}(y).

It follows that if κeweκewe\kappa_{e}^{\prime}w_{e}^{\prime}\geq\kappa_{e}w_{e},

|μe(w)μe(w)|=κeweκewe=inffLip(1)f(y)f(x)=wxy(Δf(x)Δf(y))(Δf(x)Δf(y))(Δf(x)Δf(y))(Δf(x)Δf(y))|Δf(y)Δf(y)|+|Δf(x)Δf(x)|.\displaystyle\begin{split}|\mu_{e}(\vec{w}^{\prime})-\mu_{e}(\vec{w})|&=\kappa_{e}^{\prime}w_{e}^{\prime}-\kappa_{e}w_{e}\\ &=\displaystyle\inf_{\begin{subarray}{c}f\in Lip(1)\\ f(y)-f(x)=w_{xy}^{\prime}\end{subarray}}(\Delta f(x)-\Delta f(y))-\Big{(}\Delta f(x)-\Delta f(y)\Big{)}\\ &\leq(\Delta f^{\prime}(x)-\Delta f^{\prime}(y))-(\Delta f(x)-\Delta f(y))\\ &\leq|\Delta f^{\prime}(y)-\Delta f(y)|+|\Delta f^{\prime}(x)-\Delta f(x)|.\end{split} (11)

While if κeweκewe\kappa_{e}^{\prime}w_{e}^{\prime}\leq\kappa_{e}w_{e}, let gg^{\prime} be the function that achieves

inf{(Δg(x)Δg(y)):gLip(1),g(y)g(x)=wxy}.\inf\{(\Delta g^{\prime}(x)-\Delta g^{\prime}(y)):g^{\prime}\in Lip(1),g^{\prime}(y)-g^{\prime}(x)=w^{\prime}_{xy}\}.

Note for these gg^{\prime}, g(y)g(x)=wxy>wxyg^{\prime}(y)-g^{\prime}(x)=w^{\prime}_{xy}>w_{xy}. By Lemma 3, it follows that there exists gLip(1)g\in Lip(1) such that g(y)g(x)=wxyg(y)-g(x)=w_{xy} and |g(z)g(z)|ϵ0|g(z)-g^{\prime}(z)|\leq\epsilon_{0}. Then

|μe(w)μe(w)|=κeweκewe=inffLip(1)f(y)f(x)=wxy(Δf(x)Δf(y))(Δg(x)Δg(y))(Δg(x)Δg(y))(Δg(x)Δg(y))|Δg(y)Δg(y)|+|Δg(x)Δg(x)|.\displaystyle\begin{split}|\mu_{e}(\vec{w}^{\prime})-\mu_{e}(\vec{w})|&=\kappa_{e}w_{e}-\kappa_{e}^{\prime}w_{e}^{\prime}\\ &=\displaystyle\inf_{\begin{subarray}{c}f\in Lip(1)\\ f(y)-f(x)=w_{xy}\end{subarray}}(\Delta f(x)-\Delta f(y))-\Big{(}\Delta g^{\prime}(x)-\Delta g^{\prime}(y)\Big{)}\\ &\leq(\Delta g(x)-\Delta g(y))-(\Delta g^{\prime}(x)-\Delta g^{\prime}(y))\\ &\leq|\Delta g^{\prime}(y)-\Delta g(y)|+|\Delta g(x)-\Delta g^{\prime}(x)|.\end{split} (12)

Next, we evaluate the right side of inequality (11), the result for (12) is similar and we omit the details.

As |f(z)f(z)|ϵ0|f(z)-f^{\prime}(z)|\leq\epsilon_{0} for all zV(G)z\in V(G), we have f(z)f(y)f(z)f(y)+2ϵ0f^{\prime}(z)-f^{\prime}(y)\leq f(z)-f(y)+2\epsilon_{0} and f(z)f(y)f(z)f(y)2ϵ0f^{\prime}(z)-f^{\prime}(y)\geq f(z)-f(y)-2\epsilon_{0}.

If Δf(y)Δf(y)0\Delta f^{\prime}(y)-\Delta f(y)\geq 0, then

|Δf(y)Δf(y)|\displaystyle|\Delta f^{\prime}(y)-\Delta f(y)| =zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))\displaystyle=\displaystyle\sum_{z\in N(y)}\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}(f^{\prime}(z)-f^{\prime}(y))-\displaystyle\sum_{z\in N(y)}\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}(f(z)-f(y))
zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y)+2ϵ0)zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))\displaystyle\leq\displaystyle\sum_{z\in N(y)}\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}(f(z)-f(y)+2\epsilon_{0})-\displaystyle\sum_{z\in N(y)}\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}(f(z)-f(y))
=2ϵ0+zN(y)(γ(wyz)uN(y)γ(wyu)γ(wyz)uN(y)γ(wyu))(f(z)f(y))\displaystyle=2\epsilon_{0}+\displaystyle\sum_{z\in N(y)}(\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}-\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})})(f(z)-f(y))

If Δf(y)Δf(y)0\Delta f^{\prime}(y)-\Delta f(y)\leq 0, then

|Δf(y)Δf(y)|\displaystyle|\Delta f^{\prime}(y)-\Delta f(y)| =zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))\displaystyle=\displaystyle\sum_{z\in N(y)}\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}(f(z)-f(y))-\displaystyle\sum_{z\in N(y)}\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}(f^{\prime}(z)-f^{\prime}(y))
zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y)2ϵ0)\displaystyle\leq\displaystyle\sum_{z\in N(y)}\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}(f(z)-f(y))-\displaystyle\sum_{z\in N(y)}\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}(f(z)-f(y)-2\epsilon_{0})
=2ϵ0zN(y)(γ(wyz)uN(y)γ(wyu)γ(wyz)uN(y)γ(wyu))(f(z)f(y))\displaystyle=2\epsilon_{0}-\displaystyle\sum_{z\in N(y)}(\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}-\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})})(f(z)-f(y))

Let CC be the Lipschitz constant for the γ\gamma function. Since γ\gamma is a positive Lipschitz continuous function over [δ,1][\delta,1], then there exist M>0M>0 so that γM\gamma\geq M.

For both cases, we have

|Δf(y)Δf(y)|\displaystyle|\Delta f^{\prime}(y)-\Delta f(y)| =|zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))zN(y)γ(wyz)uN(y)γ(wyu)(f(z)f(y))|\displaystyle=\left\lvert\displaystyle\sum_{z\in N(y)}\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}(f^{\prime}(z)-f^{\prime}(y))-\displaystyle\sum_{z\in N(y)}\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}(f(z)-f(y))\right\rvert
2ϵ+zN(y)|γ(wyz)uN(y)γ(wyu)γ(wyz)uN(y)γ(wyu)||f(z)f(y)|\displaystyle\leq 2\epsilon+\displaystyle\sum_{z\in N(y)}|\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}-\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}||f(z)-f(y)|
2ϵ+zN(y)|γ(wyz)uN(y)γ(wyu)γ(wyz)uN(y)γ(wyu)|wyz\displaystyle\leq 2\epsilon+\displaystyle\sum_{z\in N(y)}|\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}-\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}|w_{yz}
2ϵ+zN(y)|γ(wyz)uN(y)γ(wyu)γ(wyz)uN(y)γ(wyu)|\displaystyle\leq 2\epsilon+\displaystyle\sum_{z\in N(y)}\left\lvert\frac{\gamma(w^{\prime}_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})}-\frac{\gamma(w_{yz})}{\displaystyle\sum_{u\in N(y)}\gamma(w_{yu})}\right\rvert
2ϵ+zN(y)|γ(wyz)γ(wyz)|min(uN(y)γ(wyu),uN(y)γ(wyu))\displaystyle\leq 2\epsilon+\displaystyle\sum_{z\in N(y)}\frac{|\gamma(w^{\prime}_{yz})-\gamma(w_{yz})|}{\min\left(\displaystyle\sum_{u\in N(y)}\gamma(w_{yu}),\displaystyle\sum_{u\in N(y)}\gamma(w^{\prime}_{yu})\right)}
2ϵ+Cϵdy1dyM\displaystyle\leq 2\epsilon+C\epsilon d_{y}\frac{1}{d_{y}M}
2ϵ+CϵM.\displaystyle\leq 2\epsilon+\frac{C\epsilon}{M}.

Similarly, |Δf(x)Δf(x)|2ϵ+CϵM|\Delta f^{\prime}(x)-\Delta f(x)|\leq 2\epsilon+\frac{C\epsilon}{M}. It follows that

|μe(w)μe(w)|(4+2CM)|wewe|.|\mu_{e}(\vec{w}^{\prime})-\mu_{e}(\vec{w})|\leq(4+\frac{2C}{M})|w_{e}^{\prime}-w_{e}|.

This completes the proof that (9) has a unique solution in time interval (0,T)(0,T) for some T>0T>0. Now we further prove that TT can be extended to infinity. It is enough to prove the right-hand side of (9) is linearly bounded by ww.

For any edge hh, we have 0<wh10<w_{h}\leq 1, by Lemma 1, we then have 2wh<κh2-\frac{2}{w_{h}}<\kappa_{h}\leq 2. Then we get

wet>2we+wehE(G)(2wh)wh>(22|E(G)|)we\displaystyle\begin{split}\frac{\partial w_{e}}{\partial t}&>-2w_{e}+w_{e}\sum_{h\in E(G)}(-\frac{2}{w_{h}})w_{h}\\ &>(-2-2|E(G)|)w_{e}\end{split}

and

wet<(2we)we+wehE(G)2wh<2we+2.\displaystyle\begin{split}\frac{\partial w_{e}}{\partial t}&<-(-\frac{2}{w_{e}})w_{e}+w_{e}\sum_{h\in E(G)}2w_{h}\\ &<2w_{e}+2.\end{split}

Since for all edges ee, we(0)e(22|E(G)|)tw_{e}(0)e^{(-2-2|E(G)|)t} and 1+(we(0)+1)e2t-1+(w_{e}(0)+1)e^{2t} are right-bottom and right-top solutions to the component problem of (9) and both of them exist in time interval [0,)[0,\infty), then so does the solution of (9). This completes the proof of Theorem 3.

Eliminating the second additive term in the derivative equation of (9), we have the unnormalized continuous Ricci flow system of equations:

{dwedt=κeweX(0)=X0.\begin{cases}\frac{dw_{e}}{dt}=-\kappa_{e}w_{e}\\ X(0)=X_{0}.\end{cases} (13)

Replace (9) in algorithm 1 by (13), we have the following corollary.

Corollary 1.

For any initial weighted graph GG, by fixing the exit condition (I), there exists a unique solution X(t)X(t), for all time t[0,)t\in[0,\infty), to the system of ordinary differential equations in (13).

4 Solutions to the continuous Ricci flow

In this section we will exhibit some of the solutions of the continuous Ricci flow. On general graphs, there is no explicit expression for Ricci curvature, for each edge, κe(t)\kappa_{e}(t) can be expressed independently as a infimum of expression involved continuous functions. In addition, the Right-hand-side of (9) is non-linear, these make it not easy to study the convergence result of Ricci flow.

In order to reduce the number of metric variables wew_{e} evolved in the ODE, we solve the Ricci flow on path of length 22. Let G=(V,w,μ)G=(V,w,\mu) be defined on a weighted path of length 22. Denote the vertices in VV as {x,z,y}\{x,z,y\}. By definitions, the function μ={μxα,μyα,μzα}\mu=\{\mu_{x}^{\alpha},\mu_{y}^{\alpha},\mu_{z}^{\alpha}\} is as follows:

μxα(v)={αifv=x1αifv=z,μyα(v)={αifv=y1αifv=z,μzα(v)={αifv=zax(1α)ifv=xay(1α)ifv=y,\displaystyle\mu_{x}^{\alpha}(v)=\begin{cases}\alpha&\mbox{if}\ v=x\\ 1-\alpha&\mbox{if}\ v=z,\end{cases}\ \ \ \mu_{y}^{\alpha}(v)=\begin{cases}\alpha&\mbox{if}\ v=y\\ 1-\alpha&\mbox{if}\ v=z,\end{cases}\ \ \ \mu_{z}^{\alpha}(v)=\begin{cases}\alpha&\mbox{if}\ v=z\\ a_{x}(1-\alpha)&\mbox{if}\ v=x\\ a_{y}(1-\alpha)&\mbox{if}\ v=y,\end{cases}

where ax=γ(wzx)/(γ(wzx)+γ(wzy))a_{x}=\gamma(w_{zx})/(\gamma(w_{zx})+\gamma(w_{zy})) and ay=γ(wzy)/(γ(wzx)+γ(wzy))a_{y}=\gamma(w_{zy})/(\gamma(w_{zx})+\gamma(w_{zy})), simply speaking, ax,aya_{x},a_{y} are functions of w(t)w(t).

zzxxyyaxa_{x}aya_{y}1α1-\alpha1α1-\alpha
Figure 1: Path of length 22.

The Ollivier-Lin-Lu-Yau curvature κ\kappa is then as follows:

κxz=1+axaywyzwxz,κyz=1+ayaxwxzwyz.\kappa_{xz}=1+a_{x}-a_{y}\frac{w_{yz}}{w_{xz}},\ \kappa_{yz}=1+a_{y}-a_{x}\frac{w_{xz}}{w_{yz}}.

By (9), we have that

wxzt=wyzax,wyzt=wxzay.\frac{\partial w_{xz}}{\partial t}=w_{yz}-a_{x},\ \frac{\partial w_{yz}}{\partial t}=w_{xz}-a_{y}.

4.1 Unnormalized continuous Ricci flow

First we give an example showing that the unnormalized Ricci flow (13) would converge to a point if the initial metric satisfies a certain conditions. On the path graph of length 22, for arbitrary choice of γ\gamma, we have a system of homogeneous linear differential equations:

{wxzt=(1+ax)wxz+aywyz,wyzt=axwxz(1+ay)wyz.\displaystyle\begin{cases}\frac{\partial w_{xz}}{\partial t}&=-(1+a_{x})w_{xz}+a_{y}w_{yz},\\ \frac{\partial w_{yz}}{\partial t}&=a_{x}w_{xz}-(1+a_{y})w_{yz}.\end{cases} (14)

Since ax+ay=1a_{x}+a_{y}=1, then the associated matrix always has eigenvalues λ1=1,λ2=2\lambda_{1}=-1,\lambda_{2}=-2. If we set ax=ay=12a_{x}=a_{y}=\frac{1}{2}, then corresponding eigenvectors are (0.7071,0.7071)T(0.7071,0.7071)^{T}, and (0.7071,0.7071)T(0.7071,-0.7071)^{T}, then (14) has solution of form:

{wxz(t)=0.7071(c1et+c2e2t),wyz(t)=0.7071(c1etc2e2t).\displaystyle\begin{cases}w_{xz}(t)=0.7071(c_{1}e^{-t}+c_{2}e^{-2t}),\\ w_{yz}(t)=0.7071(c_{1}e^{-t}-c_{2}e^{-2t}).\end{cases}

If the initial metric satisfies wxz(0)=wyz(0)w_{xz}(0)=w_{yz}(0), i.e. c2=0c_{2}=0, then wxz(t)=wyz(t)=0.7071×c1etw_{xz}(t)=w_{yz}(t)=0.7071\times c_{1}e^{-t}. Thus both wxz(t),wyz(t)w_{xz}(t),w_{yz}(t) are decreasing functions with time tt which implies that the edge length converge to zero, in this case the graph converges to a point.

4.2 Normalized continuous Ricci flow

Although by Theorem 3 we are guaranteed to have an unique solution to the system of ODEs in (9), the types of solutions we obtain depend on the choice of γ\gamma in (8) and sometimes the initial condition. In this subsection, we give examples of different solutions to (9) defined on the same path graph of length 22. The results also answers the question asked at the beginning of the paper, we will see that the limit of the Ricci-flow on path exists, and it is possible to have a constant curvature although the initial graph does not have.

Example 1.
Constant solution:

If we pick ax=wyza_{x}=w_{yz} and ay=wxza_{y}=w_{xz}, note γ\gamma is the function satisfying γ(wzx)/γ(wzy)=wzy/wzx\gamma(w_{zx})/\gamma(w_{zy})=w_{zy}/w_{zx}, then |γ(wzx)γ(wzy)|=C|wzywzx||\gamma(w_{zx})-\gamma(w_{zy})|=C|w_{zy}-w_{zx}| for some constant CC. It follows that dwxzdt=dwyzdt=0\frac{dw_{xz}}{dt}=\frac{dw_{yz}}{dt}=0. Hence wxz(t)=wxz(0)w_{xz}(t)=w_{xz}(0) and wyz(t)=wyz(0)w_{yz}(t)=w_{yz}(0) for all tt and

κxz(t)=κyz(t)=1.\kappa_{xz}(t)=\kappa_{yz}(t)=1.
Stable solution without collapsing:

If we pick ax=wxza_{x}=w_{xz} and ay=wyza_{y}=w_{yz}, note γ\gamma is the function satisfying γ(wzx)/γ(wzy)=wzx/wzy\gamma(w_{zx})/\gamma(w_{zy})=w_{zx}/w_{zy}, then |γ(wzx)γ(wzy)|=C|wzxwzy||\gamma(w_{zx})-\gamma(w_{zy})|=C|w_{zx}-w_{zy}| for some constant CC. Then

wxzt\displaystyle\frac{\partial w_{xz}}{\partial t} =wyzwxz=12wxz,\displaystyle=w_{yz}-w_{xz}=1-2w_{xz},
wyzt\displaystyle\frac{\partial w_{yz}}{\partial t} =wxzwyz=12wyz.\displaystyle=w_{xz}-w_{yz}=1-2w_{yz}.

It follows that

wxz(t)\displaystyle w_{xz}(t) =12(12wxz(0))e2t,\displaystyle=\frac{1}{2}-\left(\frac{1}{2}-w_{xz}(0)\right)e^{-2t},
wyz(t)\displaystyle w_{yz}(t) =12(12wyz(0))e2t.\displaystyle=\frac{1}{2}-\left(\frac{1}{2}-w_{yz}(0)\right)e^{-2t}.

Thus wxz(t)12w_{xz}(t)\to\frac{1}{2} and wyz(t)12w_{yz}(t)\to\frac{1}{2} as tt\to\infty, and

κxz(t)1,κyz(t)1.\kappa_{xz}(t)\to 1,~{}\kappa_{yz}(t)\to 1.
Stable solution with collapsing:

Suppose WLOG that wxz(0)>wyz(0)w_{xz}(0)>w_{yz}(0). If we pick ax=wyz2wxz2+wyz2a_{x}=\frac{w_{yz}^{2}}{w_{xz}^{2}+w_{yz}^{2}} and ay=wxz2wxz2+wyz2a_{y}=\frac{w_{xz}^{2}}{w_{xz}^{2}+w_{yz}^{2}}, note γ\gamma is the function satisfying γ(wzx)/γ(wzy)=wzy2/wzx2\gamma(w_{zx})/\gamma(w_{zy})=w_{zy}^{2}/w_{zx}^{2}, then |γ(wzx)γ(wzy)|=C|wzy2wzx2|=C|wzywzx||\gamma(w_{zx})-\gamma(w_{zy})|=C|w_{zy}^{2}-w_{zx}^{2}|=C|w_{zy}-w_{zx}| for some constant CC.

Then

wxzt\displaystyle\frac{\partial w_{xz}}{\partial t} =wyzwyz2wxz2+wyz2=wyz(wxz2wyzwxz)wxz2+wyz2>0,\displaystyle=w_{yz}-\frac{w_{yz}^{2}}{w_{xz}^{2}+w_{yz}^{2}}=\frac{w_{yz}(w_{xz}^{2}-w_{yz}w_{xz})}{w_{xz}^{2}+w_{yz}^{2}}>0,
dwyzdt\displaystyle\frac{dw_{yz}}{dt} =wxzwxz2wxz2+wyz2=wxz(wyz2wyzwxz)wxz2+wyz2<0.\displaystyle=w_{xz}-\frac{w_{xz}^{2}}{w_{xz}^{2}+w_{yz}^{2}}=\frac{w_{xz}(w_{yz}^{2}-w_{yz}w_{xz})}{w_{xz}^{2}+w_{yz}^{2}}<0.

It follows that wxz(t)1w_{xz}(t)\to 1 and wyz(t)0w_{yz}(t)\to 0 as tt\to\infty and

κxz(t)2.\kappa_{xz}(t)\to 2.

The edge yzyz converges to point zz eventually.

5 Convergence of Ricci flow

In this section, we prove convergence result of Ricci flow on path and star graphs equipped with any initial weight. A graph minor is obtained from a given graph by repeatedly removing or contracting edges. From the path instance, we will also see its graph minors under Ricci flow accompanied with appropriated edge operations.

5.1 Ricci flow on path

Let P be a finite path of length n3n\geq 3, denote the edge set of P as {ei}i=1n\{e_{i}\}_{i=1}^{n} where e1,ene_{1},e_{n} are leave edges. We prove the following result:

Theorem 4.

Let γ(x)=1x\gamma(x)=\frac{1}{x}. Ricci flow (9) on path P converges. By contracting edges with small weights, any initial weighted path will converge to a path of length 22.

Proof.

Recall Du=xN(u)γ(wux)D_{u}=\sum_{x\in N(u)}\gamma(w_{ux}), by calculation, if e=(u,v)e=(u,v) is non-leaf edge with xuvyx-u-v-y, the Ollivier-Lin-Lu-Yau Ricci curvature is

κuv\displaystyle\kappa_{uv} =γ(wuv)Du+γ(wvu)Dvwuxwuvγ(wux)Duwvywuvγ(wvy)Dv\displaystyle=\frac{\gamma(w_{uv})}{D_{u}}+\frac{\gamma(w_{vu})}{D_{v}}-\frac{w_{ux}}{w_{uv}}\frac{\gamma(w_{ux})}{D_{u}}-\frac{w_{vy}}{w_{uv}}\frac{\gamma(w_{vy})}{D_{v}}
=1wuv1wuv+1wux+1wuv1wuv+1wvy1wuv(1wuv+1wux)1wvu(1wuv+1wvy)\displaystyle=\frac{\frac{1}{w_{uv}}}{\frac{1}{w_{uv}}+\frac{1}{w_{ux}}}+\frac{\frac{1}{w_{uv}}}{\frac{1}{w_{uv}}+\frac{1}{w_{vy}}}-\frac{1}{w_{uv}(\frac{1}{w_{uv}}+\frac{1}{w_{ux}})}-\frac{1}{w_{vu}(\frac{1}{w_{uv}}+\frac{1}{w_{vy}})}
=0.\displaystyle=0.

If e=(x,u)e=(x,u) is a leaf edge with with xuvx-u-v and dx=1d_{x}=1, then

κux\displaystyle\kappa_{ux} =1+γ(wux)Duwuvwuxγ(wuv)Du\displaystyle=1+\frac{\gamma(w_{ux})}{D_{u}}-\frac{w_{uv}}{w_{ux}}\frac{\gamma(w_{uv})}{D_{u}}
=1+1wux1wuv+1wux1wux(1wuv+1wux)\displaystyle=1+\frac{\frac{1}{w_{ux}}}{\frac{1}{w_{uv}}+\frac{1}{w_{ux}}}-\frac{1}{w_{ux}(\frac{1}{w_{uv}}+\frac{1}{w_{ux}})}
=1.\displaystyle=1.

Then hEκhwh=we1+wen,\sum_{h\in E}\kappa_{h}w_{h}=w_{e_{1}}+w_{e_{n}}, since ewe(t)=1\sum_{e}w_{e}(t)=1 for all time tt, then

weit=wei(1+we1+wen)<0,i{1,n},\frac{\partial w_{e_{i}}}{\partial t}=w_{e_{i}}(-1+w_{e_{1}}+w_{e_{n}})<0,~{}~{}i\in\{1,n\},
weit=wei(0+we1+wen)>0,i{2,,n1}.\frac{\partial w_{e_{i}}}{\partial t}=w_{e_{i}}(0+w_{e_{1}}+w_{e_{n}})>0,~{}~{}i\in\{2,\ldots,n-1\}.

Then we1(t),wen(t)w_{e_{1}}(t),w_{e_{n}}(t) decrease monotonically and weight weiw_{e_{i}} on non-leave edges eie_{i} increase monotonically, and by calculation, we1(t)+wen(t)=11(1(we1(0)wen(0)))etw_{e_{1}}(t)+w_{e_{n}}(t)=\frac{1}{1-(1-(w_{e_{1}}(0)-w_{e_{n}}(0)))e^{t}}, then we1(t)+wen(t)w_{e_{1}}(t)+w_{e_{n}}(t) tends to zero as tt goes to infinity. By repeatedly contracting edges with small enough weight (leaves), eventually the path converge to a path of length two. Refer to the constant solution of Example 1, weights on these two edges will not change any more. ∎

5.2 Ricci flow on star

A star is a tree with one internal node and a number of leaves. By choosing γ(x)=1x\gamma(x)=\frac{1}{x}, we can prove that Ricci flow on any initial weighted star graph converges.

Refer to caption
Refer to caption
Figure 2: An non-constant weighted star converges to constant-weighted star
Theorem 5.

Let GG be a star with initial weight w(0)\vec{w}(0) and γ(x)=1/x\gamma(x)=1/x. Denote the internal node of SnS_{n} as uu, leave as xx, 3du<3\leq d_{u}<\infty. The Ricci flow (9) deforms the weight w(0)\vec{w}(0) of GG to a weight of constant value 1du\frac{1}{d_{u}}.

To prove Theorem 5, we first need to prove we(t)w_{e}(t) is a monotone function for any edge ee in the graph, thus the limtwe(t)\lim_{t\to\infty}w_{e}(t) exists and is finite, then we prove the limit of we(t)w_{e}(t) is 1du\frac{1}{d_{u}}. Let Fe(t)=κh(t)wh(t)κe(t)F_{e}(t)=\sum\kappa_{h}(t)w_{h}(t)-\kappa_{e}(t). Then wet=we(t)Fe(t)\frac{\partial w_{e}}{\partial t}=w_{e}(t)F_{e}(t). We will prove the following result.

Lemma 6.

Let GG be a star and γ(x)=1/x\gamma(x)=1/x. For any edge eE(G)e\in E(G), if Fe(0)0F_{e}(0)\geq 0, then for all t(0,)t\in(0,\infty), Fe(t)0F_{e}(t)\geq 0; if Fe(0)0F_{e}(0)\leq 0, then for all t(0,)t\in(0,\infty), Fe(t)0F_{e}(t)\leq 0. Thus we(t)w_{e}(t) is a monotone function over time tt.

To prove above lemma, we need a result |Fe(t)t|C|Fe(t)||\frac{\partial F_{e}(t)}{\partial t}|\leq C|F_{e}(t)| for all tt, where CC is finite. This can be seen from the following facts.

Lemma 7 (Hamilton’s lemma).

[8] Let ff be a locally Lipschitz continuous function on [a,b][a,b]. (1): f(a)0f(a)\leq 0, and when f0f\geq 0 we have df/dt0df/dt\leq 0, then f(b)0f(b)\leq 0.

Conversely, (2): let f(a)0f(a)\geq 0, and when f0f\leq 0 we have df/dt0df/dt\geq 0, then f(b)0f(b)\geq 0.

Proof.

We prove item (1). By contradiction, assume f(b)>0f(b)>0. We need an auxiliary function g=exfg=e^{-x}f defined on [a,b][a,b]. Since f(b)>0f(b)>0, then g(b)>0g(b)>0. Consider the maximal point t0[a,b]t_{0}\in[a,b] of gg, then g(t0)g(b)>0g(t_{0})\geq g(b)>0, thus f(t0)>0f(t_{0})>0. We have dg/dt=exf+exdf/dt=ex(df/dtf)dg/dt=-e^{-x}f+e^{-x}df/dt=e^{-x}(df/dt-f), and dgdt(t0)\frac{dg}{dt}(t_{0}) is strictly less than 0 as f(t0)f(t_{0}) is strictly greater than 0 and dfdt(t0)0\frac{df}{dt}(t_{0})\leq 0.

Note gg is a locally Lipschitz continuous function, then dgdt(t0)<0\frac{dg}{dt}(t_{0})<0 means limh0supg(t0+h)g(t0)h<0\lim_{h\to 0}sup\frac{g(t_{0}+h)-g(t_{0})}{h}<0. Since t0bt_{0}\leq b, thus for any h<0h<0 such that g(t0+h)g(t0)>0g(t_{0}+h)-g(t_{0})>0, which is a contradiction to the maximal point t0t_{0} of gg.

Proof of item (2) is similar. ∎

The following uses Lemma 7 directly.

Corollary 2 (Hamilton’s Corollary).

Let ff be a locally Lipschitz continuous function on [a,b][a,b]. Let cc represent a finite positive value. If |df/dt|c|f||df/dt|\leq c|f|, then f(b)0f(b)\leq 0 if f(a)0f(a)\leq 0; f(b)0f(b)\geq 0 if f(a)0f(a)\geq 0.

Proof.

We prove the first result. Let g=ectfg=e^{-ct}f, then g(a)0g(a)\leq 0 as f(a)0f(a)\leq 0 by condition. If for all at<ba\leq t<b, f(t)0f(t)\leq 0, then f(b)0f(b)\leq 0 by continuity of ff. Assume there exist t<bt<b such that f(t)>0f(t)>0, then we have dfdtcf\frac{df}{dt}\leq cf. Further we have dgdt=cectf+ectdfdt0\frac{dg}{dt}=-ce^{-ct}f+e^{-ct}\frac{df}{dt}\leq 0. By Lemma 7, g(b)0g(b)\leq 0, thus f(b)0f(b)\leq 0. ∎

We need the following lemma [11] to obtain a bound for expression Du(t)Du(t)\frac{D_{u}^{\prime}(t)}{D_{u}(t)}. Its proof can be found in [18].

Lemma 8.

If q ,1q2,,qn{}_{1},q_{2},\ldots,q_{n} are positive numbers, then

min1inpiqip1+p2++pnq1+q2++qnmax1inpiqi\min_{1\leq i\leq n}\frac{p_{i}}{q_{i}}\leq\frac{p_{1}+p_{2}+\cdots+p_{n}}{q_{1}+q_{2}+\cdots+q_{n}}\leq\max_{1\leq i\leq n}\frac{p_{i}}{q_{i}}

for any real numbers p1,p2,,pn.p_{1},p_{2},\ldots,p_{n}. Equality holds on either side if and only if all the ratios pi/qip_{i}/q_{i} are equal.

Proof of Lemma 6.

By calculation, for any edge e=uxe=ux we have κux=1+2duwuxDu\kappa_{ux}=1+\frac{2-d_{u}}{w_{ux}D_{u}}and Fux=du2Du(1wuxdu).F_{ux}=\frac{d_{u}-2}{D_{u}}(\frac{1}{w_{ux}}-d_{u}). Since 1wuxDu<1,1Du<1du\frac{1}{w_{ux}D_{u}}<1,\frac{1}{D_{u}}<\frac{1}{d_{u}} then FuxF_{ux} is bounded at any finite time.

The derivative of Fux(t)F_{ux}(t) respect to tt is

Fux\displaystyle F_{ux}^{\prime} =(du2)wuxwux2DuDuDudu2Du(1wuxdu)\displaystyle=-\frac{(d_{u}-2)w_{ux}^{\prime}}{w_{ux}^{2}D_{u}}-\frac{D_{u}^{\prime}}{D_{u}}\frac{d_{u}-2}{D_{u}}(\frac{1}{w_{ux}}-d_{u})
=(du2)wuxFuxwux2DuDuDuFux\displaystyle=-\frac{(d_{u}-2)w_{ux}F_{ux}}{w_{ux}^{2}D_{u}}-\frac{D_{u}^{\prime}}{D_{u}}F_{ux}
=((du2)wuxDuDuDu)Fux.\displaystyle=(-\frac{(d_{u}-2)}{w_{ux}D_{u}}-\frac{D_{u}^{\prime}}{D_{u}})F_{ux}.

Using Lemma 8, we get

|DuDu|\displaystyle|\frac{D_{u}^{\prime}}{D_{u}}| =|zuwuzwuz2zu1wuz|=|zuFuzwuzzu1wuz|zu|Fuz|wuzzu1wuzmaxzu|Fuz|.\displaystyle=|\frac{\sum_{z\sim u}-\frac{w_{uz}^{\prime}}{w_{uz}^{2}}}{\sum_{z\sim u}\frac{1}{w_{uz}}}|=|\frac{\sum_{z\sim u}\frac{F_{uz}}{w_{uz}}}{\sum_{z\sim u}\frac{1}{w_{uz}}}|\leq\frac{\sum_{z\sim u}\frac{|F_{uz}|}{w_{uz}}}{\sum_{z\sim u}\frac{1}{w_{uz}}}\leq\max_{z\sim u}{|F_{uz}|}.

Thus

|Fux(t)|(du2+maxzu|Fuz|)|Fux|.\displaystyle|F_{ux}^{\prime}(t)|\leq(d_{u}-2+\max_{z\sim u}{|F_{uz}|})|F_{ux}|. (15)

Let C=(du2+maxzu|Fuz|)C=(d_{u}-2+\max_{z\sim u}{|F_{uz}|}), clearly, it is a finite number. Since Fe(t)F_{e}(t) is differentiable and Fe(t)F_{e}^{\prime}(t) is bounded, by Hamilton’s Corollary 2, if Fe(0)0F_{e}(0)\geq 0, then Fe(t)0F_{e}(t)\geq 0 for all t>0t>0; if Fe(0)0F_{e}(0)\leq 0, then Fe(t)0F_{e}(t)\leq 0 for all t>0t>0. Thus we(t)w_{e}(t) is a monotone function over time t[0,)t\in[0,\infty).

Lemma 9 (Barbalat’s Lemma).

[2] If f(t)f(t) has a finite limit as tt\to\infty and if f(t)f^{\prime}(t) is uniformly continuous (or f′′(t)f^{\prime\prime}(t) is bounded), then limtf(t)=0\lim_{t\to\infty}f^{\prime}(t)=0.

Proof of Theorem 5.

By Lemma 6, we have limtwe(t)\lim_{t\to\infty}w_{e}(t) exists and is finite, both we(t)w_{e}(t) and Fe(t)F_{e}(t) are uniformly continuous, thus we(t)=weFew_{e}^{\prime}(t)=w_{e}F_{e} is uniformly continuous. By Barbalat’s Lemma 9, limtwe(t)=0\lim_{t\to\infty}w_{e}^{\prime}(t)=0.

Note limtwe(t)=0\lim_{t\to\infty}w_{e}(t)=0 if and only if Fe(t)F_{e}(t) is negative for large tt, equivalently, we(t)>1duw_{e}(t)>\frac{1}{d_{u}} a contradiction.

Thus, it must be limtFe(t)=0\lim_{t\to\infty}F_{e}(t)=0, implies limtwe(t)=1du\lim_{t\to\infty}w_{e}(t)=\frac{1}{d_{u}}. Therefore, the Ricci flow (9) on star graph SnS_{n} with n3n\geq 3 converges to constant-weighted star of same size. ∎

6 Conclusions

In this study, we propose an normalized continuous Ricci flow for weighted graphs, based on Ollivier-Lin-Lu-Yau Ricci curvature and prove that the Ricci flow metric X(t)X(t) with initial data X(0)X(0) exists and is unique for all t0t\geq 0 by fixing the violation of distance condition. We also show some explicit, rigorous examples of Ricci flows on tree graphs. Future work already underway, we expect results of more general Ricci flows evolved on various graphs.

References

  • [1] S. Bai, A. Huang, L. Lu, and S.T. Yau. On the sum of ricci-curvatures for weighted graphs. Pure Appl. Math. Q., 17(5):1599–1617, 2021.
  • [2] I. Barbălat. Systèmes d’équations différentielles d’oscillations non linéaires. Rev. Math. Pures Appl, 4:267–270, 1959.
  • [3] D. Bourne, D. Cushing, S. Liu, F. Münch, and N. Peyerimhoff. Ollivier–ricci idleness functions of graphs. SIAM J. Discrete Math, 32, 04 2017.
  • [4] S. Brendle and R. Schoen. Manifolds with 1/4-pinched curvature are space forms. J. Amer. Math. Soc., 22:287–307, 2009.
  • [5] R. Fang, D.and Xue. Ordinary Differential Equation. Higher Education Press, 2017.
  • [6] S. Fortunato. Community detection in graphs. Phys. Rep, 486, 06 2009.
  • [7] R. Hamilton. Three-manifolds with positive ricci curvature. J. Differ. Geom, 17:255–362, 06 1982.
  • [8] R. Hamilton. Four-manifolds with positive curvature operator. J. Differ. Geom, 24(2):153–179, 1986.
  • [9] J. Leskovec, K. Lang, and M. Mahoney. Empirical comparison of algorithms for network community detection. Proceedings of the 19th International Conference on World Wide Web, WWW ’10, 04 2010.
  • [10] Y. Lin, L. Lu, and S.T. Yau. Ricci curvature of graphs. Tohoku Math. J., 63, 12 2011.
  • [11] H. Minc. Nonnegative matrices. Wiley, New York, 1988.
  • [12] Florentin Münch and Radoslaw K. Wojciechowski. Ollivier ricci curvature for general graph laplacians: Heat equation, laplacian comparison, non-explosion and diameter bounds. Adv. Math., 356, 11 2019.
  • [13] Chien-Chun Ni, Yu-Yao Lin, Feng Luo, and Jie Gao. Community detection on networks with ricci flow. Sci. Rep, 9, 07 2019.
  • [14] Y. Ollivier. Ricci curvature of markov chains on metric spaces. J. Funct. Anal., 256:810–864, 02 2009.
  • [15] Yann Ollivier. Ricci curvature of metric spaces. C. R. Math., 345(11):643 – 646, 2007.
  • [16] L. Peel, D. Larremore, and A. Clauset. The ground truth about metadata and community detection in networks. Science Advances, 3, 08 2016.
  • [17] G. Perelman. The entropy formula for the ricci flow and its geometric applications. arXiv preprint math/0211159, 2002.
  • [18] R. Varga. Matrix iterative analysis. Prentice-Hall, Englewood Cliffs, NJ, 1962.

Beijing Yanqi Lake Institute of Mathematical Sciences and Applications, China. Email address: sbai@seu.edu.cn. This author is supported by NSFC grant number 12301434.

Yau Mathematical Science Center, Tsinghua University, Beijing, 100084, China; Department of Matheatics, Tsinghua University, Beijing, 100084,China. Email address: yonglin@tsinghua.edu.cn. This author was supported in part by NSFC grant number 12071245.

University of South Carolina, Columbia, SC 29208, USA. Email address: lu@math.sc.edu. The author was supported in part by NSF grant DMS 2038080.

Georgia Institute of Technology, Atlanta, GA, 30332, USA. Email address: zwang672@gatech.edu.

Yau Mathematical Sciences Center, Tsinghua University, Beijing, 100084, China; Yanqi Lake Beijing Institute of Mathematical Sciences and Applications, Beijing, 101408, China. Email address: styau@tsinghua.edu.cn, yau@math.harvard.edu