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

The deformation Space of Geodesic Triangulations and Generalized Tutte’s Embedding Theorem

Yanwen Luo, Tianqi Wu, Xiaoping Zhu
Abstract.

We proved the contractibility of the deformation space of the geodesic triangulations on a closed surface of negative curvature. This solves an open problem proposed by Connelly et al. in 1983 [7], in the case of hyperbolic surfaces. The main part of the proof is a generalization of Tutte’s embedding theorem for closed surfaces of negative curvature.

Key words and phrases:
geodesic triangulations, Tutte’s embedding
Acknowledgement: The author was supported in part by NSF 1760471, NSF DMS FRG 1760527 and NSF DMS 1811878.

1. Introduction

In this paper, we study the deformation space of geodesic triangulations of a surface within a fixed homotopy type. Such space can be viewed as a discrete analogue of the space of surface diffeomorphisms homotopic to the identity. Our main theorem is the following.

Theorem 1.1.

For a closed orientable surface of negative curvature, the space of geodesic triangulations in a homotopy class is contractible. In particular, it is connected.

The group of diffeomorphisms of a smooth surface is one of the fundamental objects in the study of low dimensional topology. Determining the homotopy types of diffeomorphism groups has profound implications to a wide range of problems in Teichmuller spaces, mapping class groups, and geometry and topology of 3-manifolds. Smale [16] proved that the group of diffeomorphisms of a closed 2-disk which fix the boundary pointwisely is contractible. This enables him to show that the group of orientation-preserving diffeomorphisms of the 2-sphere is homotopic equivalent to SO(3)SO(3) [16]. Earle-Eells [9] identified the homotopy type of the group of the diffeomorphisms homotopic to the identity for any closed surface. In particular, such topological group is contractible for a closed orientable surface with genus greater than one, consisting with our Theorem 1.1 for the discrete analogue.

Cairns [5] initiated the investigation of the topology of the space of geodesic triangulations, and proved that if the surface is a geometric triangle in the Euclidean plane, the space of geodesic triangulations with fixed boundary edges is connected. A series of further developments culminated in a discrete version of Smale’s theorem proved by Bloch-Connelly-Henderson [2] as follows.

Theorem 1.2.

The space of geodesic triangulations of a convex polygon with fixed boundary edges is homeomorphic to some Euclidean space. In particular, it is contractible.

A simple proof of the contractibility of the space above is provided in [15] using Tutte’s embedding theorem [17]. It also provides examples showing that the homotopy type of this space can be complicated if the boundary of the polygon is not convex. For closed surfaces, it is conjectured in [7] that

Conjecture 1.3.

The space of geodesic triangulations of a closed orientable surface with constant curvature deformation retracts to the group of isometries of the surface homotopic to the identity.

The connectivity of these spaces has been explored in [5, 6, 14]. Awartani-Henderson [1] identified a contractible subspace in the space of geodesic triangulations of the 2-sphere. Hass-Scott [14] showed that the space of geodesic triangulation of a surface with a hyperbolic metric is contractible if the triangulation contains only one vertex. The main result of this paper affirms conjecture 1.3 in the case of hyperbolic surfaces.

1.1. Set Up and the Main Theorem

Assume MM is a connected closed orientable smooth surface with a smooth Riemannian metric gg of non-positive Gaussian curvature. A topological triangulation of MM can be identified as a homeomorphism ψ\psi from |T||T| to MM, where |T||T| is the carrier of a 2-dimensional simplicial complex T=(V,E,F)T=(V,E,F) with the vertex set VV, the edge set EE, and the face set FF. For convenience, we label the vertices as 1,2,,n1,2,...,n where n=|V|n=|V| is the number of vertices. The edge in EE determined by vertices ii and jj is denoted as ijij. Each edge is identified with the closed unit interval [0,1][0,1].

Let T(1)T^{(1)} be the 1-skeleton of TT, and denote X=X(M,T,ψ)X=X(M,T,\psi) as the space of geodesic triangulations homotopic to ψ|T(1)\psi|_{T^{(1)}}. More specifically, XX contains all the embeddings φ:T(1)M\varphi:T^{(1)}\rightarrow M satisfying that

  1. (1)

    The restriction φij\varphi_{ij} of φ\varphi on the edge ijij is a geodesic parameterized with constant speed, and

  2. (2)

    φ\varphi is homotopic to ψ|T(1)\psi|_{T^{(1)}}.

It has been proved by Colin de Verdière [8] that such X(M,T,ψ)X(M,T,\psi) is always non-empty. Further, XX is naturally a metric space, with the distance function

dX(φ,ϕ)=maxxdg(φ(x),ϕ(x)).d_{X}(\varphi,\phi)=\max_{x}d_{g}(\varphi(x),\phi(x)).

Then our main theorem is formally stated as follows.

Theorem 1.4.

If (M,g)(M,g) has strictly negative Gaussian curvature, then X(M,T,ψ)X(M,T,\psi) is contractible. In particular, it is connected.

1.2. Generalized Tutte’s Embedding

Let X~=X~(M,T,ψ)\tilde{X}=\tilde{X}(M,T,\psi) be the super space of XX, containing all the continuous maps φ:T(1)M\varphi:T^{(1)}\rightarrow M satisfying that

  1. (1)

    The restriction φij\varphi_{ij} of φ\varphi on the edge ijij is geodesic parameterized with constant speed, and

  2. (2)

    φ\varphi is homotopic to ψ|T(1)\psi|_{T^{(1)}}.

Notice that elements in X~\tilde{X} may not be embeddings of T(1)T^{(1)} to MM. The space X~\tilde{X} is also naturally a metric space, with the same distance function

dX~(φ,ϕ)=maxxdg(φ(x),ϕ(x)).d_{\tilde{X}}(\varphi,\phi)=\max_{x}d_{g}(\varphi(x),\phi(x)).

We call an element in X~\tilde{X} a geodesic mapping. A geodesic mapping is determined by the positions qi=φ(i)q_{i}=\varphi(i) of the vertices and the homotopy classes of φij\varphi_{ij} relative to the endpoints qiq_{i} and qjq_{j}. In particular, this holds for geodesic triangulations. Since we can perturb the vertices of a geodesic triangulation to generate another, XX is a 2n2n dimensional manifold.

Let (i,j)(i,j) be the directed edge starting from the vertex ii and ending at the vertex jj. Denote E={(i,j):ijE}\vec{E}=\{(i,j):ij\in E\} as the set of directed edges of TT. A positive vector w>0Ew\in\mathbb{R}^{\vec{E}}_{>0} is called a weight of TT. For any weight ww and geodesic mapping φX~\varphi\in\tilde{X}, we say φ\varphi is ww-balanced if for any iVi\in V,

j:ijEwijvij=0.\sum_{j:ij\in E}w_{ij}{v}_{ij}=0.

Here vijTqiMv_{ij}\in T_{q_{i}}M is defined with the exponential map exp:TMM\exp:TM\to M such that expqi(tvij)=φij(t)\exp_{q_{i}}(t{v}_{ij})=\varphi_{ij}(t) for t[0,1]t\in[0,1].

The main part of the proof of Theorem 1.4 is to generalize Tutte’s embedding theorem (see Theorem 9.2 in [17] or Theorem 6.1 in [10]) to closed surfaces of negative curvature. Specifically, we will prove the following two theorems.

Theorem 1.5.

Assume (M,g)(M,g) has strictly negative Gaussian curvature. For any weight ww, there exists a unique geodesic mapping φX~(M,T,ψ)\varphi\in\tilde{X}(M,T,\psi) that is ww-balanced. Such induced map Φ(w)=φ\Phi(w)=\varphi is continuous from >0E\mathbb{R}^{\vec{E}}_{>0} to X~\tilde{X}.

Theorem 1.6.

If φX~\varphi\in\tilde{X} is ww-balanced for some weight ww, then φX\varphi\in X.

Theorem 1.6 can be regarded as a generalization of the embedding theorems by Colin de Verdière (see Theorem 2 in [8]) and Hass-Scott (see Lemma 10.12 in [14]), which imply that the minimizer of the following discrete Dirichlet energy

E(φ)=12ijEwijlij2E(\varphi)=\frac{1}{2}\sum_{ij\in E}w_{ij}l^{2}_{ij}

among the maps φ\varphi in the homotopy class of ψ|T(1)\psi|_{T^{(1)}} is a geodesic triangulation. Here lijl_{ij} is the geodesic length of φij\varphi_{ij} in MM. The minimizer is a ww-balanced geodesic mapping with wij=wjiw_{ij}=w_{ji} for ijEij\in E. Hence, Theorem 1.6 extends the previous results from the cases of symmetric weights to non-symmetric weights. We believe that, the proofs in Colin de Verdière [8] and Hass-Scott [14] could be easily modified to work with our non-symmetric case. Nevertheless, we will give a new proof in Section 3 to make the paper self-contained.

1.3. Mean Value Coordinates and the Proof of Theorem 1.4

Theorem 1.5 and 1.6 give a continuous map Φ\Phi from >0E\mathbb{R}^{\vec{E}}_{>0} to XX. For the oppositie direction, we can construct a weight ww for a geodesic embedding φX\varphi\in X, using mean value coordinates which was firstly introduced by Floater [11]. Given φX\varphi\in X, the mean value coordinates are defined to be

wij=tan(αij/2)+tan(βij/2)|vij|,w_{ij}=\frac{\tan(\alpha_{ij}/2)+\tan(\beta_{ij}/2)}{|v_{ij}|},

where |vij||v_{ij}| equals to the geodesic length of φij([0,1])\varphi_{ij}([0,1]), and αij\alpha_{ij} and βij\beta_{ij} are the two inner angles in φ(T(1))\varphi(T^{(1)}) at the vertex φ(i)\varphi(i) sharing the edge φij([0,1])\varphi_{ij}([0,1]). The construction of mean value coordinates gives a continuous map Ψ\Psi from XX to >0E\mathbb{R}^{\vec{E}}_{>0}. Further, by Floater’s mean value theorem (see Proposition 1 in [11]), any φX\varphi\in X is Ψ(φ)\Psi(\varphi)-balanced. Namely, ΦΨ=idX\Phi\circ\Psi=id_{X}. Then Theorem 1.4 is a direct consequence of Theorem 1.5 and 1.6.

Proof of Theorem 1.4.

Since >0E\mathbb{R}^{\vec{E}}_{>0} is contractible, ΨΦ\Psi\circ\Phi is homotopic to the identity map. Since ΦΨ=idX\Phi\circ\Psi=id_{X}, XX is homotopy equivalent to the contractible space >0E\mathbb{R}^{\vec{E}}_{>0}. ∎

In the remaining of the paper, we will prove Theorem 1.5 in Section 2 and Theorem 1.6 in section 3.

2. Proof of Theorem 1.5

Theorem 1.5 consists of three parts: the existence of ww-balanced geodesic mapping, the uniqueness of ww-balanced geodesic mapping and the continuity of the map Φ\Phi. In this section, we will first parametrize X~\tilde{X} by M~n\tilde{M}^{n}, where M~\tilde{M} is the universal covering of MM, and then prove the three parts in Subsection 2.1 and 2.2 and 2.3 respectively.

Assume that pp is the covering map from M~\tilde{M} to MM, and Γ\Gamma is the corresponding group of deck transformations of the covering so that M~/Γ=M\tilde{M}/\Gamma=M. For any iVi\in V, fix a lifting q~iM~\tilde{q}_{i}\in\tilde{M} of qiMq_{i}\in M. For any edge ijij, denote φ~ij(t)\tilde{\varphi}_{ij}(t) as the lifting of φij(t)\varphi_{ij}(t) such that φ~ij(0)=q~i\tilde{\varphi}_{ij}(0)=\tilde{q}_{i}. Then p(φ~ij(1))=φij(1)=qj=p(q~j)p(\tilde{\varphi}_{ij}(1))=\varphi_{ij}(1)=q_{j}=p(\tilde{q}_{j}), and there exists a unique deck transformation AijΓA_{ij}\in\Gamma such that φ~ij(1)=Aijq~j\tilde{\varphi}_{ij}(1)=A_{ij}\tilde{q}_{j}. It is easy to see that Aij=Aji1A_{ij}=A_{ji}^{-1} for any edge ijij.

Equip M~\tilde{M} with the natural pullback Riemannian metric g~\tilde{g} of gg with negative Gaussian curvature. This metric is equivariant with respect to Γ\Gamma. For any x,yM~x,y\in\tilde{M}, there exists a unique geodesic with constant speed parameterization γx,y:[0,1]M~\gamma_{x,y}:[0,1]\rightarrow\tilde{M} such that γx,y(0)=x\gamma_{x,y}(0)=x and γx,y(1)=y\gamma_{x,y}(1)=y. We can naturally parametrize X~\tilde{X} as follows.

Theorem 2.1.

For any (x1,,xn)M~n(x_{1},...,x_{n})\in\tilde{M}^{n}, define φ=φ[x1,,xn]\varphi=\varphi[x_{1},...,x_{n}] as

φij(t)=pγxi,Aijxj(t)\varphi_{ij}(t)=p\circ\gamma_{x_{i},A_{ij}x_{j}}(t)

for any ijEij\in E and t[0,1]t\in[0,1]. Then such φ\varphi is a well-defined geodesic mapping in X~\tilde{X}, and the map (x1,,xn)φ[x1,,xn](x_{1},...,x_{n})\mapsto\varphi[x_{1},...,x_{n}] is a homeomorphism from M~n\tilde{M}^{n} to X~\tilde{X}.

Here we omit the proof of Theorem 2.1 which is routine but lengthy. In the remaining of this section, for any x,y,zM~x,y,z\in\tilde{M} and u,vTxM~u,v\in T_{x}\tilde{M}, we denote

  1. (1)

    d(x,y)d(x,y) as the intrinsic distance between x,yx,y in (M~,g~)(\tilde{M},\tilde{g}), and

  2. (2)

    v(x,y)=expx1yTxM~v(x,y)=\exp_{x}^{-1}y\in T_{x}\tilde{M}, and

  3. (3)

    xyz\triangle xyz as the geodesic triangle in M~\tilde{M} with vertices x,y,zx,y,z, which could possibly be degenerate, and

  4. (4)

    yxz\angle yxz as the inner angle of xyz\triangle xyz at xx if d(x,y)>0d(x,y)>0 and d(x,z)>0d(x,z)>0, and

  5. (5)

    |v||v| as the norm of vv under the metric g~x\tilde{g}_{x}, and

  6. (6)

    uvu\cdot v as the inner product of uu and vv under the metric g~x\tilde{g}_{x}.

By scaling the metric if necessary, we may assume that the Gaussian curvatures of (M,g)(M,g) and (M~,g~)(\tilde{M},\tilde{g}) are bounded above by 1-1.

2.1. Proof of the Uniqueness

We first prove the following Lemma 2.2 using CAT(0) geometry. See Theorem 4.3.5 in [4] and Theorem 1A.6 in [3] for the well-known comparison theorems.

Lemma 2.2.

Assume x,y,zM~x,y,z\in\tilde{M}, then

  1. (1)

    |v(z,x)v(z,y)|d(x,y)|v(z,x)-v(z,y)|\leq d(x,y), and

  2. (2)

    v(x,y)v(x,z)+v(y,x)v(y,z)d(x,y)2v(x,y)\cdot v(x,z)+v(y,x)\cdot v(y,z)\geq d(x,y)^{2},

and the equality holds if and only if xyz\triangle xyz is degenerate.

Proof.

If xyz\triangle xyz is degenerate, then there exists a geodesic γ\gamma in M~\tilde{M} such that x,y,zγx,y,z\in\gamma, and then the proof is straightforward. So we assume that xyz\triangle xyz is non-degenerate.

(1) Three points v(z,x),v(z,y)v(z,x),v(z,y), and 0 in TzM~T_{z}\tilde{M} determine a Euclidean triangle, where |v(z,x)|=d(x,z)|v(z,x)|=d(x,z), and |v(z,y)|=d(z,y)|v(z,y)|=d(z,y) and the angle between v(z,x)v(z,x) and v(z,y)v(z,y) is equal to xzy\angle xzy. Then by the CAT(0) comparison theorem,

|v(z,x)v(z,y)|<d(x,y).|v(z,x)-v(z,y)|<d(x,y).

(2) Let x,y,z2x^{\prime},y^{\prime},z^{\prime}\in\mathbb{R}^{2} be such that

|xz|2=|v(x,z)|,|yz|2=|v(y,z)|,and|xy|2=|v(x,y)|.|x^{\prime}-z^{\prime}|_{2}=|v(x,z)|,\quad\quad\quad|y^{\prime}-z^{\prime}|_{2}=|v(y,z)|,\quad\text{and}\quad|x^{\prime}-y^{\prime}|_{2}=|v(x,y)|.

Then by the CAT(0)(0) comparison theorem, yxz<yxz\angle yxz<\angle y^{\prime}x^{\prime}z^{\prime}, and xyz<xyz\angle xyz<\angle x^{\prime}y^{\prime}z^{\prime}. Hence,

v(x,y)v(x,z)+v(y,x)v(y,z)>(yx)(zx)+(xy)(zy)=|xy|22=d(x,y)2.v(x,y)\cdot v(x,z)+v(y,x)\cdot v(y,z)>(y^{\prime}-x^{\prime})\cdot(z^{\prime}-x^{\prime})+(x^{\prime}-y^{\prime})\cdot(z^{\prime}-y^{\prime})=|x^{\prime}-y^{\prime}|_{2}^{2}=d(x,y)^{2}.

Proof of the uniqueness part in Theorem 1.5 .

If not, assume φ[x1,,xn]\varphi[x_{1},...,x_{n}] and φ[x1,,xn]\varphi[x_{1}^{\prime},...,x_{n}^{\prime}] are two different geodesic mappings that are both ww-balanced for some weight ww. We are going to prove a discrete maximum principle for the function jd(xj,xj)j\mapsto d(x_{j},x_{j}^{\prime}). Assume iVi\in V is such that d(xi,xi)=maxjVd(xj,xj)>0d(x_{i},x_{i}^{\prime})=\max_{j\in V}d(x_{j},x_{j}^{\prime})>0. By lifting the ww-balanced assumption to M~\tilde{M}, we have that

(1) j:ijEwijv(xi,Aijxj)=0,\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j})=0,

and

(2) j:ijEwijv(xi,Aijxj)=0.\sum_{j:ij\in E}w_{ij}v(x_{i}^{\prime},A_{ij}x_{j}^{\prime})=0.

Then by part (1) of Lemma 2.2 and equation (1),

|j:ijEwijv(xi,Aijxj)|\displaystyle\left|\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j}^{\prime})\right|
=\displaystyle= |j:ijEwijv(xi,Aijxj)j:ijEwijv(xi,Aijxj)|\displaystyle\left|\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j}^{\prime})-\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j})\right|
\displaystyle\leq j:ijEwijd(Aijxj,Aijxj)\displaystyle\sum_{j:ij\in E}w_{ij}d(A_{ij}x_{j},A_{ij}x_{j}^{\prime})
=\displaystyle= j:ijEwijd(xj,xj)\displaystyle\sum_{j:ij\in E}w_{ij}d(x_{j},x_{j}^{\prime})
\displaystyle\leq d(xi,xi)j:ijEwij.\displaystyle d(x_{i},x_{i}^{\prime})\sum_{j:ij\in E}w_{ij}.

By part (2) of Lemma 2.2, equation (2), and the Cauchy-Schwartz inequality,

d(xi,xi)|j:ijEwijv(xi,Aijxj)|\displaystyle d(x_{i},x_{i}^{\prime})\cdot\left|\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j}^{\prime})\right|
\displaystyle\geq v(xi,xi)j:ijEwijv(xi,Aijxj)+v(xi,xi)j:ijEwijv(xi,Aijxj)\displaystyle v(x_{i},x_{i}^{\prime})\cdot\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j}^{\prime})+v(x_{i}^{\prime},x_{i})\cdot\sum_{j:ij\in E}w_{ij}v(x_{i}^{\prime},A_{ij}x_{j}^{\prime})
\displaystyle\geq j:ijEwijd(xi,xi)2.\displaystyle\sum_{j:ij\in E}w_{ij}\cdot d(x_{i},x_{i}^{\prime})^{2}.

Therefore, the equalities hold in both inequalities above. Then for any neighbor jj of ii, d(xj,xj)=d(xi,xi)=maxkVd(xk,xk)d(x_{j},x_{j}^{\prime})=d(x_{i},x_{i}^{\prime})=\max_{k\in V}d(x_{k},x_{k}^{\prime}), and AijxjA_{ij}x_{j} is on the geodesic determined by xix_{i} and xix_{i}^{\prime}. Then the one-ring neighborhood of p(xi)p(x_{i}) in φ[x1,,xn](T(1))\varphi[x_{1},...,x_{n}](T^{(1)}) degenerates to a geodesic arc. By the connectedness of the surface, we can repeat the above argument and deduce that d(xj,xj)=d(xi,xi)d(x_{j},x_{j}^{\prime})=d(x_{i},x_{i}^{\prime}) for any jVj\in V. Further, for any triangle σF\sigma\in F, φ[x1,,xn](σ)\varphi[x_{1},...,x_{n}](\partial\sigma) degenerates to a geodesic arc.

It is not difficult to extend φ[x1,,xn]\varphi[x_{1},...,x_{n}] to a continuous map φ~\tilde{\varphi} from |T||T| to MM, such that for any triangle σF\sigma\in F, φ~(σ)=φ[x1,,xn](σ)\tilde{\varphi}(\sigma)=\varphi[x_{1},...,x_{n}](\partial\sigma) being a geodesic arc.

It is also not difficult to prove that φ~\tilde{\varphi} is homotopic to ψ\psi. Therefore, φ~\tilde{\varphi} is degree-one and surjective. This is contradictory to that φ~(|T|)\tilde{\varphi}(|T|) is a finite union of geodesic arcs. ∎

2.2. Proof of the Existence

Here we prove a stronger existence result.

Theorem 2.3.

Given a compact subset KK of >0E\mathbb{R}^{\vec{E}}_{>0}, there exists a compact subset K=K(M,T,ψ,K)K^{\prime}=K^{\prime}(M,T,\psi,K) of X~\tilde{X} such that for any wKw\in K, there exists a ww-balanced geodesic mapping φK\varphi\in K^{\prime}.

We first introduce a topological Lemma 2.4 and then reduce Theorem 2.3 to Lemma 2.5.

Lemma 2.4.

Suppose Bn={xn:|x|1}B^{n}=\{x\in\mathbb{R}^{n}:|x|\leq 1\} is the unit ball in n\mathbb{R}^{n}, and f:Bnnf:B^{n}\rightarrow\mathbb{R}^{n} is a continuous map such that xf(x)/|f(x)|x\neq f(x)/|f(x)| for any xBn=Sn1x\in\partial B^{n}=S^{n-1} with f(x)0f(x)\neq 0. Then ff has a zero in BnB^{n}.

Proof.

If not, g(x)=f(x)/|f(x)|g(x)=f(x)/|f(x)| is a continuous map from BnB^{n} to Bn\partial B^{n}. Since BnB^{n} is contractible, g(x)g(x) is null-homotopic, and thus g|Sn1g|_{S^{n-1}} is also null-homotopic. Since g(x)xg(x)\neq x, it is easy to verify that

H(x,t)=tg(x)+(1t)(x)|tg(x)+(1t)(x)|H(x,t)=\frac{tg(x)+(1-t)(-x)}{|tg(x)+(1-t)(-x)|}

is a homotopy between g|Sn1g|_{S^{n-1}} and id|Sn1-id|_{S^{n-1}}. This contradicts to that id|Sn1-id|_{S_{n-1}} is not null-homotopic. ∎

Lemma 2.5.

We fix an arbitrary point qM~q\in\tilde{M}. If w>0Ew\in\mathbb{R}^{\vec{E}}_{>0} and (x1,,xn)M~n(x_{1},...,x_{n})\in\tilde{M}^{n} satisfies that

(3) v(xi,q)j:ijEwijv(xi,Aijxj)0v(x_{i},q)\cdot\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j})\leq 0

for any iVi\in V, then

iVd(xi,q)2<R2\sum_{i\in V}d(x_{i},q)^{2}<R^{2}

for some constant R>0R>0 which depends only on M,T,ψ,qM,T,\psi,q and

λw:=maxijEwijminijEwij.\lambda_{w}:=\frac{\max_{ij\in E}w_{ij}}{\min_{ij\in E}w_{ij}}.

The vector in Figure 1

ri=j:ijEwijv(xi,Aijxj)r_{i}=\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j})

is defined as the residue vector rir_{i} at xix_{i} of φ[x1,,xn]\varphi[x_{1},\cdots,x_{n}] with respect to the weight ww. Notice that a geodesic mapping φ\varphi is ww-balanced if and only if all its residue vectors vanish with respect to ww. Lemma 2.5 means that if all the residue vectors are dragging xix_{i}’s away from qq, then all the xix_{i}’s must stay not far away from qq.

The definition of residue vector is similar to the concept of discrete tension field in [12].

Refer to caption
Figure 1. The residue vector and Lemma 2.5.
Proof of Theorem 2.3.

Fix an arbitrary base point qM~q\in\tilde{M}, and then by Lemma 2.5 we can pick a sufficiently large constant R=R(M,T,ψ,K)>0R=R(M,T,\psi,K)>0 such that if

i=1nd(xi,q)2=R2,\sum_{i=1}^{n}d(x_{i},q)^{2}=R^{2},

there exists iVi\in V such that

v(xi,q)j:ijEwijv(xi,Aijxj)>0.v(x_{i},q)\cdot\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j})>0.

We will prove that the compact set

K={φ[x1,,xn]:i=1nd(xi,q)2R2}K^{\prime}=\{\varphi[x_{1},...,x_{n}]:\sum_{i=1}^{n}d(x_{i},q)^{2}\leq R^{2}\}

is satisfactory.

For any xM~x\in\tilde{M}, let Px:TxM~TqM~P_{x}:T_{x}\tilde{M}\rightarrow T_{q}\tilde{M} be the parallel transport along the geodesic γx,q\gamma_{x,q}. Set

B={(v1,,vn)(TqM~)n:i=1n|vi|21}B=\{(v_{1},...,v_{n})\in(T_{q}\tilde{M})^{n}:\sum_{i=1}^{n}|v_{i}|^{2}\leq 1\}

as a Euclidean 2n2n-dimensional unit ball, and construct a map F:B(TqM~)nF:B\rightarrow(T_{q}\tilde{M})^{n} in the following three steps. Firstly, we construct nn points x1,,xnM~x_{1},...,x_{n}\in\tilde{M} as xi(v1,,vn)=expq(Rvi).x_{i}(v_{1},...,v_{n})=\exp_{q}(Rv_{i}). Secondly, we compute the residue vector at each xix_{i} as

ri=j:ijEwijv(xi,Aijxj)TxiM~.r_{i}=\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j})\in T_{x_{i}}\tilde{M}.

Lastly, we pull back the residues to TqM~T_{q}\tilde{M} as F(v1,,vn)=(Px1(r1),,Pxn(rn)).F(v_{1},...,v_{n})=\left(P_{x_{1}}(r_{1}),...,P_{x_{n}}(r_{n})\right).

Notice that the map (v1,,vn)φ[x1,,xn](v_{1},...,v_{n})\mapsto\varphi[x_{1},...,x_{n}] is a homeomorphism from BB to KK^{\prime}, and F(v1,,vn)=0F(v_{1},...,v_{n})=0 if and only if the corresponding φ[x1,,xn]\varphi[x_{1},...,x_{n}] in KK^{\prime} is ww-balanced map. Hence, it suffices to prove that FF has a zero in BB. By Lemma 2.4 it suffices to prove that for any (v1,,vn)B(v_{1},...,v_{n})\in\partial B,

(v1,,vn)F(v1,..,vn)|F(v1,,vn)|.(v_{1},...,v_{n})\neq\frac{F(v_{1},..,v_{n})}{|F(v_{1},...,v_{n})|}.

Suppose (v1,,vn)(v_{1},...,v_{n}) is an arbitrary point on B\partial B, and then it suffices to prove that there exists iVi\in V such that viFi(v1,,vn)=viPxi(ri)<0v_{i}\cdot F_{i}(v_{1},...,v_{n})=v_{i}\cdot P_{x_{i}}(r_{i})<0.

Notice that x1(v1,,vn),,xn(v1,,vn)x_{1}(v_{1},...,v_{n}),...,x_{n}(v_{1},...,v_{n}) satisfy that i=1nd(q,xi)2=R2\sum_{i=1}^{n}d(q,x_{i})^{2}=R^{2}, so by our assumption on RR, there exists iVi\in V such that

v(xi,q)j:ijEwijv(xi,Aijxj)=v(xi,q)ri>0,v(x_{i},q)\cdot\sum_{j:ij\in E}w_{ij}v(x_{i},A_{ij}x_{j})=v(x_{i},q)\cdot r_{i}>0,

and thus,

viPxi(ri)=1d(q,xi)Pxi(v(xi,q))Pxi(ri)=1d(q,xi)v(xi,q)ri<0.v_{i}\cdot P_{x_{i}}(r_{i})=-\frac{1}{d(q,x_{i})}P_{x_{i}}\left(v(x_{i},q)\right)\cdot P_{x_{i}}(r_{i})=-\frac{1}{d(q,x_{i})}v(x_{i},q)\cdot r_{i}<0.

In the rest of this subsection, we will prove Lemma 2.5 by contradiction. Let us first sketch the idea of the proof. Assume iVd(xi,q)2\sum_{i\in V}d(x_{i},q)^{2} is very large, then by a standard compactness argument, there exists a long edge ijij in the geodesic mapping φ[x1,,xn]\varphi[x_{1},...,x_{n}]. Assume d(q,xi)d(q,xj)d(q,x_{i})\geq d(q,x_{j}), then the corresponding long edge γxi,Aijxj\gamma_{x_{i},A_{ij}x_{j}} in M~\tilde{M} is pulling xix_{i} towards qq. This implies that there exists another long edge γxi,Aikxk\gamma_{x_{i},A_{ik}x_{k}} dragging xix_{i} away from qq, otherwise the residue vector rir_{i} would not drag xix_{i} away from qq. It can be shown that d(q,xk)>d(q,xi)d(q,x_{k})>d(q,x_{i}). Repeating the above steps, we can find an arbitrary long sequence of vertices such that the distance from each of these vertices to qq is increasing. This is impossible as we only have finitely many vertices.

Refer to caption
Figure 2. Triangles in Step (b), (c), and (d).

Here is a listing of properties serving as the building blocks of the proof of Lemma 2.5.

Lemma 2.6.

(a) For any constant C>0C>0, there exists a constant C1=C1(M,T,ψ,C)>0C_{1}=C_{1}(M,T,\psi,C)>0 such that if

iVd(xi,q)2C1,\sum_{i\in V}d(x_{i},q)^{2}\geq C_{1},

then

maxijEd(xi,Aijxj)C.\max_{ij\in E}d(x_{i},A_{ij}x_{j})\geq C.

(b) There exists a constant C2=C2(M,T,ψ)>0C_{2}=C_{2}(M,T,\psi)>0 such that if

d(Aijxj,q)C2,d(A_{ij}x_{j},q)\geq C_{2},

then

(Ajiq)xjq=q(Aijxj)(Aijq)π8.\angle(A_{ji}q)x_{j}q=\angle q(A_{ij}x_{j})(A_{ij}q)\leq\frac{\pi}{8}.

(c) There exists a constant C3>0C_{3}>0 such that if x,yM~x,y\in\tilde{M} satisfy that

d(y,q)d(x,q)+C3,d(y,q)\geq d(x,q)+C_{3},

then

xyqπ4.\angle xyq\leq\frac{\pi}{4}.

(d) There exists a constant C4>0C_{4}>0 such that if x,yM~x,y\in\tilde{M} satisfy that

d(x,y)C4, and d(x,q)d(y,q),d(x,y)\geq C_{4},\quad\text{ and }\quad d(x,q)\geq d(y,q),

then

yxqπ8.\angle yxq\leq\frac{\pi}{8}.

(e) For any constant C>0C>0, there exists a constant C5=C5(M,T,ψ,C)>0C_{5}=C_{5}(M,T,\psi,C)>0 such that if

maxijEd(xi,Aijxj)C5,\max_{ij\in E}d(x_{i},A_{ij}x_{j})\geq C_{5},

then there exists ijEij\in E such that

v(xi,q)|v(xi,q)|v(xi,Aijxj)C.\frac{v(x_{i},q)}{|v(x_{i},q)|}\cdot v(x_{i},A_{ij}x_{j})\geq C.

(f) For any constant C>0C>0, there exists a constant C6=C6(M,T,ψ,λw,C)>0C_{6}=C_{6}(M,T,\psi,\lambda_{w},C)>0 such that if

v(xi,q)|v(xi,q)|v(xi,Aijxj)C6\frac{v(x_{i},q)}{|v(x_{i},q)|}\cdot v(x_{i},A_{ij}x_{j})\geq C_{6}

for some edge ijEij\in E, then there exists ikEik\in E such that

v(xi,q)|v(xi,q)|v(xi,Aikxk)C.\frac{v(x_{i},q)}{|v(x_{i},q)|}\cdot v(x_{i},A_{ik}x_{k})\leq-C.

(g) For any constant C>0C>0, there exists a constant C7=C7(M,T,ψ,C)>0C_{7}=C_{7}(M,T,\psi,C)>0 such that if

v(xi,q)|v(xi,q)|v(xi,Aikxk)C7,\frac{v(x_{i},q)}{|v(x_{i},q)|}\cdot v(x_{i},A_{ik}x_{k})\leq-C_{7},

then

d(xk,q)d(xi,q)+C.d(x_{k},q)\geq d(x_{i},q)+C.

(h) For any constant C>0C>0, there exists a constant C8=C8(M,T,ψ,C)>0C_{8}=C_{8}(M,T,\psi,C)>0 such that if

d(xj,q)d(xi,q)+C8,d(x_{j},q)\geq d(x_{i},q)+C_{8},

then

v(xj,q)|v(xj,q)|v(xj,Ajixi)C.\frac{v(x_{j},q)}{|v(x_{j},q)|}\cdot v(x_{j},A_{ji}x_{i})\geq C.
Refer to caption
Figure 3. Vertices leaving the point qq.
Proof of Lemma 2.5 assuming Lemma 2.6.

For any C>0C>0, there exists a sufficiently large constant C~=C~(M,T,ψ,λw,C)\tilde{C}=\tilde{C}(M,T,\psi,\lambda_{w},C) determined from (a), (e), (f) and (g) in Lemma 2.6 such that if

iVd(xi,q)2C~,\sum_{i\in V}d(x_{i},q)^{2}\geq\tilde{C},

then there exist three vertices xix_{i}, xjx_{j}, and xkx_{k} shown in Figure 3 with

d(xk,q)d(xj,q)+C.d(x_{k},q)\geq d(x_{j},q)+C.

Moreover, by (g), (e) and (f) of Lemma 2.6, we can find another vertex xlx_{l} such that

d(xl,q)d(xk,q)+Cd(xj,q)+2C,d(x_{l},q)\geq d(x_{k},q)+C\geq d(x_{j},q)+2C,

if the constant C~(M,T,ψ,λw,C)\tilde{C}(M,T,\psi,\lambda_{w},C) is sufficiently large.

Inductively, we can find a sequence i1,,in+1Vi_{1},...,i_{n+1}\in V such that

d(xi1,q)>d(xi2,q)>>d(xin+1,q).d(x_{i_{1}},q)>d(x_{i_{2}},q)>...>d(x_{i_{n+1}},q).

This contradicts to the fact that VV only has nn different elements. ∎

Proof of Lemma 2.6.

(a) By a standard compactness argument, the set

{φX~:maxijElength(φij([0,1]))C}\{\varphi\in\tilde{X}:\max_{ij\in E}~{}~{}length(\varphi_{ij}([0,1]))\leq C\}

is a compact subset of X~\tilde{X}. Notice that (x1,,xn)φ[x1,,xn](x_{1},...,x_{n})\mapsto\varphi[x_{1},...,x_{n}] is a homeomorphism from M~n\tilde{M}^{n} to X~\tilde{X} and

length(φij([0,1]))=d(xi,Aijxj).length(\varphi_{ij}([0,1]))=d(x_{i},A_{ij}x_{j}).

Therefore,

{(x1,,xn)M~n:maxijEd(xi,Aijxj)C}\{(x_{1},...,x_{n})\in\tilde{M}^{n}:\max_{ij\in E}~{}~{}d(x_{i},A_{ij}x_{j})\leq C\}

is compact and the conclusion follows.

(b) We claim that the constant C2C_{2}, which is determined by

sinhC2=maxijEsinhd(Aijq,q)sinπ8,\sinh C_{2}=\frac{\max_{ij\in E}\sinh d(A_{ij}q,q)}{\sin\frac{\pi}{8}},

is satisfactory. Let ABC\triangle ABC be the hyperbolic triangle with the corresponding edge lengths

a=d(Aijxj,q),b=d(Aijxj,Aijq),c=d(Aijq,q).a=d(A_{ij}x_{j},q),\quad b=d(A_{ij}x_{j},A_{ij}q),\quad c=d(A_{ij}q,q).

Since M~\tilde{M} is a CAT(1-1) space, it suffices to show that Cπ/8\angle C\leq\pi/8. By the hyperbolic law of sine,

sinC=sinhcsinAsinhamaxijEsinhd(Aijq,q)1sinhC2=sinπ8.\sin\angle C=\frac{\sinh c\cdot\sin\angle A}{\sinh a}\leq\frac{\max_{ij\in E}\sinh d(A_{ij}q,q)\cdot 1}{\sinh C_{2}}=\sin\frac{\pi}{8}.

(c) We claim that the constant C3C_{3} determined by

sinhC3=1sinπ8\sinh C_{3}=\frac{1}{\sin\frac{\pi}{8}}

is satisfactory. Let ABC\triangle ABC be the hyperbolic triangle with the corresponding edge lengths

a=d(x,y),b=d(y,q),c=d(x,q).a=d(x,y),\quad b=d(y,q),\quad c=d(x,q).

Since M~\tilde{M} is a CAT(1-1) space, it suffices to show that Cπ/8\angle C\leq\pi/8. By the hyperbolic law of sine

sinC=sinhcsinBsinhbsinhcsinhbsinhcsinh(c+C3)sinhcsinhcsinhC3=sinπ8.\sin\angle C=\frac{\sinh c\cdot\sin\angle B}{\sinh b}\leq\frac{\sinh c}{\sinh b}\leq\frac{\sinh c}{\sinh(c+C_{3})}\leq\frac{\sinh c}{\sinh c\cdot\sinh C_{3}}=\sin\frac{\pi}{8}.

(d) We claim that the constant C4C_{4} determined by

sin2π8coshC4=2\sin^{2}\frac{\pi}{8}\cdot\cosh C_{4}=2

is satisfactory. Let ABC\triangle ABC be the hyperbolic triangle with the corresponding edge lengths

a=d(x,y),b=d(y,q),c=d(x,q).a=d(x,y),\quad b=d(y,q),\quad c=d(x,q).

Since M~\tilde{M} is a CAT(1-1) space, it suffices to show that Bπ/8\angle B\leq\pi/8. By the hyperbolic law of cosine,

cosA=cosBcosC+sinBsinCcosha.\cos A=-\cos B\cos C+\sin B\sin C\cosh a.

Then,

2sinBsinCcoshasinBsinCcoshC4=2sinBsinCsin2π82sin2Bsin2π8.2\geq\sin B\sin C\cosh a\geq\sin B\sin C\cosh C_{4}=2\cdot\frac{\sin B\sin C}{\sin^{2}\frac{\pi}{8}}\geq 2\cdot\frac{\sin^{2}B}{\sin^{2}\frac{\pi}{8}}.

Thus, Bπ/8\angle B\leq\pi/8.

Refer to caption
Figure 4. Triangles in Step (5).

(e) We claim that the constant C5C_{5} determined by

C5=max{C4,2C2,2C}C_{5}=\max\{C_{4},2C_{2},\sqrt{2}C\}

is satisfactory. Assume ijEij\in E and d(xi,Aijxj)C5d(x_{i},A_{ij}x_{j})\geq C_{5}, and we have two cases shown in Figure 4.

If d(xi,q)d(Aijxj,q)d(x_{i},q)\geq d(A_{ij}x_{j},q), then by part (d)

(Aijxj)xiqπ8π4,\angle(A_{ij}x_{j})x_{i}q\leq\frac{\pi}{8}\leq\frac{\pi}{4},

and

v(xi,q)|v(xi,q)|v(xi,Aijxj)=cos((Aijxj)xiq)d(xi,Aijxj)12C5C.\frac{v(x_{i},q)}{|v(x_{i},q)|}\cdot v(x_{i},A_{ij}x_{j})=\cos(\angle(A_{ij}x_{j})x_{i}q)\cdot d(x_{i},A_{ij}x_{j})\geq\frac{1}{\sqrt{2}}C_{5}\geq C.

If d(xi,q)d(Aijxj,q)d(x_{i},q)\leq d(A_{ij}x_{j},q), then d(Aijxj,q)C2d(A_{ij}x_{j},q)\geq C_{2}. By part (b) and part (d),

(Ajiq)xjqπ8, and (Ajiq)xj(Ajixi)=q(Aijxj)xiπ8,\angle(A_{ji}q)x_{j}q\leq\frac{\pi}{8},\quad\text{ and }\quad\angle(A_{ji}q)x_{j}(A_{ji}x_{i})=\angle q(A_{ij}x_{j})x_{i}\leq\frac{\pi}{8},

and qxj(Ajixi)π/4\angle qx_{j}(A_{ji}x_{i})\leq\pi/4. Therefore,

v(xj,q)|v(xj,q)|v(xj,Ajixi)=cos((Ajixj)xjq)d(xj,Ajixi)12C5C.\frac{v(x_{j},q)}{|v(x_{j},q)|}\cdot v(x_{j},A_{ji}x_{i})=\cos(\angle(A_{ji}x_{j})x_{j}q)\cdot d(x_{j},A_{ji}x_{i})\geq\frac{1}{\sqrt{2}}C_{5}\geq C.

(f) We claim that the constant C6C_{6} determined by

C6=nλwCC_{6}=n\lambda_{w}\cdot C

is satisfactory. If not, for any ikEik\in E, we have

v(xi,q)|v(xi,q)|v(xi,Aikxk)>C.\frac{v(x_{i},q)}{|v(x_{i},q)|}\cdot v(x_{i},A_{ik}x_{k})>-C.

Then

0v(xi,q)|v(xi,q)|ikEwikv(xi,Aikxk)>wijC6+ikEwik(C)0\geq\frac{v(x_{i},q)}{|v(x_{i},q)|}\cdot\sum_{ik\in E}w_{ik}v(x_{i},A_{ik}x_{k})>w_{ij}C_{6}+\sum_{ik\in E}w_{ik}(-C)
wijC6+ikEλwwij(C)wij(C6nλwC)0,\geq w_{ij}C_{6}+\sum_{ik\in E}\lambda_{w}w_{ij}(-C)\geq w_{ij}(C_{6}-n\lambda_{w}C)\geq 0,

and it is a contradiction.

(g) We claim that C7=C+maxijEd(Aijq,q)C_{7}=C+\max_{ij\in E}d(A_{ij}q,q) is satisfactory. Notice that

d(Aijxj,q)=d(xj,Ajiq)d(xj,q)+d(q,Ajiq)d(xj,q)+maxijEd(Aijq,q).d(A_{ij}x_{j},q)=d(x_{j},A_{ji}q)\leq d(x_{j},q)+d(q,A_{ji}q)\leq d(x_{j},q)+\max_{ij\in E}d(A_{ij}q,q).

By part (1) of Lemma 2.2,

d(Aijxj,q)|v(xi,Aijxj)v(xi,q)|(v(xi,Aijxj)v(xi,q))v(xi,q)|v(xi,q)|d(A_{ij}x_{j},q)\geq|v(x_{i},A_{ij}x_{j})-v(x_{i},q)|\geq-\left(v(x_{i},A_{ij}x_{j})-v(x_{i},q)\right)\cdot\frac{v(x_{i},q)}{|v(x_{i},q)|}
=C7+|v(xi,q)|=C7+d(xi,q).=C_{7}+|v(x_{i},q)|=C_{7}+d(x_{i},q).

Then

d(xj,q)d(xi,q)C7maxijEd(Aijq,q)=C.d(x_{j},q)-d(x_{i},q)\geq C_{7}-\max_{ij\in E}d(A_{ij}q,q)=C.

(h) We claim that the constant C8C_{8} determined by

C8=max{C3,2C}+maxijEd(Aijq,q)C_{8}=\max\{C_{3},\sqrt{2}C\}+\max_{ij\in E}d(A_{ij}q,q)

is satisfactory. Notice that

d(xj,q)d(xi,q)+C8d(xi,Aijq)d(Aijq,q)+C8d(Ajixi,q)+max{C3,2C}.d(x_{j},q)\geq d(x_{i},q)+C_{8}\geq d(x_{i},A_{ij}q)-d(A_{ij}q,q)+C_{8}\geq d(A_{ji}x_{i},q)+\max\{C_{3},\sqrt{2}C\}.

Then by part (c), (Ajixi)xjqπ/4\angle(A_{ji}x_{i})x_{j}q\leq\pi/4, and by the triangle inequality,

d(xj,Ajixi)d(xj,q)d(Ajixi,q)2C.d(x_{j},A_{ji}x_{i})\geq d(x_{j},q)-d(A_{ji}x_{i},q)\geq\sqrt{2}C.

Therefore,

v(xj,q)|v(xj,q)|v(xj,Ajixi)=cos((Ajixi)xjq)d(xj,Ajixi)122C=C.\frac{v(x_{j},q)}{|v(x_{j},q)|}\cdot v(x_{j},A_{ji}x_{i})=\cos(\angle(A_{ji}x_{i})x_{j}q)\cdot d(x_{j},A_{ji}x_{i})\geq\frac{1}{\sqrt{2}}\cdot\sqrt{2}C=C.

2.3. Proof of the Continuity

Proof of the continuity part of Theorem 1.5.

If not, there exists ϵ>0\epsilon>0 and a weight ww and a sequence of weights w(k)w^{(k)} such that

  1. (1)

    w(k)w^{(k)} converge to ww, and

  2. (2)

    dX~(Φ(w(k)),Φ(w))ϵd_{\tilde{X}}(\Phi(w^{(k)}),\Phi(w))\geq\epsilon for any k1k\geq 1.

By the stronger existence result Theorem 2.3, the sequence Φ(w(k))\Phi(w^{(k)}) are in some fixed compact subset KK^{\prime} of X~\tilde{X}. By picking a subsequence, we may assume that Φ(w(k))\Phi(w^{(k)}) converge to some φX~\varphi\in\tilde{X}. Since Φ(w(k))\Phi(w^{(k)}) is w(k)w^{(k)}-balanced, then by the continuity of the residue vectors rir_{i}, φ\varphi is ww-balanced, and thus Φ(w)=φ\Phi(w)=\varphi, which is contradictory to that Φ(w(k))\Phi(w^{(k)}) does not converge to Φ(w)\Phi(w). ∎

3. Proof of Theorem 1.6

3.1. Set up and preparations

Assume φX~\varphi\in\tilde{X} is ww-balanced for some weight ww, and we will prove that φ\varphi is an embedding. Recall that qi=φ(i)q_{i}=\varphi(i) for each iVi\in V, and denote lijl_{ij} as the length of φij([0,1])\varphi_{ij}([0,1]) for any ijEij\in E. It is not difficult to show that φ\varphi has a continuous extension φ~\tilde{\varphi} defined on |T||T|, such that for any triangle σF\sigma\in F a continuous lifting map Φσ\Phi_{\sigma} of φ~|σ\tilde{\varphi}|_{\sigma} from σ\sigma to M~\tilde{M} will

  1. (1)

    map σ\sigma to a geodesic triangle in M~\tilde{M} homeomorphically if φ(σ)\varphi(\partial\sigma) does not degenerate to a geodesic, and

  2. (2)

    map σ\sigma to Φσ(σ)\Phi_{\sigma}(\partial\sigma) if φ(σ)\varphi(\partial\sigma) degenerates to a geodesic.

The main tool to prove Theorem 1.6 is the Gauss-Bonnet formula. We will need to define the inner angles for each triangle in φ(T(1))\varphi(T^{(1)}), even for the degenerate triangles. A convenient way is to assign a “direction” to each edge, even for the degenerate edges with zero length.

Definition 3.1.

A direction field is a map v:ETMv:\vec{E}\rightarrow TM satisfying that

  1. (1)

    vijTqiMv_{ij}\in T_{q_{i}}M for any (i,j)E(i,j)\in\vec{E}, and

  2. (2)

    |vij|=1|v_{ij}|=1 for any (i,j)E(i,j)\in\vec{E}.

Given a direction field vv, define the inner angle of the triangle σ=ijk\sigma=\triangle ijk at the vertex ii as

θσi=θσi(v)=vij0vik=arccos(vijvik),\theta^{i}_{\sigma}=\theta^{i}_{\sigma}(v)=\angle v_{ij}0v_{ik}=\arccos(v_{ij}\cdot v_{ik}),

where 0 is the origin and vij0vik\angle v_{ij}0v_{ik} is the angle between vijv_{ij} and vikv_{ik} in TqiMT_{q_{i}}M.

A direction field vv assigns a unit tangent vector in TqiMT_{q_{i}}M to each directed edge starting from ii, and determines the inner angles in TT.

Definition 3.2.

A direction field vv is admissible if

  1. (1)
    vij=φij(0)lijv_{ij}=\frac{\varphi_{ij}^{\prime}(0)}{l_{ij}}

    if lij>0l_{ij}>0, and

  2. (2)

    vij=vjiv_{ij}=-v_{ji} in TqiM=TqjMT_{q_{i}}M=T_{q_{j}}M if lij=0l_{ij}=0, and

  3. (3)

    for a fixed vertex iVi\in V, if lij=0l_{ij}=0 for any neighbor jj of ii, then there exist neighbors jj and kk of ii such that vij=vikv_{ij}=-v_{ik}, and

  4. (4)

    if σ=ijkF\sigma=\triangle ijk\in F and lij=ljk=lik=0l_{ij}=l_{jk}=l_{ik}=0, then θσi(v)+θσj(v)+θσk(v)=π\theta^{i}_{\sigma}(v)+\theta^{j}_{\sigma}(v)+\theta^{k}_{\sigma}(v)=\pi.

An admissible direction field encodes the directions of the non-degenerate edges in φ(T(1))\varphi(T^{(1)}), and the induced angle sum of a degenerate triangle is always π\pi. Then for any admissible vv and triangle σF\sigma\in F, by the Gauss-Bonnet formula

(4) π=iσθσi(v)Φσ(σ)K𝑑Aiσθσi(v)φ~(σ)K𝑑A.\pi=\sum_{i\in\sigma}\theta^{i}_{\sigma}(v)-\int_{\Phi_{\sigma}(\sigma)}KdA\geq\sum_{i\in\sigma}\theta^{i}_{\sigma}(v)-\int_{\tilde{\varphi}(\sigma)}KdA.

Here dAdA is the area form on (M~,g~)(\tilde{M},\tilde{g}) or (M,g)(M,g).

The concept of the direction field is similar to the discrete one form defined in [13].

3.2. Proof of Theorem 1.6

The proof of Theorem 1.6 uses the four lemmas below. We will postpone their proofs to the the subsequent subsections.

Lemma 3.3.

If vv is admissible and θ=θ(v)\theta=\theta(v), then for any iVi\in V,

σ:iσθσi=2π,\sum_{\sigma:i\in\sigma}\theta^{i}_{\sigma}=2\pi,

and for any σ,σF\sigma,\sigma^{\prime}\in F, φ~(σ)φ~(σ)\tilde{\varphi}(\sigma)\cap\tilde{\varphi}(\sigma^{\prime}) has area 0.

Based on Lemma 3.3, if admissible direction fields exist, the image of the star of each vertex determined by φ~\tilde{\varphi} does not contain any flipped triangles overlapping with each other. If φ~(σ)\tilde{\varphi}(\sigma) does not degenerate to a geodesic arc for any triangle σF\sigma\in F, then φ~\tilde{\varphi} is locally homeomorphic and thus globally homeomorphic as a degree-one map. Therefore, we only need to exclude the existence of degenerate triangles.

Define an equivalence relation on VV as follows. Two vertices i,ji,j are equivalent if there exists a sequence of vertices i=i0,i1,,ik=ji=i_{0},i_{1},...,i_{k}=j such that li0i1==lik1ik=0l_{i_{0}i_{1}}=...=l_{i_{k-1}i_{k}}=0. This equivalence relation introduces a partition V=V1Vm.V=V_{1}\cup...\cup V_{m}. Denote ykMy_{k}\in M as the only point in φ(Vk)\varphi(V_{k}). For any xMx\in M and u,vTxMu,v\in T_{x}M, denote uvu\|v as uu and vv are parallel, i.e., there exists (α,β)(0,0)(\alpha,\beta)\neq(0,0) such that αu+βv=0\alpha u+\beta v=0.

The following Lemma 3.4 shows that there are plenty of choices of admissible direction fields.

Lemma 3.4.

For any v1Ty1M,,vmTymMv_{1}\in T_{y_{1}}M,...,v_{m}\in T_{y_{m}}M, there exists an admissible vv such that vijvkv_{ij}\|v_{k} if iVki\in V_{k} and lij=0l_{ij}=0.

The following Lemma 3.5 shows that for any VkV_{k} with at least two vertices, the image of its “neighborhood” lies in a geodesic.

Lemma 3.5.

If |Vk|2|V_{k}|\geq 2, then there exists vkTykMv_{k}\in T_{y_{k}}M such that vkφij(0)v_{k}\|\varphi_{ij}^{\prime}(0) if iVki\in V_{k} and lij>0l_{ij}>0.

Now let vkv_{k} be as Lemma 3.5 if |Vk|2|V_{k}|\geq 2, and arbitrary if |Vk|=1|V_{k}|=1. Then construct an admissible direction field vv as in Lemma 3.4, with induced inner angles θσi=θσi(v)\theta^{i}_{\sigma}=\theta^{i}_{\sigma}(v). If the image of a triangle σ\sigma under φ\varphi degenerates to a geodesic, then its inner angles θσi\theta^{i}_{\sigma} are π\pi or 0. Let FF^{\prime}\neq\emptyset be the set of degenerate triangles under φ\varphi.

Lemma 3.6.

If σF\sigma\in F^{\prime}, iσi\in\sigma, and θσi=π\theta^{i}_{\sigma}=\pi, then σF\sigma^{\prime}\in F^{\prime} for any σ\sigma^{\prime} in the star neighborhood of the vertex ii.

Let Ω\Omega be a connected component of the interior of {σ:σF}|T|\cup\{\sigma:\sigma\in F^{\prime}\}\subset|T|, and Ω~\tilde{\Omega} be the completion of Ω\Omega under the natural path metric on Ω\Omega. Notice that Ω~\tilde{\Omega} could be different from the closure of Ω\Omega in MM.

Since φ~\tilde{\varphi} is surjective, FFF^{\prime}\neq F and Ω|T|\Omega\neq|T| and Ω~\tilde{\Omega} has non-empty boundary. Then Ω~\tilde{\Omega} is a connected surface with a natural triangulation T=(V,E,F)T^{\prime}=(V^{\prime},E^{\prime},F^{\prime}), and

χ(Ω~)=22×(genus of Ω~)#{boundary components of Ω~}1.\chi(\tilde{\Omega})=2-2\times(\text{genus of }\tilde{\Omega})-\#\{\text{boundary components of $\tilde{\Omega}$}\}\leq 1.

Assume VIV^{\prime}_{I} is the set of interior vertices, and VBV^{\prime}_{B} is the set of boundary vertices, and EIE^{\prime}_{I} is the set of interior edges, and EBE^{\prime}_{B} is the set of boundary edges of Ω~\tilde{\Omega}. Then |VB|=|EB||V^{\prime}_{B}|=|E^{\prime}_{B}|, and by Lemma 3.6, if iVBi\in V^{\prime}_{B} and iσi\in\sigma, then θσi=0\theta^{i}_{\sigma}=0. Therefore,

π|F|=σF,iσθσi=iVIσF:iσθσi=2π|VI|.\pi|F^{\prime}|=\sum_{\sigma\in F^{\prime},i\in\sigma}\theta^{i}_{\sigma}=\sum_{i\in V^{\prime}_{I}}\sum_{\sigma\in F^{\prime}:i\in\sigma}\theta^{i}_{\sigma}=2\pi|V_{I}^{\prime}|.

Thus,

1χ(Ω~)=\displaystyle 1\geq\chi(\tilde{\Omega})= |V||E|+|F|=|VI|+|VB||EI||EB|+|F|\displaystyle|V^{\prime}|-|E^{\prime}|+|F^{\prime}|=|V^{\prime}_{I}|+|V^{\prime}_{B}|-|E^{\prime}_{I}|-|E^{\prime}_{B}|+|F^{\prime}|
=\displaystyle= |VB||EI||EB|+32|F|=|EI|+32|F|\displaystyle|V^{\prime}_{B}|-|E^{\prime}_{I}|-|E^{\prime}_{B}|+\frac{3}{2}|F^{\prime}|=-|E^{\prime}_{I}|+\frac{3}{2}|F^{\prime}|
=\displaystyle= |EI|+12(|EB|+2|EI|)=12|EB|.\displaystyle-|E^{\prime}_{I}|+\frac{1}{2}(|E^{\prime}_{B}|+2|E^{\prime}_{I}|)=\frac{1}{2}|E_{B}^{\prime}|.

Therefore, |VB|=|EB|2|V^{\prime}_{B}|=|E^{\prime}_{B}|\leq 2. Since Ω~\tilde{\Omega} has non-empty boundary, |EB|=1|E_{B}^{\prime}|=1 or 22. In either case, it contradicts with the fact that TT is a simplicial complex.

3.3. Proof of Lemma 3.3

We claim that for any iVi\in V,

σ:iσθσi2π.\sum_{\sigma:i\in\sigma}\theta^{i}_{\sigma}\geq 2\pi.

If lij=0l_{ij}=0 for any neighbor jj of ii, this is a consequence of condition (3) in the definition of an admissible direction field. If lij0l_{ij}\neq 0, by the ww-balanced condition, {φij(0)/lij:ijE}\{\varphi_{ij}^{\prime}(0)/l_{ij}:ij\in E\} should not be contained in any open half unit circle, and the angle sum around ii should be at least 2π2\pi.

By the fact that φ~\tilde{\varphi} is surjective and equation (4), we have

iV(2πσ:iσθσi)+σFφ~(σ)K𝑑AσFφ~(σ)K𝑑AMK𝑑A=2πχ(M),\sum_{i\in V}(2\pi-\sum_{\sigma:i\in\sigma}\theta^{i}_{\sigma})+\sum_{\sigma\in F}\int_{\tilde{\varphi}(\sigma)}KdA\leq\sum_{\sigma\in F}\int_{\tilde{\varphi}(\sigma)}KdA\leq\int_{M}KdA=2\pi\chi(M),

and

iV(2πσ:iσθσi)+σFφ~(σ)K𝑑A2π|V|σF(iσθσiφ~(σ)K𝑑A)=2πχ(M).\sum_{i\in V}(2\pi-\sum_{\sigma:i\in\sigma}\theta^{i}_{\sigma})+\sum_{\sigma\in F}\int_{\tilde{\varphi}(\sigma)}KdA\geq 2\pi|V|-\sum_{\sigma\in F}(\sum_{i\in\sigma}\theta^{i}_{\sigma}-\int_{\tilde{\varphi}(\sigma)}KdA)=2\pi\chi(M).

Hence, the inequalities above are equalities. This fact implies that

iV(2πσ:iσθσi)=0.\sum_{i\in V}(2\pi-\sum_{\sigma:i\in\sigma}\theta^{i}_{\sigma})=0.

Since each term in this summation is non-positive, then σ:iσθσi=2π\sum_{\sigma:i\in\sigma}\theta^{i}_{\sigma}=2\pi. The statement on the area follows similarly.

3.4. Proof of Lemma 3.4

We claim that for any kk, there exists a map h:Vkh:V_{k}\rightarrow\mathbb{R} such that

  1. (1)

    h(i)h(j)h(i)\neq h(j) if iji\neq j, and

  2. (2)

    for a fixed iVki\in V_{k}, if lij=0l_{ij}=0 for any neighbor jj of ii, then there exist neighbors jj and jj^{\prime} of ii in VkV_{k} such that h(j)<h(i)<h(j)h(j)<h(i)<h(j^{\prime}).

Given such hh, set vv as

vij={sgn[h(j)h(i)]vk, if iVk and lij=0,φij(0), if lij>0,v_{ij}=\begin{cases}\text{sgn}[h(j)-h(i)]\cdot v_{k},&\text{ if }i\in V_{k}\text{ and }l_{ij}=0,\\ \varphi_{ij}^{\prime}(0),&\text{ if }l_{ij}>0,\end{cases}

where sgn is the sign function. It is easy to verify that such vv is satisfactory.

To construct such function hh, we will prove the following lemma, which is more general than our claim.

Lemma 3.7.

Assume G=(V,E)G=(V^{\prime},E^{\prime}) is a subgraph of the 11-skeleton T(1)T^{(1)}, and EEE^{\prime}\neq E. Denote

int(G)={iV:ijEijE},int(G)=\{i\in V^{\prime}:ij\in E\Rightarrow ij\in E^{\prime}\},

and G=Vint(G)\partial G=V^{\prime}-int(G). Then there exists h:Vh:V^{\prime}\rightarrow\mathbb{R} such that

  1. (1)

    h(i)h(j)h(i)\neq h(j) if iji\neq j, and

  2. (2)

    for any iint(G)i\in int(G) there exist neighbors jj and jj^{\prime} of ii in VV^{\prime} such that h(j)<h(i)<h(j)h(j)<h(i)<h(j^{\prime}).

Proof.

We prove by the induction on the size of VV^{\prime}. The case |V|=1|V^{\prime}|=1 is trivial. For the case |V|2|V^{\prime}|\geq 2, first notice that |G|2|\partial G|\geq 2 for any proper subgraph GG of T(1)T^{(1)}. Assign distinct values h¯(i)\bar{h}(i) to each iGi\in\partial G, then solve the discrete harmonic equation

j:ijE(h¯(j)h¯(i))=0,iint(G),\sum_{j:ij\in E}(\bar{h}(j)-\bar{h}(i))=0,\quad\forall i\in int(G),

with the given Dirichlet boundary condition on G\partial G.

Let s1<<sks_{1}<...<s_{k} be the all distinct values that appear in {h¯(i):iV}\{\bar{h}(i):i\in V^{\prime}\}. Then consider the subgraphs Gi=(Vi,Ei)G_{i}=(V^{\prime}_{i},E^{\prime}_{i}) defined as

Vi={jV:h(j)=si},V_{i}^{\prime}=\{j\in V^{\prime}:h(j)=s_{i}\},

and

Ei={jjE:j,jVi}.E_{i}^{\prime}=\{jj^{\prime}\in E^{\prime}:j,j^{\prime}\in V_{i}^{\prime}\}.

Notice that |G|2|\partial G|\geq 2, so k2k\geq 2 and |Vi|<|V||V_{i}^{\prime}|<|V^{\prime}| for any i=1,,ki=1,...,k. By the induction hypothesis, there exists a function hi:Vih_{i}:V_{i}^{\prime}\rightarrow\mathbb{R} such that

  1. (1)

    hi(j)hi(j)h_{i}(j)\neq h_{i}(j^{\prime}) if jjj\neq j^{\prime}, and

  2. (2)

    for any jint(Gi)j\in int(G_{i}) there exist neighbors j,j′′j^{\prime},j^{\prime\prime} of jj in ViV^{\prime}_{i} such that hi(j)<hi(j)<hi(j′′)h_{i}(j^{\prime})<h_{i}(j)<h_{i}(j^{\prime\prime}).

Define hi(j)=0h_{i}(j)=0 if jVij\notin V_{i}^{\prime}, then for sufficiently small positive ϵ1,,ϵk\epsilon_{1},...,\epsilon_{k},

h=h¯+i=1kϵihi{h}=\bar{h}+\sum_{i=1}^{k}\epsilon_{i}h_{i}

is a desirable function. ∎

3.5. Proof of Lemma 3.5

It amounts to prove that if i,iVki,i^{\prime}\in V_{k} and ij,ijEij,i^{\prime}j^{\prime}\in E and lij>0l_{ij}>0 and lij>0l_{i^{\prime}j^{\prime}}>0, then φij(0)φij(0)\varphi^{\prime}_{ij}(0)\|\varphi^{\prime}_{i^{\prime}j^{\prime}}(0). Let

D=(i,i,i′′Vkiii′′)(i,iVkii),D=\left(\cup_{i,i^{\prime},i^{\prime\prime}\in V_{k}}\triangle ii^{\prime}i^{\prime\prime}\right)\cup\left(\cup_{i,i^{\prime}\in V_{k}}ii^{\prime}\right),

which is a closed path-connected set in |T||T|. For any iVki\in V_{k}, iDi\in\partial D if and only if there exists ijEij\in E with lij>0l_{ij}>0. Therefore, it suffices to prove that

  1. (1)

    for any iVki\in V_{k} and edges ij,ijij,ij^{\prime} with lij>0l_{ij}>0 and lij>0l_{ij^{\prime}}>0, φij(0)φij(0)\varphi_{ij}^{\prime}(0)\|\varphi_{ij^{\prime}}^{\prime}(0), and

  2. (2)

    for any ijEij\in E satisfying that ijDij\subset\partial D, there exists mVVkm\in V-V_{k} such that ijmF\triangle ijm\in F, and thus φim(0)=φjm(0)\varphi_{im}^{\prime}(0)=\varphi_{jm}^{\prime}(0), and

  3. (3)

    D\partial D is connected.

For part (1), if it is not true, then there exists iVki\in V_{k} and ijEij\in E and ijEij^{\prime}\in E such that lij>0l_{ij}>0 and lij>0l_{ij^{\prime}}>0 and φij(0)\varphi_{ij}^{\prime}(0) is not parallel to φij(0)\varphi_{ij^{\prime}}^{\prime}(0). Assuming this claim, by the ww-balanced condition, there exists ij′′Eij^{\prime\prime}\in E with lij′′>0l_{ij^{\prime\prime}}>0, and the three vectors φij(0),φij(0),φij′′(0)\varphi_{ij}^{\prime}(0),\varphi_{ij^{\prime}}^{\prime}(0),\varphi_{ij^{\prime\prime}}^{\prime}(0) are not contained in any closed half-space in TqiMT_{q_{i}}M. Assume imEim\in E and lim=0l_{im}=0, and without loss of generality, ij,im,ij,ij′′ij,im,ij^{\prime},ij^{\prime\prime} are ordered counter-clockwise in the one-ring neighborhood of ii in TT. By Lemma 3.4, there exists an admissible vv such that vimφij′′(0)v_{im}\|\varphi_{ij^{\prime\prime}}^{\prime}(0). By possibly changing a sign, we may assume that vim=φij′′(0)/lij′′v_{im}=\varphi_{ij^{\prime\prime}}^{\prime}(0)/l_{ij^{\prime\prime}}.

Refer to caption
Figure 5. Overlapping triangles lead to angle surplus.

Then as Figure 5 shows, a contradiction follows

2π=σiσθσivij0vim+vim0vij+vij0vij′′+vij′′0vij2\pi=\sum_{\sigma i\in\sigma}\theta^{i}_{\sigma}\geq\angle v_{ij}0v_{im}+\angle v_{im}0v_{ij^{\prime}}+\angle v_{ij^{\prime}}0v_{ij^{\prime\prime}}+\angle v_{ij^{\prime\prime}}0v_{ij}
=2vij0vij′′+2vij′′0vij>2π.=2\angle v_{ij}0v_{ij^{\prime\prime}}+2\angle v_{ij^{\prime\prime}}0v_{ij^{\prime}}>2\pi.

Part (2) is straightforward and we will prove part (3). By our assumption on the extension φ~\tilde{\varphi}, φ~(D)\tilde{\varphi}(D) contains only one point, then the embedding map iD=ψ1(ψ|D)i_{D}=\psi^{-1}\circ(\psi|_{D}) from DD to |T||T| is homotopic to the constant map ψ1(φ~|D)\psi^{-1}\circ(\tilde{\varphi}|_{D}), meaning that DD is contractible in |T||T|. If D\partial D contains at least two boundary components, then it is not difficult to show that |T|D|T|-D has a connected component DD^{\prime} homeomorphic to an open disk. Let ΦD:DM~\Phi_{D}:D\to\tilde{M} be a lifting of φ~|D\tilde{\varphi}|_{D}. Then ΦD(D)ΦD(D)\Phi_{D}(\partial D^{\prime})\subset\Phi_{D}(D) contains only a single point xM~x\in\tilde{M}. Then by the ww-balanced condition, it is not difficult to derive a maximum principle and show that φ~|D\tilde{\varphi}|_{D^{\prime}} equals to constant xx. Then by the definition of DD it is easy to see that DD^{\prime} should be a subset of DD. It is contradictory.

3.6. Proof of Lemma 3.6

Assume ijij and ijij^{\prime} are two edges in σ\sigma. If the conclusion is not true, then there exists ikEik\in E such that lik>0l_{ik}>0 and vikv_{ik} is not parallel to vijv_{ij}. Notice that vij=vijv_{ij}=-v_{ij^{\prime}}, and we have

2π=σF:iσθσivij0vij+vij0vik+vij0vik=2π.2\pi=\sum_{\sigma\in F:i\in\sigma}\theta^{i}_{\sigma}\geq\angle v_{ij}0v_{ij^{\prime}}+\angle v_{ij}0v_{ik}+\angle v_{ij^{\prime}}0v_{ik}=2\pi.

Then the equality holds in the above inequality, and for any ikEik^{\prime}\in E, vikv_{ik^{\prime}} should be on the half circle that contains vij,vik,vijv_{ij},v_{ik},v_{ij^{\prime}}. Let vmv_{m} be the middle point of this half circle, then

vmj:ijEwijlijvijwiklikvmvik>0.v_{m}\cdot\sum_{j:ij\in E}w_{ij}l_{ij}v_{ij}\geq w_{ik}l_{ik}v_{m}\cdot v_{ik}>0.

This contradicts to the fact that φ\varphi is ww-balanced.

References

  • [1] Marwan Awartani and David W Henderson, Spaces of geodesic triangulations of the sphere, Transactions of the American Mathematical Society 304 (1987), no. 2, 721–732.
  • [2] Ethan D Bloch, Robert Connelly, and David W Henderson, The space of simplexwise linear homeomorphisms of a convex 2-disk, Topology 23 (1984), no. 2, 161–175.
  • [3] Martin R Bridson and André Haefliger, Metric spaces of non-positive curvature, vol. 319, Springer Science & Business Media, 2013.
  • [4] Dmitri Burago, Iu D Burago, Yuri Burago, Sergei Ivanov, Sergei V Ivanov, and Sergei A Ivanov, A course in metric geometry, vol. 33, American Mathematical Soc., 2001.
  • [5] Stewart S Cairns, Isotopic deformations of geodesic complexes on the 2-sphere and on the plane, Annals of Mathematics (1944), 207–217.
  • [6] Erin Wolf Chambers, Jeff Erickson, Patrick Lin, and Salman Parsa, How to morph graphs on the torus, Proceedings of the 2021 ACM-SIAM Symposium on Discrete Algorithms (SODA), SIAM, 2021, pp. 2759–2778.
  • [7] Robert Connelly, David W Henderson, Chung Wu Ho, and Michael Starbird, On the problems related to linear homeomorphisms, embeddings, and isotopies, Continua, decompositions, manifolds, 1983, pp. 229–239.
  • [8] Y Colin de Verdiere, Comment rendre géodésique une triangulation d’une surface, L’Enseignement Mathématique 37 (1991), 201–212.
  • [9] Clifford J Earle, James Eells, et al., A fibre bundle description of teichmüller theory, Journal of Differential Geometry 3 (1969), no. 1-2, 19–43.
  • [10] Michael Floater, One-to-one piecewise linear mappings over triangulations, Mathematics of Computation 72 (2003), no. 242, 685–696.
  • [11] Michael S Floater, Mean value coordinates, Computer Aided Geometric Design 20 (2003), no. 1, 19–27.
  • [12] Jonah Gaster, Brice Loustau, and Léonard Monsaingeon, Computing discrete equivariant harmonic maps, arXiv preprint arXiv:1810.11932 (2018).
  • [13] Steven Gortler, Craig Gotsman, and Dylan Thurston, Discrete one-forms on meshes and applications to 3d mesh parameterization, Computer Aided Geometric Design 23 (2006), no. 2, 83–112.
  • [14] Joel Hass and Peter Scott, Simplicial energy and simplicial harmonic maps, Asian Journal of Mathematics 19 (2015), no. 4, 593–636.
  • [15] Yanwen Luo, Spaces of geodesic triangulations of surfaces, arXiv preprint arXiv:1910.03070. Submitted (2019).
  • [16] Stephen Smale, Diffeomorphisms of the 2-sphere, Proceedings of the American Mathematical Society 10 (1959), no. 4, 621–626.
  • [17] William Thomas Tutte, How to draw a graph, Proceedings of the London Mathematical Society 3 (1963), no. 1, 743–767.