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

The Deformation Spaces of Geodesic Triangulations of Flat Tori

Yanwen Luo, Tianqi Wu, Xiaoping Zhu Department of Mathematics, Rutgers University, New Brunswick NJ, 08817 [email protected] Center of Mathematical Sciences and Applications, Harvard University, Cambridge, MA, 02138 [email protected] Department of Mathematics, Rutgers University, New Brunswick New Jersey 08817 [email protected]
Abstract.

We prove that the deformation space of geodesic triangulations of a flat torus is homotopy equivalent to a torus. This solves an open problem proposed by Connelly et al. in 1983, in the case of flat tori. A key tool of the proof is a generalization of Tutte’s embedding theorem for flat tori. When this paper is under preparation, Erickson and Lin proved a similar result, which works for all convex drawings.

Key words and phrases:
geodesic triangulations, Tutte’s embedding
Acknowledgement: The author was supported in part by NSF 1737876, NSF 1760471, NSF DMS FRG 1760527 and NSF DMS 1811878.

1. Introduction

This paper is a continuation of the previous work [13], where we proved that the deformation space of geodesic triangulations of a surface with negative curvature is contractible. The purpose of this paper is to identify the homotopy type of the deformation space of geodesic triangulations of a flat torus. This solves an open question proposed by Connelly et al. in [4]. The main result of this paper is

Theorem 1.1.

The deformation space of geodesic triangulations of a flat torus is homotopic equivalent to a torus.

It is conjectured in [4] that the space of geodesic triangulations of a closed orientable surface SS with constant curvature deformation retracts to the group of orientation preserving isometries of SS homotopic to the identity. This paper affirms this conjecture in the case of flat tori. The case of hyperbolic surfaces has been proved in [13]. In a very recent work, Erickson-Lin [8] proved independently a generalized version of our Theorem 1.1 for general graph drawings on a flat torus.

The study of the homotopy types of spaces of geodesic triangulations stemmed from the work by Cairns [2]. A brief history of this problem can be found in [13]. These spaces are closely related to diffeomoprhism groups of surfaces. Bloch-Connelly-Henderson [1] proved that the space of geodesic triangulations of a convex polygon is contractible. The space of geodesic triangulations of a planar polygon is equivalent to the space of simplexwise linear homeomorphisms. Hence, the Bloch-Connelly-Henderson theorem can be viewed as a discrete analogue of Smale’s theorem, which states that the diffeomorphism group of the closed 22-disk fixing the boundary pointwise is contractible. Earle-Eells [7] proved that the group of orientation-preserving diffeomorphisms of a torus isotopic to the identity is homotopy equivalent to a torus. Theorem 1.1 can be regarded as a discrete version of this theorem.

Similar to the previous work [13], our key idea to prove Theorem 1.1 originates from Tutte’s embedding theorem.

1.1. Set Up and the Main Theorem

Let 𝕋2=2/2=[0,1]2/\mathbb{T}^{2}=\mathbb{R}^{2}/\mathbb{Z}^{2}=[0,1]^{2}/\sim be the flat torus constructed by gluing the opposite sides of the unit square in 2\mathbb{R}^{2}.

A topological triangulation of 𝕋2\mathbb{T}^{2} can be identified as a homeomorphism ψ\psi from |𝒯||\mathcal{T}| to 𝕋2\mathbb{T}^{2}, where |𝒯||\mathcal{T}| is the carrier of a 2-dimensional simplicial complex 𝒯=(V,E,F)\mathcal{T}=(V,E,F) with the vertex set VV, and the edge set EE, and the face set FF. For convenience, we label the vertices as v1,,vnv_{1},...,v_{n} where n=|V|n=|V| is the number of the vertices. Denote |E||E| as the number of the edges, and eij\vec{e}_{ij} as the directed edge from the vertex ii to its neighbor jj, and E={eij:ijE}\vec{E}=\{\vec{e}_{ij}:ij\in E\} as the set of directed edges, and N(i)N(i) as the indices of neighbored vertices of viVv_{i}\in V.

A geodesic triangulation with the combinatorial type (𝒯,ψ)(\mathcal{T},\psi) is an embedding φ\varphi from the one-skeleton 𝒯(1)\mathcal{T}^{(1)} to 𝕋2\mathbb{T}^{2} satisfying that

  1. (a)

    the restriction φij\varphi_{ij} of φ\varphi on each edge eije_{ij}, identified with a unit interval [0,1][0,1], is a geodesic of constant speed, and

  2. (b)

    φ\varphi is homotopic to the restriction of ψ\psi on 𝒯(1)\mathcal{T}^{(1)}.

Let X=X(𝕋2,𝒯,ψ)X=X(\mathbb{T}^{2},\mathcal{T},\psi) denote the set of all such geodesic triangulations, which is so-called a deformation space of geodesic triangulations of 𝕋2\mathbb{T}^{2}. This space can be defined for other flat tori in a similar fashion. Perturbing each vertex locally, we can construct a family of geodesic triangulations from an initial geodesic triangulation. Therefore, the space XX is naturally a 2n2n-dimensional manifold.

For any geodesic triangulation φX\varphi\in X, we can always translate φ\varphi on 𝕋2\mathbb{T}^{2} to make the image φ(v1)\varphi(v_{1}) of the first vertex v1v_{1} be at the (quotient of the) origin (0,0)(0,0). By this normalization, we can decompose XX as X=X0×𝕋2X=X_{0}\times\mathbb{T}^{2}, where

X0=X0(𝕋2,𝒯,ψ)={φX:φ(v1)=(0,0)}.X_{0}=X_{0}(\mathbb{T}^{2},\mathcal{T},\psi)=\{\varphi\in X:\varphi(v_{1})=(0,0)\}.

Since there are affine transformations between any two flat tori, and an affine transformation always preserves the geodesic triangulations, Theorem 1.1 reduces to the following.

Theorem 1.2.

Given a topological triangulation (𝒯,ψ)(\mathcal{T},\psi) of 𝕋2\mathbb{T}^{2}, the space X0=X0(𝕋2,𝒯,ψ){X_{0}}={X_{0}}(\mathbb{T}^{2},\mathcal{T},\psi) is contractible.

1.2. Key Tool: Generalized Tutte’s Embedding Theorem

Let φ\varphi be a map from 𝒯(1)\mathcal{T}^{(1)} to 𝕋2\mathbb{T}^{2}. Assume φ\varphi maps every edge in EE to a geodesic arc parametrized by constant speed on 𝕋2\mathbb{T}^{2}. A positive assignment w+Ew\in\mathbb{R}_{+}^{\vec{E}} on the set of directed edges is called a weight of 𝒯\mathcal{T}. We say φ\varphi is ww-balanced at viv_{i} if

jN(i)wijφ˙ij=0,\sum_{j\in N(i)}w_{ij}\dot{\varphi}_{ij}=0,

where φ˙ij=φ˙ij(0)Tφ(vi)𝕋22\dot{\varphi}_{ij}=\dot{\varphi}_{ij}(0)\in T_{\varphi(v_{i})}\mathbb{T}^{2}\cong\mathbb{R}^{2}. Then φ˙ij\dot{\varphi}_{ij} indicates the direction of the edge φ(eij)\varphi(e_{ij}) and φ˙ij\|\dot{\varphi}_{ij}\| equals to the length of φ(eij)\varphi(e_{ij}). A map φ\varphi is called ww-balanced if it is ww-balanced at each vertex in VV. We have the following version of Tutte’s embedding theorem, which is a special case of Gortler-Gotsman-Thurston’s embedding result in [11] and Theorem 1.6 in [13].

Theorem 1.3.

Assume (𝒯,ψ)(\mathcal{T},\psi) is a topological triangulation of 𝕋2\mathbb{T}^{2}, and φ\varphi is a map from 𝒯(1)\mathcal{T}^{(1)} to 𝕋2\mathbb{T}^{2} such that φ\varphi is homotopic to ψ|𝒯(1)\psi|_{\mathcal{T}^{(1)}} and the restriction φij\varphi_{ij} of φ\varphi on each edge eije_{ij} is a geodesic parametrized by constant speed. If φ\varphi is ww-balanced for some weight w+Ew\in\mathbb{R}^{\vec{E}}_{+}, then φ\varphi is an embedding, or equivalently φ\varphi is a geodesic triangulation.

To be self-contained, we will give a simple proof for Theorem 1.3 , which is adapted from Gortler-Gotsman-Thurston’s argument in [11].

The classical Tutte’s embedding theorem [14] states that a straight-line embedding of a simple 3-vertex-connected planar graph can be constructed by fixing an outer face as a convex polygon and solving interior vertices on the condition that each vertex is in the convex hull of its neighbors. Various new proofs of Tutte’s embedding theorem have been proposed by Floater [9], Gortler-Gotsman-Thurston [11], et al.

Tutte’s embedding theorem has been generalized by Colin de Verdière [5], Delgado-Friedrichs [6], and Hass-Scott [12] to surfaces with non-positive Gaussian curvatures. They showed that the minimizer of a discrete Dirichlet energy is a geodesic triangulation. Here the fact that φ\varphi is a minimizer of a discrete Dirichlet energy means that φ\varphi is ww-balanced for some symmetric weight ww in +E\mathbb{R}^{\vec{E}}_{+} with wij=wjiw_{ij}=w_{ji}. Their result also implies that X=X(𝕋2,𝒯,ψ)X=X(\mathbb{T}^{2},\mathcal{T},\psi) is not an empty set for any topological triangulation (𝒯,ψ)(\mathcal{T},\psi). Recently, Luo-Wu-Zhu [13] proved a new version of Tutte’s embedding theorem, for nonsymmetric weights and triangulations of orientable closed surfaces with non-positive Gaussian curvature.

Gortler-Gotsman-Thurston [11] generalized Tutte’s embedding theorem to flat tori. In contrast to the case of convex polygons and surfaces of negative curvatures, it is not always possible to construct a geodesic triangulation of 𝕋2\mathbb{T}^{2} such that it is ww-balanced with respect to a given non-symmetric weight ww. See [3] Section 1.1 for a detailed discussion.

1.3. Outline of the Proof

Fix a lifting (x~i,y~i)2(\tilde{x}_{i},\tilde{y}_{i})\in\mathbb{R}^{2} of ψ(vi)𝕋2\psi(v_{i})\in\mathbb{T}^{2} for each i=1,,ni=1,...,n. Then for any eijE\vec{e}_{ij}\in\vec{E}, there exists a unique lifting ψ~ij:[0,1]2\tilde{\psi}_{ij}:[0,1]\rightarrow\mathbb{R}^{2} of ψij=ψ|eij:[0,1]𝕋2\psi_{ij}=\psi|_{e_{ij}}:[0,1]\rightarrow\mathbb{T}^{2} such that ψ~ij(0)=(x~i,y~i)\tilde{\psi}_{ij}(0)=(\tilde{x}_{i},\tilde{y}_{i}). Then

ψ~ij(1)=(x~j,y~j)+(bijx,bijy)\tilde{\psi}_{ij}(1)=(\tilde{x}_{j},\tilde{y}_{j})+(b_{ij}^{x},b_{ij}^{y})

for some lattice point (bijx,bijy)2(b_{ij}^{x},b_{ij}^{y})\in\mathbb{Z}^{2}.

A geodesic triangulation φ\varphi in X0X_{0} can be represented by (xi,yi)2(x_{i},y_{i})\in\mathbb{R}^{2} for i=1,,ni=1,...,n with (x1,y1)=(0,0)(x_{1},y_{1})=(0,0). Under this representation, φij:[0,1]𝕋2\varphi_{ij}:[0,1]\rightarrow\mathbb{T}^{2} is the quotient of the linear map

φij(t)=t(xj+bijx,yj+bijy)+(1t)(xi,yi),\varphi_{ij}(t)=t(x_{j}+b_{ij}^{x},y_{j}+b_{ij}^{y})+(1-t)(x_{i},y_{i}),

and the equations for ww-balanced conditions at all the vertices can be written as

jN(i)wij(xjxi+bijx)=\displaystyle\sum_{j\in N(i)}w_{ij}(x_{j}-x_{i}+b^{x}_{ij})= 0,\displaystyle 0,
jN(i)wij(yjyi+bijy)=\displaystyle\sum_{j\in N(i)}w_{ij}(y_{j}-y_{i}+b^{y}_{ij})= 0.\displaystyle 0.

In a closed matrix form, we can write

(1) A(w)𝐱=𝐛(w),A(w)\mathbf{x}=\mathbf{b}(w),

where the weight matrix A(w)A(w) is

A(w)=(j=1nw1jw12w13w1nw21j=1nw2jw23w2nw31w32j=1nw3jw3nwn1wn2wn3j=1nwnj),A(w)=\begin{pmatrix}-\sum_{j=1}^{n}w_{1j}&w_{12}&w_{13}&\dots&w_{1n}\\ w_{21}&-\sum_{j=1}^{n}w_{2j}&w_{23}&\dots&w_{2n}\\ w_{31}&w_{32}&-\sum_{j=1}^{n}w_{3j}&\dots&w_{3n}\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ w_{n1}&w_{n2}&w_{n3}&\dots&-\sum_{j=1}^{n}w_{nj}\end{pmatrix},

and

𝐱=(x1y1x2y2xnyn),𝐛(w)=(j=1nw1jb1jxj=1nw1jb1jyj=1nw2jb2jxj=1nw2jb2jyj=1nwnjbnjxj=1nwnjbnjy).\mathbf{x}=\begin{pmatrix}x_{1}&y_{1}\\ x_{2}&y_{2}\\ \vdots&\vdots\\ x_{n}&y_{n}\end{pmatrix},\quad\mathbf{b}(w)=\begin{pmatrix}-\sum_{j=1}^{n}w_{1j}b_{1j}^{x}&-\sum_{j=1}^{n}w_{1j}b_{1j}^{y}\\ -\sum_{j=1}^{n}w_{2j}b_{2j}^{x}&-\sum_{j=1}^{n}w_{2j}b_{2j}^{y}\\ \vdots&\vdots\\ -\sum_{j=1}^{n}w_{nj}b_{nj}^{x}&-\sum_{j=1}^{n}w_{nj}b_{nj}^{y}\end{pmatrix}.

Here we denote wij=0w_{ij}=0 if eijEe_{ij}\notin E.

A weight ww in +E\mathbb{R}^{\vec{E}}_{+} is called admissible if the equation (1) is solvable. Let WW be the space of admissible weights. For any wWw\in W, a solution 𝐱\mathbf{x} to the equation (1) uniquely determines the coordinates of the vertices, and a ww-balanced map φ\varphi that is homotopic to ψ|𝒯(1)\psi|_{\mathcal{T}^{(1)}}. By Theorem 1.3, such φ\varphi is an embedding, and φX\varphi\in X. Noticing that A(w)A(w) is weakly diagonal dominant and the graph 𝒯(1)\mathcal{T}^{(1)} is connected, the solution to equation (1) is unique up to a 2-dimensional translation, and is unique if we require (x1,y1)=(0,0)(x_{1},y_{1})=(0,0). Define the Tutte map as

Ψ:WX0\Psi:W\to{X}_{0}

sending an admissible weight ww to the unique ww-balanced geodesic triangulation in X0{X}_{0}. The Tutte map is continuous by the continuous dependence of the solutions to the coefficients in a linear system.

Refer to caption
Figure 1. The mean value coordinate.

The Tutte map is also surjective, since there exists a smooth map σ\sigma from X0{X}_{0} to WW by the “mean value coordinates”

wij=tan(αij/2)+tan(βij/2)lijw_{ij}=\frac{\tan(\alpha_{i}^{j}/2)+\tan(\beta_{i}^{j}/2)}{l_{ij}}

introduced by Floater [10]. Here αij,βij\alpha_{i}^{j},\beta_{i}^{j} are two inner angles adjacent to eije_{ij} at the vertex viv_{i}, and lijl_{ij} is the edge length of φij\varphi_{ij}. See Figure 1 for an illustration. Floater [10] showed that any geodesic triangulation φ\varphi is σ(φ)\sigma(\varphi)-balanced, i.e., Φσ=IdX0\Phi\circ\sigma=Id_{X_{0}}.

Having the knowledge of the Tutte map and the mean value coordinates, Theorem 1.2 reduces to the following proposition.

Proposition 1.4.

There exists a continuous map Φ:+EW\Phi:\mathbb{R}^{\vec{E}}_{+}\to W such that

Φ|W=IdW.\Phi|_{W}=Id_{W}.
Proof of Theorem 1.2 assuming Proposition 1.4..

By Proposition 1.4, WW is contractible since there exists a retraction from the contractible space +E\mathbb{R}^{\vec{E}}_{+} to WW. So σΦ\sigma\circ\Phi is homotopic to the identity map on WW. On the other hand Φσ=IdX0\Phi\circ\sigma=Id_{X_{0}}, and thus X0X_{0} is homotopic to WW and contractible. ∎

1.4. Organization of this Paper

In Section 2, we will prove Proposition 1.4 by constructing a flow. In Section 3, we prove Theorem 1.3 following the idea in Gortler-Gotsman-Thurston’s work [11].

1.5. Acknowledgement

We really appreciate Professor Jeff Erickson for providing valuable insights and references about this problem.

2. Proof of Proposition 1.4

Set an energy function on the weight space +E\mathbb{R}^{\vec{E}}_{+} as

(w)=min𝐱n×2A(w)𝐱𝐛(w)2=min𝐱n×2:x1=y1=0A(w)𝐱𝐛(w)2,\mathcal{E}(w)=\min_{\mathbf{x}\in\mathbb{R}^{n\times 2}}\big{\|}A(w)\mathbf{x}-\mathbf{b}(w)\big{\|}^{2}=\min_{\mathbf{x}\in\mathbb{R}^{n\times 2}:x_{1}=y_{1}=0}\big{\|}A(w)\mathbf{x}-\mathbf{b}(w)\big{\|}^{2},

where the norm is the Frobenius norm of a matrix. The second minimization problem above is a least square problem with 2(n1)2(n-1) real variables and a non-degerate coefficient matrix. By the standard formula in linear least squares (LLS) or quadratic programming (QP), the minimizer, denoted as 𝐱(w)\mathbf{x}(w), is a smooth function of ww, and thus (w)\mathcal{E}(w) is also a smooth function of ww. Note that (w)=0\mathcal{E}(w)=0 if and only if ww is admissible, and intuitively (w)\mathcal{E}(w) measures the deviation of ww from being admissible. The key idea of the proof is to construct a flow on +E\W\mathbb{R}^{\vec{E}}_{+}\backslash W to minimize (w)\mathcal{E}(w), as in the following Lemma 2.1.

Lemma 2.1.

There exists a smooth function Θ:+E\WE\Theta:\mathbb{R}^{\vec{E}}_{+}\backslash W\rightarrow\mathbb{R}^{\vec{E}} and a continuous function C(w)C(w) on +E\mathbb{R}^{\vec{E}}_{+} such that for any initial value w0+Ew^{0}\in\mathbb{R}^{\vec{E}}_{+}, the flow w(t)w(t) defined by

(2) {w˙(t)=Θ(w(t))w(0)=w0\Big{\{}\begin{array}[]{ll}\dot{w}(t)=\Theta(w(t))\\ w(0)=w^{0}\end{array}

satisfies that for any tt in the maximum existing interval [0,T)[0,T),

  1. (a)
    0w˙ij(t)wij(t),0\leq\dot{w}_{ij}(t)\leq w_{ij}(t),

    and

  2. (b)
    d(w(t))dtC(w0)(w(t)).\frac{d\mathcal{E}(w(t))}{dt}\leq-C(w^{0})\sqrt{\mathcal{E}(w(t))}.

Proposition 1.4 is proved in Section 2.1, assuming this Lemma 2.1. Then we construct a flow in Section 2.2, and in Section 2.3 show that this flow is satisfactory for Lemma 2.1.

2.1. Proof of Proposition 1.4 Assuming Lemma 2.1

Assume Θ(w)\Theta(w) and C=C(w0)C=C(w^{0}) are as in Lemma 2.1. Given w0+E\Ww^{0}\in\mathbb{R}_{+}^{\vec{E}}\backslash W, assume w(t)w(t) is the flow defined by equation (2), and [0,T)[0,T) is the maximum existing interval.

We claim that T=T(w0)<T=T(w^{0})<\infty and w(t)w(t) converge to some w¯\bar{w} as tTt\rightarrow T. Since

ddtC,\frac{d\mathcal{E}}{dt}\leq-C\sqrt{\mathcal{E}},

we have that

d()dt(t)C2,\frac{d(\sqrt{\mathcal{E}})}{dt}(t)\leq-\frac{C}{2},

and

(w(t))(w(0))Ct2,\sqrt{\mathcal{E}(w(t))}\leq\sqrt{\mathcal{E}(w(0))}-\frac{Ct}{2},

which implies

T2(w0)C(w0)<.T\leq\frac{2\sqrt{\mathcal{E}(w^{0})}}{C(w^{0})}<\infty.

Since

0w˙ij(t)wij(t),0\leq\dot{w}_{ij}(t)\leq w_{ij}(t),

we have that

(3) wij(t)wij0etwij0eT.w_{ij}(t)\leq w^{0}_{ij}e^{t}\leq w^{0}_{ij}e^{T}.

Then by the monotone convergence theorem, w(t)w(t) converge to some w¯\bar{w}. By the maximality of TT, w¯\bar{w} has to be in WW. Let Φ:+EW\Phi:\mathbb{R}^{\vec{E}}_{+}\rightarrow W be such that Φ(w0)=w¯\Phi(w^{0})=\bar{w} if w0Ww^{0}\notin W, and Φ(w)=w\Phi(w)=w if wWw\in W.

Now we prove that Φ\Phi is continuous, i.e., for any w+Ew\in\mathbb{R}^{\vec{E}}_{+} and ϵ>0\epsilon>0, there exists δ>0\delta>0 such that |Φ(w)Φ(w)|ϵ|\Phi(w^{\prime})-\Phi(w)|_{\infty}\leq\epsilon for any ww^{\prime} with |ww|<δ|w^{\prime}-w|_{\infty}<\delta. We consider the two cases wWw\in W and wWw\notin W.

2.1.1. wWw\in W

Since C(w)C(w) is continous, there exists C1>0C_{1}>0 and δ1>0\delta_{1}>0 such that C(w)C1C(w^{\prime})\geq C_{1} for any ww^{\prime} with |ww|δ1|w-w^{\prime}|_{\infty}\leq\delta_{1}. Since \mathcal{E} is continuous, there exists δ2(0,δ1)\delta_{2}\in(0,\delta_{1}) such that

(w)[C12log(1+ϵ2|w|+ϵ)]2\mathcal{E}(w^{\prime})\leq\left[\frac{C_{1}}{2}\log(1+\frac{\epsilon}{2|w|_{\infty}+\epsilon})\right]^{2}

if |ww|<δ2|w^{\prime}-w|_{\infty}<\delta_{2}. Then we will show that δ=min{δ2,ϵ/2}\delta=\min\{\delta_{2},\epsilon/2\} is satisfactory. Assume ww^{\prime} satisfies that |ww|<δ|w^{\prime}-w|_{\infty}<\delta. If wWw^{\prime}\in W, then |Φ(w)Φ(w)|=|ww|<δϵ|\Phi(w^{\prime})-\Phi(w)|_{\infty}=|w^{\prime}-w|_{\infty}<\delta\leq\epsilon. If wWw^{\prime}\notin W,

T(w)2(w)C(w)C1log(1+ϵ/(2|w|+ϵ))C1=log(1+ϵ2|w|+ϵ)log(1+ϵ2|w|).T(w^{\prime})\leq\frac{2\sqrt{\mathcal{E}(w^{\prime})}}{C(w^{\prime})}\leq\frac{C_{1}\log(1+\epsilon/(2|w|_{\infty}+\epsilon))}{C_{1}}=\log(1+\frac{\epsilon}{2|w|_{\infty}+\epsilon})\leq\log(1+\frac{\epsilon}{2|w^{\prime}|_{\infty}}).

So by the inequality (3),

|Φ(w)ijΦ(w)ij||Φ(w)ijwij|+|wijwij|<wij(eT(w)1)+ϵ2ϵ.|\Phi(w^{\prime})_{ij}-\Phi(w)_{ij}|\leq|\Phi(w^{\prime})_{ij}-w^{\prime}_{ij}|+|w^{\prime}_{ij}-w_{ij}|<w^{\prime}_{ij}(e^{T(w^{\prime})}-1)+\frac{\epsilon}{2}\leq\epsilon.

2.1.2. wWw\notin W

Assume w¯=Φ(w)W\bar{w}=\Phi(w)\in W. Then by the result of the previous case, there exists δ1>0\delta_{1}>0 such that |Φ(w)Φ(w)|<ϵ|\Phi(w^{\prime})-\Phi(w)|_{\infty}<\epsilon for any ww^{\prime} with |ww¯|<δ1|w^{\prime}-\bar{w}|_{\infty}<\delta_{1}. Assume w(t)w(t) is the flow determined by the equation (2) with the initial value w0=ww^{0}=w. Then there exists some t0t_{0}, such that |w(t0)w¯|<δ1/2|w(t_{0})-\bar{w}|_{\infty}<\delta_{1}/2. By the continuous dependence of the solutions of ODEs on the initial values, there exists δ>0\delta>0 such that if |ww|<δ|w^{\prime}-w|_{\infty}<\delta, then |w(t0)w(t0)|<δ1/2|w^{\prime}(t_{0})-w(t_{0})|<\delta_{1}/2, where w(t)w^{\prime}(t) is the flow determined by equation (2) with the initial value w0=ww^{0}=w^{\prime}. So if |ww|<δ|w-w^{\prime}|_{\infty}<\delta, we have |w(t0)w¯|<δ1|w^{\prime}(t_{0})-\bar{w}|<\delta_{1} and

|Φ(w)Φ(w)|=|Φ(w(t0))Φ(w)|<ϵ.|\Phi(w^{\prime})-\Phi(w)|_{\infty}=|\Phi(w^{\prime}(t_{0}))-\Phi(w)|_{\infty}<\epsilon.

2.2. Construction of the Flow

Denote 𝐱(w)\mathbf{x}(w) as the minimizer of the second minimization problem in the definition of the energy function (w)\mathcal{E}(w). Define the residual 𝐫(w)\mathbf{r}(w) as

𝐫(w)=A(w)𝐱(w)𝐛(w),\mathbf{r}(w)=A(w)\mathbf{x}(w)-\mathbf{b}(w),

where

𝐫(w)=(r1x(w)r1y(w)r2x(w)r2y(w)rnx(w)rny(w))=(𝐫x(w)𝐫y(w))=(𝐫1(w)𝐫2(w)𝐫n(w))n×2.\mathbf{r}(w)=\begin{pmatrix}r_{1}^{x}(w)&r_{1}^{y}(w)\\ r_{2}^{x}(w)&r_{2}^{y}(w)\\ \vdots&\vdots\\ r_{n}^{x}(w)&r_{n}^{y}(w)\end{pmatrix}=\begin{pmatrix}\mathbf{r}^{x}(w)&\mathbf{r}^{y}(w)\\ \end{pmatrix}=\begin{pmatrix}\mathbf{r}_{1}(w)\\ \mathbf{r}_{2}(w)\\ \vdots\\ \mathbf{r}_{n}(w)\end{pmatrix}\in\mathbb{R}^{n\times 2}.

The vector 𝐫i\mathbf{r}_{i} is the residual at the vertex viVv_{i}\in V, and 𝐫x\mathbf{r}^{x}, 𝐫y\mathbf{r}^{y} are the projections of the total residual in the directions of the xx-axis and the yy-axis respectively. Then by the minimality of 𝐱(w)\mathbf{x}(w),

AT(w)𝐫x(w)=0andAT(w)𝐫y(w)=0.A^{T}(w)\mathbf{r}^{x}(w)=0\quad\textrm{and}\quad A^{T}(w)\mathbf{r}^{y}(w)=0.

Equivalently,

𝐫x(w)A(w)(n),and𝐫y(w)A(w)(n).\mathbf{r}^{x}(w)\perp A(w)(\mathbb{R}^{n}),\quad\textrm{and}\quad\mathbf{r}^{y}(w)\perp A(w)(\mathbb{R}^{n}).

Since rank(A(w))=n1rank(A(w))=n-1, 𝐫x𝐫y\mathbf{r}^{x}\|\mathbf{r}^{y}. So rank(𝐫)1rank(\mathbf{r})\leq 1, and 𝐫i𝐫j\mathbf{r}_{i}\|\mathbf{r}_{j} for any 1i,jn1\leq i,j\leq n. Here uvu\|v means that vectors uu and vv are parallel, i.e., linearly dependent.

Lemma 2.2.

Assume 𝐫0\mathbf{r}\neq 0 and the following properties holds for 𝐫1,𝐫2,,𝐫n\mathbf{r}_{1},\mathbf{r}_{2},\cdots,\mathbf{r}_{n}.

  1. (a)

    The vectors have the same direction, namely,

    𝐫i,𝐫j>0,1i,jn.\langle\mathbf{r}_{i},\mathbf{r}_{j}\rangle>0,\forall 1\leq i,j\leq n.
  2. (b)

    If

    C=maxeijEwijwji,C=\max_{\vec{e}_{ij}\in\vec{E}}\frac{w_{ij}}{w_{ji}},

    then

    𝐫i2𝐫j2Cn1,1i,jn.\frac{\|\mathbf{r}_{i}\|_{2}}{\|\mathbf{r}_{j}\|_{2}}\leq C^{n-1},\forall 1\leq i,j\leq n.
Proof.

Without loss of generality, after a rotation we can assume that all the vectors 𝐫i\mathbf{r}_{i} are parallel to xx-axis, namely 𝐫y=0\mathbf{r}^{y}={0}.

To prove part (a), assume that 𝐫i,𝐫j=rixrjx0\langle\mathbf{r}_{i},\mathbf{r}_{j}\rangle=r_{i}^{x}\cdot r_{j}^{x}\leq 0 for some 1i,jn1\leq i,j\leq n. Then one can find a non-zero vector 𝐩=(p1,,pn)T0n\mathbf{p}=(p_{1},\cdots,p_{n})^{T}\in\mathbb{R}^{n}_{\geq 0} so that 𝐩𝐫x\mathbf{p}\perp\mathbf{r}^{x}. Then 𝐩A(w)(n)\mathbf{p}\in A(w)(\mathbb{R}^{n}) and there exists 𝐪=(q1,,qn)Tn\mathbf{q}=(q_{1},\cdots,q_{n})^{T}\in\mathbb{R}^{n} with 𝐩=A(w)𝐪\mathbf{p}=A(w)\mathbf{q}. Then if qi=maxjqjq_{i}=\max_{j}q_{j} for some ii,

0pi=j=1nwij(qjqi)0,0\leq p_{i}=\sum_{j=1}^{n}w_{ij}(q_{j}-q_{i})\leq 0,

and thus qj=qiq_{j}=q_{i} if jN(i)j\in N(i). By this maximum principle and the connectedness of the graph, qj=qiq_{j}=q_{i} for any jVj\in V, and 𝐩=A(w)𝐪=0\mathbf{p}=A(w)\mathbf{q}=0. This contradicts with that 𝐩\mathbf{p} is nonzero.

If part (b) is not true, ordering the set VV based on the values of rixr_{i}^{x} monotonically, one can find a non-empty proper subset V0VV_{0}\subsetneq V such that

miniV0{rix}maxiVV0{rix}>C.\frac{\underset{i\in V_{0}}{\min}\{r^{x}_{i}\}}{\underset{i\in V-V_{0}}{\max}\{r^{x}_{i}\}}>C.

Choose a vector 𝐩n\mathbf{p}\in\mathbb{R}^{n} such that pi=1p_{i}=1 if iV0i\in V_{0}, and pi=0p_{i}=0 otherwise. Then the contradiction follows from

0=\displaystyle 0= 𝐫x,A(w)𝐩=iVrixjN(i)wij(pjpi)=eijEwijrixpjiVrixpijN(i)wij\displaystyle\langle\mathbf{r}^{x},A(w)\mathbf{p}\rangle=\sum_{i\in V}r_{i}^{x}\sum_{j\in N(i)}w_{ij}(p_{j}-p_{i})=\sum_{\vec{e}_{ij}\in\vec{E}}w_{ij}r^{x}_{i}p_{j}-\sum_{i\in V}r^{x}_{i}p_{i}\sum_{j\in N(i)}w_{ij}
=\displaystyle= eijEjV0wijrixiV0rixjN(i)wij=iV0,jVV0jN(i)(rjxwjirixwij)<0.\displaystyle\sum_{\underset{j\in V_{0}}{\vec{e}_{ij}\in\vec{E}}}w_{ij}r^{x}_{i}-\sum_{i\in V_{0}}r^{x}_{i}\sum_{j\in N(i)}w_{ij}=\sum_{\underset{j\in N(i)}{i\in V_{0},j\in V-V_{0}}}(r_{j}^{x}w_{ji}-r_{i}^{x}w_{ij})<0.

Assume 𝐧=𝐧(w)2\mathbf{n}=\mathbf{n}(w)\in\mathbb{R}^{2} is the unit vector that is parallel to 𝐫1\mathbf{r}_{1} and 𝐧,𝐫1>0\langle\mathbf{n},\mathbf{r}_{1}\rangle>0. Define for each directed edge,

uij=uij(w)=𝐧(𝐱j𝐱i+(bijx,bijy)),eijE,u_{ij}=u_{ij}(w)=\mathbf{n}\cdot(\mathbf{x}_{j}-\mathbf{x}_{i}+(b_{ij}^{x},b_{ij}^{y})),\quad\forall\vec{e}_{ij}\in\vec{E},

where 𝐱i=(xi,yi)\mathbf{x}_{i}=(x_{i},y_{i}) is given by the minimizer 𝐱(w)\mathbf{x}(w). Note that

𝐫i2=𝐧𝐫i=jN(i)wijuij.\|\mathbf{r}_{i}\|_{2}=\mathbf{n}\cdot\mathbf{r}_{i}=\sum_{j\in N(i)}w_{ij}u_{ij}.
Lemma 2.3.

There exists a constant β=β(T,ψ)>0\beta=\beta(T,\psi)>0 such that for any w+Ew\in\mathbb{R}^{\vec{E}}_{+}, there exists eijE\vec{e}_{ij}\in\vec{E} such that uij(w)βu_{ij}(w)\leq-\beta.

Proof.

Since uij=ujiu_{ij}=-u_{ji} for any ijEij\in E, it suffices to find eij\vec{e}_{ij} such that |uij(w)|β|u_{ij}(w)|\geq\beta. Assume 𝐧=(n1,n2)2\mathbf{n}=(n_{1},n_{2})\in\mathbb{R}^{2}. Then |n1|1/2|n_{1}|\geq 1/\sqrt{2} or |n2|12|n_{2}|\geq 1\sqrt{2}.

If |n1|1/2|n_{1}|\geq 1/\sqrt{2}, let γ1=𝕋×{0}\gamma_{1}=\mathbb{T}\times\{0\} be a horizontal simple loop in 𝕋2\mathbb{T}^{2}. Then ψ1(γ1)\psi^{-1}(\gamma_{1}) is a simple loop in the carrier of 𝒯\mathcal{T}, and it is not difficult to show that there exists a sequence of vertices v(1),,v(k)=v(0)v(1),...,v(k)=v(0) such that v(i)v(i+1)v(i)\sim v(i+1) for any i=0,,k1i=0,...,k-1, and the union i=0k1ev(i)v(i+1)\cup_{i=0}^{k-1}e_{v(i)v(i+1)} is a piecewise linear loop in |𝒯||\mathcal{T}|, which is homotopy equivalent to ψ1(γ1)\psi^{-1}(\gamma_{1}). By choosing an appropriate orientation, we have that

i=0k1(𝐱v(i)+1𝐱v(i)+(bv(i+1)v(i)x,bv(i+1)v(i)y))=(1,0).\sum_{i=0}^{k-1}(\mathbf{x}_{v(i)+1}-\mathbf{x}_{v(i)}+(b^{x}_{v(i+1)v(i)},b^{y}_{v(i+1)v(i)}))=(1,0).

So

i=0k1uv(i)v(i+1)=𝐧(1,0)=n1,\sum_{i=0}^{k-1}u_{v(i)v(i+1)}=\mathbf{n}\cdot(1,0)=n_{1},

and there exists some ii such that |uv(i)v(i+1)||n1|/k1/(2k)|u_{v(i)v(i+1)}|\geq|n_{1}|/k\geq 1/(\sqrt{2}k). Notice that here kk is a constant depends only on 𝒯\mathcal{T} and ψ\psi.

Similarly, if |n2|1/2|n_{2}|\geq 1/\sqrt{2}, there exists some eijE\vec{e}_{ij}\in\vec{E} such that |uij|1/(2k)|u_{ij}|\geq 1/(\sqrt{2}k^{\prime}) for some constant k=k(𝒯,ψ)k^{\prime}=k^{\prime}(\mathcal{T},\psi). ∎

We define the smooth flow Θ\Theta on the domain +E\W\mathbb{R}_{+}^{\vec{E}}\backslash W on each edge as follows

(4) {w˙ij=wijg(1α(wij+wji)uij)h(wijwji),wij(0)=wij0,\Big{\{}\begin{array}[]{ll}\dot{w}_{ij}=w_{ij}\cdot g\big{(}\frac{1}{\alpha}(w_{ij}+w_{ji})u_{ij}\big{)}\cdot h(w_{ij}-w_{ji}),\\ w_{ij}(0)=w_{ij}^{0},\end{array}

where gg and hh are smooth non-increasing functions such that

  1. (a)

    The function g1g\equiv 1 on (,1)(-\infty,-1) and g0g\equiv 0 on [0,+)[0,+\infty), and

  2. (b)

    The function h1h\equiv 1 on (,1)(-\infty,1), and h0h\equiv 0 on [2,+)[2,+\infty), and

  3. (c)
    α=α(w)=β(2|E|+eijEwij1)1.\alpha=\alpha(w)=\beta\cdot\bigg{(}2|E|+\sum_{\vec{e}_{ij}\in\vec{E}}w_{ij}^{-1}\bigg{)}^{-1}.

Roughly speaking, the function gg tends be positive if uij<0u_{ij}<0, meaning that wijuijw_{ij}u_{ij} will decrease so as to reduce the residual 𝐫i2\|\mathbf{r}_{i}\|_{2}. The function hh controls the difference between wijw_{ij} and wjiw_{ji}. α(w)\alpha(w) is smooth and very small. Specifically we have that

(5) α(w)β2|E| and α(w)βL(w),\alpha(w)\leq\frac{\beta}{2|E|}\quad\text{ and }\quad\alpha(w)\leq\beta L(w),

where L(w)=mineijEwijL(w)=\min_{\vec{e}_{ij}\in\vec{E}}w_{ij} is a continuous function on w+Ew\in\mathbb{R}^{\vec{E}}_{+}. Denote

M(w)=max{2,maxeijE|wijwji|},M(w)=\max\{2,\max_{\vec{e}_{ij}\in\vec{E}}|w_{ij}-w_{ji}|\},

which is another continous function on +E\mathbb{R}^{\vec{E}}_{+}.

Lemma 2.4.

Assume the flow w(t)w(t) satisfies the equation (4). Then we have the following.

  1. (a)

    0w˙ijwij0\leq\dot{w}_{ij}\leq w_{ij}.

  2. (b)

    uij0u_{ij}\geq 0 implies w˙ij=0\dot{w}_{ij}=0. Then w˙ijuij0\dot{w}_{ij}u_{ij}\leq 0 for all directed edges.

  3. (c)

    wijwji2w_{ij}-w_{ji}\geq 2 implies w˙ij=0\dot{w}_{ij}=0.

  4. (d)

    L(w(t))L(w(t)) is non-decreasing and M(w(t))M(w(t)) is non-increasing.

  5. (e)

    For any edge ijij,

    wij(t)wji(t)1+ML.\frac{w_{ij}(t)}{w_{ji}(t)}\leq 1+\frac{M}{L}.
  6. (f)

    The residual vectors 𝐫i(t)\mathbf{r}_{i}(t) satisfies

    maxiV𝐫i(t)2miniV𝐫i(t)2(1+ML)n1,1i,jn.\frac{\max_{i\in V}\|\mathbf{r}_{i}(t)\|_{2}}{\min_{i\in V}\|\mathbf{r}_{i}(t)\|_{2}}\leq(1+\frac{M}{L})^{n-1},\quad\forall 1\leq i,j\leq n.
  7. (g)

    The residual vectors 𝐫i(t)\mathbf{r}_{i}(t) satisfies

    (6) (w(t))n(1+M/L)n1𝐫i(t)2,1in.\frac{\sqrt{\mathcal{E}(w(t))}}{\sqrt{n}(1+M/L)^{n-1}}\leq\mathbf{\|}\mathbf{r}_{i}(t)\|_{2},\quad\forall 1\leq i\leq n.
Proof.

Part (a) - (d) are straightforward from equation (4) and the defining properties of smooth functions gg and hh. Part (e) follows from

wij(t)wji(t)=1+wij(t)wji(t)wji(t)1+ML.\frac{w_{ij}(t)}{w_{ji}(t)}=1+\frac{w_{ij}(t)-w_{ji}(t)}{w_{ji}(t)}\leq 1+\frac{M}{L}.

Part (f) follows from Part (e) and Lemma 2.2. For Part (g), by definition

(t)=j=1n𝐫j(t)22.\mathcal{E}(t)=\sum_{j=1}^{n}\|\mathbf{r}_{j}(t)\|_{2}^{2}.

Part (f) implies that

(t)n(1+ML)2n2𝐫i(t)22and(t)n(1+M/L)n1𝐫i(t)2,1in.\mathcal{E}(t)\leq n(1+\frac{M}{L})^{2n-2}\|\mathbf{r}_{i}(t)\|_{2}^{2}\quad\text{and}\quad\frac{\sqrt{\mathcal{E}(t)}}{\sqrt{n}(1+M/L)^{n-1}}\leq\|\mathbf{r}_{i}(t)\|_{2},\quad\forall 1\leq i\leq n.

2.3. Proof of Lemma 2.1

Proof.

Let

C(w)=βL/M2n(1+M/L)n1C(w)=\frac{\beta L/M}{2\sqrt{n}(1+M/L)^{n-1}}

where L=L(w)L=L(w) and M=M(w)M=M(w) are continous functions on +E\mathbb{R}^{\vec{E}}_{+}, on which C(w)C(w) is also continuous. We claim that such function C(w)C(w) and the flow Θ\Theta defined as equation (4) are satisfactory. Assume w0+E\Ww^{0}\in\mathbb{R}^{\vec{E}}_{+}\backslash W and w(t)w(t) is a flow defined by equation (4). By part (a) of Lemma 2.4 we only need to prove the Part (b) of Lemma 2.1. By part (d) of Lemma 2.4, it is easy to see that C(w(t))C(w(t)) is non-decreasing on tt. So we only need to prove that

d(w(t))dtC(w(t))(w(t)).\frac{d\mathcal{E}(w(t))}{dt}\leq-C(w(t))\sqrt{\mathcal{E}(w(t))}.

Given wR+Ew\in R^{\vec{E}}_{+} and 𝐱n×2\mathbf{x}\in\mathbb{R}^{n\times 2}, denote

~(w,𝐱)=A(w)𝐱𝐛(w)2,\tilde{\mathcal{E}}(w,\mathbf{x})=\|A(w)\mathbf{x}-\mathbf{b}(w)\|^{2},

and then

d(w())dt(t)=\displaystyle\frac{d{\mathcal{E}}(w(\cdot))}{dt}(t)= limϵ0(w(t+ϵ))(w(t))ϵ\displaystyle\lim_{\epsilon\to 0}\frac{{\mathcal{E}}(w(t+\epsilon))-{\mathcal{E}}(w(t))}{\epsilon}
\displaystyle\leq limϵ0~(w(t+ϵ),𝐱(w(t)))~(w(t),𝐱(w(t)))ϵ\displaystyle\lim_{\epsilon\to 0}\frac{\tilde{\mathcal{E}}(w(t+\epsilon),\mathbf{x}(w(t)))-\tilde{\mathcal{E}}(w(t),\mathbf{x}(w(t)))}{\epsilon}
=\displaystyle= ~w(w(t),𝐱(w(t)))w˙.\displaystyle\frac{\partial\tilde{\mathcal{E}}}{\partial w}(w(t),\mathbf{x}(w(t)))\cdot\dot{w}.

So it suffices to show

~w(w(t),𝐱(w(t)))w˙C(w(t)).\frac{\partial\tilde{\mathcal{E}}}{\partial w}(w(t),\mathbf{x}(w(t)))\cdot\dot{w}\leq-C\sqrt{\mathcal{E}(w(t))}.

Notice that

~wij(w(t),𝐱(w(t)))\displaystyle\frac{\partial\tilde{\mathcal{E}}}{\partial w_{ij}}(w(t),\mathbf{x}(w(t)))
=\displaystyle= (wiji=1njN(i)wij(𝐱j𝐱i+(bijx,bijy))22)|(w,𝐱(w))\displaystyle\left(\frac{\partial}{\partial w_{ij}}\sum_{i=1}^{n}\big{\|}\sum_{j\in N(i)}w_{ij}(\mathbf{x}_{j}-\mathbf{x}_{i}+(b_{ij}^{x},b_{ij}^{y}))\big{\|}_{2}^{2}\right)\bigg{|}_{(w,\mathbf{x}(w))}
=\displaystyle= 2𝐫i(𝐱j𝐱i+(bijx,bijy))\displaystyle 2\mathbf{r}_{i}\cdot(\mathbf{x}_{j}-\mathbf{x}_{i}+(b_{ij}^{x},b_{ij}^{y}))
=\displaystyle= 2𝐫i2𝐧(𝐱j𝐱i+(bijx,bijy))\displaystyle 2\|\mathbf{r}_{i}\|_{2}\mathbf{n}\cdot(\mathbf{x}_{j}-\mathbf{x}_{i}+(b_{ij}^{x},b_{ij}^{y}))
=\displaystyle= 2𝐫i2uij\displaystyle 2\|\mathbf{r}_{i}\|_{2}\cdot u_{ij}

and then,

(7) ~w(w(t),𝐱(w(t)))w˙=2eijE𝐫i2uijw˙ij2(w(t))n(1+M/L)n1eijEuijw˙ij.\frac{\partial\tilde{\mathcal{E}}}{\partial w}(w(t),\mathbf{x}(w(t)))\cdot\dot{w}=2\sum_{\vec{e}_{ij}\in\vec{E}}\|\mathbf{r}_{i}\|_{2}u_{ij}\dot{w}_{ij}\leq\frac{2\sqrt{\mathcal{E}(w(t))}}{\sqrt{n}(1+M/L)^{n-1}}\sum_{\vec{e}_{ij}\in\vec{E}}u_{ij}\dot{w}_{ij}.

Here we use the fact that w˙ijuij0\dot{w}_{ij}u_{ij}\leq 0 for all directed edges (part (b) of Lemma 2.4), and the inequality (6). It remains to show that

eijEuijw˙ijβL2M.\sum_{\vec{e}_{ij}\in\vec{E}}u_{ij}\dot{w}_{ij}\leq-\frac{\beta L}{2M}.

By Lemma 2.3 there exists a directed edge eij\vec{e}_{i^{\prime}j^{\prime}} with uijβu_{i^{\prime}j^{\prime}}\leq-\beta. Then we will consider the following two cases:

2.3.1. Case 1: wijwji1w_{i^{\prime}j^{\prime}}-w_{j^{\prime}i^{\prime}}\leq 1

By the definition of the function hh,

h(wijwji)=1.h(w_{i^{\prime}j^{\prime}}-w_{j^{\prime}i^{\prime}})=1.

We also have

g(1α(wij+wji)uij)=1,g\left(\frac{1}{\alpha}(w_{i^{\prime}j^{\prime}}+w_{j^{\prime}i^{\prime}})u_{i^{\prime}j^{\prime}}\right)=1,

since

(wij+wji)uij2βLα.(w_{i^{\prime}j^{\prime}}+w_{j^{\prime}i^{\prime}})u_{i^{\prime}j^{\prime}}\leq-2\beta L\leq-\alpha.

By equation (4), w˙ij=wij\dot{w}_{i^{\prime}j^{\prime}}=w_{i^{\prime}j^{\prime}}. Notice that w˙ijuij0\dot{w}_{ij}u_{ij}\leq 0 for any eijE\vec{e}_{ij}\in\vec{E} by Part (b) of Lemma 2.4, so

eijEuijw˙ijuijw˙ij=uijwijβLβL2M.\sum_{\vec{e}_{ij}\in\vec{E}}u_{ij}\dot{w}_{ij}\leq u_{i^{\prime}j^{\prime}}\dot{w}_{i^{\prime}j^{\prime}}=u_{i^{\prime}j^{\prime}}w_{i^{\prime}j^{\prime}}\leq-\beta L\leq-\frac{\beta L}{2M}.

2.3.2. Case 2: wijwji1w_{i^{\prime}j^{\prime}}-w_{j^{\prime}i^{\prime}}\geq 1

Denote

E0={eijE|uij<0,(wijwji)uijα}.\vec{E}_{0}=\Big{\{}\vec{e}_{ij}\in\vec{E}\big{|}u_{ij}<0,(w_{ij}-w_{ji})u_{ij}\geq\alpha\Big{\}}.

If eijE0\vec{e}_{ij}\in\vec{E}_{0}, then obviously wijwji<0w_{ij}-w_{ji}<0 and

h(wijwji)=1.h(w_{ij}-w_{ji})=1.

Also,

g(1α(wij+wji)uij)=1g\left(\frac{1}{\alpha}(w_{ij}+w_{ji})u_{ij}\right)=1

since

(wij+wji)uij(wjiwij)uijα.(w_{ij}+w_{ji})u_{ij}\leq(w_{ji}-w_{ij})u_{ij}\leq-\alpha.

By equation (4), w˙ij=wij\dot{w}_{ij}=w_{ij} and

(8) eijEw˙ijuijeijE0w˙ijuij=eijE0wijuijLMeijE0(wijwji)uij.\displaystyle\sum_{\vec{e}_{ij}\in\vec{E}}\dot{w}_{ij}u_{ij}\leq\sum_{\vec{e}_{ij}\in\vec{E}_{0}}\dot{w}_{ij}u_{ij}=\sum_{\vec{e}_{ij}\in\vec{E}_{0}}{w}_{ij}u_{ij}\underset{*}{\leq}-\frac{L}{M}\sum_{{e}_{ij}\in E_{0}}({w}_{ij}-{w}_{ji})u_{ij}.

The last step (inequality ()(*)) uses the fact that wijL(wijwji)/Mw_{ij}\geq-L(w_{ij}-w_{ji})/M, which is equivalent to that wji/wij1+M/Lw_{ji}/w_{ij}\leq 1+M/L.

By the fact that uijβu_{i^{\prime}j^{\prime}}\leq-\beta, and the assumption wijwji1w_{i^{\prime}j^{\prime}}-w_{j^{\prime}i^{\prime}}\geq 1,

(wijwji)uijβ<0<α,(w_{i^{\prime}j^{\prime}}-w_{j^{\prime}i^{\prime}})u_{i^{\prime}j^{\prime}}\leq-\beta<0<\alpha,

and thus eijE0\vec{e}_{i^{\prime}j^{\prime}}\notin\vec{E}_{0}. Notice that

eijEwijuij=i=1njN(i)wijuij=i=1n𝐫i20,\sum_{\vec{e}_{ij}\in\vec{E}}w_{ij}u_{ij}=\sum_{i=1}^{n}\sum_{j\in N(i)}w_{ij}u_{ij}=\sum_{i=1}^{n}\|\mathbf{r}_{i}\|_{2}\geq 0,

and

eijEwijuij=eijE:uij<0(wijwji)uij\displaystyle\sum_{\vec{e}_{ij}\in\vec{E}}w_{ij}u_{ij}=\sum_{\vec{e}_{ij}\in\vec{E}:u_{ij}<0}(w_{ij}-w_{ji})u_{ij}
=\displaystyle= eijE0(wijwji)uij+eijEE0{eij}:uij<0(wijwji)uij+(wijwji)uij\displaystyle\sum_{\vec{e}_{ij}\in{\vec{E}_{0}}}({w}_{ij}-{w}_{ji})u_{ij}+\sum_{{e}_{ij}\in{\vec{E}}-\vec{E}_{0}-\{{e}_{i^{\prime}j^{\prime}}\}:u_{ij}<0}({w}_{ij}-{w}_{ji})u_{ij}+(w_{i^{\prime}j^{\prime}}-w_{j^{\prime}i^{\prime}})u_{i^{\prime}j^{\prime}}
\displaystyle\leq eijE0(wijwji)uij+|E|αβ.\displaystyle\sum_{\vec{e}_{ij}\in{\vec{E}_{0}}}({w}_{ij}-{w}_{ji})u_{ij}+|E|\alpha-\beta.

Then

eijE0(wijwji)uijβ|E|αβ/2,\sum_{\vec{e}_{ij}\in{\vec{E}_{0}}}({w}_{ij}-{w}_{ji})u_{ij}\geq\beta-|E|\alpha\geq\beta/2,

and

eijEuijw˙ijβL2M.\sum_{\vec{e}_{ij}\in\vec{E}}u_{ij}\dot{w}_{ij}\leq-\frac{\beta L}{2M}.

3. Proof of Theorem 1.3

We will first introduce the concept of discrete one forms, and the Index Theorem proposed by Gortler-Gotsman-Thurston [11].

3.1. Discrete One Forms and the Index Theorem

A discrete one form is a real-valued function η\eta on the set of directed edges such that it is anti-symmetric on each undirected edge. Specifically, let ηij=η(eij)\eta_{ij}=\eta(\vec{e}_{ij}) be the value of η\eta on the directed edge from viv_{i} to vjv_{j}, then we have ηij=ηji.\eta_{ij}=-\eta_{ji}.

For a discrete one form, an edge is degenerate (resp. non-vanishing) if the one form is zero (resp. non-zero) on it. A vertex is degenerate (resp. non-vanishing) if all of edges connected to it are degenerate (resp. non-vanishing). A face is degenerate (resp. non-vanishing) if all of its three edges are degenerate (resp. non-vanishing). A one-form is degenerate (resp. non-vanishing) if all the edges are degenerate (resp. non-vanishing). Each edge is either degenerate or non-vanishing. However, vertices or faces can be degenerate, non-degenerate but vanishing on some edges, or non-vanishing.

Refer to caption
Figure 2. Typical vertex with (1) positive, (2) zero, and (3) negative index.

Assumet η\eta is a discrete one form. Deonote sc(η,v)sc(\eta,v) as the number of sign changes of the non-zero values of η\eta on the directed edges starting from vv, counted in counter-clockwise order. For a vertex vVv\in V, define the index of vv as Ind(η,v)=(2sc(η,v))/2Ind(\eta,v)=(2-sc(\eta,v))/2. Similarly, for a non-degenerate face tFt\in F, the index of tt is Ind(η,t)=(2sc(η,t))/2Ind(\eta,t)=(2-sc(\eta,t))/2, where sc(η,t)sc(\eta,t) is the number of sign changes of the non-zero values of η\eta on the three edges of tt, counted in counter-clockwise order.

The following theorem is a special case of the Index Theorem from [11], which is a discrete version of the Poincare-Hopf theorem for discrete one forms.

Theorem 3.1.

Let η\eta be a non-vanishing discrete one form on a triangulation of a torus. Then

viVInd(η,vi)+tijkFInd(η,tijk)=0.\sum_{v_{i}\in V}Ind(\eta,v_{i})+\sum_{t_{ijk}\in F}Ind(\eta,t_{ijk})=0.

Assume φ\varphi satisfies the assumption in Theorem 1.3, then for any unit vector 𝐧2\mathbf{n}\in\mathbb{R}^{2} we can naturally construct a discrete one form η\eta, by letting ηij=φ˙ij𝐧\eta_{ij}=\dot{\varphi}_{ij}\cdot\mathbf{n}. If φX\varphi\in X, a generic unit vector determines a non-vanishing discrete one form η\eta. Further, if φX\varphi\in X and such constructed η\eta is non-vanishing, it is not difficult to show that all the indices of the vertices and faces are zero. Figure 2 illustrates how the neighborhood of vv looks like if it has positive, or zero, or negative index, for the case 𝐧=(1,0)\mathbf{n}=(1,0).

Based on this construction, we have

Lemma 3.2.

Given a triangulation (𝒯,ψ)(\mathcal{T},\psi) of 𝕋2\mathbb{T}^{2}, denote tijkFt_{ijk}\in F as the triangle with three vertices viv_{i}, vjv_{j}, and vkv_{k}. There exists a non-vanishing discrete one form η\eta such that ηij>0\eta_{ij}>0 and ηjk>0\eta_{jk}>0. Moreover, all the indices of the vertices and faces of η\eta are zero.

Proof.

By the result of Colin de Verdière [5] and Hass-Scott [12], the space X(𝒯,ψ)X(\mathcal{T},\psi) is not empty for any (𝒯,ψ)(\mathcal{T},\psi). Let φ\varphi be a geodesic triangulation in XX. Then it is not difficult to find a unit vector 𝐧\mathbf{n} such that φ˙ij𝐧>0\dot{\varphi}_{ij}\cdot\mathbf{n}>0 and φ˙jk𝐧>0\dot{\varphi}_{jk}\cdot\mathbf{n}>0. Define the discrete one form η\eta as ηij=φ˙ij𝐧\eta_{ij}=\dot{\varphi}_{ij}\cdot\mathbf{n}. We can perturb the unit vector 𝐧\mathbf{n} a little bit to make η\eta non-vanishing, and then such η\eta is satisfactory. ∎

3.2. The Proof of Theorem 1.3

Assume φ:𝒯(1)𝕋2\varphi:\mathcal{T}^{(1)}\to\mathbb{T}^{2} satisfies the assumption of Theorem 1.3, then there exists a unique extension φ¯:|𝒯|𝕋2\bar{\varphi}:|\mathcal{T}|\to\mathbb{T}^{2} such that the restriction of φ¯\bar{\varphi} to every face is linear. Such φ¯\bar{\varphi} is homotopic to ψ\psi, and φ\varphi is a geodesic triangulation in XX if and only if φ¯\bar{\varphi} is a homeomorphism.

For any triangle tijkXt_{ijk}\in X, we say that φ¯(tijk)\bar{\varphi}(t_{ijk}) is degenerate if φ¯(tijk)\bar{\varphi}(t_{ijk}) is contained in some geodesic λ\lambda. If φ¯(tijk)\bar{\varphi}(t_{ijk}) is not degenerate, we can naturally define its inner angle θjki\theta^{i}_{jk} at φ(vi)\varphi(v_{i}). We claim that

  1. (a)

    φ¯(tijk)\bar{\varphi}(t_{ijk}) is not degenerate for any tijkFt_{ijk}\in F, and

  2. (b)

    φ¯\bar{\varphi} is locally a homeomorphism.

Then φ¯\bar{\varphi} is a proper local homeomorphism, and thus is a covering map. Since φ¯\bar{\varphi} is homotopic to the homeomorphism ψ\psi, φ¯\bar{\varphi} is indeed a degree-one covering map, i.e., a homeomorphism.

3.2.1. Proof of Claim (a)

Assume there is some triangle tFt\in F such that φ¯(t)\bar{\varphi}(t) is degenerate and hence contained in a geodesic λ\lambda. Here λ\lambda is assumed to be a closed geodesic, or a densely immersed complete geodesic. Let 𝒞\mathcal{C} be the union of all triangles tt such that φ¯(t)λ\bar{\varphi}(t)\subset\lambda. Then 𝒞\mathcal{C} is not the whole complex 𝒯\mathcal{T}, otherwise φ¯\bar{\varphi} is not homotopic to the homeomorphism ψ\psi. So, we can find a vertex v0𝒞v_{0}\in\partial\mathcal{C}. Denote star(v0)star(v_{0}) as the star-neighborhood of v0v_{0} in 𝒯\mathcal{T}. Then φ¯(star(v0))\bar{\varphi}(star(v_{0})) is not in λ\lambda, but φ¯(t0)λ\bar{\varphi}(t_{0})\subset\lambda for some triangle t0t_{0} in star(v0)star(v_{0}).

Let 𝐧\mathbf{n} be a unit vector that is orthogonal to the geodesic λ\lambda, and define η\eta as ηij=φ˙ij𝐧\eta_{ij}=\dot{\varphi}_{ij}\cdot\mathbf{n}. Then the vertex v0v_{0} is non-degenerate with respect to η\eta, but the face t0t_{0} is degenerate. Let ξ\xi be a discrete one form in Lemma 3.2 with the triangle t0t_{0} and vj=v0v_{j}=v_{0}. Scale ξ\xi to make it very small so that η+ξ\eta+\xi has the same signs with η\eta on the non-degenerate edges of η\eta.

Notice that sc(η,v0)0sc(\eta,v_{0})\neq 0, or equivalently sc(η,v0)2sc(\eta,v_{0})\geq 2, otherwise all the edges connecting v0v_{0} lie on a half-space, which contradicts to the assumption that φ\varphi is balanced. Since t0t_{0} is degenerate in η\eta, taking the opposite ξ-\xi instead of ξ\xi if necessary, we can assume that sc(η,v0)<sc(η+ξ,v0)sc(\eta,v_{0})<sc(\eta+\xi,v_{0}), and thus Ind(η+ξ,v0)<0Ind(\eta+\xi,v_{0})<0.

Noticing that η+ξ\eta+\xi is non-vanishing, we will derive a contradiction with the Index Theorem 3.1, by showing that the index of η+ξ\eta+\xi is nonpositive for any vertex and face.

If a face tt is degenerate in η\eta, then Ind(η+ξ,t)=Ind(ξ,t)=0Ind(\eta+\xi,t)=Ind(\xi,t)=0. If a face tt is non-degenerate in η\eta, then Ind(η+ξ,t)=Ind(η)=0Ind(\eta+\xi,t)=Ind(\eta)=0. In fact, the index of any non-degenerate face is zero.

If a vertex vv is degenerate in η\eta, then Ind(η+ξ,v)=Ind(ξ,v)=0Ind(\eta+\xi,v)=Ind(\xi,v)=0. If a vertex vv is non-vanishing, then Ind(η+ξ,v)=Ind(η,v)Ind(\eta+\xi,v)=Ind(\eta,v). Since φ\varphi is balanced, Ind(η,v)0Ind(\eta,v)\leq 0.

If a vertex vv is non-degenerate but vanishing at some edges in η\eta, then

Ind(η+ξ,v)Ind(η,v)0,Ind(\eta+\xi,v)\leq Ind(\eta,v)\leq 0,

since adding ξ\xi can only introduce more sign changes.

3.2.2. Proof of Claim (b)

Since φ\varphi is ww-balanced, it is not difficult to show that for any vertex ii,

jk:tijkFθjki2π,\sum_{jk:t_{ijk}\in F}\theta_{jk}^{i}\geq 2\pi,

and the equality holds if and only if all the edges around viv_{i} do not “fold” under the map φ¯\bar{\varphi}. By the Gauss-Bonnet theorem,

i=1n(2πjk:tijkFθjki)=0.\sum_{i=1}^{n}(2\pi-\sum_{jk:t_{ijk}\in F}\theta^{i}_{jk})=0.

So

jk:tijkFθjki=2π,\sum_{jk:t_{ijk}\in F}\theta_{jk}^{i}=2\pi,

for any vertex viv_{i}, and all the edges in EE do not fold. Thus, φ¯\bar{\varphi} is a local homeomorphism.

References

  • [1] 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.
  • [2] Stewart S Cairns, Isotopic deformations of geodesic complexes on the 2-sphere and on the plane, Annals of Mathematics (1944), 207–217.
  • [3] Erin Wolf Chambers, Jeff Erickson, Patrick Lin, and Salman Parsa, How to morph graphs on the torus, arXiv preprint arXiv:2007.07927 (2020).
  • [4] 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.
  • [5] Y Colin de  Verdière , Comment rendre géodésique une triangulation d’une surface, L’Enseignement Mathématique 37 (1991), 201–212.
  • [6] Olaf Delgado-Friedrichs, Equilibrium placement of periodic graphs and convexity of plane tilings, Discrete & Computational Geometry 33 (2005), no. 1, 67–81.
  • [7] 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.
  • [8] Jeff Erickson and Patrick Lin, Planar and toroidal morphs made easier, arXiv:2106.14086 (2021).
  • [9] Michael Floater, One-to-one piecewise linear mappings over triangulations, Mathematics of Computation 72 (2003), no. 242, 685–696.
  • [10] Michael S Floater, Mean value coordinates, Computer Aided Geometric Design 20 (2003), no. 1, 19–27.
  • [11] Steven Gortler, Craig Gotsman, and Dylan Thurston, Discrete one-forms on meshes and applications to 3d mesh parameterization, Computer Aided Geometric Design (2006).
  • [12] Joel Hass and Peter Scott, Simplicial energy and simplicial harmonic maps, arXiv preprint arXiv:1206.2574 (2012).
  • [13] Yanwen Luo, Tianqi Wu, and Zhu Xiaoping, The deformation spaces of geodesic triangulations and generalized Tutte’s embedding theorem, arXiv:2105.00612 (2021).
  • [14] William Thomas Tutte, How to draw a graph, Proceedings of the London Mathematical Society 3 (1963), no. 1, 743–767.