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

Angle structure on general hyperbolic 33-manifolds

Huabin Ge, Longsong Jia, Faze Zhang
Abstract

Let MM be a non-compact hyperbolic 33-manifold with finite volume and totally geodesic boundary components. By subdividing mixed ideal polyhedral decompositions of MM, under some certain topological conditions, we prove that MM has an ideal triangulation which admits an angle structure.

1 Introduction

Thurston’s geometrization conjecture (proved by Perelman [33]-[35], or see [6],[20],[31] for more details) tells us that any 33-manifold can be decomposed via J-S-J splitting into 88 distinct geometric pieces, the majority of which are hyperbolic. A 33-manifold with a hyperbolic geometric structure is termed as a hyperbolic 33-manifold. Thus, determining the existence of a hyperbolic structure on a given 33-manifold is a significant research topic. Thurston [41] introduced the concept of ideal triangulations on 33-manifolds and proved that a solution of the hyperbolic gluing equations on the ideal triangulation can yield a hyperbolic structure on the 33-manifold. However, solving the hyperbolic gluing equations on general ideal triangulations is extremely challenging. In the 1990s, Casson [21] and Rivin [38] discovered a powerful technique for solving Thurston’s gluing equations. By introducing the concept of angle structures on ideal triangulations, they proved that if a maximal volume angle structure exists, it provides solutions to Thurston’s gluing equations, thereby yielding a hyperbolic structure on the 33-manifold. The converse is also true, and one can find a self-contained exposition of Casson-Rivin’s program in [9].

An angle structure on an ideal triangulation refers to assigning real numbers (called dihedral angle) in the interval (0,π)(0,\pi) at each edge of each ideal tetrahedron such that the sum of the dihedral angles at each ideal vertex is π\pi and the sum of dihedral angles around each edge is 2π2\pi. Kang-Rubinstein [19] extended the real numbers assigned at each edge into [0,π][0,\pi], and called this structure as semi-angle structure. If the dihedral angles are only taken values 0 or π\pi, then the structure is termed as taut structure by Lackanby [23]. Luo-Tillmann [26] further extended the assignment of dihedral angles to the whole real numbers \mathbb{R}, which is defined as generalized angle structure.

Conversely, to study the geometry and topology of hyperbolic 33-manifolds, a natural approach is to investigate their geometric ideal tetrahedra decompositions. The existence of geometric ideal triangulations are very important for studying hyperbolic 3-manifolds. On the one hand, Thurston’s proof of the famous Hyperbolic Dehn Filling Theorem takes it for granted that each cusped hyperbolic 3-manifold has a geometric ideal triangulation, see [5], [29], [36] for example. On the other hand, geometric ideal triangulations ensures that all three-dimensional hyperbolic manifolds can be constructed from ideal tetrahedra with the help of Thurston’s gluing equations. The first breakthrough on geometric ideal decomposition is due to Epstein-Penner [7]. They introduced canonical geometric ideal polyhedral decompositions on cusped hyperbolic manifolds with finite volume. Similarly, Kojima [28] obtained truncated hyperideal polyhedral decompositions on compact hyperbolic 33-manifolds with totally geodesic boundary. Recently, we [12] obtained mixed ideal polyhedral decompositions on non-compact hyperbolic 33-manifolds with finite volume and totally geodesic boundary. In all known examples, a cusped hyperbolic manifold MM admits at least one geometric ideal triangulation, but whether this holds true for arbitrary cusped hyperbolic 33-manifold MM is still an open question. As was pointed by Guéritaud and Schleimer in [15], “it is a difficult problem in general. General results are known only when MM is restricted to belong to certain classes of manifolds: punctured-torus bundles, two-bridge link complements, certain arborescent link complements and related objects, or covers of any of these spaces” (see [1]-[3], [13]-[14], [16], [18], [22], [32] for instance). For compact three-dimensional manifolds with boundary, Feng-Ge-Hua [8] established a deep connection between the combinatorial Ricci flow and geometric ideal triangulations, and proved that any topological ideal triangulation with degree greater than or equal to 10 are geometric. Although it was difficult to obtain geometric triangulations directly, Luo-Schleimer-Tillmann [25] proved the virtual existence of geometric triangulations. Furthermore, Futer-Hamilton-Hoffman [10] proved that there are infinitely many virtual geometric triangulations.

To a certain extent, the existence of angle structure is weaker than that of geometric triangulations. Hence, a quite natural question arises: for any hyperbolic 33-manifold MM, are there topological triangulations which admits an angle structure? Luo-Tillmann [26] gave a sufficient and necessary condition for the existence of angle structure on cusped hyperbolic 33-manifolds using the theory of normal surface. By using Epstein-Penner’s ideal geometric polyhedral decompositions and the results of Luo-Tillmann [26], Hodgson-Rubinstein-Segerman [17] proved that there exists an ideal triangulation which admits an angle structure on cusped hyperbolic 33-manifolds with a certain topological condition. By using Kojima’s hyperideal (truncated) polyhedral decompositions, Qiu-Zhang-Yang [37] obtained a hyperideal (truncated) triangulation which admits an angle structure on compact hyperbolic 33-manifolds with totally geodesic boundary. In this paper we consider a hyperbolic 33-manifold MM with both cusps and totally geodesic boundaries. By Moise’s work [30], there is always an ideal triangulation 𝒯\mathcal{T} on MM. We want to know whether 𝒯\mathcal{T} admits an angle structure. Based on our previous work on mixed ideal polyhedral decompositions [12] and Luo-Tillmann’s arguments in [26], we get a natural ideal triangulation 𝒯\mathcal{T} (see Corollary 3.2) on MM directly without using Moise’s work [30]. Moreover, under some topological conditions, 𝒯\mathcal{T} admits an angle structure. To be precise, we have:

Theorem 1.1.

Let MM be a non-compact finite-volume hyperbolic 33-manifold with totally geodesic boundary. Denote M¯\overline{M} by the compact 33-manifold with boundary, with each torus (or Klein bottle) boundary component corresponds to a cusp of MM. In other words, after subtracting the torus (or Klein bottle) boundary components, M¯\overline{M} is homeomorphic to MM. If H1(M¯;2)H1(M¯,M¯;2)H_{1}(\overline{M};\mathbb{Z}_{2})\rightarrow H_{1}(\overline{M},\partial\overline{M};\mathbb{Z}_{2}) is the zero map, then there is a topological ideal triangulation on MM such that it admits an angle structure.

To prove Theorem 1.1, the polyhedral decomposition theory [12], which parallels [7] and [28], will play a crucial role. In addition, we follow the spirits in Luo-Tillmann [26], Kang-Rubinstein [19] and Hodgson-Rubinstein-Segerman [17]. We outline the proof in three steps. Step 1, using our geometric polyhedral decompositions [12], we first decompose MM into hyperbolic ideal or truncated polyhedra. Then we triangulate each polyhedron. The triangulation of the common faces of adjacent polyhedra may not necessarily match. On the mismatched faces, by inserting flat ideal tetrahedra, we are able to achieve an ideal triangulation 𝒯\mathcal{T}. Step 2, following Kang-Rubinstein [19] and Luo-Tillmann [26], there is a strong relationship between the angle structure of (M,𝒯)(M,\mathcal{T}) and the theory of normal surfaces in 𝒯\mathcal{T}. On each individual σi\sigma_{i} of 𝒯\mathcal{T}, there exist four types of normal triangles and three types of normal quadrilaterals, each of which is characterized by coordinates. These coordinates satisfy a system of compatibility equations with solution space denoted by 𝒞(M,𝒯)\mathcal{C}(M,\mathcal{T}). By introducing the generalized Euler characteristic function χ\chi^{*} defined on 𝒞(M,𝒯)\mathcal{C}(M,\mathcal{T}) and using Farkas’s lemma, we derive a sufficient combinatorial condition for the existence of angle structures, i.e.

Proposition 1.2.

Let MM be a non-compact hyperbolic 33-manifold with finite volume and totally geodesic boundary. There exists an ideal triangulation 𝒯\mathcal{T} on MM, and if χ(s)<0\chi^{*}(s)<0 for all s𝒞(M,𝒯)s\in\mathcal{C}(M,\mathcal{T}) with all quadrilateral coordinates non-negative and at least one quadrilateral coordinates positive, then (M,𝒯)(M,\mathcal{T}) admits an angle structure.

Step 3, by excluding cases that do not satisfy the combinatorial conditions in Proposition 1.2, we get the topological conditions in Theorem 1.1 for the existence of angle structures.

The paper is organized as follows. We give some basic notions in Section 2, including the polyhedral decomposition theory, ideal triangulation, angle structure, normal surface and Farkas’s lemma. In Section 3, we derive an ideal triangulation directly and prove Proposition 1.2. In Section 4, we prove Theorem 1.1.

Acknowledgements: The authors are very grateful to Professor Ruifeng Qiu, Feng Luo, Tian Yang for many discussions on related problems in this paper. The first two authors would like to thank Professor Gang Tian for his long-term support and encouragement. Huabin Ge is supported by NSFC, no.12341102 and no.12122119. Faze Zhang is supported by NSFC, no.12471065.

2 Preliminaries

2.1 Polyhedral decomposition theory

Let us recall the definitions of the ideal tetrahedra and partially truncated tetrahedra first.

Definition 2.1.

An ideal tetrahedron in 3\mathbb{H}^{3} is a finite-volume region in 3\mathbb{H}^{3} bounded by four geodesic faces. Any two faces intersect each other, and any three faces intersect at the infinite boundary of 3\mathbb{H}^{3}. See the left picture of Figure 1 for example.

Refer to caption
Figure 1: an ideal tetrahedron (left) / a type 1-3 partially truncated tetrahedron (right)

Denote 3RP3\mathbb{H}^{3}\subset RP^{3} by the open unite ball representing the hyperbolic 3-space via the projective Klein model. Following the terminology in [4], we have

Definition 2.2.

A hyperideal tetrahedron P^\hat{P} is a non-compact tetrahedron in 3\mathbb{H}^{3} which is just the intersection of 3\mathbb{H}^{3} with a projective tetrahedron P~\tilde{P} whose vertices are all outside 3\mathbb{H}^{3} and whose edges all meet 3\mathbb{H}^{3}.

Given that the hyperideal tetrahedron has been extensively studied and used by several mathematicians, we would like to emphasize that the definition of a hyperideal tetrahedron varies from author to author. Some authors require that all four vertices of a hyperideal tetrahedron must be hyperideal vertices, i.e. points outside the closure of 3\mathbb{H}^{3}, e.g. [27]. It is called strictly hyperideal by Schlenker [39].

Definition 2.3.

A hyperideal tetrahedron of 1-3 type P^\hat{P} means that there are only one vertex vv of P~\tilde{P} lies outside of the closure of 3\mathbb{H}^{3} and the other three vertices are all lie in 3\partial\mathbb{H}^{3}. The truncation of a type 1-3 hyperideal tetrahedron P^\hat{P} at the vertex vv is defined as cutting off the thick end towards vv from P^\hat{P} by a geodesic plane which is perpendicular to the three adjacent geodesic faces at vv. After the truncation of P^\hat{P} at vv, the resulting polyhedron PP is called a partially truncated tetrahedron of 1-3 type. The unique face of PP induced by the truncation is called an external face and the other faces of PP are called internal faces. See the right picture of Figure 1 for example.

The concept of ideal tetrahedron, hyperideal tetrahedron and partially truncated tetrahedra of 1-3 type can be generalized to polyhedra case easily.

Definition 2.4.

An ideal polyhedron in 3\mathbb{H}^{3} is a finite-volume polyhedral region bounded by several geodesic faces. Any two faces intersect each other, and any three faces intersect at the infinite boundary of 3\mathbb{H}^{3}.

Definition 2.5.

A hyperideal polyhedron P^\hat{P} is a non-compact polyhedron in 3\mathbb{H}^{3} which is just the intersection of 3\mathbb{H}^{3} with a projective polyhedron P~\tilde{P} whose vertices are all outside 3\mathbb{H}^{3} and whose edges all meet 3\mathbb{H}^{3}.

Definition 2.6.

A hyperideal polyhedron of 1-k type P^\hat{P} is defined as there is exactly one vertex vv (called hyperideal vertex) of the corresponding P~\tilde{P} lies outside of 33\mathbb{H}^{3}\cup\partial\mathbb{H}^{3} and the other kk vertices (called ideal vertices) are all lie in 3\partial\mathbb{H}^{3}. After the truncation of P^\hat{P} at the hyperideal vertex vv, the resulting polyhedron PP is defined as a partially truncated polyhedron of 1-k type. The unique face of PP induced by the truncation is called an external face and the other faces of PP are called internal faces.

Similarly, one can define the truncation of a hyperideal tetrahedron (or polyhedron resp.), which is truncated at all hyperideal vertices, i.e. points outside the closure of 3\mathbb{H}^{3}. After the truncating, one get a partially truncated tetrahedron (or polyhedron resp.).

Similar to the role of Epstein-Penner’s decomposition [7] played in [17], we developed a theory of decomposition of general hyperbolic 3-manifolds, which will paly essential role in the proof of our main theorem. The theory says any hyperbolic 3-manifold admits a geometric polyhedral decomposition.

Theorem 2.7 (Ge-Jiang-Zhang, [12]).

Let MM be a volume finite, non-compact, complete hyperbolic 33-dimensional manifold with totally geodesic boundary. Then MM admits a mixed geometric ideal polyhedral decomposition 𝒟\mathcal{D} such that each cell is either an ideal polyhedron or a partially truncated polyhedron with only one hyperideal vertex.

Theorem 2.7 may be considered as a combination of Epstein-Penner’s decomposition [7] and Kojima’s decomposition [28]. However, to some extent, our conclusion is slightly stronger, and only ideal polyhedra and 1-kk type polyhedra appear in our decomposition. It is worth emphasizing that there are several advantages to choosing type 1-3 tetrahedra and type 1-kk polyhedra. Firstly, for polyhedral decomposition, it is more suitable for the algorithm. Secondly, it is very suitable for the proof of the main theorem in this paper, which can highlight the essence of the proof, and can also greatly simplify the expression, making the details clearer and more readable. See Section 3-4 for details.

2.2 Ideal triangulation

Strictly speaking, the ideal tetrahedron, as well as the hyperideal tetrahedron, have no real existent vertices. But for the sake of convenience, we often think that they all have four vertices. For example, a truncated type 1-3 tetrahedral σ\sigma has three real vertices, which are the three vertices of a hyperbolic triangle obtained by truncating a strictly hyperideal vertex, and four non-existent vertices, of which three are ideal vertices and one is a strictly hyperideal vertex. But we often think that these four unreal vertices are all the vertices of σ\sigma, while those three real vertices are not considered vertices of σ\sigma.

Additionally, it should be noted that the geometry of ideal tetrahedra, hyperideal tetrahedra, and truncated tetrahedra is hyperbolic. For each ideal tetrahedron, it corresponds to a topological tetrahedron with no geometry (i.e., on the boundary of a topological 3-ball, four marked vertices are selected, and each of the two marked points is connected to a marked edge, which does not intersect each other). It can even be considered equivalently that these topological tetrahedra are combinatorial tetrahedra with only combinatorial structures, where combinatorial structures refer to vertices, edges, faces, and connection relationships between them. We can also think of it as corresponding to a type of Euclidean tetrahedra, which may not be the same in shape, but have the same combinatorial structure. Similarly, each truncated hyperbolic tetrahedron corresponds to a combinatorial tetrahedron and a type of truncated Euclidean tetrahedron. For example, the following Figure 2 depicts the correspondence between a tetrahedron with only one truncated vertex (left) and a tetrahedron with four truncated vertices (right). If a (truncated) Euclidean tetrahedron σ\sigma corresponds to a (truncated) hyperbolic tetrahedron according to the above rules, then a vertex vv in σ\sigma is called an ideal vertex or a hyperideal vertex, defined according to its corresponding vertex type in the (truncated) hyperbolic tetrahedron.

Refer to caption
Figure 2: correspondence of tetrahedra

The surface generated by truncating a vertex is called an external surface. The remaining faces are called internal faces, or in other words, faces containing at least one ideal vertex are called internal faces. In Euclidean tetrahedra and combinatorial tetrahedra, external and internal faces can be defined similarly. The edges in the external faces are called the external edges, and the remaining edges are called the internal edges. Equivalently, the edges that contain at least one ideal vertex are called the internal edges. Thus, each tetrahedron, whether it is an ideal tetrahedron or a truncated tetrahedron, always has exactly six internal edges. Below we provide the precise meaning of an ideal triangulation.

Definition 2.8.

Given a topological three dimensional manifold MM with both topological cusps and boundaries whose connected components all have negative Euler characteristic numbers. For example, MM is derived from a volume finite, non-compact, complete hyperbolic 33-manifold with totally geodesic boundary, after forgetting the hyperbolic structure. Let X=i=1nσiX=\sqcup_{i=1}^{n}\sigma_{i} be a topological space with each σi\sigma_{i} either an Euclidean tetrahedron or a truncated Euclidean tetrahedron. Let Φ\Phi be a collection of affine homeomorphisms between the internal faces, and X(0)X^{(0)} be the set of ideal vertices in XX. If (X\X(0))/Φ(X\backslash X^{(0)})/\Phi is homomorphic to MM, then the cell decomposition 𝒯=X\X(0)\mathcal{T}=X\backslash X^{(0)} is called an ideal triangulation of MM.

Note that the quotient of external faces in XX constitutes a triangulation of M\partial M. The 11-simplices (22-simplices resp.) in MMM\setminus\partial M are called internal edges (faces resp.) of 𝒯\mathcal{T}.

2.3 Angle structure

By definition, an angle structure on a single topological, or combinatorial ideal tetrahedron σ\sigma assigns to each edge of σ\sigma a positive number, called dihedral angle, such that the sum of the angles associated to the three edges meeting in a vertex is π\pi for each vertex of the tetrahedron. After a simple calculation, it can be found that the dihedral angles of opposite sides are equal. For a truncated tetrahedron with at least one hyperideal vertex, the definition of an angle structure is a little different. The main difference is that the sum of the dihedral angles associated to the three edges meeting in a strictly hyperideal vertex (or equivalently, a truncated face) is less than π\pi. Combining the above two definitions, the concept of an angle structure can be extended to the most general 3-dimensional pseudo-manifolds with ideal triangulations, see [11], [19], [21], [24], [26], [27] for instance.

Definition 2.9.

Let 𝒯\mathcal{T} be an ideal triangulation on a topological 3-manifold MM with 3-simplices σ1,,σn\sigma_{1},...,\sigma_{n}. An angle structure on (M,𝒯)(M,\mathcal{T}) is a function α\alpha that, for each internal edge eije_{ij} of σi\sigma_{i} (1in1\leq i\leq n, 1j61\leq j\leq 6), assigns a values α(eij)(0,π)\alpha(e_{ij})\in(0,\pi) which is called the dihedral angle, so that:

  1. 1.

    for any internal edge ee in 𝒯\mathcal{T}, the sum of all dihedral angles around ee is 2π2\pi;

  2. 2.

    if vv is a vertex of some σi\sigma_{i} with 1in1\leq i\leq n, and is not in the external faces of σi\sigma_{i}, the sum of all dihedral angles adjacent to vv in σi\sigma_{i} is π\pi;

  3. 3.

    if ff is an external face of some truncated tetrahedron σj\sigma_{j} with 1jn1\leq j\leq n, the sum of all dihedral angles adjacent to ff in σj\sigma_{j} is less than π\pi.

In the above definition, if all the dihedral angles are allowed to be taken in the closed interval [0,π][0,\pi], it is said to be a semi-angle structure. By the work of Bao-Bonahon [4] ( or see [27] and [39]), any angle structure on (M,𝒯)(M,\mathcal{T}) endows each single 3-simplex σi\sigma_{i} with a hyperbolic geometry, making it an ideal tetrahedron or hyperideal truncated tetrahedron according to its type. However, it must be pointed out that due to the possible differences in edge lengths (even after choosing decorations), these hyperbolic tetrahedra generally cannot be glued together to obtain a hyperbolic structure on MM.

2.4 Normal surface and generalised Euler characteristic

Let 𝒯\mathcal{T} be an ideal triangulation on a topological 3-manifold MM with 3-simplex σ1,,σn\sigma_{1},...,\sigma_{n}.

Definition 2.10.

Let ff be an internal face. A normal arc is an embedding of an arc ll into ff such that l\partial l lie on different internal edges of ff. A normal disk in σi\sigma_{i} (1in1\leq i\leq n) is an embedding of a disk dd into σi\sigma_{i} (1in1\leq i\leq n) such that dσid\cap\partial\sigma_{i} is either empty or consists of disjoint normal arcs.

For each 1in1\leq i\leq n, there are seven types of normal disks on σi\sigma_{i}, of which 4 are normal triangles, corresponding to 4 vertices; 3 normal quadrilaterals, corresponding to 3 sets of opposing edges. For example, Figure 3 shows a special normal triangle, as well as an example of a special normal quadrilateral.

Refer to caption
Figure 3: normal disks
Definition 2.11.

A compact surface FF in MM is called normal with respect to 𝒯\mathcal{T}, if FσiF\cap\sigma_{i} is either empty or consists of disjoint normal disks of σi\sigma_{i} for each 1in1\leq i\leq n.

Recall nn is the number of all types of tetrahedra in the triangulation 𝒯\mathcal{T}. We fix an ordering of all normal disc types (q1,,q3n,t1,,t4n)(q_{1},...,q_{3n},t_{1},...,t_{4n}) in 𝒯\mathcal{T}, where qiq_{i} denotes a normal quadrilateral type and tjt_{j} a normal triangle type. For any given normal surface FF, its normal coordinate F¯=(x1,,x3n,y1,,y4n)\overline{F}=(x_{1},...,x_{3n},y_{1},...,y_{4n}) is a vector in 7n\mathbb{R}^{7n}, where xix_{i} is the number of normal discs of type qiq_{i} in FF, and yjy_{j} is the number of normal discs of type tjt_{j} in FF. A properly embedded normal surface FF is uniquely determined up to normal isotopy by its normal coordinate ([26], [40]). (x1,,x3n)(x_{1},...,x_{3n}) is called the quadrilateral coordinate of FF.

For a normal disk DD, let b(D)b(D) be the number of normal arcs in DMD\cap\partial M. If tt is a normal triangle in FF meeting a particular tetrahedra at edges eie_{i} with edge valence did_{i}, 1i31\leq i\leq 3, then its contribution to the Euler characteristic of FF is taken to be

χ(t)=12(1+b(t))+i=131di.\chi^{*}(t)=-\frac{1}{2}\big{(}1+b(t)\big{)}+\sum_{i=1}^{3}\frac{1}{d_{i}}.

If qq is a normal quadrilateral in FF meeting a particular tetrahedra at edges eie_{i} with edge valence did_{i}, 1i41\leq i\leq 4, then its contribution to the Euler characteristic of FF is taken to be

χ(q)=12(2+b(q))+i=141di.\chi^{*}(q)=-\frac{1}{2}\big{(}2+b(q)\big{)}+\sum_{i=1}^{4}\frac{1}{d_{i}}.

Then the generalized Euler characteristic function χ:7n\chi^{*}:\mathbb{R}^{7n}\rightarrow\mathbb{R} is defined as

χ(x1,,x3n,y1,,y4n)=i=13nxiχ(qi)+j=14nyjχ(tj),\chi^{*}(x_{1},...,x_{3n},y_{1},...,y_{4n})=\sum_{i=1}^{3n}x_{i}\chi^{*}(q_{i})+\sum_{j=1}^{4n}y_{j}\chi^{*}(t_{j}), (2.1)

where qiq_{i} and tjt_{j} are the normal quadrilaterals and normal triangles in FF respectively. The generalized Euler characteristic function is linear, and coincides with the classical Euler characteristic for any embedded or immersed normal surface FF in (M,𝒯)(M,\mathcal{T}). See Section 2.6, [26] for details.

There are some linear constraints between the normal coordinate of a normal surface FF. Let ff be an internal face shared by two tetrahedra, say σi\sigma_{i} and σj\sigma_{j}. Let aa be a normal isotopy class of a normal arc in ff. There are two types of normal disks in σi\sigma_{i}, one is type xix_{i} for normal triangles, the other is type yjy_{j} for normal quadrilaterals, such that when restricted to ff, the corresponding normal arcs are of the same type as aa. Similarly, there are two types of normal disks in σj\sigma_{j}, one is type xix_{i^{\prime}} for normal triangles, the other is type yjy_{j^{\prime}} for normal quadrilaterals with similar properties. Thus, the normal coordinate F¯\overline{F} satisfies a linear equation at the face ff as follows, which is called the compatibility equation:

xi+yj=xi+yj.x_{i}+y_{j}=x_{i^{\prime}}+y_{j^{\prime}}. (2.2)

Denote 𝒞(M,𝒯)\mathcal{C}(M,\mathcal{T}) by the set of all vectors (x1,,x3n,y1,,y4n)7n(x_{1},...,x_{3n},y_{1},...,y_{4n})\in\mathbb{R}^{7n} satisfying (2.2) at any internal face. For this linear space, Kang-Rubinstein [19] once introduced a basis

{Wσi,Wej|i=1,,n,j=1,,m},\big{\{}W_{\sigma_{i}},\,W_{e_{j}}\,|\,i=1,...,n,\,j=1,...,m\big{\}},

where nn is the number of 33-simplex in 𝒯\mathcal{T} and mm is the number of internal edges in 𝒯\mathcal{T}. For concrete meanings of WσiW_{\sigma_{i}} and WejW_{e_{j}}, see [19] and [26]. Then for any s𝒞(M,𝒯)s\in\mathcal{C}(M,\mathcal{T}), there exists a unique coordinate (ω1,,ωn,z1,,zm)(\omega_{1},...,\omega_{n},z_{1},...,z_{m}) such that

s=i=1nωiWσi+j=1mzjWej.s=\sum_{i=1}^{n}\omega_{i}W_{\sigma_{i}}+\sum_{j=1}^{m}z_{j}W_{e_{j}}. (2.3)

Given a normal disk dd in some σi\sigma_{i}, if dd is a kk-sided polygon, and intersects the internal edges ei1,,eike_{i1},...,e_{ik} of some particular σi\sigma_{i}, then its combinatorial area A(d)A(d) is defined to be

j=1kα(eij)(k2)π.\sum_{j=1}^{k}\alpha(e_{ij})-(k-2)\pi. (2.4)
Definition 2.12.

If s=i=1nωiWσi+j=1mzjWejs=\sum\limits_{i=1}^{n}\omega_{i}W_{\sigma_{i}}+\sum\limits_{j=1}^{m}z_{j}W_{e_{j}}, then define

χ(A)(s)=12πtxt(s)A(t),\chi^{(A)}(s)=\frac{1}{2\pi}\sum_{t\in\bigtriangleup}x_{t}(s)A(t), (2.5)

where \bigtriangleup is the set of all the normal triangles of ss, and xtx_{t} is the corresponding normal triangle coordinate of ss.

Remark 2.13.

For any function α\alpha defined on the internal edges of (M,𝒯)(M,\mathcal{T}), which is called “angle” and may not satisfy the conditions for defining an angle structure, one can define the combinatorial area A(d)A(d) without any difference. By definition, for an internal edge ee, its combinatorial curvature k(e)k(e) is 2π2\pi minus the sum of the angles surrounding it. For the area-curvature function (A,k)(A,k), one may define [26] its Euler characteristic function as

χ(A,k)(s)=12π(txt(s)A(t)+j=1m2zjk(ej)),\chi^{(A,\,k)}(s)=\frac{1}{2\pi}\Big{(}\sum_{t\in\bigtriangleup}x_{t}(s)A(t)+\sum_{j=1}^{m}2z_{j}k(e_{j})\Big{)},

where ss is expressed in Definition 2.12. However, only k=0k=0 case is used in our paper.

Luo-Tillmann ([26], Lemma 15) got the following relationship between χ\chi^{*} and χ(A)\chi^{(A)}:

Lemma 2.14.

Suppose (M,𝒯)(M,\mathcal{T}) admits an angle (or semi-angle) structure α\alpha with (A)(A). If s𝒞(M,𝒯)s\in\mathcal{C}(M,\mathcal{T}), then

χ(A)(s)=χ(s)12πqA(q)xq(s),\chi^{(A)}(s)=\chi^{*}(s)-\frac{1}{2\pi}\sum_{q\in\square}A(q)x_{q}(s), (2.6)

where \square is the set of all the normal quadrilaterals of ss and A(q)A(q) is the combinatorial area of qq induced by the angle (or semi-angle) structure α\alpha.

2.5 Farkas’s lemma

If x=(x1,,xk)kx=(x_{1},...,x_{k})\in\mathbb{R}^{k}, we use x>0x>0 (x0x\geq 0, x<0x<0, x0x\leq 0 resp.) to mean that all components of xix_{i} are positive (non-negative, negative, non-positive resp.). To approach our main results, we need the following duality result from linear programming, which is known as Farkas’s lemma, and can be found, for instance, in [42]. In the following lemma, vectors in k\mathbb{R}^{k} and l\mathbb{R}^{l} are considered to be column vectors.

Lemma 2.15.

Let AA be a real k×lk\times l matrix, bkb\in\mathbb{R}^{k}, and \cdot be the inner product on k\mathbb{R}^{k}.

  1. 1.

    {xl|Ax=b}\{x\in\mathbb{R}^{l}|Ax=b\}\neq\emptyset if and only if for all yky\in\mathbb{R}^{k} such that ATy=0A^{T}y=0, one has yb=0y\cdot b=0.

  2. 2.

    {xl|Ax=b,x0}\{x\in\mathbb{R}^{l}|Ax=b,x\geq 0\}\neq\emptyset if and only if for all yky\in\mathbb{R}^{k} such that ATy0A^{T}y\leq 0, one has yb0y\cdot b\leq 0.

  3. 3.

    {xl|Ax=b,x>0}\{x\in\mathbb{R}^{l}|Ax=b,x>0\}\neq\emptyset if and only if for all yky\in\mathbb{R}^{k} such that ATy0A^{T}y\neq 0 and ATy0A^{T}y\leq 0, one has yb<0y\cdot b<0.

3 The proof of Proposition 1.2

It must be pointed out that in the proof of our main theorems, in addition to ideal ones, we only use truncated type 1-3 tetrahedra and type 1-kk polyhedra. Therefore, whenever a hyperideal polyhedron (or partially truncated polyhedron, hyperideal tetrahedron, partially truncated tetrahedron resp.) appears later, it always refers to a hyperideal polyhedron of 1-kk type (partially truncated polyhedron of 1-kk type, hyperideal tetrahedron of 1-3 type, partially truncated tetrahedron of 1-3 type resp.). And from now on, MM always refers to a hyperbolic 3-manifold with both cusps and totally geodesic boundaries.

3.1 The ideal triangulation of MM

Definition 3.1.

Let QQ be an ideal quadrilateral with four ideal vertices v1v_{1}, v2v_{2}, v3v_{3}, v4v_{4} in turn. We connect the two pairs of vertices, v1v_{1} and v3v_{3}, v2v_{2} and v4v_{4}, with geodesic segments e1e_{1} and e2e_{2}, respectively. Then the ideal quadrilateral QQ with diagonals e1e_{1} and e2e_{2} is called a flat ideal tetrahedron, see Figure 4.

Refer to caption
Figure 4: a flat ideal tetrahedron

By Theorem 2.7, we obtain the following (topological) ideal triangulation of MM:

Corollary 3.2.

Let MM be a volume finite, non-compact, complete hyperbolic 33-manifold with totally geodesic boundary. Then MM admits an ideal triangulation 𝒯\mathcal{T} such that each cell is either an (may be flat) ideal tetrahedron or a partially truncated tetrahedron.

Proof.

Using Theorem 2.7, we get a mixed ideal polyhedra decomposition MM such that each cell PP is either an ideal polyhedron or a partially truncated polyhedron with only one truncated hyperideal vertex.

To obtain the ideal triangulation of MM, two more steps are required. The first step is to triangulate eacn polyhedron PP as follows, and there are two cases to consider:

Case 11, assume PP is a partially truncated polyhedron induced by a convex polyhedron P~\tilde{P} in RP3RP^{3}. Recall a pyramid is a polyhedron with one face (known as the “base”) a nn-polygon and all the other nn faces triangles meeting at a common polygon vertex (known as the “tip”) vv. Consider the polyhedron P~\tilde{P}, by taking cones at the unique hyperideal vertex vv, which is called the cone vertex, with each face of P~\tilde{P} disjoint from vv, we get a decomposition of P~\tilde{P} into pyramids. These pyramids have a common tip vv, and each face of P~\tilde{P} disjoint from vv is the base of a pyramid. For any base polygon DD in a pyramid, we arbitrarily pick a vertex ww of DD and decompose DD into triangles by taking cones at ww. The triangulation of DD extends to a triangulation of the pyramid into tetrahedra, and no vertices are added throughout the process. In this way, P~\tilde{P} is decomposed into a union of convex tetrahedra, each of which has only one common hyperideal vertex vv in RP3RP^{3}. Then after the truncation at the hyperideal vertex vv from the intersection of 3\mathbb{H}^{3} with P~\tilde{P}, PP has a triangulation whose cells are all partially truncated tetrahedron.

Case 22, assume PP is an ideal polyhedron. By taking cones at any ideal vertex vv, which is also called cone vertex, we get a decomposition of PP into pyramids with tips vv and bases the faces of PP disjoint from vv. Similar to case 11, for any base polygon DD in a pyramid, we arbitrarily pick a vertex ww of DD and decompose DD into ideal triangles by taking cones at ww. The ideal triangulation of DD extends to a decomposition of the pyramid into ideal tetrahedra. In this way, PP is decomposed into a union of ideal tetrahedra with no vertices are added throughout the process.

Consider a common face DD of two adjacent polyhedra PP and PP^{\prime}. If both PP and PP^{\prime} are partially truncated, by the constructions in case 1, the triangulation of DD in PP matches perfectly with the triangulation of DD in PP^{\prime}. For the remaining possibilities, at least one of PP and PP^{\prime} is an ideal polyhedron, the decompositions of the ideal polygon DD from PP and PP^{\prime} are not always match together to get a triangulation. Such face DD is called a pillow.

The second step is to insert flat tetrahedra into these pillows to get a layered triangulation as follows.

Let vv and vv^{\prime} be the vertices of a face DD where the cone is taken at separately from PP and PP^{\prime}. Let v1,,viv_{1},\dots,v_{i} and ω1,,ωj\omega_{1},\dots,\omega_{j} be the remaining vertices of DD which are arranged sequentially along the boundaries of DD from vv to vv^{\prime}, see Figure 5 for example. For each diagonal switch from vkvv_{k}v^{\prime} to vk+1vv_{k+1}v, k=1,,i1k=1,...,i-1, and each diagonal switch from ωkv\omega_{k}v^{\prime} to ωk+1v\omega_{k+1}v, k=1,,j1k=1,...,j-1, we insert a flat tetrahedron with vertices {v,vi,vi+1,v}\{v,v_{i},v_{i+1},v^{\prime}\} and a flat tetrahedron with vertices {v,ωk,ωk+1,v}\{v,\omega_{k},\omega_{k+1},v^{\prime}\} respectively.

Refer to caption
Figure 5: triangulating a pillow

After the above two steps, we finally get an ideal triangulation 𝒯\mathcal{T} of MM such that each cell is an (may be flat) ideal tetrahedron or a partially truncated tetrahedron. ∎

Note at each internal edge in the ideal triangulation 𝒯\mathcal{T}, the value of its dihedral angle is in [0,π][0,\pi], hence we obtain a naturally semi-angle structure on 𝒯\mathcal{T}.

3.2 Sufficient conditions for the existence of angle structures

Let 𝒯\mathcal{T} be the ideal triangulation derived in the previous section, and all its (flat) ideal or partially truncated tetrahedra are σ1,,σn\sigma_{1},...,\sigma_{n}. Suppose α\alpha is an angle structure on (M,𝒯)(M,\mathcal{T}). For each σi\sigma_{i} and each internal edge eij(1j6)e_{ij}~{}(1\leq j\leq 6) in σi\sigma_{i}, consider the dihedral angle αij=α(eij)\alpha_{ij}=\alpha(e_{ij}) at eije_{ij} as a variable of the equations in Farkas’ Lemma 2.15, see Figure 6.

Refer to caption
Figure 6: the edge labelling and normal triangles in σi\sigma_{i}

At the four normal triangles ti1t_{i}^{1}, ..., ti4t_{i}^{4} of σi\sigma_{i}, we have the following equations:

{αi1+αi2+αi3=ai1αi1+αi5+αi6=ai2αi3+αi4+αi5=ai3αi2+αi4+αi6=ai4.\begin{cases}\alpha_{i1}+\alpha_{i2}+\alpha_{i3}=a_{i}^{1}\\ \alpha_{i1}+\alpha_{i5}+\alpha_{i6}=a_{i}^{2}\\ \alpha_{i3}+\alpha_{i4}+\alpha_{i5}=a_{i}^{3}\\ \alpha_{i2}+\alpha_{i4}+\alpha_{i6}=a_{i}^{4}.\\ \end{cases} (3.1)

If σi\sigma_{i} is a partially truncated tetrahedron, then only one of its aika_{i}^{k} is less then π\pi, and the the remaining three are equal to π\pi. If σi\sigma_{i} is an (flat) ideal tetrahedron, then all its aika_{i}^{k} are equal to π\pi. The elements of {aik|1in,1k4}\{a_{i}^{k}|1\leq i\leq n,1\leq k\leq 4\} are divided into two classes, one of them satisfies {ai1k1<π}\{a_{i_{1}}^{k_{1}}<\pi\} with the corresponding normal triangles denoted by {ti1k1}\{t_{i_{1}}^{k_{1}}\}, the other satisfies {ai2k2=π}\{a_{i_{2}}^{k_{2}}=\pi\} with the corresponding normal triangles denoted by {ti2k2}\{t_{i_{2}}^{k_{2}}\}.

Recall nn is the number of 33-simplex in 𝒯\mathcal{T} and mm is the number of internal edges in 𝒯\mathcal{T}. At each internal edge eje_{j} of 𝒯\mathcal{T}, where 1jm1\leq j\leq m, there holds:

iαij=2π.\sum_{i}\alpha_{ij}=2\pi. (3.2)

where the sum traverses all the tetrahedra surround the edge eje_{j}.

Consider the vector

(a,b)=(,aik,,bj,)(a,b)=(...,a_{i}^{k},...,b_{j},...) (3.3)

where aika_{i}^{k} (i=1,,ni=1,...,n, k=1,,4k=1,...,4) comes from (3.1) and bj=2πb_{j}=2\pi (j=1,,mj=1,...,m) comes from (3.2). Writing the system of equations (3.1) and (3.2) in a matrix form as

Bx=(a,b)T.Bx=(a,b)^{T}.

Consider the transpose BTB^{T} of BB, which is also the dual of BB. It has one variable hikh_{i}^{k} for each normal triangle type tikt_{i}^{k} and one variable zjz_{j} for each internal edge eje_{j}. If we denote the variables corresponding to BTB^{T} by (h,z)=(,hik,,,zj,)(h,z)=(...,h_{i}^{k},...,...,z_{j},...), where i=1,,ni=1,...,n, k=1,,4k=1,...,4 and j=1,,mj=1,...,m, then BT(h,z)T=(,zj+hik+hil,)TB^{T}(h,z)^{T}=(...,z_{j}+h_{i}^{k}+h_{i}^{l},...)^{T}.

The following useful formula (i.e., Formula (4.9) in [26]) belongs to Luo-Tillmann:

1π(h,z)(a,b)\displaystyle\frac{1}{\pi}(h,z)\cdot(a,b) =χ(Wω,z)χ(A)(Wω,z)\displaystyle=\chi^{*}(W_{\omega,z})-\chi^{(A)}(W_{\omega,z}) (3.4)
+12π(zj+hik+hil)(aik+ail2π).\displaystyle\;\;\;\;+\frac{1}{2\pi}\sum(z_{j}+h_{i}^{k}+h_{i}^{l})(a_{i}^{k}+a_{i}^{l}-2\pi).

where Wω,z=i=1nωiWσi+j=1mzjWejW_{\omega,z}=\sum\limits_{i=1}^{n}\omega_{i}W_{\sigma_{i}}+\sum\limits_{j=1}^{m}z_{j}W_{e_{j}}, and the summation runs over all internal edges in the whole triangulation 𝒯\mathcal{T}.

Note the vector (a,b)(a,b) looks like this

(,ai1k1,,π,,πa,2π,,2πb),(~{}\underbrace{...~{},~{}a_{i_{1}}^{k_{1}},~{}...~{},~{}\pi,~{}...~{},\pi}_{a},~{}\underbrace{2\pi,~{}...~{},2\pi}_{b}~{}),

where ai1k1<πa_{i_{1}}^{k_{1}}<\pi. In order to obtain an angle structure on (M,𝒯)(M,\mathcal{T}), we only need to prove that there exists a vector (a,b)=(,ai1k1,,π,,2π,)(a,b)=(...,a_{i_{1}}^{k_{1}},...,\pi,...,2\pi,...) such that

{Bx=(a,b)T,x>0}.\big{\{}Bx=(a,b)^{T},\,x>0\big{\}}\neq\emptyset. (3.5)

Further using the third part of Farkas’s lemma, i.e., Lemma 2.15, we just need to show that for all (h,z)(h,z) with BT(h,z)T0B^{T}(h,z)^{T}\neq 0 and BT(h,z)T0B^{T}(h,z)^{T}\leq 0, one has (h,z)(a,b)<0(h,z)\cdot(a,b)<0.

Claim 3.3.

There exists a vector (,ai1k1,,π,,2π,)=(a,b)(...,a_{i_{1}}^{k_{1}},...,\pi,...,2\pi,...)=(a,b) with ai1k1<πa_{i_{1}}^{k_{1}}<\pi such that

(h,z)(a,b)<0(h,z)\cdot(a,b)<0

for all (h,z)(h,z) with BT(h,z)T0B^{T}(h,z)^{T}\neq 0 and BT(h,z)T0B^{T}(h,z)^{T}\leq 0.

Proof.

By Corollary 3.2, 𝒯\mathcal{T} is an ideal triangulation of MM. Then (M,𝒯)(M,\mathcal{T}) admits a semi-angle structure according to the decomposition operation in Section 3.1 (when inserting a flat polyhedron, a 0 angle or π\pi angle may occur). Then by the second part of Farkas’s Lemma 2.15, there exists a a¯i1k1<π\bar{a}_{i_{1}}^{k_{1}}<\pi such that (h,z)(a¯,b)0(h,z)\cdot(\bar{a},b)\leq 0 for all BT(h,z)T0B^{T}(h,z)^{T}\leq 0, where (a¯,b)=(,a¯i1k1,,π,,2π,)(\bar{a},b)=(...,\bar{a}_{i_{1}}^{k_{1}},...,\pi,...,2\pi,...).

According to (3.4), as well as the known conditions in Theorem 1.2, if ai1k1=πa_{i_{1}}^{k_{1}}=\pi, then

1π(h,z)(a,b)\displaystyle\;\;\;\;\frac{1}{\pi}(h,z)\cdot(a,b)
=χ(Wω,z)χ(A)(Wω,z)+12π(zj+hik+hil)(aik+ail2π)\displaystyle=\chi^{*}(W_{\omega,z})-\chi^{(A)}(W_{\omega,z})+\frac{1}{2\pi}\sum(z_{j}+h_{i}^{k}+h_{i}^{l})(a_{i}^{k}+a_{i}^{l}-2\pi)
=χ(Wω,z)χ(A)(Wω,z)\displaystyle=\chi^{*}(W_{\omega,z})-\chi^{(A)}(W_{\omega,z})
=χ(Wω,z)<0.\displaystyle=\chi^{*}(W_{\omega,z})<0.

Next, we use the continuity method to prove the above, and for this, we introduce a function F(t)F(t) defined for all t[0,1]t\in[0,1]:

F(t)(h,z)\displaystyle F(t)(h,z) =(h,z)(,a¯i1k1+t(πa¯i1k1),,π,,2π,)\displaystyle=(h,z)\cdot(...,\bar{a}_{i_{1}}^{k_{1}}+t(\pi-\bar{a}_{i_{1}}^{k_{1}}),...,\pi,...,2\pi,...)
=(a¯i1k1+t(πa¯i1k1))hi1k1+πhi2k2+j=1m2πzj.\displaystyle=\sum(\bar{a}_{i_{1}}^{k_{1}}+t(\pi-\bar{a}_{i_{1}}^{k_{1}}))h_{i_{1}}^{k_{1}}+\sum\pi h_{i_{2}}^{k_{2}}+\sum_{j=1}^{m}2\pi z_{j}.

Here (h,z)(h,z) are the variables corresponding to BTB^{T}, and hi2k2h_{i_{2}}^{k_{2}} is the variable corresponding to aik=πa_{i}^{k}=\pi in the equation (3.1).

Thus we have F(0)0F(0)\leq 0, and by Equation (3.2), F(1)<0F(1)<0 for all BT(h,z)T0B^{T}(h,z)^{T}\neq 0 and BT(h,z)T0B^{T}(h,z)^{T}\leq 0.

Lemma 3.4.

F(t)(h,z)<0F(t)(h,z)<0 for any t(0,1)t\in(0,1), and for any variable (h,z)(h,z).

Proof.

We prove it by contradiction. Assume there is a t0(0,1)t_{0}\in(0,1) and (h0,z0)(h_{0},z_{0}) such that

F(t0)(h0,z0)0.F(t_{0})(h_{0},z_{0})\geq 0.

On the one hand F(0)(h0,z0)0F(0)(h_{0},z_{0})\leq 0, derived from the seme-angle structure on (M,𝒯M,\mathcal{T}), and on the other hand, F(t)(h0,z0)F(t)(h_{0},z_{0}) as a function of tt is monotonic because of the linearity of F(t)(h0,z0)F(t)(h_{0},z_{0}). Then F(1)(h0,z0)0F(1)(h_{0},z_{0})\geq 0, which contradicts the fact that F(1)(h0,z0)<0F(1)(h_{0},z_{0})<0. Then we proved the above lemma.

Finally, if set ai1k1=a¯i1k1+t0(πa¯i1k1)<πa_{i_{1}}^{k_{1}}=\bar{a}_{i_{1}}^{k_{1}}+t_{0}(\pi-\bar{a}_{i_{1}}^{k_{1}})<\pi, then we have

(h,z)(a,b)=F(t0)(h,z)<0.(h,z)\cdot(a,b)=F(t_{0})(h,z)<0.

Hence we get the above Claim 3.3. ∎

By the third part of Farkas’s Lemma 2.15,

{Bx=(a,b)T,x>0}.\big{\{}Bx=(a,b)^{T},x>0\big{\}}\neq\emptyset.

Hence (M,𝒯)(M,\mathcal{T}) admits an angle structure, and Proposition 1.2 is proved.

4 The proof of Theorem 1.1

Let MM be a cusped hyperbolic 33-manifold with totally geodesic boundary and 𝒯\mathcal{T} be the ideal triangulation of MM as in Corollary 3.2. Recall that there exists a natural semi-angle structure α\alpha on (M,𝒯)(M,\mathcal{T}) by the analysis in Section 3. For a quadrilateral type normal disk qq in 𝒯\mathcal{T}, the combinatorial area of qq by definition is

A(q)=α1+α2+α3+α42π,A(q)=\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}-2\pi,

where α1,,α4\alpha_{1},...,\alpha_{4} are the dihedral angles assigned by α\alpha in the vertices (which are the intersections of qq with the four edges of a particular tetrahedron σi\sigma_{i}) of qq.

Definition 4.1.

A quadrilateral type normal disk qq is called vertical if A(q)=0A(q)=0.

From Proposition 1.2 and Section 3, we get the following proposition:

Proposition 4.2.

If there is no such s𝒞(M,𝒯)s\in\mathcal{C}(M,\mathcal{T}), all of its quadrilateral coordinates are non-negative, all non-vertical quadrilateral coordinates are zero, and at least one quadrilateral coordinate is positive, then (M,𝒯)(M,\mathcal{T}) admits an angle structure.

Proof.

Let α\alpha be the semi-angle structure on (M,𝒯)(M,\mathcal{T}), and qq be a quadrilateral type normal disk in a particular tetrahedron σi\sigma_{i} of 𝒯\mathcal{T}.

If σi\sigma_{i} is a flat tetrahedron, then A(q)=0+π+0+π2π=0A(q)=0+\pi+0+\pi-2\pi=0. Hence qq is vertical.

If σi\sigma_{i} is an ideal or partially truncated tetrahedron, in this case it is not flat, then the combinatorial area

A(q)=α1+α2+α3+α42π<α1+α2+α3+α4+2α52π=(α1+α2+α5π)+(α3+α4+α5π)0,\begin{split}A(q)&=\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}-2\pi\\ &<\alpha_{1}+\alpha_{2}+\alpha_{3}+\alpha_{4}+2\alpha_{5}-2\pi\\ &=(\alpha_{1}+\alpha_{2}+\alpha_{5}-\pi)+(\alpha_{3}+\alpha_{4}+\alpha_{5}-\pi)\\ &\leq 0,\end{split}

where α5\alpha_{5} is the angle of the edge in σi\sigma_{i} which faces qq, and is positive as a real dihedral angle induced from the hyperbolic structure.

According to the conditions in the proposition, for any s𝒞(M,𝒯)s\in\mathcal{C}(M,\mathcal{T}) with all quadrilateral coordinates non-negative and at least one quadrilateral coordinate positive, there exist quadrilateral type normal disk qq such that qq is non-vertical and xq>0x_{q}>0. Note χ(0)(s)=0\chi^{(0)}(s)=0, hence by Lemma 2.14,

χ(s)=χ(s)χ(0)(s)=12πqA(q)xq(s)<0.\chi^{*}(s)=\chi^{*}(s)-\chi^{(0)}(s)=\frac{1}{2\pi}\sum_{q\in\square}A(q)x_{q}(s)<0.

Then by Theorem 1.2, we get an angle structure on 𝒯\mathcal{T}. ∎

Recall 𝒞(M,𝒯)7\mathcal{C}(M,\mathcal{T})\subset\mathbb{R}^{7} is the solution space of the compatibility equations (2.2). Denote 𝒩(M,𝒯)\mathcal{N}(M,\mathcal{T}) by a subset of 𝒞(M,𝒯)\mathcal{C}(M,\mathcal{T}) whose element has all non-negative integer coordinates. With a similar argument of Hodgson-Rubinstein-Segerman [17], we have:

Proposition 4.3.

If there is no such x𝒩(M,𝒯)x\in\mathcal{N}(M,\mathcal{T}), all of its quadrilateral coordinates are non-negative, all non-vertical quadrilateral coordinates are zero, and at least one quadrilateral coordinate is positive, then (M,𝒯)(M,\mathcal{T}) admits an angle structure.

Proof.

If there is a x𝒞(M,𝒯)x\in\mathcal{C}(M,\mathcal{T}) satisfying the properties assumed in Proposition 4.2, then the coordinates of xx are all rational numbers since the coefficients in the compatibility equations (2.2) are all integers. So there is a positive integer NN such that Nx𝒞(M,𝒯)Nx\in\mathcal{C}(M,\mathcal{T}), i.e., NxNx has integer coordinates. Because all the quadrilateral coordinates of xx are non-negative, so is NxNx. If all the triangle coordinates of NxNx are non-negative, we are done. If some triangle coordinates of NxNx are negative, we can always add enough normal copies of peripheral M\partial M, tori or Klein bottles. Then we can get a (Nx)𝒞(M,𝒯)(Nx)^{\prime}\in\mathcal{C}(M,\mathcal{T}) whose triangle coordinates are all none-negative. Hence (Nx)𝒩(M;𝒯)(Nx)^{\prime}\in\mathcal{N}(M;\mathcal{T}), we are done.

According to Proposition 4.3, to get an angle structure on (M,𝒯)(M,\mathcal{T}), we should exclude the cases that “x𝒩(M,𝒯)x\in\mathcal{N}(M,\mathcal{T}) with all its quadrilateral coordinates non-negative, all non-vertical quadrilateral coordinates zero, and at least one quadrilateral coordinate positive”.

Let 𝒟\mathcal{D} be the mixed polyhedron decomposition described in Theorem 2.7. Let D^\hat{D} be the decomposition synthesized by 𝒟\mathcal{D} with the ideal polyhedra pillows defined in the proof of Corollary 3.2. Let {Pj,j=1,,k}\{P_{j},j=1,...,k^{\prime}\} and {Dj,j=k+1,,n}\{D_{j^{\prime}},j^{\prime}=k^{\prime}+1,...,n^{\prime}\} be the ideal polyhedra (including partially truncated polyhedra) and polygonal pillows respectively in D^\hat{D}, and let {σi,i=1,,k}\{\sigma_{i^{\prime}},i^{\prime}=1,...,k\} and {σi′′,i′′=k+1,,n}\{\sigma_{i^{\prime\prime}},i^{\prime\prime}=k+1,...,n\} be the non-planar 3-simplices and planar tetrahedra respectively in 𝒯\mathcal{T}. Notably, according to Lemma 7.4 of Hodgson-Rubinstein-Segerman [17], each polygonal pillow represented by DjD_{j^{\prime}} can only be a quadrilateral or a hexagon.

Consider the normal surface SS corresponds to a s𝒩(M,𝒯)s\in\mathcal{N}(M,\mathcal{T}). If SσiS\cap\sigma_{i^{\prime}} does not contain any normal quadrilaterals for each i=1,,ki^{\prime}=1,...,k, and Sσi′′S\cap\sigma_{i^{\prime\prime}} contains at least one normal quadrilateral for some k+1i′′nk+1\leq i^{\prime\prime}\leq n, then

  1. 1.

    for each j=1,,kj=1,...,k^{\prime} so that SPjS\cap P_{j}\neq\varnothing, then SPjS\cap P_{j} are vertex-linking polygons;

  2. 2.

    for each j=k+1,,nj^{\prime}=k^{\prime}+1,...,n^{\prime} so that SDjS\cap D_{j^{\prime}}\neq\varnothing, then SDjS\cap D_{j^{\prime}} are either twisted surfaces or vertex-linking bigons.

See Figure 7 for example.

Refer to caption
Figure 7: (a)(a) is a vertex-linking normal disk; (b)(b) is a vertex-linking bigon in a pillow; (c)(c) is a twisted 66-gon in a pillow

The following are the proof of Theorem 1.1:

Proof.

Let MM be a non-compact, volume-finite hyperbolic three-dimensional manifold. Denote M¯\overline{M} by the compact 33-manifold with boundary, with each torus (or Klein bottle) boundary component corresponds to a cusp of MM. After subtracting the torus (or Klein bottle) boundary components, M¯\overline{M} is homeomorphic to MM, and each boundary of M¯\overline{M} is a torus or Klein bottle or a closed surface of high genus. Consider the mixed polyhedral decomposition 𝒟\mathcal{D} of MM described in Theorem 2.7. It determines a graph Γ\Gamma with one vertex for each polyhedron and an edge joining two vertices for each face incident to the two corresponding polyhedra

Lemma 4.4.

There exists an ideal triangulation 𝒯\mathcal{T}^{\prime} of MM, which is a subdivision of 𝒟\mathcal{D}, such that Γ\Gamma contains a maximal tree TT with the property that the gluing faces corresponding to the edges of TT are not polygon pillows in (M,𝒯)(M,\mathcal{T}^{\prime}).

Proof.

For each vertex of degree 11 in the maximal tree TT, if the corresponding polyhedron is a truncated one, take the hyperideal vertex as the cone vertex; if the corresponding polyhedron is an ideal one, take the ideal vertex which is not on the face and is represented by an edge of TT as the cone vertex. Then remove these vertices and the corresponding edges from TT, and continue the above operation in the remaining tree until either there is an empty set left, or a single point set left.

If there is an empty set left, then the proof is complete. If there is a single point set left, there are two scenarios to consider:

Case 11, if the corresponding polyhedron is ideal, then any triangulation of this polyhedron satisfies the condition.

Case 22, if the corresponding polyhedron is truncated, then take the unique hyperideal vertex of this truncated polyhedron as the cone vertex, and take any triangulation at the other ideal faces. Since each truncated face is shared by two truncated polyhedra, the inherited triangulations of the truncated faces match perfectly. Hence the final triangulation of this truncated polyhedron also satisfies the condition. ∎

Note that the construction of 𝒯\mathcal{T}^{\prime} and 𝒯\mathcal{T} is consistent as in Corollary 3.2. Hence (M;𝒯)(M;\mathcal{T}^{\prime}) also satisfies Proposition 1.2.

For the ideal triangulation 𝒯\mathcal{T}^{\prime}, denote 𝒟\mathcal{D}^{\prime} by the union of 𝒟\mathcal{D} and polygon pillows generated by 𝒯\mathcal{T}^{\prime}. Now assume that there exists a s𝒩(M,𝒯)s\in\mathcal{N}(M,\mathcal{T}^{\prime}), and without loss of generality, further assume that the corresponding normal surface SS (of ss) is not a Haken sum and contains some twisted disks. By the conditions of Theorem 1.1 and the construction of 𝒯\mathcal{T}^{\prime}, SS does not intersect the maximal tree TT.

Next we show the number of twisted disks in SS is even. If there is an odd number of twisted disks of SS at some polygon pillow in 𝒟\mathcal{D}^{\prime}. Denote the edge on Γ\Gamma corresponding to the polygon pillow by ee. Then TeT\cup e contains a closed curve cc whose intersection with SS is odd. Thus, cc represents a nontrivial element [c][c] in the homology group H1(M¯;Z2)H_{1}(\overline{M};Z_{2}).

Consider the following long exact sequence:

H2(M¯,M¯;Z2)H1(M¯;Z2)fH1(M¯;Z2)gH1(M¯,M¯;Z2)...\rightarrow H_{2}(\overline{M},\partial\overline{M};Z_{2})\rightarrow H_{1}(\partial\overline{M};Z_{2})\stackrel{{\scriptstyle f}}{{\rightarrow}}H_{1}(\overline{M};Z_{2})\stackrel{{\scriptstyle g}}{{\rightarrow}}H_{1}(\overline{M},\partial\overline{M};Z_{2})\rightarrow...

Since the closed curve representing any non-trivial element in H1(M¯;Z2)H_{1}(\partial\overline{M};Z_{2}) must have an even geometric intersection number with any closed surface in MM, [c][c] is not in the image of the map ff, and therefore not in the kernel of the map gg, contradicting the known conditions in Theorem 1.1. Therefore, the number of twisted disks in SS is even at all polygon pillows.

If the number of all vertex-linking disks and vertex-linking bigons are even, then SS is a double of some surface, contradicting the assumption that SS is not a Haken sum. Hence, there must be at least one vertex-linking bigons or vertex-linking disks whose number is odd. Let SS^{\prime} be a copy of all such odd-count vertex-linking bigons or vertex-linking disks.

Claim 4.5.

SS^{\prime} is a solution to the compatibility equations.

Proof.

Consider the normal arc type α\alpha at a face FF in the triangulation (M,𝒯)(M,\mathcal{T}^{\prime}).

Case 11: Both tetrahedra σ1\sigma_{1}, σ2\sigma_{2} adjacent to FF are ideal, or truncated. If SS^{\prime} has a normal disk at FF, it must be a vertex-linking disk in some ideal polyhedron or partially truncated polyhedron of 𝒟\mathcal{D}^{\prime}. Since the contributions of σ1\sigma_{1}, σ2\sigma_{2} to α\alpha are the same, it satisfies the compatibility equations.

Case 22: Both tetrahedra σ1\sigma_{1}, σ2\sigma_{2} adjacent to FF are flat ideal. If SS^{\prime} has a normal disk at FF, it must be a vertex-linking bigon in some polygonal pillows of 𝒟\mathcal{D}^{\prime}. Again, the contributions of σ1\sigma_{1}, σ2\sigma_{2} to α\alpha are the same, satisfying the compatibility equations.

Case 33: One tetrahedron σ1\sigma_{1} is ideal or truncated, and the other σ2\sigma_{2} is flat ideal. Since the possible twisted disks on SS at α\alpha are even, the parity of bigons and linking disks at α\alpha is the same. If SS^{\prime} has a normal disk at FF, it must be a bigon and a linking disk in some polygonal pillows of 𝒟\mathcal{D}^{\prime}. The contributions of σ1\sigma_{1}, σ2\sigma_{2} to α\alpha are the same, also satisfying the compatibility equations. ∎

Therefore, SS is the Haken sum of SS^{\prime} and some surface which contains at least one twisted disk. This contradicts with the assumption that SS is not a Haken sum. Therefore Theorem 1.1 is proved. ∎

References

  • [1] H. Akiyoshi, On the Ford domains of once-punctured torus groups, from: “Hyperbolic spaces and related topics (Japanese) (Kyoto, 1998)”, Sūrikaisekikenkyūsho Kōkyūroku 1104 (1999), 109-121.
  • [2] H. Akiyoshi, M. Sakuma, M. Wada, Y. Yamashita, Jørgensen’s picture of punctured torus groups and its refinement, from: “Kleinian groups and hyperbolic 3-manifolds (Warwick, 2001)”, (Y. Komori, V. Markovic, C. Series, editors), London Math. Soc. Lecture Note Ser. 299, Cambridge Univ. Press (2003), 247-273.
  • [3] H. Akiyoshi, M. Sakuma, M. Wada, Y. Yamashita, Punctured torus groups and 2-bridge knot groups. I, Lecture Notes in Math. 1909, Springer, Berlin (2007).
  • [4] X. Bao, F. Bonahon, Hyperideal polyhedra in hyperbolic 3-space, Bull. Soc. Math. France 130 (2002), no. 3, 457-491.
  • [5] R. Benedetti, C. Petronio, Lectures on hyperbolic geometry, Springer, Berlin, 1992.
  • [6] H. Cao, X. Zhu, A complete proof of the Poincaré and geometrization conjectures-application of the Hamilton-Perelman theory of the Ricci flow, Asian J. Math. 10 (2006), 165-492, Erratum in Asian J. Math. 10 (2006), 663-664.
  • [7] D. Epstein, R. Penner, Euclidean decomposition of non-compact hyperbolic manifolds, J. Differential Geom. 27 (1988), 67-80.
  • [8] K. Feng, H. Ge, B. Hua, Combinatorial Ricci flows and the hyperbolization of a class of compact 3-manifolds, Geom. Topol. 26 (2022), no. 3, 1349-1384.
  • [9] D. Futer, F. Guéritaud, From angled triangulations to hyperbolic structures, Contemp. Math., 541 American Mathematical Society, Providence, RI, 2011, 159-182.
  • [10] D. Futer, E. Hamilton, N. R. Hoffman, Infinitely many virtual geometric triangulations, J. Topol. 15 (2022), no. 4, 2352-2388.
  • [11] S. Garoufalidis, C. D. Hodgson, J. H. Rubinstein, H. Segerman, 1-efficient triangulations and the index of a cusped hyperbolic 3-manifold, Geom. Topol. 19 (2015), 2619-2689.
  • [12] H. Ge, L. Jia, F. Zhang, The polyhedral decomposition of cusped hyperbolic 33-manifolds with totally geodesic boundary, arxiv, 2024.
  • [13] F. Guéritaud, On canonical triangulations of once-punctured torus bundles and two-bridge link complements, Geom. Topol. 10 (2006), 1239-1284, With an appendix by D. Futer.
  • [14] F. Guéritaud, Triangulated cores of punctured-torus groups, J. Differential Geom. 81 (2009), 91-142.
  • [15] F. Guéritaud, S. Schleimer, Canonical triangulations of Dehn fillings, Geom. Topol. 14 (2010), no. 1, 193-242.
  • [16] S. L. Ham, J. S. Purcell, Geometric triangulations and highly twisted links, Algebr. Geom. Topol. 23 (2023), 1399-1462.
  • [17] C. D. Hodgson, J. H. Rubinstein, H. Segerman, Triangulations of hyperbolic 3-manifolds admitting strict angle structures, J. Topol. 5 (2012), no. 4, 887-908.
  • [18] T. Jørgensen, On pairs of once-punctured tori, from: “Kleinian groups and hyperbolic 3-manifolds (Warwick, 2001)”, (Y Komori, V Markovic, C Series, editors), London Math. Soc. Lecture Note Ser. 299, Cambridge Univ. Press (2003), 183-207.
  • [19] E. Kang, J. H. Rubinstein, Ideal triangulations of 33-manifolds. II. Taut and angle structures, Algebr. Geom. Topol.5 (2005), 1505-1533.
  • [20] B. Kleiner, J. Lott, Notes on Perelman’s papers, Geom. Topol. 12 (2008), no.5, 2587-2855.
  • [21] M. Lackenby, Word hyperbolic Dehn surgery, Invent. Math. 140 (2000), 243-282.
  • [22] M. Lackenby, The canonical decomposition of once-punctured torus bundles, Comment. Math. Helv. 78 (2003), 363-384.
  • [23] M. Lackenby, Taut ideal triangulations of 3-manifolds, Geom. Topol. 4 (2000), 369-395.
  • [24] F. Luo, A combinatorial curvature flow for compact 3-manifolds with boundary, Electron. Res. Announc. Amer. Math. Soc. 11 (2005) 12-20.
  • [25] F. Luo, S. Schleimer, S. Tillmann, Geodesic ideal triangulations exist virtually, Proc. Amer. Math. Soc. 136 (2008), no. 7, 2625-2630.
  • [26] F. Luo, S. Tillmann, Angle structures and normal surfaces, Trans. Amer. Math. Soc. 360 (2008), no. 6, 2849-2866.
  • [27] F. Luo, T. Yang, Volume and rigidity of hyperbolic polyhedral 3-manifolds, J. Topol. 11 (2018) 1-29.
  • [28] S. Kojima, Polyhedral decomposition of hyperbolic 3-manifolds with totally geodesic boundary, Aspects of low-dimensional manifolds, 93-112, Adv. Stud. Pure Math., 20 (1992), 93-112.
  • [29] B. Martelli, An Introduction to Geometric Topology, eprint arXiv:1610.02592.
  • [30] E. E. Moise, Affine structures in 3-Manifolds. V. The triangulation theorem and hauptvermutung, Ann. Math. 56 (1952), 96-114.
  • [31] J. Morgan, G. Tian, Ricci flow and the Poincaré conjecture, Clay Mathematics Monographs 3, Amer. Math. Soc. (2007).
  • [32] B. Nimershiem, Geometric triangulations of a family of hyperbolic 3-braids, Algebr. Geom. Topol. 23 (2023), no.9, 4309-4348.
  • [33] G. Perelman, The entropy formula for the Ricci flow and its geometric applications, arXiv:math/0211159.
  • [34] G. Perelman, Ricci flow with surgery on three-manifolds, arXiv:math/0303109.
  • [35] G. Perelman, Finite extinction time for the solutions to the Ricci flow on certain three-manifolds, arXiv:math/0307245.
  • [36] C. Petronio, J. Porti, Negatively oriented ideal triangulations and a proof of Thurston’s hyperbolic Dehn filling theorem, Expo. Math. 18 (2000), no. 1, 1-35.
  • [37] R. Qiu, F. Zhang, T. Yang, Angle structures and hyperbolic 33-manifolds with totally geodesic boundary, arXiv:1405.1545v2, 2014.
  • [38] I. Rivin, Euclidean structures on simplicial surfaces and hyperbolic volume, Ann. of Math. 139 (1994), 553-580.
  • [39] J. M. Schlenker, Hyperideal polyhedra in hyperbolic manifolds, preprint (2002) arXiv math/0212355.
  • [40] S. Tillmann, Normal surfaces in topologically finite 3-manifolds, Enseign. Math. 54(3-4) (2008), 329-380.
  • [41] W. P. Thurston, The geometry and topology of three-manifolds, lecture notes, Princeton Univ. (1980), available at http://library.msri.org/books/gt3m/.
  • [42] G. M. Ziegler, Lectures on polytopes, Springer-Verlag, New York, 1995.

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

Longsong Jia, [email protected]
School of Mathematical Sciences, Peking University, Beijing, 100871, P.R. China

Faze Zhang, [email protected]
School of Mathematics and Statistics, Northeast Normal University, Changchun, Jilin, 130024, P.R.China