A discrete uniformization theorem for decorated piecewise Euclidean metrics on surfaces
Abstract.
In this paper, we introduce a new discretization of the Gaussian curvature on surfaces, which is defined as the quotient of the angle defect and the area of some dual cell of a weighted triangulation at the conic singularity. A discrete uniformization theorem for this discrete Gaussian curvature is established on surfaces with non-positive Euler number. The main tools are Bobenko-Lutz’s discrete conformal theory for decorated piecewise Euclidean metrics on surfaces and variational principles with constraints.
Key words and phrases:
Discrete uniformization theorem; Decorated piecewise Euclidean metrics; Discrete Gaussian curvature; Variational principle1. Introduction
Bobenko-Lutz [2] recently introduced the decorated piecewise Euclidean metrics on surfaces. Suppose is a connected closed surface and is a finite non-empty subset of , we call a marked surface. A piecewise Euclidean metric (PE metric) on the marked surface is a flat cone metric with the conic singularities contained in . A decoration on a PE surface is a choice of circle of radius at each point . These circles in the decoration are called vertex-circles. We denote a decorated PE surface by and call the pair a decorated PE metric. In this paper, we focus on the case that for all and each pair of vertex-circles is separated.
For a PE surface , denote as the cone angle at . The angle defect
(1) |
is used to describe the conic singularities of the PE metric. Let be a triangulation of , where are the sets of vertices, edges and faces respectively. The triangulation for a PE surface is a geodesic triangulation if the edges are geodesics in the PE metric . We use one index to denote a vertex (such as ), two indices to denote an edge (such as ) and three indices to denote a face (such as ) in the triangulation . For any decorated geodesic triangle , there is a unique circle simultaneously orthogonal to the three vertex-circles at the vertices [10]. We call this circle as the face-circle of the decorated geodesic triangle and denote its center by and radius by . The center of the face-circle of the decorated geodesic triangle is the geometric center introduced by Glickenstein [9] and Glickenstein-Thomas [11] for general discrete conformal structures on surfaces. Denote as the interior intersection angle of the face-circle and the edge . Please refer to Figure 1 (left) for the angle . The edge , shared by two adjacent decorated triangles and , is called weighted Delaunay if
(2) |
The triangulation is called weighted Delaunay in the decorated PE metric if every edge in the triangulation is weighted Delaunay. Connecting the center with the vertices by geodesics produces a cellular decomposition of the decorated triangle . Denote as the sum of the signed area of the two triangles adjacent to in the cellular decomposition of the decorated triangle . The area of the triangle with the vertices , , is positive if it is on the same side of the edge as the decorated triangle , otherwise it is negative (or zero if lies in ). Please refer to the shaded domain in Figure 1 (left) for . Gluing these cells of all decorated triangles isometrically along edges in pairs leads a cellular decomposition of the decorated PE surface . Set
Please refer to Figure 1 (right) for .
Definition 1.1.
Suppose is a decorated PE surface and is a weighted Delaunay triangulation of . The discrete Gaussian curvature at the vertex is the quotient of the angle defect and the area of the dual cell at the vertex , i.e.,
(3) |
Remark 1.2.
In the literature, the discrete curvature is usually defined by the angle defect in (1). However, the angle defect is scaling invariant and does not approximate the smooth Gaussian curvature pointwisely on smooth surfaces as the triangulations of the surface become finer and finer. This is supported by the discussions in [3, 7]. For the discrete Gaussian curvature in (3), it scales by a factor upon a global rescaling of the decorated PE metric by a factor . This property is paralleling to that of the smooth Gaussian curvature on surfaces. On the other hand, the definition of the discrete Gaussian curvature coincides with the original definition of the Gaussian curvature on smooth surfaces. This implies that the discrete Gaussian curvature is a good candidate as a discretization of the smooth Gaussian curvature on surfaces.
Remark 1.3.
According to Definition 1.1, the discrete Gaussian curvature defined by (3) seems to depend on the choice of weighted Delaunay triangulations of the decorated PE surface . We will show that is an intrinsic geometric invariant of the decorated PE surface in the sense that it is independent of the weighted Delaunay triangulations of . Note that the angle defect defined by (1) is an intrinsic geometric invariant of a decorated PE surface, we just need to prove that is independent of the choice of weighted Delaunay triangulations. This is true by Lemma 2.8.
Remark 1.4.
The weighted Delaunay triangulation is a natural generalization of the classical Delaunay triangulation. When the weighted Delaunay triangulation is reduced to the classical Delaunay triangulation, i.e. for all , the area is exactly twice the area of the Voronoi cell at the vertex . Thus the area is a generalization of the area of the Voronoi cell at the vertex . As a result, the discrete Gaussian curvature in Definition 1.1 generalizes Kouřimská’s definition of discrete Gaussian curvature in [17, 16].
The discrete Yamabe problem for a decorated PE metric on asks if there exists a discrete conformal equivalent decorated PE metric on with constant discrete Gaussian curvature. The following discrete uniformization theorem solves this problem affirmatively for the discrete Gaussian curvature in Definition 1.1.
Theorem 1.5.
For any decorated PE metric on a marked surface with Euler number , there is a discrete conformal equivalent decorated PE metric with constant discrete Gaussian curvature .
By the relationships of the discrete Gaussian curvature and the classical discrete Gaussian curvature , the case in Theorem 1.5 is covered by Bobenko-Lutz’s work [2]. Therefore, we just need to prove the case in Theorem 1.5.
Remark 1.6.
The discrete Yamabe problem on surfaces for different types of discrete conformal structures with respect to the classical discrete Gaussian curvature has been extensively studied in the literature. For Thurston’s circle packings on surfaces, the solution of discrete Yamabe problem gives rise to the famous Koebe-Andreev-Thurston Theorem. See also the work of Beardon-Stephenson [1] for the discrete uniformization theorems for circle packings on surfaces. For the vertex scalings introduced by Luo [18] on surfaces, Gu-Luo-Sun-Wu [13], Gu-Guo-Luo-Sun-Wu [12], Springborn [23] and Izmestiev-Prosanov-Wu [15] give nice answers to this problem in different background geometries. Recently, Bobenko-Lutz [2] established the discrete conformal theory for decorated PE metrics and prove the corresponding discrete uniformization theorem. Since Bobenko-Lutz’s discrete conformal theory of decorated PE metrics also applies to the Euclidean vertex scalings and thus generalizes Gu-Luo-Sun-Wu’s result [13] and Springborn’s result [23], Theorem 1.5 also generalizes Kouřimská’s results in [17, 16]. It should be mentioned that Kouřimská [17, 16] constructed counterexamples to the uniqueness of PE metrics with constant discrete Gaussian curvatures. We conjecture that the decorated PE metric with constant discrete Gaussian curvature in Theorem 1.5 is not unique.
The main tools for the proof of Theorem 1.5 are Bobenko-Lutz’s discrete conformal theory for decorated PE metrics on surfaces [2] and variational principles with constraints. The main ideas of the paper come from reading of Bobenko-Lutz [2] and Kouřimská [17, 16].
The paper is organized as follows.
In Section 2,
we briefly recall Bobenko-Lutz’s discrete conformal theory for decorated PE metrics on surfaces.
Then we show that is independent of the choice of weighted Delaunay triangulations, i.e., Lemma 2.8.
We also give some notations and a variational characterization of the area .
In this section, we also extend the energy function and the area function .
In Section 3,
we translate Theorem 1.5 into an optimization problem with constraints, i.e., Lemma 3.2.
Using the classical result from calculus, i.e., Theorem 3.3, we translate Lemma 3.2 into Theorem 3.4.
By analysing the limit behaviour of sequences of discrete conformal factors, we get an asymptotic expression of the function , i.e., Lemma 3.12.
In the end, we prove Theorem 3.4.
Acknowledgements
The first author thanks Professor Feng Luo for his invitation to the workshop
“Discrete and Computational Geometry, Shape Analysis, and Applications” taking place
at Rutgers University, New Brunswick from May 19th to May 21st, 2023.
The first author also thanks Carl O. R. Lutz for helpful communications during the workshop.
2. Preliminaries on decorated PE surfaces
2.1. Discrete conformal equivalence and Bobenko-Lutz’s discrete conformal theory
In this subsection, we briefly recall Bobenko-Lutz’s discrete conformal theory for decorated PE metrics on surfaces. Please refer to Bobenko-Lutz’s original work [2] for more details on this. The PE metric on a PE surface with a geodesic triangulation defines a length map such that satisfy the triangle inequalities for any triangle . Conversely, given a function satisfying the triangle inequalities for any face , one can construct a PE metric on a triangulated surface by isometrically gluing Euclidean triangles along edges in pairs. Therefore, we use to denote a PE metric and use to denote a decorated PE metric on a triangulated surface .
Definition 2.1 ([2], Proposition 2.2).
Let be a triangulation of a marked surface . Two decorated PE metrics and on are discrete conformal equivalent if and only if there exists a discrete conformal factor such that
(4) |
(5) |
for all .
Remark 2.2.
Note that the inversive distance
(6) |
between two vertex-circles is invariant under Möbius transformations [6]. Combining (4) and (5) gives . Since each pair of vertex-circles is required to be separated, we have . Therefore, Definition 2.1 can be regarded as a special case of the inversive distance circle packings introduced by Bowers-Stephenson [4]. One can refer to [5, 14, 19, 24, 25] for more properties of the inversive distance circle packings on triangulated surfaces.
In general, the existence of decorated PE metrics with constant discrete Gaussian curvatures on triangulated surfaces can not be guaranteed if the triangulation is fixed. In the following, we work with a generalization of the discrete conformal equivalence in Definition 2.1, introduced by Bobenko-Lutz [2], which allows the triangulation of the marked surface to be changed under the weighted Delaunay condition.
Definition 2.3 ([2], Definition 4.11).
Two decorated PE metrics and on the marked surface are discrete conformal equivalent if there is a sequence of triangulated decorated PE surfaces such that
- (i):
-
the decorated PE metric of is and the decorated PE metric of is ,
- (ii):
-
each is a weighted Delaunay triangulation of the decorated PE surface ,
- (iii):
- (iv):
-
if , then and are two different weighted Delaunay triangulations of the same decorated PE surface.
Definition 2.3 defines an equivalence relationship for decorated PE metrics on a marked surface. The equivalence class of a decorated PE metric on is also called as the discrete conformal class of and denoted by .
Lemma 2.4 ([2]).
The discrete conformal class of a decorated PE metric on the marked surface is parameterized by .
For simplicity, for any , we denote it by for some . Set
For any decorated PE surface, there exists a unique complete hyperbolic surface , i.e., the hyperbolic surface induced by any triangular refinement of its unique weighted Delaunay tessellation. It is homeomorphic to and called as the fundamental discrete conformal invariant of the decorated PE metric . The decoration of is denoted by and here the height is related to by . The canonical weighted Delaunay tessellation of is denoted by . Bobenko-Lutz [2] defined the following set
and proved the following proposition.
Proposition 2.5 ([2], Proposition 4.3).
Given a complete hyperbolic surface with ends .
- (1):
-
Each is either empty or the intersection of with a closed polyhedral cone.
- (2):
-
There is only a finite number of geodesic tessellations of such that is non-empty. In particular, .
Let be the polyhedral cusp corresponding to the triangulated surface with fundamental discrete conformal invariant . The polyhedral cusp is convex if and only if is a weighted Delaunay triangulation. The set of all heights of convex polyhedral cusps over the triangulated hyperbolic surface is denoted by .
Proposition 2.6 ([2], Proposition 4.9).
Given a decorated PE metric on the marked surface . Then , and are homeomorphic.
Lemma 2.7 ([2]).
The set
is a finite set, , and each is homeomorphic to a polyhedral cone (with its apex removed) and its interior is homeomorphic to .
2.2. A decorated triangle
Denote as half of the distance of the two intersection points of the face-circle and the edge . Denote as the signed distance of the center to the edge , which is defined to be positive if the center is on the same side of the line determined by as the triangle and negative otherwise (or zero if the center is on the line). Note that is symmetric in the indices and . By Figure 3, we have
(7) |
Since and , if , then . The equality (7) implies that (2) is equivalent to
(8) |
for any adjacent triangles and sharing a common edge . Therefore, the equality (8) also characterizes a weighted Delaunay triangulation for a decorated PE metric on . Due to this fact, the equality (8) is usually used to define the weighted Delaunay triangulations of decorated PE surfaces. See [5, 8] and others for example. Then can be written as
(9) |
Since are the signed distances, thus is an algebraic sum of the area of triangles, i.e. a signed area.
Lemma 2.8.
The area is independent of the choice of weighted Delaunay triangulations of a decorated PE surface.
Proof.
Suppose a decorated quadrilateral is in a face of the weighted Delaunay tessellation of a decorated PE surface, then there exist two weighted Delaunay triangulations and of the decorated PE surface with an edge in flipped to another edge in . Please refer to Figure 2. We just need to prove the signed area in is equal to the signed area in . In , the signed area at the vertex in is . In , the signed area at the vertex in is
Since and are two weighted Delaunay triangulations of the same decorated PE metric on , then by (8). One can also refer to [2] (Proposition 3.4) for this. Moreover, and . Then .
Q.E.D.
Denote as the center of the edge , which is obtained by projecting the center to the line determined by . Denote as the signed distance of to the vertex , which is positive if is on the same side of as along the line determined by and negative otherwise (or zero if is the same as ). In general, . Since the face-circle is orthogonal to the vertex-circle at the vertex , we have
(10) |
Please refer to Figure 3 for this. Moreover, we have the following explicit expressions of and due to Glickenstein [9], i.e.,
(11) |
and
(12) |
where is the inner angle of the triangle at the vertex . The equality (11) implies that is independent of the existence of the center . Since each pair of vertex-circles is required to be separated, then . This implies
The following lemma gives some useful formulas.
Lemma 2.9 ([14, 24, 25]).
Let be a decorated triangle with the edge lengths defined by (5). If the decorated triangle is non-degenerate, then
(13) |
where
(14) |
with , , and
(15) |
As a direct application of Lemma 2.9, we have the following result.
Lemma 2.10.
The area of each decorated triangle is an analytic function with
(16) |
2.3. The extended energy function and the extended area function
There exists a geometric relationship between the decorated triangle and the geometry of hyperbolic polyhedra in -dimensional hyperbolic space. Specially, there is a generalized hyperbolic tetrahedra in with one ideal vertex and three hyper-ideal vertices corresponding to a decorated triangle . Please refer to [2] for more details on this fact. Springborn [22] found the following explicit formula for the truncated volume of this generalized hyperbolic tetrahedra
(19) | ||||
where
(20) |
is Milnor’s Lobachevsky function. Milnor’s Lobachevsky function is bounded, odd, -periodic and smooth except at integer multiples of . Please refer to [20, 21] for more information on Milnor’s Lobachevsky function .
Set
(21) | ||||
where . Then and
(22) |
for . Furthermore, on a decorated PE surface with a weighted Delaunay triangulation , Bobenko-Lutz [2] defined the following function
(23) |
where is the convex polyhedral cusp defined by the heights , and . Note that the function defined by (23) differs from its original definition in [2] (Equation 4-9) by some constant. By (22), for , we have
Using the function , we define the following energy function
which is well-defined on with . Moreover, for , we have
(24) | ||||
where is used in the last line.
Theorem 2.12 ([2], Proposition 4.13).
For a discrete conformal factor , let be a weighted Delaunay triangulation of the decorated PE surface . The map
(25) | ||||
is well-defined, concave, and twice continuously differentiable over .
Therefore, the function defined on can be extended to be
(26) |
defined on .
Definition 2.13.
Suppose is a triangulated surface with a decorated PE metric . The area function on is defined to be
By Lemma 2.10, we have the following result.
Corollary 2.14.
The function is an analytic function with
(27) |
Lemma 2.8 and Corollary 2.14 imply the following result, which shows the function defined on can be extended.
Theorem 2.15.
For a discrete conformal factor , let be a weighted Delaunay triangulation of the decorated PE surface . The map
(28) | ||||
is well-defined and once differentiable.
3. The proof of Theorem 1.5
3.1. Variational principles with constraints
In this subsection, we translate Theorem 1.5 into an optimization problem with inequality constraints by variational principles, which involves the function defined by (26).
Proposition 3.1.
The set
is an unbounded closed subset of .
Proof.
By Theorem 2.15, the set is a closed subset of . Since , thus . Then is equivalent to . This implies that the ray stays in the set . Hence the set is unbounded. Q.E.D.
According to Proposition 3.1, we have following result.
Lemma 3.2.
If and the function attains a minimum in the set , then the minimum value point of lies at the boundary of , i.e.,
Furthermore, there exists a decorated PE metric with constant discrete Gaussian curvature in the discrete conformal class.
Proof.
Suppose the function attains a minimum at . Taking , then by . By the proof of Proposition 3.1, . Hence, by the additive property of the function in (24), we have
This implies by .
Then and .
Therefore, the minimum value point of lies in the set .
The conclusion follows from the following claim.
Claim : Up to scaling, the decorated PE metrics with constant discrete Gaussian curvature in the discrete conformal class
are in one-to-one correspondence with the critical points of the function under the constraint .
We use the method of Lagrange multipliers to prove this claim. Set
where is a Lagrange multiplier. If is a critical point of the function under the constraint , then by (27) and the fact , we have
This implies
Since the anger defect defined by (1) satisfies the following discrete Gauss-Bonnet formula
we have
under the constraint . Therefore, the discrete Gaussian curvature
for any . Q.E.D.
3.2. Reduction to Theorem 3.4
By Lemma 3.2, we just need to prove that the function attains the minimum in the set . Recall the following classical result from calculus.
Theorem 3.3.
Let be a closed set and be a continuous function. If every unbounded sequence in has a subsequence such that , then attains a minimum in .
One can refer to [16] (Section 4.1) for a proof of Theorem 3.3. The majority of the conditions in Theorem 3.3 is satisfied, including the set is a closed subset of by Proposition 3.1 and the function is continuous by Theorem 2.12. To prove Theorem 1.5, we just need to prove the following theorem.
Theorem 3.4.
If and is an unbounded sequence in , then there exists a subsequence of such that .
3.3. Behaviour of sequences of discrete conformal factors
Let be an unbounded sequence in . Denote its coordinate sequence at by . Motivated by [17], we call the sequence with the following properties as a “good” sequence.
- (1):
-
It lies in one cell of ;
- (2):
-
There exists a vertex such that for all and ;
- (3):
-
Each coordinate sequence either converges, diverges properly to , or diverges properly to ;
- (4):
-
For any , the sequence either converges or diverges properly to .
By Lemma 2.7, it is obvious that every sequence of discrete conformal factors in possesses a “good” subsequence. Hence, the “good” sequence could be chosen without loss of generality.
In the following arguments, we use the following notations
(29) |
(30) |
(31) |
For a decorated triangle in , set
(32) |
Let be a coordinate sequence in . Then the edge lengths satisfy the triangle inequalities for all .
Lemma 3.5.
There exists no sequence in such that as ,
where and is a constant.
Proof.
Without loss of generality, we assume , and the sequence . The equality (30) implies , and the sequence . Here are constants. By (29), we have
Note that , then
Therefore, there exists such that , i.e., . This contradicts the triangle inequality . Q.E.D.
Combining Lemma 3.5 and the connectivity of the triangulation , we have the following result.
Corollary 3.6.
For a discrete conformal factor , let be a weighted Delaunay triangulation of the decorated PE surface . For any decorated triangle in , at least two of the three sequences , , converge.
To characterize the function in (21), we need the following lemmas.
Lemma 3.7.
Assume that the sequence diverges properly to and the sequences and converge. Then the sequence converges to zero. Furthermore, if the sequences and converge to non-zero constants, then
- (1):
-
the sequences , and converge;
- (2):
-
the sequences , and converge.
Proof.
By the assumption, we have , and , where are positive constants. The equality (29) implies
(33) |
where is a positive constant. By the cosine law, we have
This implies .
Suppose the sequences and converge to non-zero constants. Then
(34) |
for some constant .
(1): Since , then , and . By (15), we have
where are constants. Note that may be non-positive. The equalities (14) and (34) imply
Hence the sequences , and converge.
(2): The equality (11) implies
(35) |
By (10), we have
where is a constant. Note that . Hence,
Then the sequences , and converge. Q.E.D.
Lemma 3.8.
Assume that the sequence diverges properly to and the sequences and converge. If the sequence converge to zero, then
Proof.
Lemma 3.7 shows that , thus . Since , then . Then
(36) |
By the proof of Lemma 3.7, we have
where (36) is used and are positive constants. Similar to (35), we have
Here are positive constants. By (10), we have
This implies , and . Therefore, we have the following four cases
- :
-
and ;
- :
-
and ;
- :
-
and ;
- :
-
and .
For the case , since , and . This implies that the center of the face-circle lies in the interior of the triangle by the definition of . However, in this case, are bounded. This is a contradiction.
Both the cases and imply . This contradicts with the fact that for any . Indeed, the center lies in the red region in Figure 4 in the case and lies in the blue region in Figure 4 in the case . By projecting the center to the line determined by , we have . This completes the proof. Q.E.D.
Remark 3.9.
Similar to the proof of Lemma 3.8, if the sequence converges to zero, then
Consider a star-shaped -sided polygon in the marked surface with boundary vertices ordered cyclically (). Please refer to Figure 5. Let be a vertex such that the sequence diverges properly to and the sequences converge for .
Lemma 3.10.
The sequences of inner angles at the boundary vertices in the triangles of a star-shaped polygon converge to non-zero constants.
Proof.
As and for . By Lemma 3.7, for any , we have and hence . We prove the result by contradiction. Without loss of generality, we assume and in the triangle . Then for large enough, we have
By Lemma 3.8, we have . Since the edge is weighted Delaunay, thus by (8), we have
This implies .
In the triangle , suppose the sequences and converge to non-zero constants. By Lemma 3.7, the sequences and converge. This contradicts . Hence the sequences or converge to zero. By Lemma 3.8 and Remark 3.9, we have , and . Then for large enough, we have
Please refer to Figure 5. By induction, for large enough, we have
This is a contradiction. Q.E.D.
Corollary 3.11.
Assume that the sequence diverges properly to and the sequences and converge. Then the sequence converges.
Proof.
By the definition of in (21), we have
Combining Lemma 3.7 and Lemma 3.10 gives that , the sequences and converge to non-zero constants and the sequences , and converge. Combining the continuity of Milnor’s Lobachevsky function defined by (20) and the definition of the truncated volume defined by (19), we have that the sequence converges. Note that keeps invariant. Hence,
for some constant . By (30), we have . Then
where the equalities (33) and (34) is used in the second line and is used in the third line. Therefore, . Q.E.D.
The following lemma gives an asymptotic expression of the function .
Lemma 3.12.
There exists a convergent sequence such that the function satisfies
Proof.
By (26), we have
where , the equation (22) is used in the second line and is used in the third line. The sequence converges by Corollary 3.6 and Corollary 3.11. Q.E.D.
The following lemma gives the influence of the sequence on the area of a decorated triangle .
Lemma 3.13.
For a discrete conformal factor , let be a weighted Delaunay triangulation of the decorated PE surface . Assume the sequences , converge for in with edge lengths .
- (a):
-
If converges, there exists a convergent sequence of real numbers such that
(37) - (b):
-
If diverges to , there exists a convergent sequence of real numbers such that
(38)
3.4. Proof of Theorem 3.4
Let be an unbounded “good” sequence. Suppose and is an unbounded sequence in . Combining and Lemma 3.12, we just need to prove that . By the definition of “good” sequence, the sequence converges to a finite number or diverges properly to .
If converges to a finite number, then the sequence converges for all . Since the sequence lies in , the area of each triangle is bounded from above. This implies is bounded from above by (37). Then converges to a finite number or diverges properly to . Suppose converges to a finite number. Since converges for all , then are bounded for all , which implies is bounded. This contradicts the assumption that is unbounded. Therefore, the sequence diverges properly to .
If diverges properly to , then there exists at least one vertex such that the sequence diverges properly to . By Corollary 3.6, the sequences and converge for and . Since the area of each triangle is bounded from above, thus and by (38), where is a constant. Then . This implies diverges properly to . Q.E.D.
Remark 3.14.
For the case , Kouřimská [17, 16] gave the existence of PE metrics with constant discrete Gaussian curvatures. However, we can not get similar results. The main difference is that the edge length defined by (5) involves the square term of discrete conformal factors, such as , while the edge length defined by the vertex scalings only involves the mixed product of the first order terms, i.e., . Indeed, in this case, we can define the set , which is an unbounded closed subset of . Under the conditions that and the function attains a minimum in the set , Lemma 3.2 still holds. Using Theorem 3.3, we just need to prove Theorem 3.4 under the condition . However, we can not get a good asymptotic expression of the area . The asymptotic expression of the area in (38) involves , which is not enough for this case.
References
- [1] A. Beardon, K. Stephenson, The uniformization theorem for circle packings. Indiana Univ. Math. J. 39 (1990), no. 4, 1383-1425.
- [2] A. Bobenko, C. Lutz, Decorated discrete conformal maps and convex polyhedral cusps. arXiv:2305.10988v1[math.GT].
- [3] A. Bobenko, U. Pinkall, B. Springborn, Discrete conformal maps and ideal hyperbolic polyhedra. Geom. Topol. 19 (2015), no. 4, 2155-2215.
- [4] P. L. Bowers, K. Stephenson, Uniformizing dessins and Belyĭ maps via circle packing. Mem. Amer. Math. Soc. 170 (2004), no. 805.
- [5] Y. Chen, Y. Luo, X. Xu, S. Zhang, Bowers-Stephenson’s conjecture on the convergence of inversive distance circle packings to the Riemann mapping, arXiv:2211.07464 [math.MG].
- [6] H. S. M. Coxeter. Inversive distance. Annali di Matematica, 71(1):73-83, December 1966.
- [7] H. Ge, X. Xu, A combinatorial Yamabe problem on two and three dimensional manifolds, Calc. Var. Partial Differential Equations 60 (2021), no. 1, 20.
- [8] D. Glickenstein, A monotonicity property for weighted Delaunay triangulations. Discrete Comput. Geom. 38 (2007), no. 4, 651-664.
- [9] D. Glickenstein, Discrete conformal variations and scalar curvature on piecewise flat two and three dimensional manifolds, J. Differential Geom. 87 (2011), no. 2, 201-237.
- [10] D. Glickenstein, Geometric triangulations and discrete Laplacians on manifolds, arXiv:math/0508188 [math.MG].
- [11] D. Glickenstein, J. Thomas, Duality structures and discrete conformal variations of piecewise constant curvature surfaces, Adv. Math. 320 (2017), 250-278.
- [12] X. D. Gu, R. Guo, F. Luo, J. Sun, T. Wu, A discrete uniformization theorem for polyhedral surfaces II, J. Differential Geom. 109 (2018), no. 3, 431-466.
- [13] X. D. Gu, F. Luo, J. Sun, T. Wu, A discrete uniformization theorem for polyhedral surfaces, J. Differential Geom. 109 (2018), no. 2, 223-256.
- [14] R. Guo, Local rigidity of inversive distance circle packing, Trans. Amer. Math. Soc. 363 (2011) 4757-4776.
- [15] I. Izmestiev, R. Prosanov, T. Wu, Prescribed curvature problem for discrete conformality on convex spherical cone-metrics, arXiv:2303.11068 [math.MG].
- [16] H. Kouřimská, Polyhedral surfaces of constant curvature and discrete uniformization. PhD thesis, Technische Universität Berlin, 2020.
- [17] H. Kouřimská, Discrete Yamabe problem for polyhedral surfaces, Discrete Computational Geometry 70 (2023), 123-153.
- [18] F. Luo, Combinatorial Yamabe flows on surfaces, Commun. Contemp. Math. 6 (2004), no. 5, 765-780.
- [19] F. Luo, Rigidity of polyhedral surfaces, III, Geom. Topol. 15 (2011), 2299-2319.
- [20] J. Milnor, Hyperbolic geometry: The first 150 years, Bull. Amer. Math. Soc. 6 (1982) 9-24.
- [21] John G. Ratcliffe, Foundations of hyperbolic manifolds. Second edition. Graduate Texts in Mathematics, 149. Springer, New York, 2006. xii+779 pp. ISBN: 978-0387-33197-3; 0-387-33197-2.
- [22] B. Springborn, A variational principle for weighted Delaunay triangulations and hyperideal polyhedra. J. Differential Geom. 78 (2008), no. 2, 333-367.
- [23] B. Springborn, Ideal hyperbolic polyhedra and discrete uniformization. Discrete Comput. Geom. 64 (2020), no. 1, 63-108.
- [24] X. Xu, Rigidity of inversive distance circle packings revisited, Adv. Math. 332 (2018), 476-509.
- [25] X. Xu, A new proof of Bowers-Stephenson conjecture, Math. Res. Lett. 28 (2021), no. 4, 1283-1306.