Number of Triangulations of a Möbius Strip
Abstract.
Consider a Möbius strip with chosen points on its edge. A triangulation is a maximal collection of arcs among these points and cuts the strip into triangles. In this paper, we proved the number of all triangulations that one can obtain from a Möbius strip with chosen points on its edge is given by , then we made the connection with the number of clusters in the quasi-cluster algebra arising from the Möbius strip.
1. Introduction
We consider the Möbius strip as a marked surface, that is, a two-dimensional Riemann surface with a set of marked points on the boundary of . We cut this surface into triangles with arcs, which are curves on the marked surface. A maximal collections of arcs without intersections is called a triangulation. Such an example of triangulation is given at Figure 1.

We initiate a study about the number of triangulations of the Möbius strip, the only non-oriented surface with a finite number of triangulations.
In the theory of quasi-cluster algebras developed by Dupont and Palesi [DP15], the triangulations of a Möbius strip correspond naturally to the clusters in the quasi-cluster algebra from a Möbius strip, which is explained in section 4.
Our main result is the following theorem:
Theorem 1.1.
The number of triangulations of a Möbius strip with marked points is given by
To prove this result, we count the number of triangulations in an iterative way. Then we simplify the equation into a more succinct form.
2. Preliminaries
In this section, we start by reviewing some notions about non-orientable surfaces, and aim to define the triangulations of these surfaces. The review combines results from [FST08] and [DP15].
Let be a compact 2-dimensional manifold with boundary and let be a finite set of marked points of such that each boundary component contains at least one marked point. We denote the surface . In the scope of this paper, we only consider the case when all the points of are on the boundary. Moreover, to exclude pathological cases, we do not allow to be a monogon or digon.
Intuitively, a surface is orientable if moving continuously any figure on this surface cannot result in the mirror image of the figure when it is back to its starting point. Else, it is non-orientable.
Definition 2.1.
Let be a surface. A closed curve on is defined as two-sided if it has a regular, orientable neighbourhood. If any neighbourhood is non-orientable, the curve is one-sided.
A surface is non-orientable if and only if there is a one-sided curve on it. A Möbius strip is such an example. This will be an important notion in the proof of Proposition 3.3.
Definition 2.2.
A cross-cap is obtained from an annulus by identifying antipodal points on the inner boundary. This is represented (as in Figure 2) by drawing a cross inside the inner circle.

Definition 2.3.
The isotopy class of a curve is the set of embeddings of that can be deformed into each other through a continuous path of homomorphisms such that:
-
•
the endpoints of each embedding are always the same;
-
•
except for its endpoints, each embedding is disjoint from and the boundary.
We define two types of curves without self-intersections in . The first type of curves without self-intersections consists of closed one-sided curves in the interior of . The second type consists of curves with both endpoints in . We refer to the union of isotopy classes of these two types as arcs.
Two arcs are compatible if there exist representatives in their isotopy class not intersecting with each other.

Example 2.4.
In figure 4, even if they intersect each other, the curves and in (a) are representatives of compatible arcs, because and in (b) are representatives of the same respective isotopy classes and they do not intersect.

Note that what we refer to as arcs, Dupont and Palesi refer to as quasi-arcs, [DP15, Definition 1].
Let denote the set of boundary segments, that is, the set of connected components of the boundary of with the points of removed.
With the above definitions, we are now able to introduce triangulations of a surface.
Definition 2.5.
A triangulation of is a maximal collection of compatible arcs. Each area of cut by this maximal collection of compatible arcs is a triangle.
Example 2.6.

Remark that the definition implies that if another arc can be added into the current collection of arcs, then this is not a triangulation.
Once again, our terminology does not totally coincide with that of Dupont and Palesi. What we call a triangulation, they would call a quasi-triangulation, [DP15, Defintion 6].
Let denote the Möbius strip with marked points on its boundary.
Example 2.7.
Figure 6 shows a triangulation of .
Remark 2.8.
In some cases, the triangles formed as a result of triangulation can have two edges identified; these triangles are called anti-self-folded triangles. The identified edges must go through the cross-cap and all the three vertices are also identified.
Example 2.9.
There is an anti-self-folded triangle in the Figure 6 formed by the arcs and . The three edges are represented with different colors, with the yellow one and the green one identified.

Definition 2.10.
A quasi-triangle consists in a monogon encircling only a cross-cap and a one-sided curve through the cross-cap in the interior of , as in Figure 7.

Proposition 2.11.
[DP15, Proposition 7] A triangulation cuts into a finite union of triangles and quasi-triangles.
Proposition 2.12.
[DP15, Example 13] The number of arcs in a triangulation of is equal to the number of marked points.
Theorem 2.13.
[DP15, Theorem 45] Let be a surface with finite number of triangulations and marked points on the boundary of . If is orientable, it is an -gon (); if is non-orientable, it is a Möbius strip.
3. Number of Triangulations of A Möbius Strip
Recall that the Catalan numbers are defined recursively as:
(3.1) |
and is defined explicitly by
(3.2) |
Lemma 3.1.
The number of triangulations of a -gon is given by , [Lam38].
We provide this simple proof because we will use a similar approach to the Möbius strip case.
Proof.
We prove by induction that is the number of triangulations of a -gon.
Label the vertices of the -gon from to . For each triangulation of the -gon, there is a single triangle containing the edge between the vertices and . Denote the third vertex of this triangle by . By the induction hypothesis, there are triangulations of the -gon on one side of this triangle and triangulations of the -gon on the other side of this triangle, see Figure 8. Here . Therefore, the number of triangulations of a -gon is
∎

Example 3.2.
The pentagon has 5 triangulations shown in the Figure 9.

Proposition 3.3.
The number of triangulations of a Möbius strip with marked points is given by , where
Proof.
Let be a marked surface and let be a triangulation of .
Recall that in any triangulation, each boundary edge is part of one and only one triangle. Therefore, there is a single triangle containing the edge between the marked points and ; denote this triangle . It must be of one of the following three forms shown in Figure 10.

Indeed, a triangle cannot contain a cross-cap: a surface with one boundary component, three marked points, and a cross-cap is a Möbius strip with three marked points instead of a triangle.
Moreover, a triangle cannot have an edge going through a cross-cap. On the contrary, suppose that a triangle has such an edge. Then, since we can add another arc in , as depicted by the dotted arc in Figure 11, was not part of maximal collection of compatible arcs, i.e a triangulation, so was not a triangle.

Therefore, we proceed with our proof in three cases. Since Type 1 and Type 2 are mirror images, the number of triangulations occurring in each type are the same, and we discuss them in one case.
Case 1: Consider the triangle of the form Type 1 or Type 2. Without loss of generality, we consider the case of Type 2.
We count the number of triangulations by considering the three areas cut out by the two arcs, see Figure 12.

Area 1 includes marked points numbered from 1 to . Since Area 1 is orientable, it is isotopic to an -gon. Thus by the Lemma 3.1, Area 1 has triangulations.
There is no more compatible arc in Area 2; alternatively, it is a triangle. Therefore, there is only one triangulation in this area.
Area 3 is a Möbius strip with marked points. Hence, there are triangulations in Area 3.
Thus, the number of triangulations in this case is the product of triangulations coming from these three areas. The number of triangulations containing a triangle of Type 2 is
Thus, the total number of triangulations containing triangle of Type 1 and 2 in Case 1 is
Case 2: In this case, we consider of Type 3, in which both arcs go through the cross-cap. The Möbius Strip is thus divided into two areas, see Figure 13.

Area 1 is a triangle, and is thus already a triangulation and no other arcs can be added.
Consider the number of triangulations in Area 2.
We first notice that the Area 2 is isotopic to an -gon, because:
-
•
edges between vertices and ,
-
•
one edge between vertices and going through the cross-cap,
-
•
edges between vertices and ,
-
•
one edge between vertices and going through the cross-cap.
This gives us a total of edges; since the interior of Area 2 contains no cross-cap, it is an -gon. Therefore, it has triangulations.
Since in this case, , the total number of triangulations can be found by
(3.3) |
Therefore, taking all three cases into consideration, the total number of triangulations on a Möbius strip with marked points is the sum:
∎
We now want to simplify this formula and obtain a direct formula for the number of triangulations of a Möbius strip.
Proposition 3.4.
The number of triangulations of a Möbius strip with marked points is
Proof.
We proceed by induction to show that
(3.4) |
Clearly, when , we have
which verifies Equation 3.4. Now, suppose Equation 3.4 is true for all for some positive integer and prove that it is also verified for . Using both our induction hypothesis and Proposition 3.3, for , we know that
By Equation 3.1, we obtain, for ,
Using this equation, we compute:
Therefore,
Since we have now proven equation 3.4, we have now shown that
4. Connection to Quasi-Cluster Algebras
In this section, we apply our previous results to quasi-cluster algebras, which are generalizations of cluster algebras.
Introduced in 2002 by Fomin and Zelevinsky [FZ02, FZ03, FZ07], cluster algebras create an algebraic framework for canonical bases and total positivity. They related to various areas of mathematics such as combinatorics, representation theory of algebras [BMRT07], mathematical physics [AHBHY18], Poisson geometry [GSV10], Lie theory [GLS13], and, of course, triangulation of surfaces [FST08], just to name a few. Indeed, in 2008, Fomin, Shapiro and Thurston introduced a particular class of cluster algebras, called cluster algebras from surfaces, [FST08]. Later work from Dupont and Palesi introduced the quasi-cluster algebras from non-orientable surfaces [DP15]. Then, Wilson studied the shellability and sphericity of the arc complex [Wil18], and proved the positivity theorem for quasi-cluster algebras [Wil19].
In this section, we will not define cluster algebras, but rather quasi-cluster algebras, which generalize them.
Quasi-cluster algebras are commutative -algebras of Laurent polynomials with a distinguished set of generators, called cluster variables, grouped into sets called clusters. The set of all cluster variables is constructed recursively from a set of initial cluster variables using an involutive operation called mutation. Let’s define these concepts formally.
The difference between classical cluster algebras from surfaces, as defined by Fomin and Zelevinsky [FZ02] and Fomin, Shapiro Thurston, [FST08] and quasi-cluster algebras, as defined by Dupont and Palesi, [DP15], is that cluster algebras arise only from orientable surfaces, while quasi-cluster algebras arise from both orientable and non-orientable surfaces.
Let denote the field of rational functions in indeterminates, that is
Definition 4.1.
Let be a field and a sub-field of . A set is algebraically independent on if is not a root for any non-zero polynomial on .
Example 4.2.
The singletons and are algebraically independent on , but the set is not. Indeed, , so the non-zero polynomial is null in .
To each boundary segments , we associate an indeterminate such that is algebraically independent. The elements of are called the coefficients. The ground ring is set as
Definition 4.3.
An initial seed is a triplet such that:
-
•
is a triangulation of ;
-
•
is a cluster and together with , it forms a generating set of .
In other words, a seed is triangulation with a variable associated to each arc.
From a given seed, we can obtain a new one by a process called mutation. First, let’s define the mutation only on triangulations.
Proposition 4.4.
[DP15, Proposition 9] Let be a triangulation of a marked surface and let be an arc of . There exists a unique arc such that is still a triangulation of .
Example 4.5.
Figure 14 depicts the mutation in a Möbius strip with 3 marked points , where each line indicates a mutation.

Definition 4.6.
Let be an initial seed and . The mutation of in the direction transforms in a seed where with such that:
-
(1)
If the arc is the diagonal in a rectangle that creating two triangles as shown in Figure 15, then the relation is the following:
(4.1) This is called the Ptolemy relation because it is analogous to the relation that holds among the lengths of the sides and diagonals of a cyclic quadrilateral by Ptolemy’s theorem.
Figure 15. Mutation of type 1 -
(2)
Consider an anti-self-folded-triangle, as in Figure 16. Let be the arc corresponding to the two identified edges. Then, the relation is
(4.2) Figure 16. Mutation of type 2 -
(3)
If a one-sided closed curve is in the triangulation with boundary as in Figure 17, the relation is
(4.3) Figure 17. Mutation of type 3 -
(4)
Let be the diagonal in a rectangle with sides , as Figure 18 shows below, then the relation is
(4.4) Figure 18. Mutation of type 4
In general, a seed is obtained by a sequence of mutations from the initial seed.
Example 4.7.
Figure 19 shows a mutation of a seed in a Möbius strip with 3 marked points . is the new cluster variable.

Definition 4.8.
Let be the union of all clusters obtained by successive mutation from an initial seed . The quasi-cluster algebra is
That is to say, every element of which can be expressed as a polynomial in the elements of with coefficients in .
Example 4.9.
In figure 20, cluster variables from the quasi-cluster algebra from a Möbius Strip with 2 marked points () are calculated. In the figure, and are derived from the Ptolemy relation 4.1:
is mutated from an anti-self-folded triangle according to equation 4.3, and can be calculated as
is calculated according to formula 4.4, which yields
The cluster algebra consists of all polynomials in and with coefficients in .

Proposition 4.10.
There is a bijection between the set of triangulations of a marked surface and the set of seeds of the quasi-cluster algebra arising from this surface.
Proof.
Clearly, to any seed , we can associate the triangulation .
Moreover, to any arc, we can associate a unique cluster variable, [DP15]. Wilson also gave a formula to express the cluster variable associated to a given arc in term of the cluster variables in any cluster, [Wil19]. Therefore, to any triangulation , we can associate a unique cluster .
∎
Corollary 4.11.
There are seeds in a quasi-cluster algebra arising from a Möbius strip. Any other quasi-cluster algebra arising from a non-orientable surface has infinitely many seeds.
Acknowledgements
We would like to thank Jon Wilson for fruitful discussions. We would also like to thank Hugh Thomas for providing editorial advice of this paper and helping us. Finally, we would like to thank Canada/USA Mathcamp for giving us the opportunity to meet together and spark discussions on this topic.
References
- [AHBHY18] Nima Arkani-Hamed, Yuntao Bai, Song He, and Gongwang Yan. Scattering forms and the positive geometry of kinematics, color and the worldsheet. Journal of High Energy Physics, 2018(5), May 2018.
- [BMRT07] Aslak B. Buan, Robert J. Marsh, Idun Reiten, and Gordana Todorov. Clusters and seeds in acyclic cluster algebras. Proceedings of the American Mathematical Society, 135(10):3049–3060, 2007.
- [DP15] Grégoire Dupont and Frédéric Palesi. Quasi-cluster algebras from non-orientable surfaces. Journal of Algebraic Combinatorics, 42(2):429–472, Mar 2015.
- [FST08] Sergey Fomin, Michael Shapiro, and Dylan Thurston. Cluster algebras and triangulated surfaces. part I: Cluster complexes. Acta Mathematica, 201(1):83–146, 2008.
- [FZ02] Sergey Fomin and Andrei Zelevinsky. Cluster algebras I: Foundations. Journal of the American Mathematical Society, 15(2):497–529, 2002.
- [FZ03] Sergey Fomin and Andrei Zelevinsky. Cluster algebras II: Finite type classification. Inventiones mathematicae, 154(1):63–121, 2003.
- [FZ07] Sergey Fomin and Andrei Zelevinsky. Cluster algebras IV: Coefficients. Compositio Mathematica, 143(1):112–164, 2007.
- [GLS13] Ch. Geiss, B. Leclerc, and J. Schröer. Cluster algebras in algebraic Lie theory. Transform. Groups, 18(1):149–178, 2013.
- [GSV10] Michael Gekhtman, Michael Shapiro, and Alek Vainshtein. Cluster algebras and Poisson geometry, volume 167 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2010.
- [Lam38] M. Lamé. Extrait d’une lettre de M. Lamé à M. Liouville sur cette question: Un polygone convexe étant donné, de combien de manières peuton le partager en triangles au moyen de diagonales? Journal des mathématiques pures et appliquées, 3:505–507, 1938.
- [Wil18] Jon Wilson. Shellability and sphericity of finite quasi-arc complexes. Discrete Comput Geom, 59(3):680–706, 2018.
- [Wil19] Jon Wilson. Positivity for quasi-cluster algebras, 2019. arXiv:1912.12789.