Bounding the Chromatic Number via High Dimensional Embedding
Abstract: We present a necessary and sufficient condition for determining whether a hypergraph can be embedded in a high-dimensional space, and establish an upper bound for the chromatic number of such hypergraph, which can be viewed as the high-dimensional extension of the Heawood conjecture and Wagner’s Theorem, respectively. Based on them, for a graph of order , there exists accordingly an algorithm with a complexity of that can determine the upper bound for the chromatic number. Specifically, we prove: (i) The graph-closure of a -uniform-topological hypergraph is a closed--graph if and only if its minors include neither nor ; (ii) Let be a special triangulated -graph, be the vertex-hypergraph of , then is -choosable.
Keywords: embedding; coloring; chromatic number; hypergraph
1 Introduction
This paper aims to bounding the chromatic number via high dimensional embedding in quasilinear time. To achieve this goal, we generalize the concept of planar graph to a higher-dimensional form. A challenge arises: Unlike the plane, higher-dimensional Euclidean spaces are not intuitive. Most definitions and methods related to planar graphs cannot be directly extend to higher dimensions. To address this challenge, it is necessary to introduce concepts and theorems related to simplex, CW complexes, homotopy, and fundamental groups. Please refer to [1, 2, 9] for common used definitions and notations that are not specified in this work.
It is NP-complete to decide if a given graph admits a -coloring except for the cases . In particular, computing the chromatic number (denoted by ) is NP-hard [7]. There are many algorithms for the problem of calculating the chromatic number, such as greedy algorithms [4, 19], heuristic algorithms [11], parallel and distributed algorithms [16, 17], and graph embeddings [3, 8, 12, 13]. Previous researchers developed the theory of graph embeddings on two-dimensional surfaces and established upper bounds for the chromatic number of graphs on such surfaces, and we follow in this research line by further investigating the embedding problem in high-dimensional spaces.
1.1 Our results
We consider a -uniform hypergraph as a CW complex. If the fundamental group of a -uniform hypergraph is trivial, we refer to it as a special -graph. All related definitions are given in Section 2.
Wagner’s Theorem states that a finite graph is planar if and only if its minors include neither nor , and we derive its higher-dimensional form.
Theorem 1.1.
The graph-closure of a -uniform-topological hypergraph is a closed--graph if and only if its minors include neither nor .
The Heawood conjecture (or Ringel-Youngs Theorem) states that if is a graph which can be embedded in a surface of Euler characteristic , and the following theorem presents a kind of its higher-dimensional extension.
Theorem 1.2.
Let be a special triangulated -graph, be the vertex-hypergraph of , then is -choosable.
Theorem 1.1 provides conditions for determining whether a hypergraph can be embedded in -dimensional Euclidean space (denoted by ), and Theorem 1.2 gives an upper bound for the chromatic number of . Since a hypergraph can be regarded as a CW complex and a graph can be regarded as the -skeleton of a CW complex, we can use the aforementioned two theorems to estimate the upper bound for the chromatic number of a graph.
Let be a graph and be a -uniform-topological hypergraph, if is the -skeleton of the graph-closure of , then we can conclude from Theorem 1.1 that can be embedded in . Furthermore, we obtain an upper bound for the chromatic number of by Theorem 1.2. Since the chromatic number of is less than that of , we can obtain the following corollary. It should be noted that the method of transforming a graph into a CW complex is not unique, and there should exist an optimal embedding method that yields the better estimate, but we have not yet solved this problem, which remains one of our future work.
Corollary 1.
Let be a graph, if its minors include neither nor , then is -choosable.
Corollary 1 transforms the problem of determining the upper bound for the chromatic number into the problem of minor testing. Analogous to the embedding problem for two-dimensional closed surfaces [14, 12, 13], if there exists an algorithm that can determine whether a given minor exists in a graph, we can use binary search to obtain an upper bound for the chromatic number. Given a graph of order , Robertson and Seymour [15] proved that for every graph there exists a algorithm with a complexity of that decides whether contains as a minor. Later, Kawarabayashi and Reed [10] improved it to .
Since the complexity of binary search is after noting that the chromatic number of any graph does not exceed its order and the complexity of minor testing is [10], by multiplying the two complexities, we can naturally obtain an algorithm with a complexity of based on the binary search to compute the upper bound for the chromatic number of any graph.
Remark. Several remarks related to above results follow.
-
•
We observe that Corollary 1 establish a connection between graph minors and chromatic numbers, which is very similar to Hedwiger’s Conjecture [2]. In [18], Thomason proved that if a graph does not contain a -minor, then the chromatic number of is at most . Substituting , we obtain the following conclusion: If a graph does not contain a -minor, then the chromatic number of is at most . Now, substituting into Corollary 1, we know that if contains neither a -minor nor a -minor, then the chromatic number of is at most . Clearly, even without optimize the bound of Corollary 1, our conclusion is partially superior to previous results in low dimensions. Furthermore, if we can introduce the Discharging method into -graph to improve the conclusion of Corollary 1, it may help us find another approach to solving Hedwiger’s conjecture. Note that the bound in Theorem 1.2 is estimated only by the relationship between the number of vertices and edges; therefore, there remains a gap between Theorem 1.2 and the sharp upper bound for the chromatic number. By introducing mathematical tools from the field of planar graphs, such as Discharging, we can further refine the bound of Theorem 1.2. However, due to space limitations, we will not discuss this issue in detail here, and relevant conclusions will be presented in our future work. (We conjecture that if is a special triangulated -graph, then there exists a polynomial such that is -choosable.)
-
•
Taking planar graph as an example, from Wagner’s Theorem, we know that if its minors include neither nor , it can be embedded in the plane, thereby implying that the upper bound for the chromatic number of is . For any arbitrary two-dimensional closed surface, there is a similar conclusion. In 1979, Filotti et al. [6] presented a polynomial-time algorithm with a complexity of , capable of determining whether a graph can be embedded in an orientable surface. This algorithm was subsequently optimized, with the most efficient version provided by Mohar [12, 13] in 1996, achieving a complexity of .
-
•
Since there exists an algorithm with a complexity of that can determine whether a graph can be embedded in any given two-dimensional closed surface, and since the chromatic number of a graph does not exceed the number of its vertices , we can naturally obtain an algorithm with a complexity of based on binary search to compute an upper bound for the chromatic number of any graph. Unfortunately, for many classes of graphs, the upper bound for the chromatic number obtained by the aforementioned algorithm is not good enough, even for special graph classes such as bipartite graphs. For example, the bipartite graph cannot be embedded in a two-dimensional closed surface with Euler characteristic , which suggests that when using this algorithm to estimate the upper bound for the chromatic number for bipartite graphs, the result can be arbitrarily large. However, it is easy to construct a closed--graph such that is the -skeleton of it. Therefore, the upper bound for the chromatic number of is also an upper bound for the chromatic number of . According to Theorem 1.2, this bound is , which reduces the gap between the upper bound of the chromatic number and the actual chromatic number of from infinite to finite.
-
•
Over the past few decades, many researchers have attempted to study the embedding of graphs in higher-dimensional spaces. In 1973, Bothe defined the linkless embedding: A linkless embedding of an undirected graph is an embedding of the graph into three-dimensional Euclidean space in such a way that no two cycles of the graph are linked (the knotless embedding can also be defined in a similar manner). In 1995, Cohen et al. [5] proved that any finite graph can be embedded into three-dimensional Euclidean space. We focus on hypergraphs and use polytopes in higher-dimensional spaces to represent the hyperedges of hypergraphs, then extend the problem of graph embedding to higher dimensions.
1.2 Overview of the paper
Section 2 mainly talks about the definitions of hypergraphs that can be embedded in . In Section 3, we extend the concept of triangulation to higher-dimensional spaces. In Section 4, we prove Theorem 1.2. In Sections 5 and 6, we prepare the groundwork for proving Theorem 1.1, and the proof is laid out in Section 7.
2 -graph: Definition and property
By considering -dimensional simplices as hyperedges of a -uniform hypergraph, the concept of planar graphs can be extended to higher-dimensional spaces. We note that the line segments (-dimensional simplices) in a planar graph are allowed to undergo topological deformation. Therefore, we allow hyperedges embedded in higher-dimensional space to undergo topological deformation as well. To avoid confusion, we refer to these hyperedges as simplexoid, defined as follows.
Definition 1 (Simplexoid).
If is a simplex of dimension with vertex set , is a face of , is homeomorphic to , the images of under homeomorphism is , then we say is a simplexoid of dimension , and is the vertex set of . Furthermore, if the image of under homeomorphism is , then we say is a -sub-simplexoid of (or -face) if contains exactly vertices.
Definition 2 (-Uniform-Topological hypergraph).
Let be a -uniform hypergraph. If each hyperedge of is regarded as a simplexoid of dimension , we refer to such a hypergraph as a -uniform-topological hypergraph.
Definition 3 (General -graph).
Let be simplexoids of dimension which are embedded in . We say is a general -graph if the following conditions hold: If and are simplexoids in and , then is a sub-simplexoid of both and . (Each simplexoid can be regarded as a hyperedge.)
From Definitions 2 and 3, we know that if a -uniform-topological hypergraph can be embedded in , then it is a general -graph.
Definition 4 (Special -graph).
If the fundamental group of a general -graph is trivial, then this graph is called a special -graph.
Note that in the case where the fundamental group is trivial, there will be no knot or other complex structures in the manifold. According to Definition 3 and 4, it is obvious that the special -graph is a special kind of the general -graph. Since the fundamental group of a special -graph is trivial, it can be regarded, up to isomorphism, as a union of internally disjoint -dimensional spheres and -dimensional balls. (An -sphere is a topological space that is homeomorphic to a standard -sphere. The space enclosed by a -sphere is called an -ball. )
Lemma 2.1 (The Construction of Special -graph).
Every special -graph can be constructed by the following procedure.
Procedure X: are simplexoids of dimension in . Starting from , add in one by one, and this procedure satisfies the following conditions:
-
•
1. Let and be arbitrary simplexoids in and , then is a sub-simplexoid of both and .
-
•
2. Suppose our procedure has reached the -th step ( has been added in ). Let , , be the sub-simplexoid with vertex set , be the sub-simplexoid set of of dimension . satisfies the following condition when adding to : If , then such that and .
Proof.
Let be a special -graph which is constructed by Procedure X. We prove that the fundamental group of is trivial. Since is obtained by adding simplexoid be one by one, this procedure does not result in any change of the fundamental group, thus the fundamental group of is the same as that of .
Next, let be a special -graph which is constructed by Definition 4, then it is evident that there exists a simplexoid in the special -graph such that the fundamental group remains trivial after removing from . By repeatedly removing simplexoids from , we eventually obtain an empty graph. Note that the fundamental group remains trivial in this process and the process is reversible, which implies that can be constructed by Procedure X. ∎
Definition 5 (Non-special Graph).
If the fundamental group of a -uniform-topological hypergraph is trivial and cannot embedded in , then is called a non-special -graph (or non--graph for short).
Note that the generalized Poincaré conjecture ensures that the -graph can always be embedded on the -sphere.
In high-dimensional spaces, some common definitions can be generalized as follows.
Definition 6 (Multiple Simplexoids).
Let and be -dimensional simplexoids of an -graph . If , and there exists an open set such that the boundary of contains both and , then and are called -dimensional multiple simplexoids.
It should be noted that the above definition requires two simplexoids to lie on the boundary of the same open set; otherwise, we do not consider they are multiple simplexoids.
Definition 7 (-Loops).
Let be an -dimensional simplexoid of an -graph. . If there exists () such that and overlap, then is called an -loop.
Since higher-dimensional simplexoids have more than two vertices, the definition of the loops in higher-dimensional manifolds differs slightly from that in planar graphs. As long as two vertices of a simplexoids overlap, we consider it as an -loop.
Definition 8 (Simple -Graph).
If -graph does not contain multiple simplexoid and -loop, then is called a simple -graph.
Unless otherwise specified, all -graphs mentioned hereafter will be simple -graphs. Analogous to the definition of incident and adjacent in graph theory, we can define the notions of incident and adjacent in -graphs.
Definition 9 (Neighbour).
Let be an -graph, and an -dimensional simplexoid of is denoted by , and the set of -dimensional simplexoid of is denoted as . For convenience, we use to denote the vertex set of , use to denote the vertex set of . Let , we say is adjacent to if such that . The set of all vertices that adjacent to point is denoted by , the degree of is denoted by .
Definition 10 (Incident and Adjacent).
Let be an -dimensional simplexoid, and be a -dimensional simplexoid (). We say is incident to (or is incident to ) if ; we say is adjacent to if and is an -dimensional simplexoid. The set of all -dimensional simplexoid incident (adjacent) to is denoted by . We say is the -dimensional degree of .
Definition 11 (Merging of Multiple Simplexoids).
Given two multiple simplexoids and , merging of and refers to combining and into a new simplexoid , and all simplexoids incident with or are incident with .
Definition 12 (Simplexoid Deletion).
Given an -graph , there are two natural ways of deriving smaller graphs from . If is a -dimensional simplexoid of , we may obtain a graph with -dimensional simplexoids by deleting from but leaving the vertices and the remaining simplexoids intact. The resulting graph is denoted by . Similarly, if is a vertex or an -dimensional simplexoid () of , we may obtain a graph by deleting from the vertex (or simplexoid) together with all the -dimensional simplexoids incident with . The resulting graph is denoted by or .
Definition 13 (Simplexoid Contraction).
To contract a simplexoid of an -graph is to delete the simplexoid and then identify its incident vertices. The resulting graph is denoted by . It is important to note that, during the process of simplexoid contraction, if multiple simplexoids emerge, we need to merge them to ensure that the resulting graph is a simple graph.
Definition 14 (-Embedding).
Let be a -uniform-topological hypergraph, an -embedding of can be regarded as a graph isomorphic to and is embedded in .
Finally, let us provide a brief summary. The -graph graph can be considered as an extension of the definition of the planar graph into higher-dimensional space or as a special type of CW complex. Whether it is a general -graph, a special -graph, or a non--graph, they are essentially special cases of CW complex.
A general -graph requires that this CW complex can be embedded in . A special -graph requires that this CW complex can be embedded in and has a trivial fundamental group. A non--graph requires that this CW complex cannot be embedded in and has a trivial fundamental group. A thorough understanding of these definitions lays a solid foundation for subsequent proofs.
3 Triangulation of special -graph ()
A simple connected plane graph in which all faces have degree three is called a plane triangulation or, for short, a triangulation. This section extends the concept of triangulation to higher-dimensional spaces. We will divide the triangulation process into two steps: Graph-closure and triangulation. In the following part, the term -graph refers to special -graph unless otherwise specified.
3.1 Step I: Graph-closure
First, it is necessary to have a clear understanding of the structure of -graphs. From Definition 4 and Lemma 3.1, an -graph is homeomorphic to the union of a finite collection of -dimensional balls and -spheres with trivial fundamental group.
Lemma 3.1.
[1] A path connected space whose fundamental group is trivial is simply connected.
The simplexoids in -graphs can be divided into two categories which are similar to the hanging edge and non-hanging edge in planar graphs. Next, we will extend the concept of hanging edges to higher-dimensional spaces.
Definition 15 (Hanging Simplexoid).
Let be a -dimensional simplexoid of an -graph , represent the set of all -dimensional simplexoids that is incident to . If , there exists a -dimensional simplexoid such that , then is called non-hanging simplexoid, if not is called hanging simplexoid.
Since hanging simplexoids can cause some difficulties in our subsequent research, we need to find a way to eliminate hanging simplexoids in -graphs.
Before defining the graph-closure, we first need to introduce the definition of hyper-polytopes. Similar to planar graphs, the following lemma implies that an -graph can be regarded as the boundary of the maximal connected open sets in set .
Lemma 3.2.
Let be a maximal connected open set of , be the closure of , be the boundary of , then .
Proof.
Suppose to the contrary that . Let , be the minimum distance from to .
Let be a -dimensional open ball with center and radius , then .
Let , then is a connected open set with and , contradict to the fact that is a maximal connected open set of .
∎
If a planar graph does not contain hanging edges, then each face of the planar graph can be regarded as a polygon. Similarly, if an -graph does not contain hanging simplexoids, then each maximal connected open set in the -graph can be regarded as a high-dimensional polytope. Therefore, eliminating all hanging edges in -graphs can greatly facilitate subsequent research.
Let be the closure of a maximal connected open set of , then we say is a polytope of . If there is no hanging simplexoids in and is a polytope of an -graph , then the boundary of is a -sphere in by Definition 4 and Lemma 3.2.
The -dimensional degree of is denoted by by Definition 10. We say is a unit polytope of dimension if . Note that the definition of unit polytope of dimension is equivalent to -dimensional simplexoid. We will prove that each polytope of -graph can be triangulated into a finite collection of unit polytope of dimension in Section 3.2.
Definition 16 (Graph-Closure).
Let be an -graph and be a set of simplexoids such that
The minimal -graph in set is called the graph-closure of , denoted as . A minimal -graph means that for any -dimensional simplexoid in , .
Definition 17 (Closed--Graph).
Let be the graph-closure of a special -graph , then is called the closed--graph if .
Next, we prove that for any -graph , simplexoids can be added sequentially to obtain the graph-closure of .
Theorem 3.1.
Let be an -graph, we add simplexoids into one by one according to the following procedure. We prove that this procedure will inevitably terminate in a finite number of steps, and the resulting graph will be a graph-closure of .
Procedure: Let , we assume that has became into when adding the -th simplexoid .
-
•
(1). Suppose that our procedure has reached the -th step, now we need to add into .
Figure 1: graph-closure -
–
(1.1). If has no hanging-simplexoid, then we terminate the procedure, it is easy to verify that is a graph-closure of .
-
–
(1.2). Otherwise, execute (2).
-
–
-
•
(2). Let be a polytope of a -graph .
-
–
(2.1). If there exists two -dimensional hanging-simplexoids and such that each of them contains a -sub-simplexoid (denoted by and ) which is incident with no -dimensional simplexoid in and . Assume that ; . Let be a -dimensional simplexoid with , we add into and get , return to (1).
(We provide a figure of the graph-closure of an -graph to explain Srep(2.1). As shown in Figure 1, is an -graph, both and are -dimensional simplexoids, . Note that after adding the simplexoids , the number of hanging-simplexoids in decreases by one. )
-
–
(2.2). Otherwise, choose a -dimensional hanging-simplexoid and a -dimensional simplexoid such that . Let be the -sub-simplexoid of that is incident with no -dimensional simplexoid in ; be the -sub-simplexoid of such that . Assume that ; . let be a -dimensional simplexoid with , we add into and get , return to (1).
(We provide a figure of the graph-closure of an -graph to explain Srep(2.2). As shown in Figure 2, is an -graph and is a hanging-simplexoid. Note that after adding the simplexoids and , there is no hanging-simplexoid in . )
-
–

Proof.
First, we prove that the above process will certainly terminate in a finite number of steps. It is evident that through step (2), the number of hanging-simplexoids in will decrease. Since is a finite -graph with no multiple simplexoids, the process will inevitably terminate in a finite number of steps.
Since the number of hanging-simplexoids in decreases with each iteration of step (2), let denote the -graph after steps, which still contains hanging-simplexoids; let be the -graph which no longer contains any hanging-simplexoids after steps. By Definition 16, it is evident that is the closure of .
∎
Corollary 2.
Let be an -graph and be the graph-colsure of , then is homeomorphic to the union of a finite collection of -spheres.
3.2 Step II: Triangulation
In this section, we will further triangulate -graphs based on their closures, ensuring that each polytope in the -graph is a unit polytope.
Definition 18.
An -graph in which all polytopes are unit polytopes of dimension is called a triangulated -graph.
Definition 19.
Let and be -graphs. If , , and is a triangulated -graph, then is a called a triangulated -graph of .
Definition 20.
An -graph is called a special triangulated -graph if is a special -graph and all polytopes of are unit polytopes of dimension .
Next, we prove that for any -graph, simplexoids can be added sequentially to obtain a triangulated -graph.
Theorem 3.2.
Let be a closed--graph, we add simplexoids into one by one according to the following procedure. We prove that this procedure will inevitably terminate in a finite number of steps, and the resulting graph will be a triangulated -graph of .
Procedure: Let ; we assume that has become into when adding the -th simplexoid .
-
•
(1). Suppose that our procedure has reached the -th step, now we need to add into .
-
–
(1.1). If all polytopes of are unit polytopes of dimension , it is easy to verify that is a graph-closure of .
-
–
(1.2). Otherwise, execute (2).
-
–
-
•
(2). Let be a polytope of a special -graph . Choose two -dimensional simplexoids and such that . Assume that ; .
-
–
(2.1). If for any -dimensional simplexoid , then let be a -dimensional simplexoid with . Now we add into and get , return to (1). (Figure 4 is the procedure of (2.1) when is a special -graph (planar graph).)
-
–
(2.2). If there exists a -dimensional simplexoid such that , then we return to (2), reselect two different -dimensional simplexoids and , and then re-execute (2.1) and (2.2).
-
–
Proof.
Since contains no multiple simplexoid, the above process always terminates in a finite number of steps. Clearly, the necessary and sufficient condition for the above process to terminate at step is that all polytopes of are unit polytopes of dimension , which implies that is called a triangulated -graph of . ∎


In practical applications, it is necessary not only to perform a triangulation of the -graph but also to determine the number of unit polytopes in the triangulated graph. Therefore, it is required to prove the following theorem.
Theorem 3.3.
Let be a polytope of a special -graph . Each simplexoid in is the non-hang simplexoid. We use to denote the number of -dimensional simplexoids which are incident to in . We use to denote the number of -dimensional simplexoids that are in . Then can be triangulated into -dimensional simplexoids if there exists with . Such triangulation is called the -triangulation.
Proof.
Let be a -dimensional closed ball, , then is homeomorphic to the with homeomorphism . Furthermore, takes the interior of to the interior of , and the to the boundary of by Lemma 3.3.
Let be a -dimensional simplexoid which satisfy , then is also a -dimensional simplexoid. The number of such is .
, join and . All such line segments form a -dimensional manifold , and it is obvious that is a -dimensional simplexoid in (Figure 4 is an example of when ). For every such that satisfies the above conditions, we repeat this process and get a triangulation of , such triangulation is also a triangulation of since is homeomorphic to the .
Note that the number of such -dimensional simplexoids in is equal to the number of -dimensional simplexoids (denoted by ) which satisfy and , thus we can triangulate into -dimensional simplexoids.
∎
Lemma 3.3.
[9] Let be a homeomorphism between two manifolds, then takes the interior of to the interior of , and the boundary of to the boundary of .
4 Coloring of special -graph
In order to estimate the chromatic number of special -graph, we need a new definition that connects the special -graph with coloring. In fact, the vertex-hypergraph can be considered as the skeleton of the special -graph.
4.1 Relationship between the number of vertices and edges
Definition 21 (Vertex-Hypergraph).
Let be an -graph. If we regard each as a hyperedge with vertex set , then we get a -uniform hypergraph which is called the vertex-hypergraph of . A proper coloring of the hypergraph is a coloring such that for any adjacent vertices and (id est, there exists a hyperedge such that ), the chromatic number of is the smallest integer such that is -colorable. Let be a list assignment of with for any , an -coloring of is a coloring such that for any . is -choosable if has an -coloring for any list with . The choice number of , denoted by , is the least integer such that is -choosable.
We observe that performing graph-closure or triangulation on an -graph does not decrease the chromatic number of the vertex-hypergraph.
Theorem 4.1.
Let be a special triangulated -graph, be the set of -dimensional simplexoids of , then if .
Proof.
By induction on (denoted by Induction ). The theorem holds trivially for . Assume that it holds for all integers less tha , and let be a triangulated special -graph with
then
which implies that there exists a vertex such that .
Now we prove that the theorem holds for .
Claim 4.1.
Theorem 4.1 holds for .
Proof.
By induction on (denoted by Induction ). The theorem holds trivially for . Let be an integer with . Assume that the theorem holds for .
Now we prove the theorem holds for .
, let be the number of vertices which are adjacent to , be a special -graph with and . It is obvious that there is only one polytope which is not the unit polytope of , and is a special -graph (denoted by ).
Observe that is homeomorphic to , and since , then there exists a vertex such that by the induction hypothesis (Induction ).
We triangulate into a triangulated -graph , and such a triangulation is a -triangulation in Theorem 4.2. Observe that is also a special -graph, then by the induction hypothesis (Induction ).
Note that the triangulation is a -triangulation, then
On the other hand, since , and we also know that . Therefore,
and the claim follows by induction.
∎
We infer that Theorem 4.1 holds for , and the theorem follows by induction.
∎
Corollary 3.
Let be a special triangulated -graph, be the vertex-hypergraph of , then there exists a -vertex in . (An -vertex represent that . )
4.2 Proof of Theorem 1.2
Proof.
Suppose to the contrary that there exists a minimum counterexample of Theorem 1.2 with fewest vertices. Note that there exists a -vertex in by Corollary 3.
By the minimality of , is -choosable. Let , be a list assignment of with . It is obvious that has an -coloring . Observe that , we color by and get a -coloring of , a contradiction.
∎
Since an -graph is a special type of -dimensional CW complex, and a graph can be regarded as the -skeleton of an -graph, the upper bound for the chromatic number of -graphs can be easily extended to graphs.
5 Bridges
In this section, we aim to establish some lemmas of bridge in higher-dimensional spaces. Let be a proper subgraph of a connected -graph G. The set may be partitioned into classes as follows. For each component of , there is a class consisting of the -dimensional simplexoids of together with the -dimensional simplexoids linking to . Each remaining -dimensional simplexoid defines a singleton class . The subgraphs of induced by these classes are the bridges of in . It follows immediately from this definition that bridges of can intersect only in -dimensional simplexoids of with , and that any two vertices of a bridge of are connected by a path in the bridge that is internally disjoint from .
For a bridge of , the elements of are called its vertices of attachment to ; the remaining vertices of are its internal vertices. A bridge is trivial if it has no internal vertices (that is, if it is of the second type). A bridge with vertices of attachment is called a -bridge. Two bridges with the same vertices of attachment are equivalent bridges.
We are concerned here with bridges of -sphere, and all bridges are understood to be bridges of a given -sphere . Thus, to avoid repetition, we abbreviate ’bridge of ’ to ’bridge’ in the coming discussion.
Lemma 5.1.
Let be a special -graph, be a bridge of -sphere , the projection of which is denoted by is a connected special -graph.
Proof.
Let and be two distinct connected components of , and , denotes a path between vertices and , denotes another path between vertices and , then is evidently a cycle, and in graph , cycle cannot be continuously contracted to a point. Thus, it follows that the fundamental group of graph is nontrivial, leading to a contradiction. ∎
The projection of a -bridge with effects a partition of into disjoint segments, called the segments of . Two bridges avoid each other if all the vertices of attachment of one bridge lie in a single segment of the other bridge; otherwise, they overlap.
Two bridges and are skew if contains a -sphere as a subgraph which effects a partition of into two disjoint segments , and there are distinct vertices of attachment of such that and are in different segment of , note that there is a -path in such that by Lemma 5.1 and Lemma 7.6.
We give an example of skew for . As shown in Figure 5, both and are in the inner region of , the bridge induced by is denoted by , the bridge induced by is denoted by . It is obvious that effects a partition of into two disjoint segments (two hemispheres), and and lie in different segments.

Lemma 5.2.
Overlapping -hyper-bridges of a closed--graph are either skew or else equivalent -bridges.
Proof.
Suppose that bridges and overlap. Clearly, each of them must have at least vertices of attachment. If either or is a -bridge, it is easily verified that they must be skew. We may therefore assume that both and have at least vertices of attachment.
If or are not equivalent bridges, then all the vertices of attachment of one bridge cannot lie in a single segment of the other bridge. Without loss of generality, let be a -sphere which effects a partition of into disjoint segments , then there exist distinct vertices of attachment of such that and are in different segment of . It follows that and are skew.
If and are equivalent k-bridges, then . If , and are skew by Lemma 5.3; if , they are equivalent -bridges.
∎
Lemma 5.3.
Let be an special -graph which is isomorphic to with , then there is a subgraph of which is isomorphic to -sphere that effects a partition of into disjoint segments , and there are distinct vertices of attachment of such that and are in different segment of .
Proof.
Firstly, we prove that there exists a vertex such that , if not, we konw that for every vertex , which implies that is a complete graph. It is impossible since is isomorphic to .
Let be the neighbor of , then is isomorphic to a -sphere, note that effects a partition of into disjoint segments , without loss of generality, let .
It is easy to verify that since , let , it follows that .
∎
6 Hyper ear decomposition
In this section, we aim to establish some lemmas of ear decomposition in higher-dimensional spaces. Let be a -connected -graph. Apart from complete graph (), the -connected closed--graph contains a subgraph which is isomorphic to . We describe here a simple recursive procedure for generating any such graph starting with an arbitrary -sphere of the -graph.
Definition 22 (Hyper Ear).
Let be a subgraph of an -graph . A hyper ear of in is a nontrivial -ball in whose boundary lies in but whose internal vertices do not.
Definition 23 (Hyper Ear Decomposition).
A nested sequence of a closed--graph is a sequence of -graphs such that , . A hyper ear decomposition of a -connected closed--graph is a nested sequence of -connected subgraphs of such that:
-
•
is isomorphic to .
-
•
where is a hyper ear of in , .
-
•
.
Lemma 6.1.
The -connected closed--graph with has a hyper ear decomposition.
Proof.
Lemma 6.2.
In a -connected closed--graph with , each maximal connected region or is bounded by a -sphere.
Proof.
Note that has a hyper ear decomposition by Lemma 6.1. Consider an ear decomposition of , where is isomorphic to , , and, for , is a -connected subgraph of , where is an ear of in . Since is isomorphic to , the two maximal connected regions of are clearly bounded by a -sphere. Assume, inductively, that all maximal connected regions of are bounded by -spheres, where . Because is a -connected -graph, the ear of is contained in some maximal connected region of . Each region of other than is a region of as well, and so, by the induction hypothesis, is bounded by a -sphere. On the other hand, the region of is divided by into two regions of , and it is easy to see that these regions are also bounded by -spheres.
∎
Lemma 6.3.
In a loopless -connected closed--graph , the neighbours of any vertex lie on a common -sphere.
Proof.
Let be a loopless -connected -graph and be a vertex of , then is -connected, so each maximal connected region of is bounded by a sphere by Lemma 6.2. If is the region of in which the vertex was situated, the neighbours of lie on its bounding sphere .
∎
7 Recognizing -Graph
We extend the definition of -component for graphs to higher-dimensional spaces before proving our main result.
7.1 -component
Definition 24 (-component).
Let be a connected graph which is not complete, let be a vertex cut of , and let be the vertex set of a component of . The subgraph of induced by is called an -component of . In the case where is a -connected -graph and is a -vertex cut of , we find it convenient to modify each -component by adding a new -dimensional simplexoid with vertex set . We refer to this simplexoid as a marker simplexoid and the modified -components as marked -components. The set of marked -components constitutes the marked -decomposition of . The graph can be recovered from its marked -decomposition by taking the union of its marked -components and deleting the marker edge.
As shown in Figure 6, be a 3-cut of an -graph , we provide an example of the -decomposition and marked -decomposition of an -graph. The only difference between -decomposition and marked -decomposition is that marked -decomposition must contain a simplexoid with vertex set . If this simplex does not exist in the original graph, it needs to be added during the construction.

We need to establish some lemmas before proving our main results.
Lemma 7.1.
Let be a closed--graph with a -vertex cut , then each marked -component of is isomorphic to a minor of .
Proof.
Let be an -component of , with marker simplexoid . Let be another -component of , with marker simplexoid , then there is a -ball such that and is a -sphere by Lemma 6.1. It is easy to verify that is isomorphic to a minor of by contract into a single simplexoid .
∎
Lemma 7.2.
Let and be closed -graphs whose intersection is isomorphic to with , then is a closed -graph.
Proof.
Let be a hyperplane, . At this point, the hyperplane divides into two disconnected regions, denoted by and , respectively. We embed into and into in such a way that and intersect only at .
By contradiction, suppose the fundamental group of is nontrivial, then there must exist a curve that cannot be continuously contracted to the base point. If curve belongs to either or , then it can be contracted continuously to base point, which leads to a contradiction. Therefore, must intersect both and . Let and , respectively. We first transform into by homotopy, such that belongs to . It is easy to verify that and belong to , thus they can be continuously contracted to the base point. By combining the two homotopy transformations, we obtain that can be continuously contracted to the base point, a contradiction. In conclusion, the assumption is invalid, and the theorem is proven.
∎
Lemma 7.3.
Let be a closed--graph with a -vertex cut , then is a closed--graph if and only if each of its marked -components is a closed--graph.
Proof.
Suppose, first, that is a closed -graph. By Lemma 7.1, each marked -component of is isomorphic to a minor of , hence is closed--graph.
Conversely, suppose that has marked -components each of which is a closed--graph. Let denote their common marker simplexoid. Applying Lemma 7.2 and induction on , it follows that is a closed -graph, hence so is . ∎
By Lemma 7.3, we know that to prove a closed -graph can be embedded in , it is sufficient to show that all of its marked -components can be embedded in .
7.2 Connectivity
Before proving Theorem 1.1, we need a lemma regarding connectivity.
Lemma 7.4.
Let be a -connected graph on at least vertices, then contains an edge such that is -connected.
Proof.
Suppose that the theorem is false. Then, for any edge of , the contraction is not -connected. By Lemma 7.5, there exists vertex set such that is a -vertex cut of .
Choose the edge and the vertex set in such a way that has a component with as many vertices as possible. Consider the graph . Because is -connected, is -connected. Moreover has the -vertex cut . It follows that the -component is -connected.
Let be a neighbour of in a component of different from . Since is an edge of , and is a counterexample to Lemma 7.4, there is a vertex set such that is a -vertex cut of , too. (The vertices might or might not lie in .) Moreover, because H is -connected, is connected (where, if such that , we set ), and thus is contained in a component of . But this component has more vertices than (because has more vertices than ), contradicting the choice of the edge and the vertex .
∎
Lemma 7.5.
Let be a -connected graph on at least vertices, and let be an edge of such that is not -connected. Then there exists a vertex such that is a -vertex cut of .
Proof.
Let be a -vertex cut of . At least of these vertices, say , is not the vertex resulting from the contraction of . Set . Because is -connected, is certainly -connected. However has a cut vertex, namely .
If is not the vertex resulting from the contraction of , then must be a -vertex cut of , a contradiction. Hence must be the vertex resulting from the contraction of . Therefore is disconnected, in other words, is a -vertex cut of . ∎
7.3 Anti--dimension minor
The following proof demonstrates that there is a conclusion that holds in that similar to Wagner’s Theorem. It is necessary to generalize the concepts of complete graphs and bipartite graphs to higher-dimensional spaces.
Definition 25 (Complete -Uniform-Topological Hypergraph ).
Let be an -uniform-topological hypergraph, be the vertex set of with order , be the collection of all subsets of containing elements. If for any , there exists a simplexoids in such that , then we call a complete -uniform-topological hypergraph, which is denoted by . (Figure 8 is an example of . )


Definition 26 (Complete Bipartite -Uniform-Topological Hypergraph ).
Let be an -uniform-topological hypergraph, be the vertex set of , be a partition of in which and . If satisfies the following properties, we call a complete bipartite -uniform-topological hypergraph, which is denoted by . (Figure 8 is an example of . )
-
•
is a complete -uniform-topological hypergraph .
-
•
For any and , there exists a simplexoid in such that and is minimal.
Lemma 7.6.
(Jordan-Brouwer Separation Theorem) Let be a -dimensional topological sphere in the -dimensional Euclidean space (), i.e. the image of an injective continuous mapping of the -sphere into , then the complement of in consists of exactly two connected components. One of these components is bounded (the interior) and the other is unbounded (the exterior). The set is their common boundary.
Lemma 7.7.
is a non--graph.
Proof.
Suppose to the contrary that is an -graph.
Let . Let be the collection of all subsets of containing elements.
Note that for any , is isomorphic to .


Let . We assume that has already been embedded in . By Lemma 7.6, each will divide into two disconnected regions, with one of the regions being empty. We designate the non-empty region as the external region of and the empty region as the internal region. Let be the internal region corresponding to and be the external region corresponding to . (As shown in the Figure 10, when embedded in , it is obvious that line intersects with at least at one point. )
Without loss of generality, we assume that is in , and it follows that is in . Since is a complete -uniform-topological hypergraph, there must exist a simplexoid whose vertex set includes both and . By Lemma 7.6, this simplexoid must intersect with . Therefore cannot be embedded in .
∎
Lemma 7.8.
is a non--graph.
Definition 27 (Anti--dimension Minor).
If a -uniform-topological hypergraph has a -minor or -minor, then we call an anti--dimension minor.
7.4 Proof of Theorem 1.1
Proof.
Let be the graph-closure of a -uniform-topological hypergraph. If its minors include either or , it is obvious that is a non--graph by Lemma 7.7-7.8.
Now we assume that is a -connected non--graph and is simple. Because all graphs on or fewer vertices can be embedded in , we have . We proceed by induction on . By Lemma 7.4, contains an edge such that is -connected. If is non-, it has an anti--dimension minor, by induction. Since every minor of is also a minor of , we deduce that too has an anti--dimension minor. So we may assume that is an -graph.
Consider an -embedding of . Denote by the vertex of formed by contracting . Because is -connected, by Lemma 6.3 the neighbours of lie on a -sphere , the boundary of some polytope of . Denote by and , respectively, the hyper-bridges of in that contain the vertices and .
Suppose, first, that and avoid each other. In this case, and can be embedded in the polytope of in such a way that the vertices and belong to the same polytope of the resulting -graph . The edge and the simplexoids that incident with can now be drawn in that polytope so as to obtain an -embedding of itself, contradicting the hypothesis that is non-.
It follows that and do not avoid each other, that is, they overlap. By Lemma 5.2, they are therefore either skew or else equivalent -bridges. In the latter case, has a -minor; In the former case, has a -minor.
-
•
In higher dimensions, the former case is not intuitive. Here, we provide an example of minimum skew is equivalent to in , which can be analogized to higher-dimensional cases.
-
•
As shown in Figure 11(1), simplexoids , are joined together to form a two-dimensional sphere . Vertices and are inside .
-
•
As shown in Figure 11(2), simplexoids and form a bridge , which effect a partition of into disjoint segments.
-
•
As shown in Figure 11(3), there is another bridge with internal vertex . Its vertics of attachment includes both and . By definition, and are skew. On the other hand, since there are no hanging simplexoid in the graph, bridge must include simplexoids .
-
•
As shown in Figure 11(4), let and , At this point, is a complete bipartite -uniform-topological hypergraph.

∎
References
- [1] Mark Anthony Armstrong. Basic topology. Springer Science & Business Media, 2013.
- [2] John Adrian Bondy and Uppaluri Siva Ramachandra Murty. Graph theory. Springer Publishing Company, Incorporated, 2008.
- [3] John M Boyer and Wendy J Myrvold. Simplified o (n) planarity by edge addition. Graph Algorithms Appl, 5:241, 2006.
- [4] Daniel Brélaz. New methods to color the vertices of a graph. Communications of the ACM, 22(4):251–256, 1979.
- [5] Robert F. Cohen, Peter Eades, Tao Lin, and Frank Ruskey. Three-dimensional graph drawing. Algorithmica, 17(2):199–208, 1997.
- [6] IS Filotti, Gary L Miller, and John Reif. On determining the genus of a graph in o (v o (g)) steps (preliminary report). In Proceedings of the eleventh annual ACM symposium on Theory of computing, pages 27–37, 1979.
- [7] Michael R Garey, David S Johnson, and Larry Stockmeyer. Some simplified np-complete problems. In Proceedings of the sixth annual ACM symposium on Theory of computing, pages 47–63, 1974.
- [8] John Hopcroft and Robert Tarjan. Efficient planarity testing. Journal of the ACM (JACM), 21(4):549–568, 1974.
- [9] R Munkres James. Topology. Prentic Hall of India Private Limited, New delhi, 7, 2000.
- [10] Ken-ichi Kawarabayashi and Paul Wollan. A simpler algorithm and shorter proof for the graph minor decomposition. In Proceedings of the forty-third annual ACM symposium on Theory of computing, pages 451–458, 2011.
- [11] Rhyd Lewis. Guide to graph colouring. Springer, 2021.
- [12] Bojan Mohar. Embedding graphs in an arbitrary surface in linear time. In Proceedings of the twenty-eighth annual ACM symposium on Theory of computing, pages 392–397, 1996.
- [13] Bojan Mohar. A linear time algorithm for embedding graphs in an arbitrary surface. SIAM Journal on Discrete Mathematics, 12(1):6–26, 1999.
- [14] Bojan Mohar and Carsten Thomassen. Graphs on surfaces. Johns Hopkins University Press, 2001.
- [15] Neil Robertson and Paul D Seymour. Graph minors. xiii. the disjoint paths problem. Journal of combinatorial theory, Series B, 63(1):65–110, 1995.
- [16] Johannes Schneider and Roger Wattenhofer. A log-star distributed maximal independent set algorithm for growth-bounded graphs. In Proceedings of the twenty-seventh ACM symposium on Principles of distributed computing, pages 35–44, 2008.
- [17] Johannes Schneider and Roger Wattenhofer. A new technique for distributed symmetry breaking. In Proceedings of the 29th ACM SIGACT-SIGOPS symposium on Principles of distributed computing, pages 257–266, 2010.
- [18] Andrew Thomason. An extremal function for contractions of graphs. In Mathematical Proceedings of the Cambridge Philosophical Society, volume 95, pages 261–265. Cambridge University Press, 1984.
- [19] Dominic JA Welsh and Martin B Powell. An upper bound for the chromatic number of a graph and its application to timetabling problems. The Computer Journal, 10(1):85–86, 1967.