Planar graphs and poset dimension
1989, Order
https://doi.org/10.1007/BF00353652…
21 pages
1 file
Sign up for access to the world's latest research
Abstract
We view the incidence relation of a graph G=(I: E) as an order relation on ~ts vertices and edges, i.e. a <6 b if and only ifa is a vertex and b is an edge incident on a. This leads to the definition of the order-dimension of G as the minimum number of total orders on I'~ E whose intersection is <G. Our main result ~s the characterization of planar graphs as the graphs whose order-dimension does not exceed three. Strong versions of several known properties of planar graphs are ~mphed by this characterizahon These properhes include: each planar graph has arboricn) at most three and each planar graph has a plane embedding ~hose edges are straight line segments. A nice feature of this embedding ~s that the coordinates of the vertices have a purely combmatoriaI meaning.
Related papers
Order, 2020
Planar posets can have arbitrarily large dimension. However, a planar poset of height h has dimension at most 192h + 96, while a planar poset with t minimal elements has dimension at most 2t + 1. In particular, a planar poset with a unique minimal element has dimension at most 3. In this paper, we extend this result by showing that a planar poset has dimension at most 6 if it has a plane diagram in which every minimal element is accessible from below.
The Electronic Journal of Combinatorics, 2020
In 1981, Kelly showed that planar posets can have arbitrarily large dimension. However, the posets in Kelly's example have bounded Boolean dimension and bounded local dimension, leading naturally to the questions as to whether either Boolean dimension or local dimension is bounded for the class of planar posets. The question for Boolean dimension was first posed by Nešetřil and Pudlák in 1989 and remains unanswered today. The concept of local dimension is quite new, introduced in 2016 by Ueckerdt. Since that time, researchers have obtained many interesting results concerning Boolean dimension and local dimension, contrasting these parameters with the classic Dushnik-Miller concept of dimension, and establishing links between both parameters and structural graph theory, path-width and tree-width in particular. Here we show that local dimension is not bounded on the class of planar posets. Our proof also shows that the local dimension of a poset is not bounded in terms of the maxi...
European Journal of Combinatorics, 2008
The vertex-arboricity a(G) of a graph G is the minimum number of subsets into which the set of vertices of G can be partitioned so that each subset induces a forest. It is well-known that a(G) ≤ 3 for any planar graph G. In this paper we prove that a(G) ≤ 2 whenever G is planar and either G has no 4-cycles or any two triangles of G are at distance at least 3.
Complexity
The number of edges in a shortest walk (without repetition of vertices) from one vertex to another vertex of a connected graph G is known as the distance between them. For a vertex x and an edge e = a b in G , the minimum number from distances of x with a and b is said to be the distance between x and e . A vertex x is said to distinguish (resolves) two distinct edges e 1 and e 2 if the distance between x and e 1 is different from the distance between x and e 2 . A set X of vertices in a connected graph G is an edge metric generator for G if every two edges of G are distinguished by some vertex in X . The number of vertices in such a smallest set X is known as the edge metric dimension of G . In this article, we solve the edge metric dimension problem for certain classes of planar graphs.
arXiv (Cornell University), 2019
We show that height h posets that have planar cover graphs have dimension O(h 6). Previously, the best upper bound was 2 O(h 3). Planarity plays a key role in our arguments, since there are posets such that (1) dimension is exponential in height and (2) the cover graph excludes K 5 as a minor.
Ars Mathematica Contemporanea
The acyclic number a(G) of a graph G is the maximum order of an induced forest in G. The purpose of this short paper is to propose a conjecture that a(G) ≥ 1 − 3 2g n holds for every planar graph G of girth g and order n, which captures three known conjectures on the topic. In support of this conjecture, we prove a weaker result that a(G) ≥ 1 − 3 g n holds. In addition, we give a construction showing that the constant 3 2 from the conjecture cannot be decreased.
Journal of Combinatorial Theory, Series A, 2019
It has been known for more than 40 years that there are posets with planar cover graphs and arbitrarily large dimension. Recently, Streib and Trotter proved that such posets must have large height. In fact, all known constructions of such posets have two large disjoint chains with all points in one chain incomparable with all points in the other. Gutowski and Krawczyk conjectured that this feature is necessary. More formally, they conjectured that for every k 1, there is a constant d such that if P is a poset with a planar cover graph and P excludes k + k, then dim(P) d. We settle their conjecture in the affirmative. We also discuss possibilities of generalizing the result by relaxing the condition that the cover graph is planar.
arXiv: Combinatorics, 2019
A planar order is a special linear extension of the edge poset (partially ordered set) of a processive plane graph. The definition of a planar order makes sense for any finite poset and is equivalent to the one of a conjugate order. Here it was proved that there is a planar order on the vertex poset of a processive planar graph naturally induced from the planar order of its edge poset.
Connections in Discrete Mathematics
Motivated by quite recent research involving the relationship between the dimension of a poset and graph-theoretic properties of its cover graph, we show that for every d 1, if P is a poset and the dimension of a subposet B of P is at most d whenever the cover graph of B is a block of the cover graph of P , then the dimension of P is at most d + 2. We also construct examples which show that this inequality is best possible. We consider the proof of the upper bound to be fairly elegant and relatively compact. However, we know of no simple proof for the lower bound, and our argument requires a powerful tool known as the Product Ramsey Theorem. As a consequence, our constructions involve posets of enormous size.
arXiv (Cornell University), 2022
In 1989, Nešetřil and Pudlák posed the following challenging question: Do planar posets have bounded Boolean dimension? We show that every poset with a planar cover graph and a unique minimal element has Boolean dimension at most 13. As a consequence, we are able to show that there is a reachability labeling scheme with labels consisting of O(log n) bits for planar digraphs with a single source. The best known scheme for general planar digraphs uses labels with O(log 2 n) bits [Thorup JACM 2004], and it remains open to determine whether a scheme using labels with O(log n) bits exists. The Boolean dimension result is proved in tandem with a second result showing that the dimension of a poset with a planar cover graph and a unique minimal element is bounded by a linear function of its standard example number. However, one of the major challenges in dimension theory is to determine whether dimension is bounded in terms of standard example number for all posets with planar cover graphs.
Planar Graphs and Poset Dimension
WALTER SCHNYDER
Department of Mathematics, Louisiana State L’nversity, Baton Rouge, LA 70803-4918, L’S A
Communicated by W. T. Trotter
(Received: 13 June 1987; accepted: 25 August 1988)
Abstract
We view the incudence relation of a graph G=(V,E) as an order relation on its vertices and edges, i.e. a<Gb if and only if a is a vertex and b is an edge incident on a. This leads to the definition of the order-dimension of G as the minimum number of total orders on 1∪E whose intersection is <G. Our main result is the characterization of planar graphs as the graphs whose order-dimension does not exceed three. Strong versions of several known properties of planar graphs are implied by this characterization These properties include: each planar graph has arboricity at most three and each planar graph has a plane embedding whose edges are straight line segments. A nice feature of this embedding is that the coordinates of the vertices have a purely combınatorial meaning.
AMS subject classifications (1980). Prımary 06A10: secondary 05C10, 05C75.
Key words. Planar graph. poset dimension, straight line embedding.
1. Introduction
For a poset (X,<) consisting of a set X and a (transitive, irreflexive) orderrelation < on X, a realizer is a nonempty set of total orders on X whose intersection is the relation <.
E. Szpilrajn [17] has proved that each poset has a realizer. B. Dushnik and E. W. Miller [3] have defined the dimension of a poset P(dimP) as the minimum cárdinality of its realizers.
Fig. 1.
For example, the diagram shown in Figure 1 defines an order < (of height 1) on the set X={a,b,c,d,e,f} by x<y if and only if ’ x is in the lower part, y is in the upper part and x,y are adjacent’. The dimension of (X,<) cannot be 1 ( < is not total) and must therefore be 2 as < is the intersection of the two
total orders
<1:abdcef and <2:cbfaed
(Here as everywhere else in this paper, the order is from left to right.)
The importance of orders in mathematics has motivated numerous investigations of the dimension concept. For example, Komm [11] established very early that the dimension of the power-set of a set is the cardinality of this set and Hiraguchi [8] proved that, for n⩾4, the dimension of a n-element poset is at most n/2. Later, Gysin [7] and Trotter, Moore, and Sumner [20] applied results of Gallai [5] to show that the dimension of a poset (X,<) depends only of the underlying comparability graph (the graph where vertices x,y∈X are adjacent if and only if x<y or y<x ), i.e. two posets with the same comparability graph have the same dimension. These are but a few of many results in this area. For more information we refer to the overviews in [18, 12, 22].
Although two-dimensional posets are well understood [3, 10, 19] and can be recognized in polynomial time [6], no simple characterization of d-dimensional posets is known for d⩾3. Indeed, Yannakakis [23] has shown that deciding whether a poset has dimension d is a NP-complete problem, for all d⩾3. Even the recognition of four-dimensional posets of height 1 is NPcomplete. The complexity of the recognition of three-dimensional posets of height 1 remains an open problem.
One of the simplest examples of posets of height 1 is the poset P(n,k) induced by the inclusion relation on the class of all 1 - and k-element subsets of a n-element set. The study of this poset dates back to [3] proving dimP(n,n−1)=n and the lower bound dimP(n,2)⩾log2log2(n). These investigations were continued in [2] with a formula for dimP(n,k) when 2⌊χnˉ⌋⩽k⩽n. Eventually, Spencer [14] investigating the asymptotic behavior of dimP(n,k) has shown that dimP(n,2) is asymptotically equal to log2log2(n).
In this paper, posets of height 1 are viewed as hypergraphs whose incidence relation is interpreted as order relation. For example, P(n,k) is the complete k-uniform hypergraph on n vertices and P(n,2) is the complete graph Kn. We propose to extend the investigation of P(n,2) to general graphs, defining the (order-) dimension of a graph as the dimension of its incidence relation, and to try to relate the dimension of a graph to other graph-properties.
Our main result is that a graph has dimension at most three if and only if this graph is planar.
From this characterization we deduce that, given a maximal planar graph G on at least four vertices the poset consisting of the vertices, the edges and the faces of G ordered by the inclusion relation has dimension four. Removal of any face from this poset results in a three-dimensional poset. The hypergraph whose vertices are the vertices of G and whose hyperedges are the edges and faces of G has the same property.
Our result also implies strong versions of several known properties of planar graphs. For example, the property of having embeddings in the plane where edges are segments of straight lines [4,16,21]; here we can give a purely combinatorial meaning to the vertex coordinates of the embedding. Another property is the decomposition of maximal planar graphs in three edge disjoint trees [13,9] that correspond to three-dimensional realizers.
Notice that the planarity of three-dimensional graphs (‘only if’ part of our characterization) has also been shown by Babai and Duffus [1] in a different context.
2. The Order-Dimension of a Graph
Throughout this paper, V is a finite nonempty set and ∣V∣ denotes the cardinality of V. Graphs are finite simple graphs, their edges are identified with 2 element sets. A triangle of a graph G is a set of three pairwise adjacent vertices of G.
The symbols G and H are reserved for graphs, V(G) is the vertex set of G and E(G) is the edge set of G; a formula G=(V,E) means that V=V(G) and E=E(G).G is a subgraph of H if V(G)⊆V(H) and E(G)⊆E(H).
If R is a binary relation on V(G), an R-path in G is a path v0,v1,…,vn of G such that R(vi,vi+1) holds for all i,0⩽i<n. Tlte outdegree in R of a vertex x of G is the number of neighbors y of x that satisfy R(x,y).
DEFINITION. For a graph G=(V,E) the partial order <G on V∪E is defined by a<Gb⇔a∈V and b∈E and a∈b.
The (order-) dimension dim(G) of G is the dimension of the poset (V∪E,<G).
Remark. If H is a subgraph of G, the relation <H is the restriction of <G to V(H)∪E(H). Therefore dim(H)⩽dim(G), i.e., the dimension is a monotone function of graphs.
The definition of dim(G) relates <G to its realizers, thus to total orders on V(G)∪E(G). It will be useful to consider the restrictions of the latter orders to V(G). Their essential properties are stated in the following lemma whose straightforward proof is left to the reader. An analogous statement for P(n.k) can be found in [2].
LEMMA 2.1. A graph G with vertex set V has dim(G)⩽d if and only if there exists a sequence <1,<2,…,<d of total orders on V satisfying the following conditions:
(1) the intersection of <1,<2,…,<d is empty;
(2) for each edge {x,y} and each vertex z∈/{x,y} of G. there is at least one order <i in the sequence such that x<iz and y<iz.
Condition (1) is a consequence of condition (2) when the minimum degree of
G is at least 2. In this case, already the existence of d partial orders satisfying condition (2) implies dim(G)⩽d.
In the remainder of this paper, the names vertex property and edge property will refer to (1) and (2), respectively.
DEFINITION. A d-dimensional representation on a set V is a sequence <1, <2,…,<d of total orders on V that has the vertex property. This sequence represents all graphs with vertex set V for which the edge property holds. Among these graphs, the graph G with maximal edge set is the graph induced by <1,<2,…,<d.
The following examples and Proposition 2.4 illustrate the use of Lemma 2.1 and show some elementary properties of the dimension of graphs.
EXAMPLE 2.2. A graph has dimension less than 3 if and only if it is a subgraph of a path (notice that the only 1-dimensional graph is the isolated vertex): a two-dimensional representation on an n-element set V has the form
<1:v1v2…vn−1vn,<2:vnvn−1…v2v1
and represents a graph G=(V,E) iff E⊆{{vi,vi+1}∣1⩽i<n}.
EXAMPLE 2.3. A three-dimensional representation of the triangle with vertices a,b,c is any sequence <1,<2,<3 of total orders on {a,b,c} such that each of a,b and c is the maximum of one of these orders.
PROPOSITION 2.4. Each 4-colorable graph has dimension at most 4.
Proof. Let G=(V,E) be a four-colorable graph and consider a partition X,Y,Z,W of V in four color classes. Let $X^{+}, Y^{+}, Z^{+}, W^{+}$denote any total orderings of X,Y,Z,W and $X^{-}, Y^{-}, Z^{-}, W^{-}$denote the inverse orderings. It is easy to see that the following four orderings of V form a four-dimensional representation of G.
<1:X+<2:Y−<3:W+<4:Z−Y+X−Z+W−Z+W+W−Z−X−Y−Y+X+
(the superscripts chosen in column 1 are irrelevant).
The converse of Proposition 2.4 is false. For example, it can be shown that dim(K12)=4. ( K13 is the first complete five-dimensional graph).
It will be proved in Section 4 that every three-dimensional graph is planar. Together with Proposition 2.4, this implies the existence of graphs of arbitrarily high genus in dimension 4.
The following observations (Lemma 2.5 and Corollary 2.6) show that each d-dimensional graph has standard representations. This fact has also been used in [2].
LEMMA 2.5. Let <1,<2,…,<d be a d-dimensional representation of G= (V,E) and x be the maximum of <k. For i=k, let <i′ be the total order on V where x precedes all elements of V−{x} and the elements of V−{x} are ordered by <t. Then <1′,…,<k−1′,<k,<k+1′,…,<d′ is also a d-dimensional representation of G.
DEFINITION. A d-dimensional representation <1,<2,…,<d on a set V is standard if ∣V∣⩾d and, for all i=j, the maximum element of <t is one of the d−1 smallest elements of <t. The maxima of the orders of a standard d dimensional representation are the exterior elements of this representation. The other elements of V are the interior elements.
Notice that, for d⩾3, any two exterior elements of a standard d-dimensional representation are adjacent in the graph induced by this representation.
COROLLARY 2.6. Each Graph G with at least d vertices and dim(G)⩽d has a standard d-dimensional representation.
Proof. By d-fold application of Lemma 2.5 to an initial d-dimensional representation of G.
3. Three Dimensional Representations
For a sequence R1,R2,R3 of binary relations on a set V, we define the dual sequence R1∗,R2∗,R3∗ by
Rk∗(x,y)⇔Rk(x,y)&[Ri(y;x) for i=k]
In particular, if <1,<2,<3 is a three-dimensional representation on V, the relations <1∗,<2∗,<3∗ are partial orders on V and any two distinct elements x and y of V are comparable in exactly one of these orders. Each edge of the graph G induced by <1,<2,<3 receives therefore a unique label and direction. This decomposes G in three arc disjoint digraphs whose arc-sets are denoted A1,A2, A3:
Ak={(x,y)∣{x,y}∈E(G)&x<k∗y}
and E(G) is the disjoint union of the three underlying edge sets E1,E2,E3 defined by Ek:={{x,y}∣(x,y)∈Ak}. An immediate consequence of the edge and vertex properties is then Lemma 3.1.
LEMMA 3.1. For a 3-dimensional representation <1,<2,<3 on the set V, there holds: Ak={(x,y)∣y is the minimum in <k of {v∈V∣x<k∗v}}.
EXAMPLE 3.2. The three-dimensional representation <1:b c xy≤a,<2 : cazxyb,<3:abyzxc on V={a,b,c,x,y,z} induces the graph shown as Figure 2, where the thick lines correspond to A1.
Fig. 2.
THEOREM 3.3. For k=1,2,3, the digraph (V,Ak) determined by a threedimensional representation <1,<2,<3 on V is a rooted forest (whose arcs are directed toward the roots). If the representation is standard and a1,a2,a3 denote the respective maxima of <1,<2,<3, then the component Tk of ak in (V,Ak) has ak as root and includes at least all interior elements of the representation as further vertices. The other components, if any, are trivial.
Proof. By Lemma 3.1, the outdegree in (V,Ak) of a vertex x∈V is at most 1 (it is 1 iff there exists a vertex y∈V such that x<k∗y ). Hence, to each undirected cycle of the underlying graph (V,Ek) would correspond a directed cycle of (V,Ak). Since <k∗ is an (acyclic) order relation, (V,Ek) is a forest. The same outdegree property of (V,Ak) implies that each connected component of (V,Ek) contains exactly one vertex whose outdegree in (V,Ak) is 0 . This vertex is the root of the given component.
If <1,<2,<3 is standard, the only vertices that may have outdegree 0 in (V,Ak) are the respective maxima a1,a2,a3 of <1,<2,<3. Of these, only ak can have indegree greater than 0 in (V,Ak). Hence, (V,Ak) has at most one nontrivial component (rooted in ak ). This tree includes all interior elements of the representation.
Remark. For i=k the element ai is a vertex of Tk (the notation is the same as in Theorem 3.3) if and only if the edge {ai,ak} is directed from ai to ak. Therefore, if ∣V∣=n, either the trees T1,T2,T3 have each exactly n−1 vertices, or the cardinalities of their vertex sets form a permutation of the numbers n,n−1 and n−2.
We observe that Theorems 1 and 3 of [9] are a consequence of the last remark (combined with the characterization of planar graphs as graphs of dimension at most three).
COROLLARY 3.4. A three-dimensional representation <1,<2,<3 on a set V of cardinality n⩾3 is standard if and only if the induced graph G has exactly 3n−6 edges.
Proof. (1) The ‘only if’ statement is a direct consequence of the last remark. (2) Conversely, let a1,a2,a3 denote the respective maxima of <1,<2,<3. Suppose that <1,<2,<3 is not standard. For example, there exist vertices x,y∈V such that x<2y<2a1. The vertex condition implies then that a1=a3. One of the vertices x,y must be different from a3. This vertex is in no order greater than both a1 and a3. Consequently {a1,a3} is not an edge of G. However, applications of Lemma 2.5 to <1 and <3 result in a supergraph H= ( V,E′ ) of G such that {a1,a3}∈E′. Part (1) of this proof and Corollary 2.6 imply then that ∣E(G)∣<∣E′∣⩽3n−6.
4. Three Dimensional Graphs
The following notation and terminology concerning planar embeddings will be used in this and the next sections.
A triangular graph G is a maximal planar graph on at least three vertices that is embedded in the plane (i.e. whose exterior face has been chosen). The triangle of G whose edges form the boundary of the exterior face of G is the exterior triangle of G; its vertices and edges are the exterior vertices and edges of G, the other vertices and edges are the interior vertices and edges of G. An elementary triangle of G is a triangle whose edges form the boundary of an interior face of G. Given a cycle Z of G, the region with boundary Z is the subgraph of G induced by the vertices of Z and the vertices lying in the interior of Z. Therefore ‘face’ and ‘region’ have different meanings.
The notation xy (for x,y∈R3 ) represents the straight line segment with endpoints x and y. A straight line embedding of a graph G in a plane is an injection f of V(G) in this plane such that for any two distinct edges {x,y} and {u,v} of G:f(x)f(y)∩f(u)f(v)=f({x,y}∩{u,v}).
THEOREM 4.1. Each graph G=(V,E) of dimension at most three is planar. Moreover, to each three-dimensional representation <1,<2,<3 of G corresponds a straight-line embedding f:v∈V→(v1,v2)∈R2 of G in the plane, such that for all u,v∈V:ui<vi⇔u<iv(i=1,2).
Proof. For v∈V and i∈{1,2}, let vi be the power of 2 whose exponent is the ordinal of v with respect to <i.
It suffices to verify that the so defined mapping f:V→R2 is a straight line embedding of G in the plane. To simplify the notation, the same symbol will be used to denote a vertex of V and its image under f.
Note that, by the definition of f, if x,y∈V satisfy y<1x and x<2y, no vertex z=x,y will be mapped by f in the (closed) triangle delimited by the points x,y and (x1,y2) shown in Figure 3, since each point z of this triangle satisfies
2x1<z1⩽x1 or 2y2<z2⩽y2
Fig. 3.
Assume, for contradiction, that there exist two disjoint edges {x,y} and {a,b} such that xy∩ab=∅. (The case of nondisjoint edges has a similar proof.)
Let, for example, x be the maximum of {x,y,a,b} with respect to <1. By the edge property, y>1a,b,a>jx,y,b>kx,y must hold for some choice of i,j,k∈{1,2,3}.
xy∩ab=∅ and the maximality of x in <1 imply i,j,k⩾2. Thus j=k and therefore j=k=3 and i=2. Then y>2a,b and xy∩ab=∅ imply y>2x.
Consequently x>1y,y>2x and y2>a2,b2,x1>a1,b1. Therefore, a and b are both on the same side (below) of the straight line through x and y. Hence, xy∩ab=∅ in contradiction to the original assumption.
By Corollary 3.4 and Theorem 4.1, a three-dimensional representation <1,<2,<3 on a set V of cardinality ∣V∣⩾3 induces a maximal planar graph if and only if this representation is standard. In this case, we refer to the graph G induced by <1,<2,<3 together with its planar embedding defined in the proof of Theorem 4.1 as the triangular graph induced by <1,<2,<3. Elementary geometric considerations show that the exterior vertices of G are precisely the exterior elements of the representation. This justifies the terminology introduced in the definition of standard representations.
As Lemma 4.2, we now prove the converse of Theorem 3.3. For later use, it is convenient to formulate this lemma in terms of acyclic relations.
LEMMA 4.2. Let G be a triangular graph whose vertex set V has cardinality ∣V∣⩾4 and <1,<2,<3 be three acyclic relations on V such that every interior vertex of G has outdegree exactly one in each of <1∗,<2∗,<3∗. Then for each edge {x,y} of G and each vertex z∈/{x,y} there exist, for some k∈{1,2,3},a<k-path from x to z and a<k-path from y to z. Therefore, any three total orders extending <1,<2,<3 on V form a three-dimensional representation of G.
Proof. Notice that for i=j, a <i∗-path and a <i∗-path starting in the same vertex v have no common vertex except v as the existence of such a common vertex u would imply a <i-cycle from v to v via u.
Let v be any interior vertex of G. By the above, the (unique longest) <1∗,<2∗ and <3∗-paths starting at v end in distinct exterior vertices that will be denoted
respectively by a1,a2,a3. In particular, this implies the existence of some <i-path from aj to ai for any choice of j=i.
If u=v is any further interior vertex of G, the (unique longest) <i∗-path starting at u must also end at ai, as an ending in aj for j=i would imply the existence of some <k∗-path (k=i) from u to ai and therefore the existence of a <i-cycle from ai to ai via u,aj and v (Figure 4).
Fig. 4.
Therefore there exists a <i-path from each vertex u=ai of G to ai. This proves the lemma for the case where z∈{a1,a2,a3}.
If z∈/{a1,a2,a3}, the <1∗,<2∗ and <3∗-paths from z to a1,a2 and a3 divide G in three regions R1,R2 and R3 (where Rk denotes the region opposite ak, including its boundary) (Figure 5).
Fig. 5.
As the end-vertices x and y of each edge {x,y} of G are both in the same region, it suffices to show that there exists for each vertex v of G,v∈Rk, a <k-path from v to z.
Let, for example, v∈R3 and consider some <3-path from v to a3. This path must intersect the <1∗-path from z to a1 or the <2∗-path from z to a2 in a vertex u. Suppose, for example, that u belongs to the <1∗-path from z to a1. Combining the two paths, there results a <3-path from v to z via u.
5. Barycentric Representations
It has been shown in Section 4 that the graph G induced by a three-dimensional representation <1,<2,<3 is always planar. It is therefore tempting to look for an embedding of G in a plane (or sphere) in R3 such that <1,<2 and <3 are precisely the orders defined on the vertices by their three coordinates. Such an embedding however, does not exist in general, as illustrated (for planes) by Example 5.1.
EXAMPLE 5.1. Let V={a,b,c,x,y,z} and <1,<2,<3 be the three-dimensional representation defined on V by
<1:axbzcy
<2:b y cxaz
<3:czaybx.
There exists no mapping f:v∈V→(v1,v2,v3)∈R3 such that
(1) f(V) is contained in a plane,
(2) for all u,v∈V:ui<vi⇔u<iv(i=1,2,3).
Since for i=1,2,3 there exist elements u,v∈V with u<i∗v, the coefficients α1,α2,α3 of the equation α1v1+α2v2+α3v3=δ of such a plane would have to be of the same sign. Replacing in each row <i of the above matrix every element v by the product αivi and comparing the sums of columns 1,3,5 and columns 2,4,6 would then lead to the contradiction 3δ=3δ.
Nevertheless, Lemma 4.2 implies that the edge and vertex properties of a standard three-dimensional representation are essentially determined by the restrictions of its orders to the edges of the induced graph. This suggests a relaxation of condition (2) of the last example and will result (Theorem 5.3 and Corollary 5.4) in the correct interpretation of three-dimensional representations by coordinates.
DEFINITION. A barycentric embedding of a graph G is an injective function v∈V(G)→(v1,v2,v3)∈R3 that satisfies the conditions:
(1) v1+v2+v3=1 for all vertices v,
(2) For each edge {x,y} and each vertex z∈/{x,y}, there is some k∈{1,2,3} such that xk<zk and yk<zk.
LEMMA 5.2. Each barycentric embedding of a graph is a straight line embedding of this graph in the plane v1+v2+v3=1.
Proof. By applying the proof of Theorem 4.1 to the ‘projection’ f:v∈V(G)→ (v1,v2)∈R2 of an initial barycentric embedding of the graph G. Observe that if {x,y} is an edge of G and y1⩽x1,x2⩽y2, no vertex z of G will be mapped by f in the convex hull of the points (x1,x2),(y1,y2),(x1,y2).
THEOREM 5.3. Let G be the triangular graph induced by a standard three-
dimensional representation <1,<2,<3 on a set V and let a1,a2,a3 be the respective maxima of <1,<2,<3. Then there exists a straight line embedding of G such that v1,v2,v3 denoting the barycentric coordinates of vertices v∈V relative to a1,a2,a3, there holds for each interior edge {u,v} of G:u<iv⇔ ui<vi(i=1,2,3).
Proof. Let n denote the cardinality of V. It suffices to define a mapping f:v∈ V→(v1,v2,v3)∈R3 with the properties:
(0) f(a1)=(2n−5,0,0),f(a2)=(0,2n−5,0),f(a3)=(0,0,2n−5),
(1) v1,v2,v3⩾0 and v1+v2+v3=2n−5 for all vertices v,
(2) u<iv⇒ui<vi for all interior edges {u,v} of G and i=1,2,3.
By Lemmas 4.2 and 5.2 , the mapping resulting from f upon division by 2n−5 satisfies the statement of the proposition.
Using the notation of the proof of Lemma 4.2, recall that each interior vertex v divides G in three regions R1(v),R2(v),R3(v) and that each vertex u=v of region Rk(v) satisfies u<kv(k=1,2,3).
f(a1),f(a2),f(a3) are defined by ( 0 ) above. For an interior vertex v of G, we let vi(i=1,2,3) be the number of elementary triangles in region Ri(v). This clearly satisfies condition (1).
To verify condition (2), consider an interior edge {u,v} of G and assume, for example, that u<3∗v. To show: u3<v3,v1<u1,v2<u2. This is trivial if v=a3. Suppose, therefore, that v=a3. Then v is an interior vertex of G and u must lie in the interior of region R2(v) (Figure 6).
Fig. 6.
For i=1 and j=2 (or i=2 and j=1 ), the <i∗ path from u to ai and the <i∗ path from v to aj have no common vertex, as the existence of such a vertex x would imply u⩽ix⩽iv, contradicting u<3∗v. The outdegree (at most 1) properties of <1∗ and <2∗ imply thus: R3(u)⊆R3(v),R1(v)⊆R1(u), and R2(v)⊆ R2(u). These inclusions being proper, there follows u3<v3,v1<u1,v2<u2.
As each three-dimensional representation of a graph G can be extended, by adjunction of three new exterior elements, to a standard representation of a triangular supergraph of G, we obtain
COROLLARY 5.4. To each three-dimensional representation <1,<2,<3 of a graph G corresponds a barycentric embedding v∈V(G)→(v1,v2,v3)∈R3 of G such that u<iv⇔ui<vi for all edges {u,v} of G and i=1,2,3.
6. Planar Graphs have Dimension at Most Three
We now show that each planar graph has dimension at most three. By Theorem 4.1 or Corollary 5.4 , this fact implies, in particular, that each planar graph has straight line embeddings, a result of Fáry [4], Stein [16] and Wagner [21]. Our proof of three-dimensionality is based on their methods, as analyzed by Kampen [9].
Two independent versions are given: construction of a straight line embedding in this section and a coordinate-free proof in Section 7.
We first review the method of edge contraction. For a vertex x of a graph G,N(x) denotes the set of neighbors of x in G. If {x,y} is an edge of G, the contracted graph G/(x,y) is obtained from G by removal of the vertex y and the edges incident on y and by introduction of an edge {x,z} for each vertex z∈ N(y)−N(x). The edge {x,y} is contractible if x and y have exactly two common neighbors. If G is a maximal planar graph on at least four vertices and {x,y} is a contractible edge of G, then G/(x,y) is a maximal planar graph.
LEMMA 6.1 [9]. Let G be a triangular graph on n⩾4 vertices. If a,b and c denote the exterior vertices of G, then there exists a neighbor x=a,b of c such that the edge {c,x} is contractible.
Thus to each triangular graph corresponds a sequence of ‘allowed contractions’ transforming this graph into a triangle.
THEOREM 6.2. Each planar graph has dimension at most three.
Proof. By the monotonicity of the dimension, it suffices to prove the theorem for triangular graphs G. Let V be the vertex set of G and let a,b,c denote the exterior vertices of G in counterclockwise order. We show the existence of an embedding of V in the plane such that (identifying the vertices and their images):
(a) a,b,c are affinely independent and each vertex v∈V lies in the triangle delimited by a,b and c.
(b) The partial orders <1,<2,<3 defined on V by the barycentric coordinates v1,v2,v3 of the vertices v∈V with respect to a,b,c(u<iv⇔ui<vi) satisfy the hypothesis of Lemma 4.2 (if ∣V∣⩾4 ).
The proof proceeds by induction on the number n=∣V∣. The claim is trivial
if n=3. Let n⩾4 and assume that the claim is true for all triangular graphs with less than n vertices.
Consider some neighbor x=a,b of c such that the edge {c,x} is contractible and let p=v1,v2,…,vr=q be the neighbors of x, distinct from c, in counterclockwise order. The vertices p and q are the two common neighbors of c and x (Figure 7).
Fig. 7.
By induction hypothesis, there is an embedding f of V−{x} in the plane with the above properties (a) and (b). Lemma 4.2 implies that f is a barycentric embedding. By Lemma 5.2, this embedding is therefore a straight line embedding of G/(c,x) (Figure 8 ).
Fig. 8.
Notice that as c3>u3 for all u∈V−{c} and f is a straight line embedding of G, condition (2) of the definition of barycentric embeddings applied to the wheel of c implies that
p2=v21<v22<⋯<v2r=q2 and q1=v1r<v1r−1<⋯<v11=p1
Let d denote the point with barycentric coordinates (q1,p2,1−q1−p2). By the above inequalities, there is a choice of x sufficiently close from d in the quarterplane u1>q1,u2>p2 that will satisfy the relations: x3>v3′(i=1,2,…,r), x1>q1,x2>p2,x1<v1′ and x2<v2′(i=2,…,r−1).
For i=1,2,3 let <i denote the partial order defined on V by u<iv⇔ui<vi. By the choice of x, we have x<1∗p,x<2∗q,x<3∗c and v′<3∗x for all i=1,r. Therefore, the outdegree of x in each of the relations <1∗,<2∗,<3∗ is exactly one. Using the induction hypothesis, it is easy to verify that the outdegrees of each interior vertex y=x of G in the relations <1∗,<2∗,<3∗ are exactly one (notice that for i=2,…,r−1 the edge {v′,c} of G/(c,x) has been replaced in G by the edge {v′,x} and that there holds v′<3∗x ).
Observe that we placed x close to d and far from c. The standard proof that G has a straight line embedding would position x arbitrarily close to the vertex c.
Remark. If <1,<2,<3 is a standard three-dimensional representation on a set V of cardinality ∣V∣⩾4 and c is the maximum of <3, then the maximum x of V−{c} in <3 is a neighbor of c and the edge {c,x} is contractible. This fact could have been used to prove Theorem 5.3 by suitably modifying the proof of Theorem 6.2. However, the resulting proof would not provide the interpretation of barycentric coordinates as counts of elementary triangles.
Theorem 3.3 has shown that each three-dimensional representation of a triangular graph decomposes this graph in three edge-disjoint directed trees whose vertices include all interior vertices of the graph. However, not every such tree-decomposition corresponds to a three-dimensional representation as the acyclicity of the induced relations <1,<2,<3 is not guaranteed. This acyclicity was ensured in the proof of Theorem 6.2 by the use of coordinates. In the next section, we present the topological properties of the tree-decomposition that correspond to the above acyclicity. Rather than using trees, however, it will be convenient to apply the equivalent notion of labelings of angles.
7. Labeling the Angles of a Triangular Graph
The angles of a triangular graph G are the angles of its elementary triangles. If G is induced by a standard three-dimensional representation <1,<2,<3, then each angle ∠(xy,xz) of G determines a unique label k∈{1,2,3} such that x>ky,z (since {x,y,z} is a triangle and the inequalities y>,x,z and z>,x,y must also be satisfied).
The representation <1,<2,<3 induces in this way a labeling of the angles of G with the labels 1,2,3.
EXAMPLE 7.1. The representation <1:bcxyza,<2:cazxyb,<3:abyzxc on V={a,b,c,x,y,z} induces the labeled graph given as Figure 9.
The essential properties of this labeling are summarized in the following definition.
DEFINITION. A normal labeling of a triangular graph G is a labeling of the angles of G with the labels 1,2,3 that satisfies the conditions:
Fig. 9 .
(1) Each elementary triangle of G has an angle labeled 1, an angle labeled 2 and an angle labeled 3. The corresponding vertices appear in counterclockwise order.
(2) The labels of the angles of an interior vertex x of G form, in counterclockwise order, an interval of 1 's followed by an interval of 2 's followed by an interval of 3 's.
In the remainder of this section, the expression labeled triangular graph will be used to denote a triangular graph together with a normal labeling of its angles.
PROPOSITION 7.2. The labeling induced by a standard three-dimensional representation is a normal labeling.
Proof. Let G be the triangular graph induced by a standard three-dimensional representation <1,<2,<3 and consider a straight line embedding of G with the properties shown in Proposition 5.3.
Define the counterclockwise orientation of the plane by the cycle a1,a2,a3 (in this order). Property (1) of normal labelings is then trivially satisfied.
Let x be an interior vertex of G and let u1,u2,u3 denote the (unique) neighbors of x with x<i∗ui. Notice that the ordering u1,u2,u3 is counterclockwise. It is easy to see that all angles at x in a sector (xui,xui+1) have the same label i+2 (indices and labels are modulo 3 ). This shows that property (2) of normal labelings is also satisfied.
A vertex v belonging to a cycle Z in a labeled triangular graph will be said to be of type i(i=1,2,3) with respect to Z if all angles at v, interior to Z, have the label i. The interpretation of three-dimensional representations by barycentric coordinates motivates the following lemma.
LEMMA 7.3. Each cycle Z in a labeled triangular graph G has a vertex of type 1, a vertex of type 2 and a vertex of type 3.
Proof. Each interior edge {u,v} of G belongs to two elementary triangles
{x,u,v} and {y,u,v}. The labels of the angles at u and v in these triangles will be called labels of u and v along {u,v}. The definition of a normal labeling implies that these labels have (up to a renaming of u and v ) the form given in Figure 10, where i,j,k is a cyclic permutation of 1,2,3. Thus, the labels of u along {u,v} are distinct from the labels of v along {u,v}.
Fig. 10 .
Assume that there exist cycles for which the statement of the lemma is false. Among these cycles, consider a cycle Z enclosing the minimum number n of elementary triangles. By definition n⩾2. Suppose, for example that Z has no vertex of type 1.
Case 1 (Figure 11): There is an edge {u,v} interior to Z whose vertices u and v are nonconsecutive vertices of Z. The edge {u,v} divides then Z in two cycles Z1 and Z2, each containing less than n elementary triangles of G.
Fig 11 .
Z1 and Z2 have vertices of type 1 . As Z does not have such a vertex, one of the vertices u,v has type 1 in Z1 and the other has type 1 in Z2. Therefore the labels of u along {u,v} and the labels of v along {u,v} are not distinct, in contradiction to the preliminary observation.
Case 2 (Figure 12): No two nonconsecutive vertices of Z are joined by an edge interior to Z. Let u,v be counterclockwise consecutive vertices of Z and let x be the vertex in the interior of Z such that {x,u,v} is an elementary triangle (observe that x is not a vertex of Z ). Let the angles at u,v and x in this triangle be labeled i,j and k, respectively.
Then Z has a vertex of type k. Indeed, let Z′ denote the cycle obtained from Z through replacement of the edge {u,v} by the two edges {u,x} and {x,v}. Z′ encloses n−1 elementary triangles and has, therefore, a vertex w of type k. The remark preceding case 1 implies that w=u,v. The definition
Fig. 12
of normal labelings implies that w=x. Therefore, w is a vertex of type k in the cycle Z.
Applying this argument to i=2 and j=3, it follows that Z has no vertex of type 2 or 3 . Therefore Z has no vertex of type 1,2,3; again a contradiction to the above argument.
In particular, the exterior cycle of a labeled triangular graph G has a vertex of type 1 , a vertex of type 2 and a vertex of type 3 ; the exterior vertices of G. It can easily be seen that these vertices appear in counterclockwise order. (For an illustration, see Example 7.1.)
PROPOSITION 7.4. Let G be a labeled triangular graph and <1,<2,<3 be the binary relations defined on Γ(G) by: x<ky⇔ there exists an elementary triangle {x,y,z} of G such that the label of ∠(yx,yz) is k. Then the relations <1,<2,<3 are acyclic and any three total orders extending <1,<2,<3 on Γ(G) form a three-dimensional representation of G.
Proof. Suppose that there exists a cycle x=x0<kx1<k⋯<kxn=x(n⩾1). The underlying cycle Z of G does not have any vertex of type k. This contradicts Lemma 7.3.
Condition (2) of the definition of a normal labeling implies that each interior vertex of G has outdegree exactly one in <k∗(k=1,2,3). Together with Lemma 4.2, this implies the second statement of Proposition 7.4.
We now give the second (coordinate-independent) version of the proof of Theorem 6.2.
THEOREM 7.5. Each triangular graph has a normal labeling. Therefore, each planar graph has dimension at most three.
Proof. The second statement of the theorem follows from the first combined with Proposition 7.4 and the monotonicity of the dimension.
We prove the first statement by induction on the number n of vertices of a triangular graph G. The case n=3 is trivial. Let n⩾4 and assume that the theorem is true for all triangular graphs having less than n vertices.
Let a,b,c denote the exterior vertices of G and x=b,c be a neighbor of a such that a and x have exactly two common neighbors. Let a,v1,v2,…,vt be the vertices of the wheel of x, listed in counterclockwise order (Figure 13).
Fig. 13
By induction hypothesis, the graph G/(a,x) has a normal labeling and we may assume that all angles at a have the label 1 (Figure 14).
Fig. 14.
Fig. 15 .
This labeling can trivially be transformed in a normal labeling of G by labeling the angles ∠(xvi,xvi+1) and ∠(av1,ax),∠(ax,avr) with 1 , as shown in Figure 15 .
Notice that all normal labelings of a triangular graph can be obtained by the above method.
8. Adding the Faces: Hypergraphs
In this section, we consider (simple) hypergraphs H=(V,E) consisting of a nonempty set V of vertices and a set E of hyperedges that are subsets of V having
cardinality at least two. We call H a closed hypergraph if the implication X⊆Y⇒∣Y−X∣=1 or Y−X∈E holds for any two distinct hyperedges X and Y of H.
Each hypergraph H=(V,E) determines two partial orders <H∞ and <H1 on V∪E. The order <H∞ corresponds to the set inclusion (vertices being identified with singleton sets). The order <H1 corresponds to the incidence relation and is defined as for graphs. The superscripts ∞ and 1 indicate the upper bounds on the heights.
We therefore have two notions of order-dimension for a hypergraph H :
dim∞(H)=dim(V∪E,<H∞) and dim1(H)=dim(V∪E,<H1)
As a variant of Lemma 2.1, we obtain Lemma 8.1.
LEMMA 8.1. A closed hypergraph H with vertex set V has dim1(H)⩽d if and only if there exists a sequence <1,<2,…,<d of total orders on V satisfying the following conditions:
(1) the intersection of <1,<2,…,<d is empty;
(2) for each hyperedge X and each vertex y∈/X of H, there is at least one order <i in the sequence such that x<iy for all x∈X.
Proof. The ‘only if’ part is trivial. The ‘if’ part is proved by extending a sequence <1,<2,…,<d of total orders on V that satisfies conditions (1) and (2) to a sequence <1′,<2′,…,<d′ of total orders on V∪E whose intersection is <H1.
The order <i′ is obtained by insertion in <i of each hyperedge X just after its maximum (with respect to <i ). Hyperedges having the same maximum are inserted in order of decreasing cardinalities.
We show, for example, that distinct hyperedges are incomparable in the intersection of <1′,<2′,…,<d′. That is, given any two distinct hyperedges X and Y, there must exist an order <i′ such that X<i′Y.
Case 1: Y−X=∅. Then there exists a vertex y∈Y with y∈/X. By (2) there is an order <i such that x<iy for all x∈X. Hence, X<i′Y.
Case 2: Y−X=∅. Therefore Y⊂X. Since H is a closed hypergraph X is a disjoint union X=Y∪Z, where Z∈E or Z={z} for some z∈V. Let y be any element of Y. By (2) or (1) there is an order <i such that z<iy for all z∈Z. Hence, X and Y have the same maximum in this order <i. Since ∣Y∣<∣X∣, there follows X<i′Y.
Remark. This lemma remains true if dim1(H) is replaced by dim∞(H) (even when H is not closed). Therefore, if H is a closed hypergraph there holds dim1(H)=dim∞(H). We denote this common value with dim(H).
Let now G be a maximal planar graph on at least four vertices and let H(G) be the hypergraph whose vertex set is V(G) and whose hyperedges are the edges and faces of G. With H−(G) we denote the hypergraph obtained from
H(G) by elimination of one of the faces of G. Then H(G) and H−(G) are closed hypergraphs and dimH(G),dimH−(G) are well defined.
THEOREM 8.2. If G be a maximal planar graph on at least four vertices, then dimH(G)=4 and dimH−(G)=3.
Proof. View G as a triangular graph whose exterior face is the face missing in H−(G). Let <1,<2,<3 be a three-dimensional representation of G. From the proof of Lemma 4.2, it is easy to see that for each elementary triangle {x,y,z} and each vertex v∈/{x,y,z} of G there is one order <i of the representation satisfying the inequalities x<iv,y<iv and z<iv.
Thus the sequence <1,<2,<3 satisfies conditions (1) and (2) of Lemma 8.1 with respect to H−(G). As dimH−(G)>2, there follows dimH−(G)=3.
Let <4 be any total order on the vertex set V of G whose three smallest elements are the exterior elements of the representation <1,<2,<3. Clearly, the sequence <1,<2,<3,<4 satisfies both conditions of Lemma 8.1 with respect to H(G). Therefore, dimH(G)⩽4.
Suppose that dimH(G)⩽3. Then there exists a sequence <1′,<2′,<3′ of total orders on V that satisfies both conditions of Lemma 8.1. This sequence is, therefore, a three-dimensional representation of G in the sense of Section 2, thus standard as G is maximal planar. Hence, no vertex of G is greater than the three exterior elements of <1′,<2′,<3′ in any of these orders. As these three elements form a face of G this contradicts condition (2) of Lemma 8.1.
Remark. The inequality dim=H(G)⩾4 holds for all polyhedra [15].
References
- L. Babai and D. Duffus (1981) Dimension and automorphism groups of lattices, Alg. Univ. 12, 279-289.
- B. Dushnik (1950) Concerning a certain set of arrangements, Proc. Amer Math. Soc 1, 788796 .
- B. Dushnik and E. W. Miller (1941) Partially ordered sets, Amer J Math 63, 600-610.
- I. Fáry (1948) On straight line representation of planar graphs, Acta Sci. Math Szeged 11, 229−233.
- T Gallai (1967) Transitiv orientierbare Graphen, Acta Math. Acad Scl Hungar 18, 25-66.
- M. C. Golumbic (1977) The complexity of comparabılity graph recognition and coloring, Computing 18, 199-203.
- R. Gysin (1977) Dimension transitv orientierbarer Graphen, Acta Math Acad Sci Hungar 29, 313−316.
- T. Hiraguchi (1951) On the dimension of orders, Sci Rep Kanazawa Univ 1, 77-94.
- G R. Kampen (1976) Orienting planar graphs, Discrete Math 14, 337-341.
- D. Kelly (1977) The 3-irreducible partially ordered sets, Canad. J Math 29, 367-383.
- H. Komm (1948) On the dimension of partially ordered sets, Amer. J Math. 70, 507-520.
- D. Kelly and W. T. Trotter Jr. (1982) Dimension theory for ordered sets, in I. Rival (ed.), Ordered Scts, D. Reidel, Dordrecht, pp. 171-211.
- C. St J. A. Nash-Williams (1961) Edge disjoint trees of finite graphs, J London Math Soc 36, 445−450.
- J. Spencer (1971) Minimal acrambling sets of simple orders. Acta Math Acad Scl Hungar 22. 349−353.
- V. Sedmak (1954) Quelques applications des ensembles ordonnés, Bull. Soc Math Phys Serhuc 6. 12-39, 131-153.
- S. K. Stein (1951) Convex maps, Proc Amer Math Soc 2, 464-466.
- E. Szpilrajn (1930) Sur l’extension de l’ordre partiel, Fund Math 16, 386-389.
- W. T. Trotter Jr. (1983) Graphs and Partially Ordered Sets, in L. Benneke (ed.), Graph Theory: Vol. 2, Academic Press, London, pp. 237-268.
- W. T. Trotter, Jr and J. I. Moore Jr. (1976) Characterization problems for graphs, partially ordered sets, lattices and families of sets, Discrete Math 16, 361-381.
- W. T. Trotter, J. I. Moore and D. P. Sumner (1976) The dimension of a comparability graph. Proc. Amer Math Soc. 60, 35-38.
- K. Wagner (1936) Bemerkungen zum Vierfarbenproblem, Jahn=her Deutsth Math-1’eren 46, 26-32.
- D. B. West (1985) Parameters of partial orders and graphs: packing, covering, and representation. in I. Rival (ed.), Graphs and Orders, D. Reidel, Dordrecht, pp. 267-350.
- M. Yannakakis (1982) The complexity of the partial order dimension problem, SI.II J Alg. Discrete Methods 3, 351-358.
References (23)
- L. Babal and D. Duffus (1981) Dimension and automorphism groups of lattices, A/g, Umv. 12, 279-289.
- B. Dushnik (1950) Concerning a certain set of arrangements, Proc. Amer Math. Soc 1, 788- 796.
- B. Dushnik and E. W. Miller ( 1941) Partially ordered sets, Amer J Math 63, 600-610.
- Ffiry (1948) On straight line representation of planar graphs, Acta Set. Math Szeged 11, 229-233.
- T Gallal (1967) Transitiv orientierbare Graphen, Acta Math. Acad Sct Hungar 18, 25-66.
- M.C. Golumbic (1977) The complexity of comparability graph recognition and coloring, Computmg 18, 199-203.
- R. G~sln (1977) Dimension transltlV orientlerbarer Graphen, Acta Math Acad Sct Hungar 29, 313-316.
- T. Hlraguchl (1951) On the dimension of orders, Sct Rep Kanazawa Umv 1, 77-94.
- G R. Kampen (1976) Orienting planar graphs, Dz,screte 34ath 14,337-341.
- D. Kell} (1977) The 3-irreducible partially ordered sets, Canad. J Math 29,367-383.
- H. Komm (1948) On the dimension of partially ordered sets, Amer. J Math. 70, 507-520.
- D. Kelly and W.T. Trotter Jr. (1982) Dimension theory for ordered sets, in I. Rival (ed.), Ordered Sets, D. Reldel, Dordrecht, pp. 171-211.
- C. St J.A. Nash-Williams (1961) Edge disjoint trees of finite graphs, J London Math Soc 36,445-450.
- J. Spencer (1971) Minimal acrambling sets of simple orders, ,4eta Math Acad Scz Hungar 22, 349-353.
- V. Sedmak (1954) Quelques apphcations des ensembles ordonn6s, Bzdl, Soc Math Ph)',~ Serhte 6, 12-39, 131-153.
- S. K. Stein (1951) Convex maps, Proc Amer Math Soc 2,464-466.
- E. Szpilrajn (1930) Sur l'extenslon de l'ordre partiel, Fund Math 16,386-389.
- W. T. Trotter Jr. (1983) Graphs and Partially Ordered Sets, m L. Bemeke (ed.), Graph Theory. Vol. 2, Academic Press, London, pp. 237-268.
- W.T. Trotter, Jr and J. 1. Moore Jr, (1976) Characterization problems for graphs, partially ordered sets, lattices and famlhes &sets, Dz,wrete Math 16, 361-38 I.
- W.T. Trotter, J. I. Moore and D. E Sumner (1976) The dimension of a comparability graph, Proc. Amer Math Soc. 60, 35-38.
- K. Wagner (1936) Bemerkungen zum Vierfarbenproblem, ,labor,her Deut,~h ltath -!'erem 46, 26-32.
- D. B. West (1985) Parameters of partial orders and graphs: packing, covering, and representatmn, in I. Rival (ed.), Graph,s and Orders, D, Reidel, Dordrecht, pp. 267-350.
- M. Yannakakls (1982) The complexity of the partial order dimension problem, SI, 1U J Alg. Dzscrete Method,~ 3, 351-358.