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

The Three Tree Theorem

Oliver Knill Department of Mathematics
Harvard University
Cambridge, MA, 02138
(Date: September 4, 2023)
Abstract.

We prove that every 2-sphere graph different from a prism can be vertex 4-colored in such a way that all Kempe chains are forests. This implies the following ”three tree theorem”: the arboricity of a discrete 2-sphere is 3. Moreover, the three trees can be chosen so that each hits every triangle. A consequence is a result of an exercise in the book of Bondy and Murty based on work of A. Frank, A. Gyarfas and C. Nash-Williams: the arboricity of a planar graph is less or equal than 3.

Key words and phrases:
Graph theory, Arboricity, Chromatic number, Planar graphs, 4-coloring

1. Preliminaries

1.1.

A finite simple graph G=(V,E)G=(V,E) is called a 2-manifold, if every unit sphere S(v)S(v), the sub-graph of GG induced by {wV|(v,w)E}\{w\in V\;|\;(v,w)\in E\} is a 1-sphere, a cyclic graph with 44 or more elements. Because every edge in a 22-manifold bounds two triangles and every triangle is surrounded by three edges, the set TT of triangles satisfies the Dehn-Sommerville relation 3|T|=2|E|3|T|=2|E|. Since GG is K4K_{4}-free, the Euler characteristic of GG is χ(G)=|V||E|+|T|\chi(G)=|V|-|E|+|T|, where TT is the set of triangles K3K_{3}, EE the set of edges K2K_{2} and VV the set of vertices K1K_{1} in GG. If a 22-manifold GG is connected and χ(G)=2\chi(G)=2, it is called a 22-sphere. The class of 22-spheres together with K4K_{4} are known to agree with the class of maximally planar, 44-connected graphs.

1.2.

GG is contractible if there is vVv\in V such that both the unit sphere graph S(v)S(v) and the graph GvG\setminus v induced by V{v}V\setminus\{v\} are contractible. This inductive definition starts by assuming that 1=K11=K_{1} is contractible. The zero graph 0, the empty graph, is declared to be the (1)(-1)-sphere. For d0d\geq 0, a d-sphere GG is is a dd-manifold for which GvG\setminus v is contractible for some vertex vv. A dd-manifold is a graph for which every unit sphere is a (d1)(d-1)-sphere. If GG is a dd-sphere, the two contractible sets, GvG\setminus v and the unit ball B(v)B(v) together cover GG, so that its category is 22 as in the continuum.

1.3.

Using induction one can show that a contractible graph satisfies χ(G)=1\chi(G)=1 and a sphere satisfies the Euler gem formula χ(G)=1+(1)d\chi(G)=1+(-1)^{d} which for d=2d=2 is 22. The Euler-Poincaré formula k=0d(1)kfk=k=0d(1)kbk\sum_{k=0}^{d}(-1)^{k}f_{k}=\sum_{k=0}^{d}(-1)^{k}b_{k} holds for a general finite abstract simplicial complex, where fkf_{k} is the number of Kk+1K_{k+1} sub-graphs of GG and bkb_{k} is a Betti number, the nullity of the kk-th Hodge matrix block. In the connected 22-dimensional case, the Euler-Poincaré formula is |V||E|+|T|=1b1+b2|V|-|E|+|T|=1-b_{1}+b_{2}, implying χ(G)2\chi(G)\leq 2. Indeed, the identity χ(G)=2\chi(G)=2 forces b1=0b_{1}=0 and b2=1b_{2}=1 so that GG is orientable of genus 0, justifying the characterization of 2-spheres using the Euler characteristic functional alone, among connected 2-manifolds.

1.4.

Examples: 1) The octahedron and icosahedron are 22-spheres. 2) The suspension of a 1-sphere CnC_{n} is a prism graph. It is a 2-sphere for n4n\geq 4. For n=4n=4, it is the octahedron. Prisms are the only 2-dimensional non-primes in the Zykov monoid [65] of all spheres. No other 2-sphere can be written as ABA\oplus B with non-zero A,BA,B. 3) The tetrahedron K4K_{4} is not a 2-manifold because S(v)=K3S(v)=K_{3}. It is 4-connected and maximally planar although. One could look at the 2-dimensional skeleton complex of K4K_{4} which is a 2-dimensional simplicial complex qualifying for a sphere. Within graph theory, where the simplicial complexes are the Whitney complexes, it is not as sphere. 4) The triakis-icosahedron is a triangulation of the Euclidean 2-sphere {|x|=1,x3}\{|x|=1,x\in\mathbb{R}^{3}\} but as it is having K4K_{4} sub-graphs, it is not a 2-manifold. Also here, one could look at the 2-skeleton complex and get as a geometric realization a triangulation of a regular sphere. We do not allow degree 3 vertices as this produces tetrahedra which are 3-dimensional. 5) The cube and dodecahedron both have dimension 11 and are not 2-manifolds. Also here, one would have to transcend the frame work to include it as a 2-sphere. One would look at CW complexes, meaning a discrete cell complex, where the quadrangles or pentagons are included as cells. There had been a lot of confusion about defining polyhedra and polytopes [49, 54]. 6) The Barycentric refinement G1=(V1,E1)G_{1}=(V_{1},E_{1}) of a 2-sphere GG is a 2-sphere. It has as vertex set V1V_{1}, the set of complete sub-graphs of GG and the edge set E1={(x,y),xy,oryx}E_{1}=\{(x,y),x\subset y,\;{\rm or}\;y\subset x\}. We currently are under the impression that the arboricity in the barycentric limit could give an upper bound for the arboricity of all d-spheres. 7) An edge refinement Ge=(Ve,Ee)G_{e}=(V_{e},E_{e}) of a 2-sphere GG with edge e=(a,b)e=(a,b) is a 2-sphere. One has Ve=V{e},Ee=E{(v,a),(v,b),(v,c),v(d)}V_{e}=V\cup\{e\},E_{e}=E\cup\{(v,a),(v,b),(v,c),v(d)\} with {c,d}=S(a)S(b)\{c,d\}=S(a)\cap S(b). 8) The reverse of 7), edge collapse can be applied to a degree 4 vertex S(v)S(v) for which all wS(v)w\in S(v) satisfy degG(w)6{\rm deg}_{G}(w)\geq 6. The result is then again a 2-sphere. 9) A vertex refinement Ga,bG_{a,b} picks two non-adjacent points {a,b}\{a,b\} in S(v)S(v) and takes Va,b=V{w}V_{a,b}=V\cup\{w\} and Ea,b=E{(v,w),(a,w),(b,w)}E_{a,b}=E\cup\{(v,w),(a,w),(b,w)\}. 10) The reverse of 9), the kite collapse, can be applied to an embedded kite K=K2+2K=K_{2}+2 in GG if the dis-connected vertex pair {a,b}\{a,b\} in KK both have degrees 55 or larger.

Refer to caption
Refer to caption
Figure 1. The icosahedron GG to the left with a 4-coloring produced the 3 Kempe forests A,B,CA,B,C. The octahedron to the right has chromatic number 3. But any 3 or 4-coloring ff of GG leads to a coloring FF which produces a closed Kempe loop. Any prism graph has arboricity 3 but in the prism case, the three trees can never be Kempe trees. Prism graphs are also the only 2-spheres which are a join of two lower dimensional spheres.

1.5.

The arboricity of a graph GG is defined as the minimal number of forests that partition the graph [5]. A forest is a triangle-free graph for which every connected component is contractible. The empty graph 0 is not a tree and has arboricity 0, the graph 1=K11=K_{1} as well as any 0-dimensional manifold (a graph without edges) has arboricity 11 because it is a forest. For a connected graph GG, the arboricity of GG agrees with the minimal number of sub trees which cover the graph. Proof: a finite set of forests F1,,FkF_{1},\dots,F_{k} partitioning GG can in a connected GG be completed to a set of covering trees TkT_{k} of GG if GG is connected. Conversely, every collection of covering trees T1,,TkT_{1},\dots,T_{k} can be morphed into a collection of disjoint partitioning forests F1,,FkF_{1},\dots,F_{k}, where FiF_{i} are obtained from TiT_{i} by deleting edges ww that are already covered by other trees. We get back to this in an appendix.

1.6.

The Nash-Williams theorem [52, 7, 17] identifies the arboricity of a positive dimensional graph as the least integer kk such that |EH|k(|VH|1)|E_{H}|\leq k(|V_{H}|-1) for all sub-graphs H=(VH,EH)H=(V_{H},E_{H}) and the understanding is that the vertex set VHV_{H} generates the graph HH, meaning that if v,wv,w are two nodes in HH and the connection (v,w)(v,w) is in EE then also (v,w)(v,w) is in HH. The understanding is also that for H=K1H=K_{1} where any kk would work for Nash-Williams the arboricity is 11 and that for H=0H=0, the empty graph, the arboricity is 0. In particular, |E|/(|V|1)|E|/(|V|-1) is always a lower bound for the arboricity for a connected graph G=(V,E)G=(V,E). Since any 4-connected planar graph different from K4K_{4} is a sub-graph of a 2-sphere, the arboricity of a 4-connected planar graph is bounded by the maximal arboricity that is possible of 2-spheres. As we will show here, 2-spheres have arboricity 33 so that all planar graphs will have arboricity 3\leq 3, a result which appears as an exercise 21.4.6 in [5] based on work of A. Frank, A. Gyarfas and C. Nash-Williams. The text [56] sees it as a consequence of the Edmond Covering Theorem in matroid theory.

1.7.

Examples: 1) The arboricity of a cyclic graph Cn=(V,E)=({0,1,,n1},{(k,k1)modn,1,n})C_{n}=(V,E)=(\{0,1,\dots,n-1\},\{(k,k-1)\;{\rm mod}\;n,1,\leq n\}) with n4n\geq 4 is 22. It is larger or equal to |E|/(|V|1)=n/(n1)>1|E|/(|V|-1)=n/(n-1)>1 and two linear trees like T1=({0,1}{(01)})T_{1}=(\{0,1\}\{(01)\}) and T2=CnK2=(V,E{(01)})T_{2}=C_{n}\setminus K_{2}=(V,E\setminus\{(01)\}) cover CnC_{n}. That the arboricity of CnC_{n} is 22 also follows from the fact that CnC_{n} itself is not a tree. 2) The arboricity of a figure 8 graph is 22 too. It can be partitioned into 2 forests F1,F2F_{1},F_{2}, where F1F_{1} is a star graph with 5 vertices and F2F_{2} is a forest consisting of 2 trees. From the two forests, one can get a tree cover by taking T1=F1T_{1}=F_{1} by taking T2T_{2} the two trees F2F_{2} with a path connecting them. This illustrates the above switch from a forest partition to a tree cover which works in the connected graph case.

1.8.

3) The smallest 22-torus that is a manifold is obtained by triangulating a 5×55\times 5 grid and identifying the left-right and bottom-top boundaries to get 16 vertices and 216=322*16=32 triangles. It has the f-vector (|V|,|E|,|T|)=(16,48,32)(|V|,|E|,|T|)=(16,48,32) so that |E|/(|V|1)=48/15>3|E|/(|V|-1)=48/15>3 and an explicit cover with 44 forests shows the arboricity is 44. This is larger than the number 33 obtained in the Barycentric limit. It is still not excluded that the arboricity of a dd-manifold is always smaller or equal than the smallest integer larger than the Barycentric limit number cdc_{d} to be defined later and which is c2=3c_{2}=3 in two dimensions. A conjecture of Albertson and Stromquist states that all 2-manifolds have cromatic number 5\leq 5 [1]. 4) An octahedron with |V|=6,|E|=12,|T|=8|V|=6,|E|=12,|T|=8 has arboricity 33, because |E|/(|V|1)=12/5>2|E|/(|V|-1)=12/5>2 and because there are three spanning trees can cover it: start with two star graphs and a circular graph partitioning the edge set, then switch one of the equator colors. 5) There is a projective plane of chromatic number 55. Its ff-vector is f=(15,42,28)f=(15,42,28). The arboricity is 3.

Refer to caption
Refer to caption
Figure 2. A torus graph which has a 44-coloring ff and arboricity 44. A 2-torus of chromatic number 4 and arboricity 4 always must have a closed Kempe loop as without a Kempe loop, we would get 3 Kempe forests and so arboricity 3. The figure shows an example of a colored 2-torus. There are lots of closed Kempe chains. To the right we see a projective plane with a 5 coloring. The Kempe chain construction does not work any more. In the picture we colored the Kempe chains of the first 4 colors.

2. The theorem

2.1.

The proof of the following theorem makes heavy use of the 4 color theorem [63, 2].

Theorem 1 (Three Tree Theorem).

The arboricity of any 2-sphere is 33.

2.2.

The arboricity of a sub-graph HH of GG is smaller or equal than the arboricity of GG and any 4-connected planar graph is contained in a maximally planar 4-connected graph and so a 2-sphere. For non-4-connected graphs, cover the 4-connected components forests and glue the forests together. The following theorem is mentioned in [5] as Exercise 21.4.6 based on results of Frank, Frank and Gyarfas and using the Nash-William theorem.

Corollary 1 (Arboricity of planar graphs).

The arboricity of any planar graph is 3\leq 3.

2.3.

We prove the stronger result, telling that any 2-sphere different from a prism can be neatly covered with forests. A neat forest cover is a cover by 3 forests such that every triangle hits each of the three forests in exactly one point. If GG is a 22-sphere, define the upper line graph as the graph which has the edges of GG as vertices and where two are connected if they are in a common triangle. This is a 4-regular graph so that the chromatic number is less or equal to 44. We will see that the chromatic number of the upper line graph is 3\leq 3, the coloring given by the Kempe chains. Just to compare, the lower line graph or simply line graph connects two edges, if they intersect.

2.4.

The proof of the theorem follows from two propositions. The first gives a lower bound:

Proposition 1.

The arboricity of any 22-manifold is larger than 22.

Proof.

The Euler-Poincaré formula tells χ(G)=b0b1+b2=1+b2b1\chi(G)=b_{0}-b_{1}+b_{2}=1+b_{2}-b_{1} so that χ(G)=2b1\chi(G)=2-b_{1} or χ(G)=1b1\chi(G)=1-b_{1}, pending on the orientability of GG. Use the Euler gem formula |V||E|+|T|2|V|-|E|+|T|\leq 2 [54] as well as the Dehn-Sommerville relation 3|T|=2|E|3|T|=2|E| to get |V||E|/32|V|-|E|/3\leq 2 or |E|3|V|6|E|\geq 3|V|-6. If there were two spanning trees covering all the edges, then the Nash-Williams formula gives the bound |E|2|V|2|E|\leq 2|V|-2. This is incompatible with |E|=3|V|6|E|=3|V|-6 for |V|>4|V|>4 as plotting the functions |E|=2|V|2,|E|=3|V|6|E|=2|V|-2,|E|=3|V|-6 shows. ∎

2.5.

The second proposition gives an upper bound:

Proposition 2 (3 trees suffice).

A 22-sphere GG can be covered with 33 trees.

2.6.

This is exercise 21.4.6 in [5] based on work of A. Frank, A. Gyarfas and C. Nash-Williams. We will prove something slightly stronger in that we also can assure that all of the three trees intersects any of the triangles, provided that the graph GG is not a prism.

2.7.

We were not aware of previous work on the arboricity of planar graph before the google AI agent ”Bard” told us about it in a personal communication. Since the agent had hallucinated during that chat with other claims like that the arboricity of a cyclic graph CnC_{n} is 11, we did not take it seriously, but it prompted us to hit the literature again and let us to find the exercise in [5]. We are not aware of a publication proving that planar graphs have arboricity 3\leq 3. We will come back to this.

2.8.

We will establish the proposition standing on the shoulders of the 4-color theorem. There is a discrete contraction map to shrink colored spheres. It uses edge collapses and kite collapses. It can be seen as a ”discrete heat flow” because edge collapse removes curvature 1/31/3 vertices and kite collapses reduce and merge vertex degrees which tends to smooth out curvature. It in general lowers places with large positive curvature and increases places with low negative curvature.

2.9.

We actually will prove a stronger theorem. A prism graph is a 2-sphere that is the suspension of a cycle graph. (In general, the join of a kk-sphere and a ll-sphere is a (k+l+1)(k+l+1)-sphere. The join with a 0-sphere is a suspension. We add some remarks about the sphere monoid in an appendix.)

2.10.

A neat 3-forest cover of a 2-sphere is a partition of GG into 3 forests A,B,CA,B,C such that every triangle tTt\in T intersects every of these graphs. The existence of a neat 3-forest cover is equivalent to a neat 4-coloring, a 44-coloring for which there are no Kempe chains. We will show:

Theorem 2 (Existence of neat forests).

Every 2-sphere GG different from a prism admits a neat 3 forest partition.

2.11.

We can now upgrade the 4 color theorem. Of course, to prove this, we will take the 4-color theorem for granted.

Corollary 2 (Neat 4 color theorem).

Every 2-sphere GG different from a prism admits a neat 4 coloring.

Here is some indication that the “stronger 3-forest-suffice” result is hard:

Corollary 3.

The neat 4 color theorem is equivalent to the neat 3 forest theorem.

As pointed out before, it is possible prove the existence of 3 forests (not necessarily neat) without the 4-color theorem (dealing with not necessarily neat colorings). But so far, we could only see this as an excercise 21.4.6 in [5] which builds on various research of directed graphs.

3. Proof that 3 trees suffice

3.1.

In this section, we give the proof of the proposition provided we have a geometric contraction map ϕ\phi on a class 𝒳\mathcal{X} of colored 22-spheres with Kempe loops. more details of the evolution will be given in the next section. The main strategy is a descent-ascent argument: reduce the area of a given colored 2-sphere G𝒳G\in\mathcal{X} until it is small enough that we can recolor it with a neat coloring. The neat coloring can then be lifted by ascent back to the original graph. The smallest sphere that is reached, the octahedron, can not be recolored neatly but before reaching it, we can recolor.

3.2.

If we can recolor some ϕn(G)\phi^{n}(G) neatly, then this color can be lifted back to GG so that also GG admits also a neat coloring. The strategy is to prove that every colored 2-sphere with a Kempe loop that is not a prism can be compressed. This compression map ϕ\phi reduces the area of the sphere and is possible if we have a Kempe loop present and not a prism. There is then an explicit graph ϕ1(Pn\phi^{-1}(P_{n} that is reached just before entering the class of prism PnP_{n}. These pre-prisms are already large enough to be colored neatly. It follows that also GG can be recolored neatly. We have built the ladder with a non-neat color function (assuring the existence of Kempe loops and so allowing to define the map). But we climb the ladder back with the neat color function.

3.3.

Any 22-sphere GG is a planar graph. By the 4-color theorem, there is a vertex 4-coloring f:V{1,2,3,4}f:V\to\{1,2,3,4\} which we simply call a 4-coloring. A colored sphere (G,f)(G,f) is a 22-sphere G=(V,E)G=(V,E) equipped with a 4-coloring ff. We also equip it with a fixed orientation. If (a,b)(a,b) is an edge so that f(a)=1,f(b)=2f(a)=1,f(b)=2, call it an edge of type (12)(12). The vertex coloring ff produces an edge coloring F:E{A,B,C}F:E\to\{A,B,C\} using the rules

  • color the edges of type (12)(12) and (34)(34) with CC

  • color the edges of type (13)(13) and (24)(24) with BB

  • color the edges of type (23)(23) and (14)(14) with AA

Sub-graphs of GG with edge colors are simply called by their color name. Each of the sub-graphs A,B,CA,B,C are now graphs which are vertex colored by 2 colors. They are called Kempe chains [22, 62, 18]. The graphs A,B,CA,B,C are 11-dimensional, meaning that they are triangle free.

3.4.

The sub-graphs A,B,CA,B,C are not forests in general. Closed curves in AA or BB or CC are called Kempe loops. If there are no Kempe loops, then all A,B,CA,B,C are forests and we do not have to work any further as we have then already a neat 3-cover. We will therefore define the map ϕ\phi only on colored spheres which have Kempe loops. Actually, the presence of a Kempe loop is important. In general, we can not shrink a colored sphere in such a way that we have compatibility with the coloring.

3.5.

Look at the set 𝒳\mathcal{X} of colored oriented 2-spheres (G,f)(G,f) for which ff has at least one Kempe loop, meaning that there are closed loops of the same edge color. This set is not empty, as it contains any colored sphere (G,f)(G,f), for which ff is a 3-color function (These graphs were characterized by Heawood as the class of Eulerian graphs, graphs for which every vertex degree is even). The set also contains prisms. Two different Kempe loops can intersect, either with different color or the same color: a 1-2 Kempe chain can intersect for example the 3-4 Kempe chain. A 3-colorable graph is an example, where every unit sphere S(v)S(v) is a Kempe loop.

Proposition 3 (Existence of a geometric evolution).

There exists a map ϕ:𝒳𝒳\phi:\mathcal{X}\to\mathcal{X} that reduces the area of GG, as long as long as GG is not an octahedron. The octahedron is considered a fixed point of ϕ\phi. If ϕn(G)\phi^{n}(G) is a prism, then ϕn1(G)\phi^{n-1}(G) admits a neat coloring.

3.6.

This proposition is proven in the next section and establishes that we have a neat coloring on GG and so a neat forest cover establishing especially “three trees are enough”. The reason is that the evolution will reach a prism and one step before is a colored 2-sphere (G,f)𝒳(G,f)\in\mathcal{X} that allows for a 4-coloring without Kempe loops. In order to see that we have to be able to define the map in each case, where we have a Kempe loop and GG is not an octahedron. Then we have to see that we can reduce prism graphs and that so the only possible GG for which ϕ2(G)\phi^{2}(G) is the octahedron can be recolored to have no Kempe loops.

Refer to caption
Refer to caption
Figure 3. This graph graph GG has the property that ϕ(G)\phi(G) is a prism. This graph has colorings with Kempe chains (left) and neat colorings without Kempe chains (right). The neat coloring can then be lifted to all the previously obtained 2-spheres ϕn(G)\phi^{-n}(G).

4. The map

4.1.

We assume that GG is a 2-sphere that is not a prism. The map ϕ\phi first remove degree-44 vertex b edge collapse, provided there are degree 4 vertices. This lowers the number of peak positive curvature points and is done by collapsing a Heawood wheel. If no degree 4 vertices exist any more, then ϕ\phi collapses a Heawood kite. A Heawood kite (H,f)(H,f) is a colored kite graph HH, where the coloring is such that the two unconnected vertices having the same color. The kite collapse will increases curvature of some vertices and decreases curvature at others. The main difficulty will be to show that except in the prism case, but with a Kempe loop present, there always exists a Heawood kite. Not all colored spheres have Heawood kites. But we will see that all colored spheres with a Kempe loop must have a Heawood kite. The collapse of a Heawood kite might introduce again degree 4 vertices, but we remain in the class of 2-spheres. Each step ϕ\phi reduces the area of GG by at least 22.

4.2.

A Heawood leaf is a vertex in a Kempe chain that has only vertex degree 11 within that chain. A Kempe seed is a vertex in a Kempe chain that is isolated, not connected to any other vertex of the chain. A Kempe tree either has a leaf or a seed. A 11-dimensional graph that has no leaf and no seed must consist of a disjoint union of loops, possibly joined by connections. It is 11-dimensional network with vertex degrees all larger than 11.

4.3.

A Heawood wheel is a unit ball B(v)B(v) of a degree-4 vertex vv. Degree 4 vertices are the points in GG with maximal curvature 14/6=1/31-4/6=1/3. A Heawood wheel has 5 vertices and is also called a wheel graph W5W_{5}. Because we can only use 3 colors in the unit sphere S(v)S(v) of such a graph, there are by the pigeon-hole principle two vertices in S(v)S(v) which have the same color. (it is also possible that the unit sphere S(v)S(v) has only 2 colors, in which case S(v)S(v) is a Kempe chain). These vertices a,ba,b with f(a)=f(b)f(a)=f(b) have to be opposite to each other.

4.4.

We have a colored 2-sphere (G,f)(G,f), where G=(V,E)G=(V,E). Define the subset V4VV_{4}\subset V of vertices in GG which have degree 44. The sub-graph of GG generated by V4V_{4} is called G4G_{4}. A linear forest is a forest in which every tree is a linear graph or a seed K1K_{1}, a graph without edges. Here is a major reason why prism graphs are special:

Lemma 1.

If GG is a 2-sphere different from a prism, then G4G_{4} is a linear forest.

Proof.

If G4G_{4} contains a cyclic graph CnC_{n}, then GG must be a prim graph. If G4G_{4} contained a triangle, then it must be the octahedron (again a prism graph). If G4G_{4} contained a branch point (a star graph with 3 or more points), then again, it would have to be the octahedron. ∎

4.5.

Given a connected component in G4G_{4}, it is the center line QQ of a Kempe diamond. We call its suspension a Kempe diamond Q+{a,b}Q+\{a,b\}. There are now two possibilities. Either f(a)=f(b)f(a)=f(b) or f(a)f(b)f(a)\neq f(b). In the case f(a)=f(b)f(a)=f(b), we just remove the diamond by identifying a,ba,b. All the triangles in that diamond will disappear. We still can get a degree 4 vertex like that but we have reduced the number of vertices. Now go back to the beginning, of this stage. After finitely many steps, we will have either only degree 4 vertices meaning an octahedron or then reduced the number of degree 4 vertices by 11. Repeat this step until no degree 1 vertices exist

4.6.

In the case when f(a)f(b)f(a)\neq f(b), we necessarily have that the center line QQ is a Kempe chain with only two colors. If we are not in the prism case, we can collapse wheels one by one until the center line has length 0 or 11. After doing so, no degree 4 vertices are left in distance 1 of the remaining center edge or point.

4.7.

All these reductions were compatible with the coloring. It can happen that a Kempe loop is removed like for example if GG contains a Kempe loop of length 4. Collapsing that Kempe wheel removes that loop. The coloring ff of GG can be adapted to a coloring ϕ(f)\phi(f) of ϕ(G)\phi(G). We now establish that under the condition of a Kempe chain and if GG is different from a prism, there must be Heawood kite.

Lemma 2.

Given a colored sphere (G,f)(G,f) with a Kempe loop and where GG is different from a prism. Then there is at least one Heawood kite.

Proof.

A Kempe graph CC intersected with the 2-ball in the interior of a minimal loop must either be empty or have a Heawood leaf. ∎

Lemma 3.

The existence of a Heawood leaf in the interior of UU forces the existence of a Heawood kite.

Proof.

We can assume without loss of generality that we deal with a minimal 121-2 loop LL and that we have a leaf ending with a vertex vv of color 11 (the case where it ends with 22 being similar). The unit sphere of vv in CC contains only one vertex. Call it ww. Look at the unit sphere S(v)S(v) in GG. It contains ww with color 22 and all other vertices have color 3,43,4. So, there are two vertices of the same color. This produces a Heawood kite. ∎

4.8.

Under the assumption that no degree 4 vertices exist, we can now collapse a Heawood kite. A collapse reduces the vertex degrees of two of the vertices by 1. The deformation preserves the presence of a Kempe loop. The conclusion is that unless are in the Octahedron case, there is a smaller colored sphere (H,g)(H,g).

Lemma 4 (3 neat forests produce a 4 colored gem).

Let GG be a 2-sphere. If GG has a neat 3 forest cover, then it has a neat 4-coloring.

Proof.

Assume A,B,CA,B,C are the forests giving the colors attached to the edges of GG. Think of a 44-coloring is a gauge field V4={0,1,2,3}V\to\mathbb{Z}_{4}=\{0,1,2,3\}. We attach now a connection, by assigning to every edge an element in the subgroup of the symmetric group S3S_{3} as the presentation {A,B,C,A2=B2=C2=ABC=1}\{A,B,C,A^{2}=B^{2}=C^{2}=ABC=1\}. This is an Abelian group with 4 elements. Now start coloring the vertices by picking a vertex vv and attaching to it the color 040\in\mathbb{Z}_{4}. Now use the connection to color all of the neighbors. This produces a coloring. The only problem we could encounter is if we have a closed loop of the same color summing up to something non-zero. But this is excluded by having trees. ∎

4.9.

This article has shown that if we can 4-color a 2-sphere GG neatly, we can constructively cover GG with 3 forests. In applications, we want to do this fast. To speed things up, it helps to assume that the coloring has the property that every ball B2(x)B_{2}(x) of radius 22 contains all 4 colors. If not, change some of the vertices S(x)S(x) with the forth color. This preconditioning breaks already many closed Kempe chains. An extreme case with many Kempe chains is if GG is Eulerian. By a theorem of Heawood, GG can then be colored with 33 colors: just start with a point, color the unit sphere, and then follow the triangles to color the entire graph. In the case when the range of ff has 33 points, every unit sphere is a Kempe chain.

Corollary 4.

The construction of 3 neat forests is at least as hard as constructing a neat 4-coloring.

Proof.

Tree data (G,F)(G,F) with 3 neat tree coloring FF produces immediately a 44 vertex coloring ff. Since the 4-coloring is hard, also the neat 3-tree covering is hard in the sense of complexity. ∎

4.10.

Constructing a general 3-tree covering is computationally simpler. We are not aware however of a procedure that upgrades any “forest 3 cover” to a “neat forest 3-cover”. What we have shown here is that we can upgrade a 4-coloring to a neat 4-coloring.

5. Footnotes

5.1.

The arboricity of a 0-dimensional graph by definition is 11. Seeds are considered trees too and a graph without edges is a collection of seeds, still a forest. This assumption matches the forest theorem, telling that the number of rooted forests in a graph is det(L+1)=1{\rm det}(L+1)=1 and the number of rooted trees in a graph is Det(L){\rm Det}(L), where LL is the Kirchhoff matrix of GG which for a zero dimensional graph is the 0 matrix. See [28]. The arboricity for H=K1H=K_{1} should be declared to be 11 too even so |E|/(|V|1)|E|/(|V|-1) technically does not make sense. In the disconnected case, the forest arboricity (the usual book definition) is not the same than the tree arboricity (which we did not use as a definition). In the connected case, these are the same notions, as indicated in the first section and in an appendix. The arboricity of a disjoint union of graphs is the maximum of the arboricity of the individual connected compoentns.

5.2.

Covering problems with trees can also focus on packing with the same ”tile”. Most famous is Ringel conjecture from 1963 which told that any tree with nn edges packs 2n+12n+1 times into K2n+1K_{2n+1}. A tree of length 11 for example fits 33 times into K3K_{3}, a tree of length 22 fits 55 times into K5K_{5}. The arboricity of K5K_{5} is 33. There are 3 trees that do the covering. This has been proven for large nn [51].

5.3.

In chapter 5 of [8] there is an exercise linking chromatic number with arboricity. Here it is: Find a function f such that every graph of arboricity at least f(k) has colouring number at least k, and a function g such that every graph of colouring number at least g(k) has arboricity at least k, for all kk\in\mathbb{N}. We show here that for 2-spheres, there is an explicit relation. In general, given a coloring with nn colors, there is a tree cover with n(n1)/2n(n-1)/2 trees. If we have kk forests, we can color each forest with 2 color so that 2k2k is an upper bound for the chromatic number. For general graphs there is no universal relation between chromatic number and arboricity.

5.4.

The classification of discrete 2-manifolds agrees with the classification of 2-dimensional differentiable manifolds. In a combinatorial setting we want to avoid the continuum. The argument that links the discrete with the is a geometric realization of the simplicial complex of GG is a 2-manifold. The geometric realization of every ball B(v)B(v) is a wheel graph which has a two-dimensional ball as geometric realization. The structure and incidence of the unit balls provides an explicit atlas for a piecewise linear 2-manifold which then could be smoothed out to have a smooth 2-manifold. Finite discrete geometry is close to PL (piecewise linear) geometry. Topological manifolds are a much larger and much different category: a double suspension of a homotopy 3-sphere for example produces there a 5-sphere. This is not possible in the PL category or the discrete.

5.5.

The question of characterizing spheres in the continuum had been pursued by Poincaré already. The thread was taken in a finitist setting by Hermann Weyl during the “Grundlagen crisis” in mathematics. A combinatorial description based on cardinalities of simplices does not work because there are complexes with the same f-vector, where one is a sphere and the other is not. The existence of holonomy spheres already constructed by Poincaré thwarted any attempts of a cohomological characterization of spheres Weyl’s problem was solved using homotopy or contractibility. A fancy version is the solution of the Poincaré conjecture that characterizes dd-spheres using homotopy for d2d\geq 2.

5.6.

Various characterizations of spheres have been proposed in discrete, combinatorial settings. The definition used here was spearheaded in digital topology by Evako [10, 20]. Forman’s discrete Morse theory [13, 15, 14] characterized spheres using Reeb’s notions, [40], seeing them as manifolds admitting a Morse function with exactly 2 critical points. The Lusternik-Schnirelmann picture [21] is to see spheres as manifolds of characteristic 2, meaning that it can be covered by two contractible balls. In two dimension, Kuratowski’s theorem identifies 2-spheres as 2-manifolds that are planar. More precisely, Whitney would have characterized a 2-sphere as a maximally planar 4-connected graph different from K4K_{4}. 22-spheres are the class of graphs among planar graphs which are hardest to color.

5.7.

Spheres can also be characterized as manifolds which have trivial homotopy type and for 3-spheres as 3-manifolds which are simply connected. Poincaré’s conjecture is now Perelmann’s theorem. The deformation ϕ\phi is a Mickey Mouse Ricci flow and is certainly motivated as such because we get rid of large curvature parts by removing cross wheels and reduce negative curvature parts by collapsing kites. In higher dimensions, a continuum deformation argument should be able to prove a discrete Perelmann theorem. But deformations in higher dimensions most likely leads to subtle bubble-off difficulties as in the continuum.

5.8.

To see a hexahedron (=cube) or dodecahedron as 2-dimensional sphere with Euler characteristic 22 would require us to go beyond simplicial complexes and use the larger category of cell complexes, a category, where 2-dimensional cells can be attached to already present 1-dimensional cyclic graphs. This requires care. If we would declare fr example every 4-loop in a hexahedron graph to be a “cell”, we would have much too many faces. Historically, the continuum was a guide seeing polyhedra as triangulations of classical spheres. The confusion was already evident to the time of Euler. See [49, 54]. Kepler polyhedra for example are regular polyhedra of Euler characteristic different from 2. An other larger category in which one can work is the category of Δ\Delta-sets, a pre-sheaf over a simplex category and generalizing simplicial complexes. This is a finite topos and would be the ultimate finite category to work in. But graphs are much more intuitive and accessible.

5.9.

The history of the 4-color theorem is discussed in chromatography textbooks like [57, 3, 53, 6, 64, 12]. Kempe’s proof [22] from 1879 had been refuted 1890 [18] but some of his ideas survived. Already in the same issue of the American Journal of Mathematics, William E. Story pointed out that Kempe’s proof needs to be made rigorous [62]. Kempe’s descent idea for the proof allows to see quickly that one can color any 2-sphere by 5 colors. More geometric approaches have been spear headed by Fisk [11, 30, 33, 38]. Every d-manifold can be colored by 2d+22d+2 colors [45]. Again, we have to stress that this is a completely different set-up as the Heawood story Ringel and Youngs finished showing that on a genus g=(n3)(n4)/12g=\lceil(n-3)(n-4)/12\rceil there is a complete subgraph KnK_{n} embedded. For n=7n=7, this gives the famous torus case g=1g=1 of Heawood. But this embedding of a 66-dimensional graph in a 22-dimensional surface is not what we call a discrete manifold [19, 55] as the unit sphere of a vertex in the graph K7K_{7} is the 55-dimensional K6K_{6}.

5.10.

For any coloring f:Vf:V\to\mathbb{R}, the Poincaré-Hopf index i(v)=1χ(S(v))i(v)=1-\chi(S^{-}(v)) for the sub-graph S(v)S^{-}(v) generated by {wS(v),f(w)<f(v)}\{w\in S(v),f(w)<f(v)\} is an integer-valued function such that vVi(v)=χ(G)\sum_{v\in V}i(v)=\chi(G) an identity that holds for all finite simple graphs. See [16, 26, 43, 42]. [The maximum vv of ff has Sf(v)=S(v)S^{-}_{f}(v)=S(v). The valuation formula χ(AB)=χ(A)+χ(B)χ(AB)\chi(A\cup B)=\chi(A)+\chi(B)-\chi(A\cap B) produces χ(G)=χ(Gv)+χ(B(v))χ(S(v))=χ(Gv)+1χ(S(v))=χ(Gv)+i(v)\chi(G)=\chi(G-v)+\chi(B(v))-\chi(S(v))=\chi(G-v)+1-\chi(S^{-}(v))=\chi(G-v)+i(v), allowing induction.] A vertex vVv\in V is a critical point of ff if S(v)S^{-}(v) is not contractible. The Reeb characterization of spheres is a manifolds for which there is a coloring with exactly two critical points.

5.11.

The expectation of the Poincaré-Hopf index if(v)i_{f}(v) over natural probability spaces like the uniform counting measure on colorings is the Levitt curvature K(v)=1k=0(1)k|fk(S(v))|k+1K(v)=1-\sum_{k=0}(-1)^{k}\frac{|f_{k}(S(v))|}{k+1} [50, 24, 35]. For K4K_{4}-free graphs, it leads to 1f0(S(v))/2+f1(S(v))/31-f_{0}(S(v))/2+f_{1}(S(v))/3 which for 2-manifolds simplifies with f1(S(v))=f0(S(v))f_{1}(S(v))=f_{0}(S(v)) to 1f0(S(v))/61-f_{0}(S(v))/6, a curvature which goes back to Victor Eberhard [9] and was used in graph coloring arguments already at the time of Birkhoff and heavily for the 4 color theorem program [4], as Gauss Bonnet χ(G)=2\chi(G)=2 implies that there must be some vertices of degree 44 or 55, points of positive curvature.

5.12.

The integral geometric approach to Gauss-Bonnet χ(G)=vVK(v)\chi(G)=\sum_{v\in V}K(v) leads in the continuum for even-dimensional manifolds MM to the Gauss-Bonnet-Chern result. See [27, 31, 39, 44]. This can be seen by Nash embedding MM in to an ambient Euclidean space EE, using that Lebesgue almost all unit vectors uEu\in E, the function fu(x)=xuf_{u}(x)=x\cdot u is a Morse function. The Euler characteristic χ(M)=vMi(v)\chi(M)=\sum_{v\in M}i(v) is a finite sum. The expectation K(v)=E[i(v)]K(v)={\rm E}[i(v)] over the probability space Ω=S(0)={uE,|u|=1}\Omega=S(0)=\{u\in E,|u|=1\} a rotationally invariant probability measure μ\mu on S(0)S(0) (parametrizing the Morse functions). By a result of Hermann Weyl, this must be the Gauss-Bonnet-Chern curvature.

5.13.

Trees are useful data structures within networks. There are many reasons: a network equipped with a tree for example, gives fast search as missing loops avoids running in circles. The data are located in the leaves of the tree, nodes with only one neighbor Decision trees in network of decisions is an other example providing causal strategies to reach goals located at leaf positions in the tree. The Nash-Williams result shows that the arboricity measures a network density. Indeed |E|/(|V|1)|E|/(|V|-1) is close to |E|/|V||E|/|V| which is half the average vertex degree of the network by the Euler handshake formula. The arboricity is essentially the maximal density of subgraphs of GG.

5.14.

If we know a finite set of rooted trees covering the graph, we know the entire network. The reason is simple: given two nodes v,wv,w in the network, then v,wv,w are connected if and only if in one of the rooted trees, the distance between vv and ww is 11. The arboricity of a graph the minimal number of trees which are needed to cover all connections. Within a network it can serve as some sort of dimension. There are lots of interesting questions, like how many trees there can be for a neat coloring of a 2-sphere.

5.15.

We have already called a graph G=(V,D)G=(V,D) to be contractible if there is a vertex vVv\in V such that the unit sphere S(v)S(v) as well as the graph generated by V{v}V\setminus\{v\} are both contractible. The Lusternik-Schnirelmannn category cat(G){\rm cat}(G) is the minimal number of contractible sub-graphs covering GG. One has cat(G)arb(G){\rm cat}(G)\leq{\rm arb}(G) in the positive dimensional connected graph case because every tree is contractible. We will see that for a 2-sphere cat(G)=2{\rm cat}(G)=2 and always arb(G)=3{\rm arb}(G)=3. With the arboricity of a 0-dimensional graph defined to be 1, the inequality cat(G)arb(G){\rm cat}(G)\leq{\rm arb}(G) only works for connected graphs.

5.16.

Whitney once characterized 2-spheres using duality. There is a theorem of Karl von Staudt [61] telling that GG and G^\hat{G} have the same number of spanning trees, because every spanning tree for GG produces a spanning tree in G^\hat{G}. The dual graph however is only 1-dimensional and has smaller arboricity than GG if GG is a 2-sphere. See [47].

5.17.

Prims pointed towards the arithmetic of graphs. The Zykov monoid can be group completed and together with the large product (introduced by Sabidussi [58]) gies the Sabidussi ring. It is isomorphic to the Shannon ring in which the disjoint union is group completed and where the Shannon multiplication (strong multiplication) is used [59]. In the Shannon ring, the connected components are the additive co-prodoct primes. In the Sabidussi ring, the graphs which are not the join of two non-zero graphs are the additive Zykov primes. Spheres form a sub-monoid of the Zykov monoid and most spheres are prime. In dimension 22 there are very few non-primes: they are the prisms. [37, 46, 41].

5.18.

For Dehn-Sommerville relations for any dd-sphere, see [60]. See also [36] for a Gauss-Bonnet approach or [40]. They are quantities which are both invariant and explode under Barycentric refinements. The only way that this can happen is that they are 0. They can be obtained from the eigenvectors of the Barycentric refinement operator. See [32, 34]. See [48] for the tree forest ratio. The Nash-Williams density f1/(f01)f_{1}/(f_{0}-1) measures some sort of density of the network, similarly as the average simplex cardinality [23] or other functionals [25, 29]. This could lead to interesting variational problems like which dd-manifolds produce the largest Nash-Williams density. We hope to work on this a bit more in the future.

5.19.

Does cohomology determine the arboricity for manifolds? In other words, if two graphs have the same Betti vector (b0,b1,)(b_{0},b_{1},\dots) and the same dimension does this imply that the arboricity is the same? This is not true: the arboricity of a sufficiently large dd-sphere can be estimated by the Perron-Frobenius eigenvector of the Barycentric refinement operator and must be at least to a constant cdc_{d} that grows exponentially. But the arboricity of a suspension of GG is less or equal than the arboricity of GG plus 22. In higher dimension, the arboricity spectrum gets wider and wider within the same class of graphs. There are dd-spheres that can be colored with 2d2d trees and there are spheres with arboricity growing like cd=1.70948dc_{d}=1.70948^{d}. We hope to explore this a bit more in the future. Especially interesting is the question whether the constants cdc_{d} define the upper bound. We currently believe that this is the case.

5.20.

We believe that the arboricity of any 2-manifold is maximally 4. We have no proof of that but it would follow if the arboricity of a dd-manifold were the smallest integer larger than the Barycentric arboricity constant cdc_{d} which is in the case d=2d=2 equal to 33. There are examples like tori or Klein bottles with arboricity 4. In particular, we believe that the arboricity of any 3-manifold is smaller or equal than the smallest integer larger than 13/2=6/513/2=6/5 which is 77. We know using Barycentric refinement that there are 3-spheres with arboricity 77. The suspension of the octahedron, the smallest 3-sphere has Nash-Williams ration 24/724/7 and so arboricity 44. We see that there are 3-spheres of arboricity 4,5,6,74,5,6,7. This is completely different than in the 2-dimensional case considered here, where the three theorem assures that the arboricity of a 2-sphere is locked to be 33. For 5-spheres, the range of possible arboricity is 5,6,,135,6,\dots,13. The lower bound grows linearly with the dimension, while the upper bound grows exponentially with the dimension.

5.21.

If cdc_{d} is integer, that it can be by 1 larger. For d=2d=2, where cd=3c_{d}=3, this happens for 2 tori as in dimension 22 the Barycentric refinement limit is 33 and the arboricity can be 44. The first constants are {c1,c2,c3,,c10}\left\{c_{1},c_{2},c_{3},\dots,c_{10}\right\} are

{1,3,132,252,75133,181345,308796543956,663513754628,4880845874762335501595,200710991888559734735}.\left\{1,3,\frac{13}{2},\frac{25}{2},\frac{751}{33},\frac{1813}{45},\frac{3087965}{43956},\frac{6635137}{54628},\frac{488084587476}{2335501595},\frac{200710991888}{559734735}\right\}\;.

This suggests that the correct maximal arboricity numbers for dd-spheres in the case d=0,1,,10d=0,1,\dots,10 are

{1,2,3,7,13,23,41,71,122,209,359}\{1,2,3,7,13,23,41,71,122,209,359\}

We do not know whether there are integer cdc_{d} besides c1,c2c_{1},c_{2}. There are lots of open questions. Are there 3-spheres of arboricity 8 for example. We believe this is not possible.

Appendix: Arboricity

5.22.

The current public material about the arboricity of graphs is rather sparse. The functional is mentioned in [17, 5]. Most graph theory textbooks do not cover it. Both in [17] and [5], it is defined as the minimal number of forests partitioning the graph. This has been mentioned in the introduction. Lets elaborate about it a bit more.

5.23.

A tree is a triangle free graph which is contractible. The smallest tree is K1K_{1} and also called a seed. Note that the zero graph is not a tree. Every tree has Euler characteristic χ(G)=|V||E|=1\chi(G)=|V|-|E|=1. A forest is a triangle free graph for which every connected component is a tree. A forest has Euler characteristic χ(G)=b0\chi(G)=b_{0}, which is the number of connected components of the graph. A forest can be contracted to a 0 dimensional manifold. A spanning tree in a connected graph GG is a subgraph that is a tree and which has the same vertex set than GG. A spanning forest is a forest which has the same vertex set than GG.

Lemma 5.

If GG is connected, the following numbers are the same:
a) The arboricity T1(G)T_{1}(G), the minimal number of forests partitioning GG.
b) The minimal number T2(G)T_{2}(G) of trees covering GG.
c) The minimal number T3(G)T_{3}(G) of spanning trees covering GG

Proof.

This follows from the fact that every spanning tree is a tree and every tree is a forest and that every forest can be completed to a spanning tree in a connected graph. ∎

5.24.

If GG is not connected, the numbers T1(G),T2(G)T_{1}(G),T_{2}(G) are not the same. For a forest GG with 2 trees for example, we have T1(G)=1T_{1}(G)=1 while T2(G)=2T_{2}(G)=2. A related interesting concept is linear arboricity T4T2(G)T_{4}\geq T_{2}(G) which asks how many linear graphs partition the graph. A lower bound is [Δ/2][\Delta/2] (where [x] is the largest integer smaller or equal to x), where Δ(G)\Delta(G) is the maximal vertex degree. The linear arboricity conjecture asks whether [(Δ+1)/2][(\Delta+1)/2] is an upper bound for linear arboricity. If a regular graph GG of even vertex degree Δ\Delta gets a vertex removed, one gets a graph HH whose linear arboricity is Δ/2\Delta/2 if and only if the graph GG has a Hamiltonian decomposition.

5.25.

A graph is called contractible, if there exists vV(G)v\in V(G) such that S(v)S(v) and GvG\setminus v are both contractible. Contractible graphs are connected. Trees are contractible, forests that are not trees are not contractible. The Lusternik-Schnirelmann category (or simply category) of a graph is the minimal number of contractible graphs that cover the graph. Since trees are contractible, the category in general is smaller or equal than the number of trees covering GG. For a 2-sphere GG, the category of GG is 22, while the arboricity of GG is 33.

5.26.

If GG is a graph with ff-vector f=(f0,f1,,fd)f=(f_{0},f_{1},\dots,f_{d}), where fkf_{k} is the number of kk dimensional simplices (=cliques=faces) Kk+1K_{k+1} in GG. The graph G1G_{1} in which the complete subgraphs of GG are the vertices and where two are connected if one is contained in the other is called the Barycentric refinement of GG. If GG was contractible then G1G_{1} is contractible. If GG is a dd-sphere, then G1G_{1} is a dd-sphere. The graph Gn=(Gn1)1G_{n}=(G_{n-1})_{1} is the nn’th Barycentric refinement.

5.27.

The Barycentric refinement matrix AA is a (d+1)×(d+1)(d+1)\times(d+1) matrix with entries

Aij=i!S(j,i),A_{ij}=i!S(j,i)\;,

where S(j,i)S(j,i) are the Stirling numbers of the second kind. The vector AfAf is the f-vector of G1G_{1}. Define

cd=limnE(Gn)/(V(Gn)1)=limnf1(Gn)/f0(Gn).c_{d}=\lim_{n\to\infty}E(G_{n})/(V(G_{n})-1)=\lim_{n\to\infty}f_{1}(G_{n})/f_{0}(G_{n})\;.

Since the normalized ff-vectors f(Gn)/f(Gn)2f(G_{n})/||f(G_{n})||_{2} (||||2||\cdot||_{2} is the Euclidean norm), converge to the Perron-Frobenius vector we can compute the limits cdc_{d}. They grow like 1.70948d1.70948^{d}. From the Nash-Williams formula, we see that the arboricity of dd-spheres grows exponentially in dd.

Proposition 4.

a) If HH is a sub-graph of GG then arb(H)arb(G){\rm arb}(H)\leq{\rm arb}(G).
b) The arboricity of the disjoint sum of two graphs H,GH,G is max(arb(H),arb(H)){\rm max}({\rm arb}(H),{\rm arb}(H)).
c) There are graphs with a Kn+1K_{n+1} subgraph with arboricity cn\geq c_{n}
d) The arboricity of KnK_{n} is n/2{n/2} so that any graph with n vertices has arboricity n/2\leq{n/2}.
e) For a connected graph, the Lusternik Schnirelmann category of GG is a lower bound for the arboricity.

Proof.

a) This follows directly from Nash Williams. The inequality also can be derived from the definition and is used in proofs of [17].
b) This follows from the definition and that forests can be disconnected.
c) This follows from the Barycentric refinement picture and Nash-Williams.
d) For a complete graph K2nK_{2n} we can get the forests explicitly. See page 92 in [17].
e) This follows because for connected graphs, we can use the tree cover picture and because trees are contractible. ∎

Appendix: The sphere monoid

5.28.

The Zykov join GHG\oplus H of two graphs G,HG,H has V(G)V(H)V(G)\cup V(H) as vertex set E(G)E(H){(a,b),aV(G),bV(H)}E(G)\cup E(H)\cup\{(a,b),a\in V(G),b\in V(H)\} as edge set. It is dual to the disjoint union of graphs G¯+H¯¯\overline{\overline{G}+\overline{H}}, where H¯\overline{H} is the graph complement. The Zykov join G+S0G+S_{0} of GG with a 0-sphere is called the suspension of GG. The suspension of a dd-sphere is a (d+1)(d+1)-sphere.

5.29.

The union of all kk-spheres, where k(1)k\geq(-1) forms a sub-monoid of the Zykov monoid of all graphs. The Dehn-Sommerville monoid is a monoid strictly containing the sphere monoid but where the unit spheres are required to have Euler characteristic of the corresponding spheres but are Dehn-Sommerville manifolds of a dimension lower. Examples are disjoint unions of odd dimensional spheres or disjoint copies of two even dimensional spheres glued along north and south pole so that the unit spheres are either spheres or unions of spheres.

Lemma 6.

If GG is a pp-sphere and HH is a qq-sphere, then GHG\oplus H is a (p+q+1)(p+q+1)-sphere. The empty graph 0 is the zero element.

Proof.

Use induction and note that for vV(G)v\in V(G), the unit sphere is SGH(v)=SG(v)HS_{G\oplus H}(v)=S_{G}(v)\oplus H which by induction is a (p+q)(p+q)-sphere. Similarly the unit sphere for vV(H)v\in V(H) is a p+qp+q sphere. ∎

5.30.

A graph is a Zykov prime if it can not be written as G=HKG=H\oplus K, where H,KH,K are both non-zero graphs. A small observation about small non-prime spheres:

Lemma 7.

a) The only non-prime 1-sphere is C4C_{4}.
b) The only non-prime 2-sphere is a prism graph CnS0C_{n}\oplus S_{0}.
c) For every n6n\geq 6 there exists exactly one non-prime 22-sphere with nn vertices.
d) For every n8n\geq 8 the number of non-prime 33-spheres with nn vertices is 22 plus the number of 2-spheres with (n2)(n-2) vertices.

5.31.

The fact that the only non-prime 2-spheres are prisms makes them special. If we make an edge refinement of a prism along an edge in the equator, we get a larger prism. If we make an edge refinement along an edge hitting one of the poles, we get an almost prism, a graph which allows a 4 coloring without Kempe loops.

References

  • [1] M.O. Albertson and W.R. Stromquist. Locally planar toroidal graphs are 55-colorable. Proc. Amer. Math. Soc., 84(3):449–457, 1982.
  • [2] K. Appel and W. Haken. A proof of the four color theorem. Discrete Mathematics, 16:179–180, 1976.
  • [3] D. Barnette. Map Coloring, Polyhedra, and the Four-Color Problem, volume 9 of Dociani Mathematical Expositions. AMS, 1983.
  • [4] H-G. Bigalke. Heinrich Heesch, Kristallgeometrie, Parkettierungen, Vierfarbenforschung. Birkhäuser, 1988.
  • [5] J. Bondy and U. Murty. Graph theory, volume 244 of Graduate Texts in Mathematics. Springer, New York, 2008.
  • [6] G. Chartrand and P. Zhang. Chromatic Graph Theory. CRC Press, 2009.
  • [7] B. Chen, M. Matsumoto, J. Wang, Z. Zhang, and J. Zhang. A short proof of nash-williams’ theorem for the arboricity of a graph. Graphs and Combinatorics, 10:27–28, 1994.
  • [8] R. Diestel. Graph theory, volume 173 of Graduate Texts in Mathematics. Springer, 5th edition, 2016.
  • [9] V. Eberhard. Morphologie der Polyeder. Teubner Verlag, 1891.
  • [10] A.V. Evako. Dimension on discrete spaces. Internat. J. Theoret. Phys., 33(7):1553–1568, 1994.
  • [11] S. Fisk. Variations on coloring, surfaces and higher-dimensional manifolds. Advances in Mathematics, pages 226–266, 1977.
  • [12] S. Fisk. Cobordism and functoriality of colorings. Adv. in Math., 37(3):177–211, 1980.
  • [13] R. Forman. Combinatorial vector fields and dynamical systems. Math. Z., 228(4):629–681, 1998.
  • [14] R. Forman. Combinatorial differential topology and geometry. New Perspectives in Geometric Combinatorics, 38, 1999.
  • [15] R. Forman. A user’s guide to discrete Morse theory. Séminaire Lotharingien de Combinatoire, 48, 2002.
  • [16] L. Glass. A combinatorial analog of the Poincare index theorem. Journal of combinatorial theory, 15:264–268, 1973.
  • [17] F. Harary. Graph Theory. Addison-Wesley Publishing Company, 1969.
  • [18] P. J. Heawood. Map colour theorem. Quarterly Journal of Pure and Applied Mathematics, 24:332–338, 1890. Kempe refutal, Cabot Science PER 6669 Library Info.
  • [19] P. J. Heawood. Map-colour theorem. Proc. London Math. Soc. (2), 51:161–175, 1949.
  • [20] A.V. Ivashchenko. Graphs of spheres and tori. Discrete Math., 128(1-3):247–255, 1994.
  • [21] F. Josellis and O. Knill. A Lusternik-Schnirelmann theorem for graphs.
    http://arxiv.org/abs/1211.0750, 2012.
  • [22] A.B. Kempe. On the geographical problem of the four colours. Amer. Jour. Math., 2, 1979.
  • [23] O. Knill. The average simplex cardinality of a finite abstract simplicial complex. https://arxiv.org/abs/1905.02118, 1999.
  • [24] O. Knill. A graph theoretical Gauss-Bonnet-Chern theorem.
    http://arxiv.org/abs/1111.5395, 2011.
  • [25] O. Knill. Dimension and Euler characteristics of graphs.
    demonstrations.wolfram.com/DimensionAndEulerCharacteristicsOfGraphs, 2012.
  • [26] O. Knill. A graph theoretical Poincaré-Hopf theorem.
    http://arxiv.org/abs/1201.1162, 2012.
  • [27] O. Knill. On index expectation and curvature for networks.
    http://arxiv.org/abs/1202.4514, 2012.
  • [28] O. Knill. Cauchy-Binet for pseudo-determinants. Linear Algebra Appl., 459:522–547, 2014.
  • [29] O. Knill. Characteristic length and clustering.
    http://arxiv.org/abs/1410.3173, 2014.
  • [30] O. Knill. Coloring graphs using topology.
    http://arxiv.org/abs/1410.3173, 2014.
  • [31] O. Knill. Curvature from graph colorings.
    http://arxiv.org/abs/1410.1217, 2014.
  • [32] O. Knill. The graph spectrum of barycentric refinements.
    http://arxiv.org/abs/1508.02027, 2015.
  • [33] O. Knill. Graphs with Eulerian unit spheres.
    http://arxiv.org/abs/1501.03116, 2015.
  • [34] O. Knill. Universality for Barycentric subdivision.
    http://arxiv.org/abs/1509.06092, 2015.
  • [35] O. Knill. Gauss-Bonnet for multi-linear valuations.
    http://arxiv.org/abs/1601.04533, 2016.
  • [36] O. Knill. On a Dehn-Sommerville functional for simplicial complexes.
    https://arxiv.org/abs/1705.10439, 2017.
  • [37] O. Knill. On the arithmetic of graphs.
    https://arxiv.org/abs/1706.05767, 2017.
  • [38] O. Knill. Eulerian edge refinements, geodesics, billiards and sphere coloring.
    https://arxiv.org/abs/1808.07207, 2018.
  • [39] O. Knill. Constant index expectation curvature for graphs or Riemannian manifolds.
    https://arxiv.org/abs/1912.11315, 2019.
  • [40] O. Knill. Dehn-Sommerville from Gauss-Bonnet.
    https://arxiv.org/abs/1905.04831, 2019.
  • [41] O. Knill. More on numbers and graphs.
    https://arxiv.org/abs/1905.13387, 2019.
  • [42] O. Knill. More on Poincaré-Hopf and Gauss-Bonnet. https://arxiv.org/abs/1912.00577, 2019.
  • [43] O. Knill. A parametrized Poincare-Hopf theorem and clique cardinalities of graphs.
    https://arxiv.org/abs/1906.06611, 2019.
  • [44] O. Knill. On index expectation curvature for manifolds. https://arxiv.org/abs/2001.06925, 2020.
  • [45] O. Knill. Coloring Discrete Manifolds. https://arxiv.org/abs/2106.14374, 2021.
  • [46] O. Knill. Remarks about the arithmetic of graphs.
    https://arxiv.org/abs/2106.10093, 2021.
  • [47] O. Knill. Analytic torsion for graphs. https://arxiv.org/abs/2201.09412, 2022.
  • [48] O. Knill. The Tree-Forest Ratio. https://arxiv.org/abs/2205.10999, 2022.
  • [49] I. Lakatos. Proofs and Refutations. Cambridge University Press, 1976.
  • [50] N. Levitt. The Euler characteristic is the unique locally determined numerical homotopy invariant of finite complexes. Discrete Comput. Geom., 7:59–67, 1992.
  • [51] R. Montgomery, A. Pokrovskiy, and B. Sudakov. A proof of Ringel’s conjecture. Geom. Funct. Anal., 31:663–720, 2021.
  • [52] C. Nash-Williams. Edge-disjoint spanning trees of finite graphs. J. London Math. Soc., 36:455–450, 1961.
  • [53] O. Ore. The Four-Color Problem. Academic Press, 1967.
  • [54] D.S. Richeson. Euler’s Gem. Princeton University Press, Princeton, NJ, 2008. The polyhedron formula and the birth of topology.
  • [55] G. Ringel. Map Color Theorem. Grundlehren der mathematischen Wissenschaften. Springer Verlag, 1974.
  • [56] R. Ruononen. Graph theory. 2016. Translation of TUT course Graafiteoria.
  • [57] T.L. Saaty and P.C. Kainen. The four color problem, assaults and conquest. Dover Publications, 1986.
  • [58] G. Sabidussi. Graph multiplication. Math. Z., 72:446–457, 1959/1960.
  • [59] C. Shannon. The zero error capacity of a noisy channel. IRE Transactions on Information Theory, 2:8–19, 1956.
  • [60] D.M.Y. Sommerville. An introduction to the geometry of n dimensions. Methuen and Co. Ltd, 1929.
  • [61] K.G.C. Von Staudt. Geometrie der Lage. Nuernberg, 1847. page 21.
  • [62] W.E. Story. Note on the preceding paper: [on the geographical problem of the four colours]. Amer. Jour. Math., 2, 1979.
  • [63] W. Stromquest. The four color theorem for locally planar graphs. In Bondy and Murty, editors, Graph theory and related topics. Academic Press, 1979.
  • [64] R. Wilson. Four Colors Suffice. Princeton Science Library. Princeton University Press, 2002.
  • [65] A.A. Zykov. On some properties of linear complexes. (russian). Mat. Sbornik N.S., 24(66):163–188, 1949.