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

On Orthogonal Vector Edge Coloring

Allen Ibiapina MSc Ana Silva PhD11footnotemark: 1 ParGO, Departamento de Matemática, Universidade Federal do Ceará, Fortaleza, Brazil
Abstract

Given a graph GG and a positive integer dd, an orthogonal vector dd-coloring of GG is an assignment ff of vectors of d\mathbb{R}^{d} to V(G)V(G) in such a way that adjacent vertices receive orthogonal vectors. The orthogonal chromatic number of GG, denoted by χv(G)\chi_{v}(G), is the minimum dd for which GG admits an orthogonal vector dd-coloring. This notion has close ties with the notions of Lovász Theta Function, quantum chromatic number, and many other problems, and even though this and related metrics have been extensively studied over the years, we have found that there is a gap in the knowledge concerning the edge version of the problem. In this article, we discuss this version and its relation with other insteresting known facts, and pose a question about the orthogonal chromatic index of cubic graphs.

keywords:
Orthogonal vector coloring; cubic graphs; quantum chromatic number; planar graphs; Kocken-Specker sets.
t1t1footnotetext: Funded by Conselho Nacional de Desenvolvimento Científico e Tecnológico - CNPq, grants: 304576/2017-4; 437841/2018-9; 425297/2016-0; 401519/2016-3.emailsemailsfootnotetext: anasilva@mat.ufc.br; allenr.roossim@gmail.com

1 Introduction

We assume that the reader knows the basic terminology in Graph Theory and Linear Algebra; for an introduction on the subjects, we refer the reader to [26, 30]. All the graphs are considered to be finite and simple, unless said otherwise.

A proper coloring of a graph G=(V,E)G=(V,E) is a function f:Vf\colon V\to\mathbb{N} such that f(u)f(v)f(u)\neq f(v) whenever uvEuv\in E. Because we only work with proper colorings in this paper, from now on we call it simply a coloring; also, if ff uses kk colors, we refer to it as a kk-coloring. The chromatic number of GG, denoted by χ(G)\chi(G), is the minimum integer kk for which GG admits a kk-coloring; and if χ(G)=k\chi(G)=k, we say that GG is kk-chromatic. Deciding whether χ(G)k\chi(G)\leq k is \NP\NP-complete even when k=3k=3 and GG is a triangle-free graph with maximum degree at most 4 [21].

An orthogonal vector dd-coloring of GG (from now on called simply orthogonal dd-coloring) is an assignment ff of vectors of d\mathbb{R}^{d} to V(G)V(G) in such a way that adjacent vertices receive orthogonal vectors. This was originally defined in terms of the non-edges (i.e., non-adjacent vertices receive orthogonal vectors), and it was introduced to approach the Shannon capacity function, a problem arising in information theory [18]. Since then, it has been extensively studied (see [19], Chapter 9 for an overview), as well as many other notions stemming from it, as for example the quantum chromatic number [10, 1, 6].

One of the many investigated aspects when dealing with these representations is the minimum dimension for which it exists. Here, we call this parameter the orthogonal number of GG and denote it by π(G)\pi(G). Observe that, given a dd-coloring ff of GG, one can obtain an orthogonal dd-coloring from ff by associating each color with a vector in the canonical base of d\mathbb{R}^{d}. Therefore,

π(G)χ(G).\pi(G)\leq\chi(G). (1)

Little is known about the orthogonal number of a graph, and some stronger constraints have been imposed on the definition, as for instance, the faithful orthogonal dd-colorings [5, 3], and the general position orthogonal dd-coloring [20].

A natural and interesting question about the dimension is whether the inequality above is tight. Examples where the inequality is strict are not easy to find, and up to our knowledge, all the known examples are related to quantum mechanics and pseudo-telepathy games. In [1], they prove that most Hadamard graphs have this property, in [6] they give an explicit example in 18 vertices, and in [25] they give a more general result that produce smaller graphs than the Hadamard graphs (the smaller Hadamard graph GG having π(G)<χ(G)\pi(G)<\chi(G) has 1609 vertices). The latter article relates the concepts of orthogonal coloring and quantum coloring to Kochen-Specker sets, a notion used initially to prove the inadequacy of using determinism to model quantum mechanics [17].

A Kochen-Specker set is a set SdS\subseteq\mathbb{R}^{d} for which there is no function f:d{0,1}f:\mathbb{R}^{d}\rightarrow\{0,1\} such that, for every orthonormal bases BSB\subseteq S, we have uBf(u)=1\sum_{u\in B}f(u)=1. We mention that these sets are actually defined for general Hilbert spaces (we refer the reader to [14] for a fine explanation), but that the restriction made will be enough for our purposes. Also, the proof given below has appeared before in more general contexts, but we reproduce it here for the sake of clarity.

Given any set SdS\subseteq\mathbb{R}^{d}, the orthogonality graph of S, denoted by K(S)K(S), has SS as vertex set and is such that uu and vv are adjacent if and only if uu is orthogonal to vv.

Theorem 1.

Let SdS\subseteq\mathbb{R}^{d} be a Kochen-Specker set containing an orthonormal base of d\mathbb{R}^{d}. Then,

χ(K(S))>π(K(S))=d.\chi(K(S))>\pi(K(S))=d.
Proof.

Let G=K(S)G=K(S). First, observe that any orthogonal base of d\mathbb{R}^{d} is a clique in GG, and must receive orthogonal vectors in any orthogonal coloring of GG; hence, because SS contains an orthonormal base of d\mathbb{R}^{d}, we get π(G)d\pi(G)\geq d.

Now, suppose by contradiction that GG has a dd-coloring cc, and let f:d{0,1}f:\mathbb{R}^{d}\rightarrow\{0,1\} assign value 11 to every uu in color class 1, and 0 to the remaining vectors. Consider an orthonormal base BB of d\mathbb{R}^{d} contained in SS. Again, we know that BB is a clique of size dd in GG, and therefore must contain exactly one vertex of each color class. This implies that uBf(u)=1\sum_{u\in B}f(u)=1, contradicting the definition of a Kochen-Specker set. ∎

As already mentioned, finding a Kochen-Specker set is no trivial task, the first being presented as a proof of the famous Kochen-Specker Theorem [17] and containing 117 vectors. Now, it is known that any such set in 3\mathbb{R}^{3} and 4\mathbb{R}^{4} have at least 18 vertices [4, 23, 22]. It is also interesting to point out that in [14] the authors show a subset S4S\subseteq\mathbb{R}^{4} of the Kochen-Specker set presented in [24] for which |S|=17|S|=17 and 4=π(K(S))<χ(K(S))=54=\pi(K(S))<\chi(K(S))=5. Because, as mentioned before, every Kochen-Specker set in dimension 4 must have at least 18 vertices, we know that SS is not a Kochen-Specker set, thus showing that being a Kochen-Specker set is not a necessary condition to have π(K(S))<χ(K(S))\pi(K(S))<\chi(K(S)).

Now, given that finite Kochen-Specker sets are known, we get that a (simple) graph GG with χ(G)>π(G)\chi(G)>\pi(G) is also known. From this graph, it is possible to construct graphs where the distance between χ(G)\chi(G) and π(G)\pi(G) is as large as possible; it suffices to iteratively apply the join operation on two copies of GG. Given graphs G1,G2G_{1},G_{2}, the join of G1G_{1} and G2G_{2} is the graph obtained from G1G2G_{1}\cup G_{2} by adding all possible edges between V(G1)V(G_{1}) and V(G2)V(G_{2}); it is denoted by G1G2G_{1}\wedge G_{2}.

Proposition 2.

Let G1,G2G_{1},G_{2} be graphs. Then,

π(G1G2)=π(G1)+π(G2).\pi(G_{1}\wedge G_{2})=\pi(G_{1})+\pi(G_{2}).
Proof.

For each i{1,2}i\in\{1,2\}, denote V(Gi)V(G_{i}) by ViV_{i}, and let fi:V(Gi)dif_{i}\colon V(G_{i})\to\mathbb{R}^{d_{i}} be an orthogonal did_{i}-coloring of GiG_{i}. Also, let G=G1G2G=G_{1}\wedge G_{2}, and define f:V(G)d1+d2f\colon V(G)\to\mathbb{R}^{d_{1}+d_{2}} as follows. Set f(v)=(f1(v),0)f(v)=(f_{1}(v),0) if vV1v\in V_{1} and f(v)=(0,f2(v))f(v)=(0,f_{2}(v)) if vV2v\in V_{2}. One can verify that ff is an orthogonal (d1+d2)(d_{1}+d_{2})-coloring of GG; hence π(G)π(G1)+π(G2)\pi(G)\leq\pi(G_{1})+\pi(G_{2}).

Now let f:V(G)df\colon V(G)\to\mathbb{R}^{d} be an orthogonal dd-coloring of GG. For every uV1u\in V_{1} and vV2v\in V_{2}, we have that f(u)f(u) is orthogonal to f(v)f(v). In particular f(v)span(f(V1))f(v)\notin\operatorname{span}(f(V_{1})) and then span(f(V1))span(f(V2))\operatorname{span}(f(V_{1}))^{\perp}\supseteq\operatorname{span}(f(V_{2})). Thus the dimension of span(f(V1))spanf(V2)\operatorname{span}(f(V_{1}))\cup\operatorname{span}f(V_{2}) is equal to the dimension of span(f(V(G))\operatorname{span}(f(V(G)), which is equal to dd. Because ff restriced to ViV_{i} is an orthogonal coloring of GiG_{i}, we have that the dimension of span(f(Vi))\operatorname{span}(f(V_{i})) is at least π(Gi)\pi(G_{i}) for each i{1,2}i\in\{1,2\}, and it follows that dπ(G1)+π(G2)d\geq\pi(G_{1})+\pi(G_{2}). ∎

The analogous equality applies to χ(G1G2)\chi(G_{1}\wedge G_{2}) and is a folklore, easy to check, result. Now, if GG is such that χ(G)>π(G)\chi(G)>\pi(G), and H=GGH=G\wedge G, then π(H)=2π(G)\pi(H)=2\pi(G) and χ(H)=2χ(G)\chi(H)=2\chi(G), i.e, χ(H)π(H)=2(χ(G)π(G))\chi(H)-\pi(H)=2(\chi(G)-\pi(G)). Therefore, everytime we take the join, we double the gap between the chromatic number and the orthogonal number.

Corollary 3.

For every positive integer kk, there exists GG such that

χ(G)π(G)2k.\chi(G)-\pi(G)\geq 2^{k}.

Although, as we have seen, these aspects have been extensively investigated, we were unable to find any results on the related edge version of the problem (which is equal to investigate the parameter restricted to line graphs), and this is our main interest in this article. An orthogonal dd-edge-coloring of a graph GG is a function f:E(G)df\colon E(G)\to\mathbb{R}^{d} such that f(e)f(e) is orthogonal to f(e)f(e^{\prime}) whenever ee and ee^{\prime} are adjacent (i.e., eee\cap e^{\prime}\neq\emptyset). The minimum value dd such that there exists an orthogonal dd-edge-coloring of GG is the orthogonal index of G and is denoted by π(G)\pi^{\prime}(G). A kk-edge-coloring, the chromatic index χ(G)\chi^{\prime}(G) and a kk-edge-chromatic graph are defined similarly on the edges, and by arguments similar to the ones applied to vertices, we get that

π(G)χ(G).\pi^{\prime}(G)\leq\chi^{\prime}(G).

Observe that the edges incident to some vertex vv must be colored with d(v)d(v) orthogonal vectors, and so Δ(G)π(G)\Delta(G)\leq\pi^{\prime}(G). Now, using Vizing’s Theorem [29] we obtain the following inequalities:

Δ(G)π(G)χ(G)Δ(G)+1.\Delta(G)\leq\pi^{\prime}(G)\leq\chi^{\prime}(G)\leq\Delta(G)+1.

This tells us that the gap between π(G)\pi(G) and χ(G)\chi(G) is at most 1 when GG is a line graph, i.e., Corollary 3 does not hold for line graphs. Nevertheless, we can again ask if there exists a graph GG such that π(G)<χ(G)\pi^{\prime}(G)<\chi^{\prime}(G). The answer remains positive and our example is the the Kochen-Specker set given in [4]. Essentially, we show that the given set SS can also be assigned to edges of a graph G(S)G(S) on 9 vertices. Each vertex of our graph represents a distinct orthogonal base BiB_{i} of 4\mathbb{R}^{4}, and two vertices are adjacent if they share a common vector. We present the bases in the table below (the symbol 1˙\dot{1} denotes 1-1), and we use the same color to represent the same vector. Observe that each vector appears exactly twice. The graph itself is drawn in Figure 1.

𝐁𝟏\mathbf{B_{1}} 𝐁𝟐\mathbf{B_{2}} 𝐁𝟑\mathbf{B_{3}} 𝐁𝟒\mathbf{B_{4}} 𝐁𝟓\mathbf{B_{5}} 𝐁𝟔\mathbf{B_{6}} 𝐁𝟕\mathbf{B_{7}} 𝐁𝟖\mathbf{B_{8}} 𝐁𝟗\mathbf{B_{9}}
0001 0001 11˙11˙1\dot{1}1\dot{1} 11˙11˙1\dot{1}1\dot{1} 0010 11˙1˙11\dot{1}\dot{1}1 111˙111\dot{1}1 111˙111\dot{1}1 1111˙111\dot{1}
0010 0100 11˙1˙11\dot{1}\dot{1}1 1111 0100 1111 1111˙111\dot{1} 1˙111\dot{1}111 1˙111\dot{1}111
1100 1010 1100 101˙010\dot{1}0 1001 1001˙100\dot{1} 11˙001\dot{1}00 1010 1001
11˙001\dot{1}00 101˙010\dot{1}0 0011 0101˙010\dot{1} 011˙001\dot{1}0 1001˙100\dot{1} 0011 0101˙010\dot{1} 011˙001\dot{1}0
Table 1: Kochen-Specker set SS on 4\mathbb{R}^{4} used to construct G(S)G(S).
Refer to caption
Figure 1: Graph G(S)G(S) related to the bases presented in Table 1. Each base is a vertex of G(S)G(S), and two bases are adjacent if and only if they have a vector in common.
Theorem 4.

There exists a graph GG such that

4=π(G)<χ(G)=5.4=\pi^{\prime}(G)<\chi^{\prime}(G)=5.
Proof.

Let GG be the graph on vertices {B1,,B9}\{B_{1},\ldots,B_{9}\} given above and edge set {BiBjBiBj}\{B_{i}B_{j}\mid B_{i}\cap B_{j}\neq\emptyset\}. First, note that GG is 4-regular, since each base has exactly four vectors and each vector appears in exactly two bases; hence we get the upper and lower bounds. Furthemore, the vectors themselves give us an orthogonal 4-edge-coloring, since the bases related to the vertices are orthogonal (i.e., vectors incident to the same vertex are orthogonal). It remains to show that E(G)E(G) cannot be properly colored with 4 colors. Suppose otherwise and let ff be a 4-edge-coloring of GG. Since the graph is 4-regular, we get that each vertex vV(G)v\in V(G) is incident to every color. In particular f1(1)f^{-1}(1) would be a perfect matching, a contradiction since |V(G)||V(G)| is an odd number. ∎

The theorem above of course also tells us of the existence of a graph HH for which 4=π(H)<χ(G)=54=\pi(H)<\chi(G)=5 (consider H=L(G)H=L(G)). A good question is whether we can have strict inequalities for graphs with orthogonal number smaller than 3. The answer is positive, as we state in the theorem below.

Theorem 5.

If GG is a bipartite graph, then π(G)=χ(G)\pi(G)=\chi(G); and if π(G)=2\pi(G)=2 then GG is bipartite. Furthermore, there exists a graph GG such that

3=π(G)<χ(G)=43=\pi(G)<\chi(G)=4

.

Proof.

For the first part, note that π(G)2\pi(G)\geq 2 when E(G)E(G)\neq\emptyset, which implies π(G)=χ(G)\pi(G)=\chi(G) when GG is bipartite. Also, observe that for every u2u\in\mathbb{R}^{2}, up to reflection, there exists a unique unit vector v2v\in\mathbb{R}^{2} orthogonal to uu, and this implies that an odd cycle cannot admit an orthogonal 2-coloring.

Now, for the second part, let 𝕊23\mathbb{S}^{2}\subset\mathbb{R}^{3} be the unit sphere and H=K(𝕊2)H=K(\mathbb{S}^{2}) be its orthogonality graph. By construction, we know that π(H)=3\pi(H)=3, and in [12] it has been proven that χ(H)=4\chi(H)=4. Applying the De Bruijn-Erdős Theorem [9], that states that χ(X)=max{χ(Y)YX is finite}\chi(X)=\max\{\chi(Y)\mid\leavevmode\nobreak\ Y\subset X\text{ is finite}\} for every infinite graph XX, we get that a graph with 3=π(G)<χ(G)=43=\pi(G)<\chi(G)=4 exists. ∎

We then know that the gap occurs starting at graphs with orthogonal number equal to 3. Now, we can ask whether the same holds for the orthogonal index. Again because every u2u\in\mathbb{R}^{2} is orthogonal to a unique line in 2\mathbb{R}^{2}, we get that if GG has an orthogonal 2-edge-coloring, then GG must be a collection of paths and even cycles, in which case GG has also a 2-edge-coloring. This means that the gap again cannot occur when π(G)2\pi^{\prime}(G)\leq 2. As we have seen, the proof of the second part of the theorem above is non-constructive, hence it is not known whether such a graph can be a line graph. We then pose the following question.

Question 6.

Is there a graph GG such that 3=π(G)<χ(G)=43=\pi^{\prime}(G)<\chi^{\prime}(G)=4?

Clearly, such a graph would also give an explicit example in the proof of the second part of Theorem 5. Therefore, if it exists, finding such a graph should not be easy. We also comment on some ties with the Four Color Theorem which suggest that finding such a graph is indeed a hard task.

Observe that, if GG is a 2-connected planar 3-regular graph, then by the Four Color Theorem and Tait’s Theorem, we get that the answer to the question above is negative considering GG, i.e., every 2-connected planar 3-regular graph is 3-edge-colorable and hence cannot give a positive answer to the question. Tait’s Theorem says that the Four Color Theorem is equivalent to stating that every 3-regular 2-connected planar graph is 3-edge-colorable. Tait initially believed that every 3-regular 2-connected planar graph had to be hamiltonian, and if true, applying his theorem would give a beautiful proof for the Four Color Theorem. Indeed, if GG is a 3-regular hamiltonian graph, then we must have χ(G)=3\chi^{\prime}(G)=3: it suffices to remove a perfect matching of GG, and coloring the hamiltonian cycle with two colors (the number of vertices in the graph must be even since GG is 3-regular and the sum of the degrees is even). Even though the gap on his proof was found around the time of announcement in 1878, it was not until 1946 that an explicit example was found by Tutte [28]. The presented graph, named after Tutte, has 46 vertices and disproved Tait’s Conjecture about all such graphs being hamiltonian. Since then, numerous examples have been found, many with the help of Grinberg’s Theorem [13], and currently it is known that the smallest such graph has 38 vertices [15].

Even though we know that π(G)=3\pi^{\prime}(G)=3 when GG is a 2-edge connected 3-regular planar graph, because of the Four Colour Theorem and Tait’s Theorem as mentioned earlier, an interesting question is whether there exists a purely algebraic proof for this. Observe that, if so, then a negative answer to our question would give an alternative proof for the Four Color Theorem.

Besides the broad research around non-hamiltonian 3-regular 2-connected graphs, a lot of work has been done on 3-regular 4-edge-chromatic graphs, which are also classified as Class 2 graphs according to Vizing’s Theorem. Observe that these graphs are candidates for a positive answer to our question. Unfortunately, though, these graphs are also very hard to find, so much so that the minimal such graphs have been baptized snarks by Martin Gardner [11], evoking a Lewis Carroll’s poem’s rare creature. The Petersen graph was the first found snark, discovered in 1898, and up until 1975, when Isaacs constructed two infinite families of these graphs [16], apart from the Petersen graph, only three other snarks were known [2, 8, 27]. Nowadays, these elusive creatures are the object of much reasearch and are known to play a central role in a number of conjectures [7].

Finally, recall that, as seen in the proof of Theorem 5, the unit sphere 𝕊23\mathbb{S}^{2}\subset\mathbb{R}^{3} is such that 3=π(H)<χ(H)=43=\pi(H)<\chi(H)=4, where H=K(𝕊2)H=K(\mathbb{S}^{2}). Observe that if ff is an orthorgonal edge-coloring of GG, and S=f(E(G))S=f(E(G)), then there is a homomorphism from L(G)L(G) to K(S)K(S); also, we know that SS can be considered to be on 𝕊2\mathbb{S}^{2}, since it suffices to normalize the vectors in SS. In [12], the authors also prove that the orthogonal graph of the rational sphere 23\mathbb{Q}^{2}\subset\mathbb{R}^{3} is 3-chromatic. Therefore, if GG is a graph that answers Question 6 positively, then L(G)L(G) is 4-chromatic and therefore cannot have a homomorphism to K(2)K(\mathbb{Q}^{2}). Because a computer can only find representations rational coordinates, we get that no positive answer can be found with computer aid. We emphasize once more that no explicit finite graph GG with 3=π(G)<χ(G)=43=\pi(G)<\chi(G)=4 is known, and that its existence is only ensured by the De Brujin-Erdős Theorem [9]. Below, we summarize the known discussed necessary conditions when restricted to cubic graphs.

Proposition 7.

Let GG be a cubic 2-connected graph such that 3=π(G)<χ(G)=43=\pi^{\prime}(G)<\chi^{\prime}(G)=4. Then GG is non-hamiltonian and non-planar, and every orthogonal 3-coloring of E(G)E(G) must use vectors with irrational components.

References

  • [1] D. Avis, J. Hagesawa, Y. Kikuchi, and Y. Sasaki. A quantum protocol to win the graph coloring game on all Hadamard graphs. IEICE Transactions on Fundamentals of Electronics, Communications and Computer Sciences, E89-A(5):1378–1381, 2006.
  • [2] D. Blanusa. Le problème des quatre couleurs (Croatian). Glasnik Mat. Fiz. Astr. Ser. II, 1:31–42, 1946.
  • [3] A. Cabello, L. Eirik Danielsen, A.J. López-Tarrida, and J.R. Portillo. Basic exclusivity graphs in quantum correlations. Physical Review, 88(3), 2012.
  • [4] A. Cabello, J.M. Estebaranz, and G. García-Alcaine. Bell-kochen-specker theorem: a proof with 18 vectors. Phys. Lett. A, 212(4):183–187, 1996.
  • [5] A. Cabello, S. Severini, and A. Winter. (non-)contextuality of physical theories as an axiom. Technical Report 8, Institut Mittag-Leffler, 2011.
  • [6] P.J. Cameron, A. Montanaro, M.W. Newman, and S. Severini. On the quantum chromatic number of a graph. The electronic journal of combinatorics, 14(1), 2006.
  • [7] A. Cavicchioli, M. Meschiari, B. Ruini, and F. Spaggiari. A survey on snarks and new results: Products, reducibility and a computer search. J. Graph Theory, 28(2):56–86, 1998.
  • [8] B. Descartes. Network-colourings. The mathematical gazette (London), 32:67–69, 1948.
  • [9] P. Erdős. Some remarks on set theory. Proc. Amer. Math. Soc., 1:127–140, 1950.
  • [10] V. Galliard and S. Wolf. Pseudo-telepathy, entanglement, and graph colorings. In IEEE International Symposium on Information Theory (ISIT), 2002.
  • [11] M. Gardner. Mathematical games: snarks, boojums and other conjectures related to the four-color-map theorem. Scientific American, 234:126–130, 1976.
  • [12] C. D. Godsil and J. Zaks. Coloring the sphere. Technical Report CORR 88-12, University of Waterloo, 1988.
  • [13] E.J. Grinberg. Plane homogeneous graphs of degree three without Hamiltonian circuits. Latvian Math. Yearbook, 4:51–58, 1968.
  • [14] G. Haynes, C. Park, A. Schaeffer, J. Webster, and L.H. Mitchell. Orthogonal vector coloring. The electronic journal of combinatorics, 17:#\#R55, 2010.
  • [15] D.A. Holton and B.D. McKay. The smallest non-Hamiltonina 3-connected cubic planar graphs have 38 vertices. J. Combinatorial Theory B, 45:305–319, 1986.
  • [16] R. Isaacs. Infinite families of non-trivial trivalent graphs which are not tait-colorable. American Mathematical Monthly, 82(3):221–239, 1975.
  • [17] S. Kochen and E. P. Specker. The problem of hidden variables inquantum mechanics. J. Math. Mech., 17:59–87, 1967.
  • [18] L. Lovász. On the Shannon capacity of graphs. IEEE Trans. Inform. Theory, 25:1–7, 1979.
  • [19] L. Lovász. Geometric representations of graphs, chapter 9. Bolyai Soc. Math. Stud., 2002.
  • [20] L. Lovász, M. Saks, and A. Schrijver. Orthogonal representations and connectivity ofgraphs. Linear Alg. Appl., 114/115:439–454, 1989.
  • [21] F. Maffray and M. Preissmann. Sequential colorings and perfect graphs. Discrete Appl. Math, 94:287–296, 1999.
  • [22] M. Pavicic, J.-P. Merlet, B. MacKay, and N.D. Megill. Corrigendum: Kochen-specker vectors. J. Phys. A, 38(16):3709, 2005.
  • [23] M. Pavicic, J.-P. Merlet, B. MacKay, and N.D. Megill. Kocher-specker vectors. J. Phys. A, 38(7):1577–1592, 2005.
  • [24] A. Peres. Two simple proofs of the Kochen-Specker Threorem. J. Phys. A, 24(4):L175–L178, 1991.
  • [25] G. Scarpa and S. Severini. Kochen–specker sets and the rank-1 quantum chromatic number. IEEE TRANSACTIONS ON INFORMATION THEORY, 58(4):2524–2529, 2012.
  • [26] G. Strang. Introduction to Linear Algebra. Wellesley-Cambridge Press, 5th edition, 2016.
  • [27] G. Szekeres. Polyhedral decompositions of cubic graphs. Bulletin of the Australian Mathematical Society, 8(3):367–387, 1973.
  • [28] W.T. Tutte. On hamiltonian circuits. J. of the London Math. Soc., 21(2):98–101, 1946.
  • [29] V.G. Vizing. On a estimate of the chromatic class of a pp-graph. Diskret. Analiz., 3:25–30, 1964.
  • [30] D. West. Introduction to Graph Theory. Pearson, 2000.