The Chromatic Number of the Euclidean Plane Except where reference is made to the work of others, the work described in this thesis is my own or was done in collaboration with my advisory committee. This thesis does not include proprietary or classi ed information. Anna E. Boro nska Certi cate of Approval: Michel Smith Professor Mathematics and Statistics Krystyna Kuperberg, Chair Professor Mathematics and Statistics Andras Bezdek Professor Mathematics and Statistics George T. Flowers Dean Graduate School The Chromatic Number of the Euclidean Plane Anna E. Boro nska A Thesis Submitted to the Graduate Faculty of Auburn University in Partial Ful llment of the Requirements for the Degree of Master of Science Auburn, Alabama August 10, 2009 The Chromatic Number of the Euclidean Plane Anna E. Boro nska Permission is granted to Auburn University to make copies of this thesis at its discretion, upon the request of individuals or institutions and at their expense. The author reserves all publication rights. Signature of Author Date of Graduation iii Vita Anna El_zbieta Boro nska was born on March 2, 1975, in Mys lowice, Poland. She grad- uated from the mathematics and physics program of Maria Konopnicka General Lyceum in Katowice, in 1994. She then attended the Catholic University of Lublin for ve years and graduated with a Master Degree of Arts in Law in May of 1999. Thereafter, she worked in Department of Legal Supervision, Silesian Voivodship O ce in Katowice. From 1999 to 2003, she completed curriculum of the doctoral program in Law at the Catholic University of Lublin. She underwent legal training in the O ce of the District Attorney in Katowice from 2001 to 2004. Subsequently, she passed Prosecutor?s Examination at Court of Appeal in Katowice. She obtained a license of a legal adviser in Poland in 2005. She entered the Graduate Program at Auburn Univerity in August of 2007, as a collaborative student in the PhD program in Public Administration and Public Policy, and the MS program in Mathematics. She is married and has a son. iv Thesis Abstract The Chromatic Number of the Euclidean Plane Anna E. Boro nska Master of Science, August 10, 2009 (M.A., Catholic University of Lublin, 1999) 30 Typed Pages Directed by Krystyna Kuperberg We discuss the chromatic number of the plane problem. We provide a detailed history of its origins, along with some of the recent progress. Then we describe a proof, based on [1], of the following result. Theorem 0.1 Every distance excluding coloring of a locally nite plane tiling, with the property that the whole interior of any tile is colored by a single color, must employ at least six colors. The result was originally proved for polygonal tilings that have convex tiles, and the area of all tiles from the tiling bounded from below by a constant. It follows from our proof that the second condition is redundant. Moreover, using the fact that any polygon can be triangulated, Coulson?s result is extended to tilings with non convex polygons. v Acknowledgments The author wishes to thank the Committee Members for support and encouragement during the author?s time of study in the Department of Mathematics and Statistics at Auburn University. The author is especially indebted to her advisor Dr. Krystyna Kuper- berg. vi Style manual or journal used Journal of Approximation Theory (together with the style known as \aums"). Bibliograpy follows van Leunen?s A Handbook for Scholars. Computer software used The document preparation package TEX (speci cally LATEX) together with the departmental style- le aums.sty. vii Table of Contents List of Figures ix 1 History of the problem 1 2 Preliminaries 9 3 Chromatic number of plane triangulations 11 Bibliography 21 viii List of Figures 1.1 The Moser Spindle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2 The example of Hadwiger . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.3 The hexagonal 7 coloring . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.4 The example of Szekely . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 3.1 Five colors are needed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3.2 Proof of lemma 3.4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 3.3 The regular hexagon H around the tri-colored vertex T. . . . . . . . . . . . 16 3.4 Proof of lemma 3.6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 3.5 Proof of lemma 3.7. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 3.6 Six colors are needed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 ix Chapter 1 History of the problem The chromatic number of the plane problem asks for the minimum number of colors that are needed to paint all points in the plane, so that no two points in a given distance are colored alike. The question seems very natural and basic, but is yet to be fully answered. The problem goes back to 1950, when Edward Nelson, a graduate student at the University of Chicago at the time, created the problem working on the well known four-color problem. He passed the problem to other mathematicians at the University of Chicago, and soon the question about the chromatic number of the plane became well known in the mathematical community (see[6]). The problem sometimes is incorrectly credited to other mathematicians such as Paul Erd os, Martin Gardner, Hugo Hadwiger, or Leo Moser. Actually, the question was probably published for the rst time in Martin?s Gardner \Mathematical Games" col- umn in \Scienti c American" in 1960 [3]. Although in last nearly 60 years the chromatic number of the plane resisted all e orts aiming at an ultimate answer, a considerable amount of research discussing partial answers to this problem, or investigating related problems, ac- cumulated in the mathematical literature. The following bounds on the chromatic number are well known: 4 chromatic number of the plane 7. We shall explain how these bounds are obtained. At this point, however, let us mention that it has been recently shown by Saharon Shelah and Alexander Soifer [7], [8] that the answer to the problem may depend on the choice of axioms of set theory. Although in the present thesis we will consider only geometric aspects of the problem and will not explore those which belong to set theory, in this section for completeness sake, we shall brie y describe the results obtained by these authors. 1 Nelson was the rst to o er a proof of the lower bound. It was published in [4], however credit was given to his peer John Isbell (see [6]). An independent proof for the lower bound comes from Leo Moser and William Moser, commonly known as the Moser Spindle [5]. We will describe the two examples from [4] and [5], starting with the Moser Spindle. a b c d e f g Figure 1.1: The Moser Spindle Consider a graph exhibited in Figure 1.1. All edges are of unit length. Suppose this graph can be painted by three colors only, excluding unit distance. First consider the equilateral triangle abc. To exclude the distance 1, all three of these vertices must be of three di erent colors. Since the equilateral triangle acg shares the edge [a;c], vertex g must be of the same color as b. On the other hand, in the triangle bed we must use the same two colors as for [a;c] in order to color the edge [e;d]. Again, the triangle fed shares an edge [e;d] with bed, and therefore f must be colored alike to b. Clearly, g and f are colored alike and are unit distance apart. This results in a contradiction. An alternative proof of the lower bound is provided by the example in [4]. Suppose only three colors are used to color the entire plane, with the unit distance excluded. Let a be a point in the plane. Consider a unit circle C around a. No point on C can be colored 2 by the same color as a. Now consider also S, a circle centered at a and of radius p3. Let a w z x ?3 1 1 1 C S Cx Figure 1.2: The example of Hadwiger x be any point on S, and let Cx be a unit circle centered at x. Clearly Cx intersects C in two points, say w and z. Now notice that w;z;x;a are vertices of a unit rhombus. Since one of its diagonals is of length p3, the other one must be of unit length. This means that w and z are unit distance apart and therefore they must be colored by two di erent colors, both distinct from the one used to paint a. Consequently, since x is a unit distance apart from w and z, x must be painted alike to a. However, x was chosen arbitrarily on S, and therefore the entire S must be painted by the same color as a. This is a contradiction, since on S there are points a unit distance apart, and they can?t be colored alike . Isbell [4] is credited as rst to estimate the upper bound using hexagonal coloring of the plane, with seven colors (see [6]). An alternative proof comes from L. Szekely employing tiling by squares [10]. We exhibit Isbell?s idea in the following gure. The description is 3 1 2 3 4 5 6 7 1 2 3 4 5 6 7 1 2 3 4 5 6 7 1 2 3 4 5 6 7 Figure 1.3: The hexagonal 7 coloring based on [4]. The idea is to color a hexagonal tilling by seven di erent colors. The hexagons from the tiling are regular and of side length 25. We choose a hexagon and color it by the rst color. Then there are six hexagons each sharing an edge with this initial one. We assign one of the remaining six colors to each of them. Extend the coloring to the hexagonal tiling of the plane as shown in Figure 1.3. The boundary points on a given edge are colored arbitrarily by any of the two adjecent colors. This coloring is unit distance excluding. To see this, rst focus on interior points of the hexagons. If we start with hexagons with a side length 25 then no two points of the same color can be d distance apart, for any d between 45 and p28 5 . Clearly, no two points from the same hexagon are in a distance grater than 45 apart (this is the length of the great diameter). To obtain the other number, consider the isosceles triangle given by dotted lines in the gure. Clearly, the side length of this triangle gives the minimal distance between 4 two points colored alike, but from two di erent hexagons (in the gure the triangle refers to blue color, or 5). The base of this triangle is the smaller diameter of the hexagon, and therefore its length equals 2 p3 5 . On the other hand, it is not hard to check that the height of this triangle is equal to the length of the great diameter, plus side length, minus the height of the isosceles triangle with sides of length 25 and base 2 p3 5 . Therefore the height of this dotted triangle is 45 + 25 15 = 1. Consequently the side length is q 12 + ( p3 5 )2 = p28 5 . As far as boundary points are concerned, by the inequality on d, there is enough cushion so that it doesn?t matter which of the two adjacent colors is assigned to each edge (see [9]). Szekely?s example is obtained by considering in nite rows of squares in the plane. Each square is of side length 1p2. We start with an arbitrary square and color it with the rst color. Moving in the row from left to right we color squares with subsequent six colors, and then we start over again with the rst one. We use a reversed order of colors when moving to the left in this row. Now, ll in the plane by identical copies of this initial row. Moving 0 1 2 3 4 5 6 0 1 2 3 4 2 3 4 5 6 0 1 3 42 3 4 4 5 6 0 1 2 3 4 5 6 0 1 0 1 2 3 4 5 6 0 1 2 36 0 1 2 3 4 5 6 0 1 25 6 Figure 1.4: The example of Szekely downward, let every subsequent row be shifted by 1p2 + 1:1 units to the left. When moving upward shift by the same factor but to the right. Upper and right boundaries are included in the color of each square, except for the square?s upper left and lower right corners (see [9]). It should be clear from the picture that this coloring of the squares, with seven di erent colors, is unit distance excluding. This is because the closest that two points colored alike, 5 but from two di erent squares, can be is 1:1. On the other hand the diagonal of each of the squares is of length 1. Now, let us move on to the results presented by Shelah and Soifer. The authors consider an equivalent formulation of the chromatic number of the plane problem. Namely, let U2 be a graph on the set of all points of the plane as its vertex set, with two points adjacent i they are 1 distance apart. The question now is: What is the minimal number of colors that need to be employed in order to color every vertex of this graph, with no two points adjacent colored alike? Finite subgraphs of of U2 are called nite unit distance plane graphs. In 1951 it was shown by Erd os and Nicolaas de Bruijn [2] that the chromatic number of the plane is attained on some nite subgraphs. This fundamental result determined much of research to go in the direction of nite unit distance graphs, but at the same time, implicitly, was dependent on the axiom of choice. Motivated by this result, Shelah and Soifer consider how the above question may depend on the following axioms of set theory. (AC)(Axiom of choice) Every family of nonempty sets has a choice function; i.e., there is a function f such that f(S)2S for every S from . (AC@0)(Countable axiom of choice) Every countable family of nonempty sets has a choice function. (DC)(Principle of dependent choices) If E is a binary relation on a nonempty set A, and for every a2A there is b2A with aEb, then there is a sequence a1;a2;::: such that anEan+1 for every natural number n. (LM) Every set of real numbers is Lebesgue measurable. (ZF) Zermelo-Fraenkel system of axioms of set theory. (ZFC) ZF with an addition of AC. Axiom DC is a weak form of axiom AC, whereas axiom DC implies axiom AC@0. In their rst paper [7] Soifer and Shelah formulate the following theorem. 6 Theorem 1.1 Assume that any nite unit distance plane graph has chromatic number not exceeding 4. Then: (i) In ZFC the chromatic number of the plane is 4. (ii) In ZF+DC+LM the chromatic number of the plane is 5,6, or 7. In the second paper [8] the authors extend the ideas from the rst paper, giving two inter- esting examples that illustrate possible correlation of the above axioms of set theory and the chromatic number of the plane. Their rst example is as follows. Let G2 be a graph with <2 as the vertex set, and the set [ 2F f(s;t) : s;t2<;s t 2Q2g as the set of edges, where Q is the set of rationals, and F =f(p2;0);(0;p2);(p2;p2);( p2;p2)g: Then (i) In ZFC the chromatic number of G2 is equal to 4. (ii) In ZF+AC@0+LM the chromatic number of G2 is not equal to any positive integer n nor even to @0. The second example is a modi cation of the rst one. Namely, G3 is de ned exactly the same as G2, with the only exception that now F =f(p2;0);(0;p2)g: Then (i) In ZFC the chromatic number of G3 is equal to 2. 7 (ii) In ZF+AC@0+LM the chromatic number of G3 is not equal to any positive integer n nor even to @0. In the present thesis we will focus our attention on yet another partial result toward a nal solution of the chromatic number problem. D. Coulson in [1] considered special type of distance excluding coloring of the plane, associated with certain types of polygonal tilings. He showed that such a coloring must employ at least six colors. In chapter 2 this result is stated in details in theorem 0.1, whereas the discussion of the proof will follow in chapter 3. This result was stated for colorings of polygonal tilings of the plane. However, the author assumed that two polygons from such a tiling meet either at a vertex, or along an entire common edge. Since any polygon can be triangulated without introducing new vertices, we will consider only colorings of triangulations of the plane; i.e., polygonal tilings where each tile is a triangle and the intersection of any two tiles is an edge of both tiles, a vertex, or empty. 8 Chapter 2 Preliminaries We will use the following notation. <2 denotes the Euclidean plane with the distance between two points x;y 2<2 given by jjx yjj = p(x1 y1)2 + (x2 y2)2, where x = (x1;x2) and y = (y1;y2). For a xed > 0 and x2<2, by S (x) we will denote the circle of radius centered at x. A closed disk is the region D (x) bounded by S (x). By a tiling of the plane <2 we will understand a collection of polygons P such that SP =<2 if P1;P22P then IntP1\IntP2 =;. The interior IntP of a triangle P is the set of all points x for which there is a closed disk G P centered at x. An interior point of a triangulation is a point that is in the interior of one of the elements ofP. A boundary point is a point that is not an interior point. Note that a boundary point is either a vertex of a triangle from P or it lies on an edge of such a triangle. A polygon P is convex if for given x;y2P the straight segment [x;y] joining x and y lies entirely in P; i.e. [x;y] P. A tiling P is locally nite if for every point x2<2 there is a circle S centered at x that intersects only a nite number of elements from P. A locally nite tiling P is a triangulation if every element of P is a triangle and the intersection of any two tiles is an edge of both tiles, a vertex, or the empty set. A coloring of the plane is a surjective function : <2 ! , where is the set of colors. We will denote these colors by greek letters such as ; ; ; ; . Given a point x in the plane, and a color 2 , we shall say that x is -colored if (x) = . We will consider only colorings with the following property: if P is an element of the tilling, 2 is a color, x;y2IntP, and x is -colored, then y is -colored as well; 9 i.e. any element of the tiling has its entire interior colored by the same color. At this point it is important to address the following issue. We will not deal with the coloring of the boundary points at all. It is important to stress out that this does not mean that the coloring function is not de ned for these points. Rather it means that the main result does not depend on how these boundary points are colored. In other words, a restriction of the coloring to interior points only already forces that the set of colors must consists of at least six elements. We say that a coloring (x) of the plane is distance excluding if there is a distance d such that jjx yjj = d implies that (x) 6= (y); i.e. no two points of the same color can be d distance apart. From now on let such a distance d be set once and for all. The main result that we are going to discuss is as follows. Theorem: If :<2 ! is a distance excluding coloring of a locally nite tiling P, then consists of at least six colors. The above result was originally proved in [1] for locally nite polygonal tilings with convex tiles, and the area of all tiles from the tiling bounded from below by a constant. However, this last assumption is not needed to prove the above theorem, when we assume local niteness, as we are going to exhibit. Alternatively, one could drop local niteness of the triangulation, and assume the lower bound for the area of all triangles instead, and still prove the theorem. Also, as mentioned earlier, using the fact that any polygon can be triangulated allows us to deal only with triangulations, instead of polygonal tilings. This also allows for a generalization of the result by Coulson, since in such a case we do not need to state that the elements of the tilings are convex. We can start with any polygonal tiling P, and after triangulating each of its elements we can consider a triangulation P0. Since such a triangulation is still locally nite and every triangle is convex, the same arguments can be applied. The rest of the present thesis is devoted to proving the above result. We shall present in detail a number of arguments that were indicated as true in [1], but an actual proof of these facts, not immediately apparent, was not given. 10 Chapter 3 Chromatic number of plane triangulations From now on, let : <2 ! be a distance excluding coloring of a triangulation P set once and for all, and let d be the excluded distance. We shall say that a vertex T is tri-colored if T belongs to three triangles from P, each of which has its interior painted by a color di erent than the interiors of the other two. Similarly, a vertex is bi-colored if it belongs to two triangles that are not colored alike. Lemma 3.1 There is a tri-colored vertex T in the triangulation P. Proof: Suppose to the contrary that there is no tri-colored vertex. Choose a triangle P and a subcollection G of the triangulation, that is maximal with respect to the following properties: 1. P 2G, 2. if ~P 2G then IntP and Int ~P are colored alike, 3. SG is connected. Notice that SG is bounded. Suppose it was not. Choose x2IntP and consider the circle Sd(x) of radius d centered at x. Since SG is not bounded by Sd(x) there must be a point y2Sd(x)\SG. If y is an interior point then we obtain a contradiction, since x and y would be colored alike, and in distance d apart. If y is a boundary point to a triangle ~P from G, then choose an and a circle S (x) around x small enough so that S (x) IntP. Also, choose 1 so that the circle S 1(y) intersects the interior of ~P. Let z2S 1(y)\Int ~P and let Sd(z) be the circle centered at z of radius d. Clearly Sd(z) intersects S 1(x) in a point, say w. Since S 1(x) S (x) therefore w and z are colored alike. This is a contradiction, since 11 jjw zjj= d. We have obtained that SG is bounded. Consider the complement of SG; i.e. consider <2nSG. There is an unbounded set U, that is a component of this complement. Notice that the boundary of U separates the plane (no point in the complement of U can be joined by a line segment with a point in U, without crossing this boundary) therefore this boundary must contain a closed loop, say L. L is a union of edges, each of which is shared by a triangle from G and by a triangle from PnG. Consider a subcollection U of all triangles R from PnG that intersect L. Notice that all elements fromU are colored alike. Indeed, suppose the contrary. Pick a triangle P1 from U with an edge contained in L. Suppose, there was a triangle P2 in U, colored not alike to P1. Choose an edge E1 L of P1, and an edge E2 L of P2. Since L is connected, there is a path Y consisting of edges that joins E1 with E2 (with possible self-intersections). By the initial assumption there is no tri-colored vertex in this path, and therefore every vertex is bi-colored. Let be the color of G, be the color of P1, and the color of P2. Walking from E1 to E2 along Y, let v1 be the rst bi-colored vertex where -colored and -colored triangles meet. Backtrack to the previous vertex v2. At this vertex -colored and -colored triangles meet. Consider the edge [v2;v1]. This edge is shared by exactly two triangles, one of which is a triangle P32G, which in turn must be -colored. Let P4 be the triangle fromU sharing [v2;v1] with P3. If P4 is -colored then v1 is tri-colored. Otherwise v2 is tri-colored. We obtained a clear contradiction, since we assumed there is no tri-colored vertex, and therefore all elements of U are colored alike. Now, complete U so that SU is maximal with respect to two properties: 1. if P1;P22U then IntP1 and IntP2 are colored alike, 2. SU is connected. Choose an interior point p2SU. If the circle Sd(p) of radius d centered at p intersects SU, then we obtain a contradiction by the same arguments that were used to exhibit that SG is bounded. Otherwise, repeat the same reasoning as before replacingG withG[U. The local niteness of the tilling assures that the region bounded by SU will be expanding (otherwise, 12 the areas of triangles would need to tend to zero, and there would be an accumulation point, by Bolzano Weirstrass theorem-cf. proof of proposition 3.2) and, if neccesary, iterating the above procedure nitely many times we will obtain a contradiction. Proposition 3.2 Let C be a circle. Then there are only nitely many boundary points on C. Proof: Notice that by local niteness of the tilling, C can have a nonempty intersection with only nite number of elements from the triangulation. A contrario, suppose C intersects in nitely many triangles from the triangulation. By Bolzano-Weirstrass theorem, applied to a circle, there must be an x2C, such that for any given circle S around x there are points from in nitely many of these triangles in S\C. Therefore the triangulation is not locally nite at x. A contradiction. Now, we will show that any tile can have only nitely many boundary points on C. Suppose to the contrary that there is a tile P that has an in nite number of boundary points on C. First, in nitely many of these points must be interior points of edges of P, since P has only nitely many vertices. Second, for given edge E of P there are at most 2 points in common for E and C, since any straight arc intersects a circle in at most 2 points. This implies that if there are in nitely many boundary points of P on C, there must be also in nitely many edges of P. But P is a triangle. Contradiction. Consequently, C has nonempty intersection with only nitely many tiles, each of which has only nitely many boundary points on C. Lemma 3.3 The set of colors must consists of at least ve colors. Proof: Let C be a circle of radius d and centered at the tri-colored vertex T. By proposition 3.2 there are only nitely many boundary points on C. Let z be any point on C which is an interior point of a triangle Q. There is an > 0 such that z is contained in Q with some closed disk G (z) of radius centered at z. Clearly G (z) is colored with the 13 T z w q Gepsilon1(z) Sepsilon1a(T) Sepsilon1a(z) Figure 3.1: Five colors are needed. same color as z. Let ; ; be the three colors meeting at T. Suppose z is colored by and let [z;T] be the straight segment of length d joining T with z. Arbitrarily close to T there are -colored points. Therefore there is such that there is an -colored point q on the circle S (T) of radius centered at T. Consider the straight segment [q;T] and along with the segment [z;T] extend it to a parallelogram [z;T;q;w]. Then w is a point on the circle S (z) of radius centered at z. Since S (z) G (z), w is -colored, and in a distance d from -colored point q. A contradiction. By the same arguments it can be shown that z can be neither -colored nor -colored. Now let v be a point on C in a distance d from z. By the same reasoning as above v cannot be either -colored, or -colored, or -colored. Since it is d distance apart from z, it cannot be of the same color as z, and therefore a fth color is needed. In Lemmas 3.4, 3.5, 3.6, 3.7 we assume that only ve colors are used. 14 Lemma 3.4 Let C be a circle of radius d and centered at the tri-colored vertex T. If x;y2C, jjx yjj= d and x is a boundary point of two distinct tiles colored by two di erent colors, then y cannot be an interior point of any tile. Proof: Suppose =f ; ; ; ; gis the set of colors, C is the circle of radius d centered at T and x;y2C,jjx yjj= d. Again, let ; ; be the three colors meeting at T. Suppose T x q y w Q W R Figure 3.2: Proof of lemma 3.4. that x is a boundary point of some two tiles, say W;R, colored by two di erent colors, but y is an interior point of a tile Q. By the reasoning in the proof of lemma 3.3, on the circle C there can be no interior point colored by ; or . Therefore, IntW and IntR must be colored by or . Without loss of generality, suppose IntW is -colored and IntR is -colored. There is > 0 such that y is contained in Q with a closed disk G (y) of radius centered at y. Since we assume there are only 5 colors used, y must be either -colored or -colored. Suppose y is -colored. x is in the boundary of W and therefore there is such that there is a -colored point w2IntW on the circle S (x) of radius centered at 15 x. Consider the straight segment [x;w] and along with the straight segment [y;x] extend it to a parallelogram [y;x;w;q]. Then q is a point on the circle S (y) of radius centered at y. Since S (y) G (y), q is in the interior of Q and therefore it is -colored, as y is. However, q is in a distance d from -colored point w. This gives a contradiction. The same arguments apply if x is -colored, and this concludes the proof. Lemma 3.5 Let C be a circle of radius d and centered at the tri-colored vertex T. Then there is a regular hexagon H inscribed into C, each vertex of which is a boundary point of some two tiles colored by two di erent colors. Proof: Let x be any point on the circle C of radius d around T, where two tiles whose T d d d d d d d x y H C Figure 3.3: The regular hexagon H around the tri-colored vertex T. interiors are of two di erent colors meet. By lemma 3.4 if y is on on the circle C in the distance d apart from x, then y is not an interior point. This process can be repeated in order to obtain four more di erent points with the same property as x and y, since in a circle of radius d one can inscribe a regular hexagon of side length d. Therefore these six points span a regular hexagon H inscribed into C and the proof is complete. 16 Let A be a circular subarc of the circle C. We will write A = to indicate the shorter of the two subarcs of C, with endpoints in v and w. We say that A invades the interior of P if A\IntP 6=;. Lemma 3.6 Assume that the boundary of C is colored with two colors only. Let H be the regular hexagon described in lemma 3.5. Let v;w be two vertices of H sharing an edge [v;w]. Let A =< v;w >. By de nition of H there are two triangles meeting at v one -colored, and the other -colored. Suppose Pv is one of the two with v in its boundary and such that the arc A invades IntPv. Let Pw be de ned analogously for w. Then (IntPv) = (IntPw); i.e. the interiors of Pv and Pw are of the same color. Proof: Let v;w;Pv;Pw be as described above. Suppose by contradiction the interiors of Pv T H C v w PF Pv u z y r Pw Figure 3.4: Proof of lemma 3.6. and Pw are not of the same color; i.e. let IntPv be -colored and IntPw be -colored. Let F be the subarc of C such that F = CnA, and let PF be a -colored triangle from the triangulation that is invaded by F and meets Pw in w. The existence of such a triangle is 17 guaranteed by de nition of H; since we assume that Pw is -colored therefore PF must be -colored. F invades the interior of PF and there exists a point u2IntPF \F. The entire subarc of F is -colored (with a possible exception for w). Let z be a point on A in the distance d from u. Since arc A invades IntPv there must be points from Pv on the arc < v;z >. Choose y 2 IntPv\< v;z >. Clearly y is of distance d from a point r on . But this is a contradiction, since both z and y are -colored. Lemma 3.7 Let H be the regular hexagon described in lemma 3.5, and v;w its two con- secutive vertices. Let [v;w] be an edge of H, and Rv be a ray starting at v, that makes a right angle with [v;w]. Then there is an > 0 and a disk D (v) around v, such that any point on Rv\D (v) is neither -colored nor -colored ( ; are colors meeting at at v). Proof: Let [u;v] be the edge of H meeting [v;w] at v. By the reasoning of lemma 3.6 T H C v w y u z t r Figure 3.5: Proof of lemma 3.7. without loss of generality we can assume that there is z2[u;w] (close to u) such that (u;z) 18 is colored by a single color. Similarly, we can assume that there is y2 [v;w] (close to w) such that (y;w) is colored by a single color. Then (u;z) and (y;w) are not colored alike, and assume that (u;z) is -colored, but (y;w) is -colored. Now, there is a point t2Rv close enough to v so that Sd(t) intersects (y;w). Clearly t cannot be -colored, nor any other point on the straight arc [v;t]. Similarly, there is r2Rv close enough to v so that Sd(r) intersects (u;z). Clearly r cannot be -colored, nor any other point on the straight arc [v;r]. The lemma follows with = minfjr vj;jt vjg. Theorem: If :<2 ! is a distance excluding coloring of a locally nite tiling P, then consists of at least six colors. Proof: Suppose to the contrary, that we can use 5 colors only and exclude the distance T B ? ?? ? ? ? ? p1 p2 R1 R2 c a b gk Figure 3.6: Six colors are needed. d. Consider an edge B with a vertex at T, that belongs to two triangles P1 and P2 such that T 2P1\P2 and IntP1 is -colored and IntP2 is -colored. The ray starting at T and containing B intersects the hexagon H in a point, say p. Let p1 and p2 be two vertices of 19 the edge of the hexagon H that contains p (possibly p = p1 or p = p2). Notice that both [p1;T] and [p2;T] make acute angles with B. Consider the ray R1 starting at p1 that makes a right angle with [p1;p2], and the ray R2 starting at p2 that makes a right angle with [p1;p2]. Close to p1 and p2, R1 and R2 are neither -colored nor -colored. Choose a point c on R1 and g on R2 close enough to p1 and p2 respectively, so that the circles Sd(c),Sd(g) of radius d centered at c and g respectively, both intersect the edge B in its interior. One can assure that jjc p1jj=jjg p2jj. Let k be in Sd(c)\B. Observe that there is a circular arc [a;b] Sd(c) containing k such that the circular segment (a;k) is -colored and the circular segment (k;b) is -colored, by de nition of B. Therefore c cannot be either -colored or -colored. Consequently c must be -colored. By the same arguments, assuming that only ve colors are to be used, g must also be -colored. However, jjc gjj = d, since both R1 and R2 make right angle with [p1;p2] (which is of length d) and the straight segment [c;g] is parallel to [p1;p2]. This contradiction implies that a sixth color, say is needed, which completes the proof. 20 Bibliography [1] Coulson, D. On the chromatic number of plane tilings. J. Aust. Math. Soc. 77 (2004), no. 2, 191-196. [2] P. Erd os, N.G. de Bruijn, A colour problem for in nite graphs and a problem in the theory of relations, Indag. Math. 13 (1951) 371-373. [3] Gardner, M. The celebrated four-color map problem of topology. Sci. Amer. 203 1960 no. 3, 218-226. [4] Hadwiger, H. Ungel oste Probleme, Nr. 11. Elem. Math. 16 1961 103-104. [5] Moser, L., Moser, W., Solution to Problem 10, Can. Math. Bull., 4,(1961), 187-189. [6] Soifer, Alexander Chromatic number of the plane & its relatives. I. The problem & its history. Geombinatorics 12 (2003), no. 3, 131-148. [7] Soifer, A. Shelah, S. Axiom of choice and chromatic number of the plane. J. Combin. Theory Ser. A 103 (2003), no. 2, 387-391. [8] Soifer, A. Shelah, S. Axiom of choice and chromatic number: examples on the plane. J. Combin. Theory Ser. A 105 (2004), no. 2, 359{364. [9] Soifer, Alexander, The mathematical coloring book. Mathematics of coloring and the colorful life of its creators. Springer, New York, 2009. [10] Sz ekely, L. A. Remarks on the chromatic number of geometric graphs. Graphs and other combinatorial topics (Prague, 1982), 312-315, Teubner-Texte Math., 59, Teubner, Leipzig, 1983. 21