The minimum degree of
minimal Ramsey graphs for cliques
Abstract.
We prove that , where is the Ramsey parameter introduced by Burr, Erdős and Lovász in 1976, which is defined as the smallest minimum degree of a graph such that any -colouring of the edges of contains a monochromatic , whereas no proper subgraph of has this property. The construction used in our proof relies on a group theoretic model of generalised quadrangles introduced by Kantor in 1980.
Key words and phrases:
Ramsey graphs, generalised quadrangles, Kantor family2010 Mathematics Subject Classification:
05C55,05D10,51E121. Introduction
A graph is called -Ramsey for another graph , denoted by , if every -colouring of the edges of contains a monochromatic copy of . Observe that if , then every graph containing as a subgraph is also -Ramsey for . Some very interesting questions arise when we study graphs which are minimal with respect to , that is, but there is no proper subgraph of such that . We call such graphs -Ramsey minimal for and we denote the set of all -Ramsey minimal graphs for by . The classical result of Ramsey [21] implies that for any finite graph and positive integer , there exists a graph that is -Ramsey for , that is, is non-empty.
Some of the central problems in graph Ramsey theory are concerned with the case where is a clique . For example, the most well studied parameter is the Ramsey number , that denotes the smallest number of vertices of any graph in . The classical work of Erdős [8] and Erdős and Szekeres [9] shows that . While these bounds have been improved since then, most recently by Sah [23] (also see [24] and [5]), the constants in the exponent have stayed the same. We refer the reader to the survey of Conlon, Fox and Sudakov [6] for more on this and other graph Ramsey problems.
Several other questions on have also been explored; for example, the well studied size-Ramsey number which is the minimum number of edges of a graph in . We refer the reader to [1, 3, 18, 22] for various results on minimal Ramsey problems. In this paper, we will be interested in the smallest minimum degree of an -Ramsey minimal graph, which is defined by
for a finite graph and positive integer , where denotes the minimum degree of . Trivially, we have , since the complete graph on vertices is -Ramsey for and has minimum degree . The study of this parameter was initiated by Burr, Erdős and Lovász [2] in 1976. They were able to show the rather surprising exact result, , which is far away from the trivial exponential bound of . The behaviour of this function is still not so well understood for colours. Fox et al. [10] determined this function asymptotically for every fixed up-to a polylogarithmic factor, and for their result was further improved by Guo and Warnke [12] who managed to obtain matching logarithmic factors.
Theorem 1.1 (Fox, Grinshpun, Liebenau, Person, Szabó).
-
(i)
There exist constants such that for all , we have
-
(ii)
For all there exist constants such that for all , we have
Theorem 1.2 (Guo, Warnke).
.
The constant in the upper bound of Theorem 1.1(ii) is rather large (), and in particular not polynomial in . To remedy this, they proved the following general upper bound which is polynomial in both and .
Theorem 1.3 (Fox, Grinshpun, Liebenau, Person, Szabó).
For all , .
For a fixed and , Hàn, Rödl and Szabó [13] determined this function up-to polylogarithmic factors by proving the following.
Theorem 1.4 (Hàn, Rödl, Szabó).
There exists a constant such that for every and
We prove the following general upper bound that improves Theorem 1.3, and thus provides the best known upper bound on outside the special ranges covered by Theorem 1.1 and 1.4.
Theorem 1.5.
There exists an absolute constant such that for all , .
Our proof uses the equivalence between and another extremal function, called the -colour -clique packing number [10], defined as follows. Let denote the minimum for which there exist -free pairwise edge disjoint graphs on a common vertex set of size such that for any -colouring of , there exists an such that contains a copy of all of whose vertices are coloured in the th colour.
Lemma 1.6 (see [10, Theorem 1.5]).
For all integers we have .
Our graphs in the packing would come from certain point-line geometries known as generalised quadrangles that we define in the next section. In Section 3, we show that any packing of ‘triangle-free’ point-line geometries implies an upper bound on , assuming certain conditions on the parameters of the geometry. In Section, 4 we give a packing of certain subgeometries of the so-called Hermitian generalised quadrangles using a group theoretic model given by Kantor in the 1980’s [15], and deduce that this packing implies our main result.
2. Background
A (finite) generalised quadrangle of order is an incidence structure of points , lines , together with a symmetric point-line incidence relation satisfying the following axioms:
-
(i)
Each point lies on lines () and two distinct points are incident with at most one line.
-
(ii)
Each line lies on points () and two distinct lines are incident with at most one point.
-
(iii)
If is a point and is a line not incident with , then there is a unique point on collinear with .
Notice that the third axiom above ensures that there are no triangles (i.e., three distinct lines pairwise meeting in three distinct points) in . The standard reference on finite generalised quadrangles is the book by Payne and Thas [20]. The collinearity graph of a generalised quadrangle is the graph on the set of points with two points adjacent when they are both incident with a common line. A collineation of , that is, an automorphism of its collinearity graph, is an elation about the point if it is either the identity collineation, or it fixes each line incident with and fixes no point not collinear with . If there is a group of elations of about the point such that acts regularly on the points not collinear with , then we say that is an elation generalised quadrangle with elation group and base point . Necessarily, has order , as there are points not collinear to a given point in any generalised quadrangle.
Now suppose we have a finite group of order where . A Kantor family of is a set of subgroups of order , and a set of subgroups of order , such that the following are satisfied:
-
(K0)
for all ;
-
(K1)
whenever ;
-
(K2)
whenever are distinct.
Due to a theorem of Kantor (c.f., [15, Theorem A.3.1]), a Kantor family as described above, gives rise to an elation generalised quadrangle of order , which we briefly describe in Table 1.
Points | Lines |
---|---|
elements of | the right cosets |
right cosets | symbols |
a symbol . |
Incidence: , where
We will simply be needing to use the Kantor family for a well-known family of generalised quadrangles, where the Heisenberg groups appear as the group in the description above. We remark that the main property we will need is (K2), since it ensures that lines of the form , never form a triangle.For self-containment, we give a proof here. Suppose are three elements of forming the vertices of a triangle. Then there are three elements such that , , . Therefore, , , , from which it follows that . Since , we have . So the condition given by (K2) ensures that there are no triangles.
In this paper, we will also need two commonly used maps on finite fields: the trace and norm maps. For a finite field, they are the analogues of ‘real part’ and ‘square-modulus’ for the complex numbers. We will not be needing the general theory of traces and norms, just two functions in particular. The relative trace map from to is defined by . If is a field-automorphism of , then for all . Also, is additive; that is, for . The relative norm map to is defined by . It is also invariant under field-automorphisms, and is surjective. We will make use of the following property: if is odd, then every element of is a square in . To see this, let . Since is surjective, there exists such that . Since is odd, the element is well defined, and then .
3. Packing Generalised Quadrangles
A partial linear space is a point-line incidence structure with the property that any two distinct points are incident to at most one common line. A triangle-free partial linear space of order is an incidence structure satisfying Axioms (i) and (ii) of a generalised quadrangle, and (iii)′ there are no three distinct lines pairwise meeting each other in three distinct points. Clearly, any subgeometry of a generalised quadrangle where the number of points on a line and the number of lines through a point are constants is a triangle-free partial linear space. We now prove the main lemma that will imply Theorem 1.5 once we have the construction outlined in Section 4. Our proof follows the same idea as in Dudek and Rödl [7], and Fox et al. [10].
Lemma 3.1.
Let be positive integers. Say there exists a family of triangle-free partial linear spaces of order , on the same point set and pairwise disjoint line-sets , such that the point-line geometry is also a partial linear space. If and , then .
Proof.
In order to show that , we will exhibit -free pairwise edge disjoint graphs on the common vertex set , such that for any -colouring of , there exists an such that contains a copy of all of whose vertices are coloured in the th colour. We start by recalling the following properties about each partial linear space , :
-
(P1)
Every point is incident with lines of .
-
(P2)
Every line contains points from .
-
(P3)
Any two points of lie on at most one line of .
-
(P4)
is triangle-free.
Furthermore, given that is a partial linear space and the line-sets are disjoint,
-
(P5)
For any , and any , , and are incident with at most one common point.
Let , and . For each line , we select uniformly at random one partition of among all , where denotes the th component of the partition, such that for some , and . Choices for distinct lines in are independent.
We define a graph on the vertex set as follows. For every , we include the edges of a complete -partite graph between the vertex sets for . Note that the graph is a collection of Turán graphs on vertices with parts. Each Turán graph comes from one line . By property (P3), any two points are incident with at most one line, therefore the different Turán graphs are edge-disjoint. Furthermore, by property (P4), is -free. Finally, by property (P5), for any , and are edge disjoint.
In order to conclude, we need to show that with positive probability, for any -colouring of , there exists an such that contains a copy of all of whose vertices are coloured in the th colour. Note that given on the vertex set , in any -colouring of , at least one of the colours occurs at least times. Therefore if for every , every set of at least vertices contains a copy of , then we get the desired property. The choices of partitions being done independently, to conclude our proof it suffices to show that for each , with positive probability every set of at least vertices contains a copy of in .
Fix . For a subset , let denote the event that the induced subgraph contains no copy of . Let with . By property (P4), any copy of can only appear from one line , i.e.
All the events are independent, therefore
For a given line , let , and let be the random partition of . Note that contains no copy of if and only if there exists such that . For a fixed ,
Therefore
Given that , and , we have
Therefore,
and therefore
(3.1) |
Note that since we have
and since we have
Therefore,
Then there exists an instance of such that every subset of with at least vertices contains a copy of in . ∎
4. The construction
Let be a prime power, and denote the finite field of order by . We will use the model of the Hermitian generalised quadrangle that appears in [14, Section 3] (see Example 3). For a definition of see [20, Chapter 3].
Let be the group defined on by the following operation:
where is the relative trace map from to . It turns out that is the Heisenberg group of order with centre of order .
We can construct a generalised quadrangle by constructing a Kantor family of . Define
Then and form a Kantor family of giving rise to a generalised quadrangle isomorphic to .111In [14] the dual of this generalised quadrangle is defined, denoted by , but it is well known that is isomorphic to the dual of the elliptic generalised quadrangle (see [20, Chapter 3], where is denoted by ).
From now on we will assume that is odd. Let be an element of . For each , define as follows. Let
By Axiom (K1), we have . So . Therefore, we can write every element as , with and . Define .
Lemma 4.1.
For every , is an automorphism of .
Proof.
Let . It suffices to show that is a homomorphism from to , since is clearly bijective. Let . Then
Using in , we get,
Using the fact that is additive, we obtain,
Therefore, is an automorphism of . ∎
Lemma 4.2.
For every odd prime power , there exists such that for all nonzero .
Proof.
Suppose that for some non-zero . Multiplying by and defining we get the following cubic equation,
Therefore if is such that the cubic is irreducible, then we get a contradiction. We show that such a choice of exists. Let be a non-square in and suppose that is an element of such that . By the Hansen-Mullen Irreducibility Conjecture (which is true, see [4, Theorem 2.7]) there exists a monic cubic of the form irreducible in for some . Let
By [16, Theorem 3], is irreducible in . ∎
Theorem 4.3.
Let be an element of such that for all non-zero . For each and , let
and let
Then:
-
(i)
Every element of is a set of subgroups and any two cosets from subgroups of different such sets intersect each other in at most one element.
-
(ii)
If we let be the underlying set of , then for every the set of lines gives rise to a triangle-free partial linear space of order .
Proof.
First, for each and , we have
Since is an automorphism of the underlying group (as per Lemma 4.1), it follows that and also form a Kantor family, and give rise to an isomorphic generalised quadrangle. Therefore, by taking cosets of the subgroups , , as lines we get a subgeometry of this generalised quadrangle which is a triangle-free partial linear space of order . We now prove the first part.
First we show that whenever . An element of is of the form for some , but it is also for some . Therefore, and hence
Now and so
Expanding this equation gives us
Suppose, by way of contradiction, that . Since , we can rewrite the equation as
Let be the relative norm function, defined by . The left-hand side has norm 1 and hence
Now and we can factor out , so
Note that by definition of the relative norm map,
Using that in , , we obtain
Therefore,
a contradiction. So and .
Now for any two subgroups of a group, the intersection of two cosets of and is either empty, or a coset of , which proves our claim. ∎
Corollary 4.4.
There exists an absolute constant such that for all , we have .
Proof.
Let , , and let be the smallest prime such that . By Lemma 4.2 and Theorem 4.3(ii), there exists a family of triangle-free partial linear spaces of order , on the same point set and pairwise disjoint line-sets , and by Theorem 4.3(i), the point-line geometry is also a partial linear space. Note that and . Combining Lemmas 1.6 and 3.1, . By Bertrand’s postulate, , and using yields the desired bound, with . Note that, in light of Conjecture 5.2, we did not try to further optimise this constant. ∎
5. Concluding remarks
While generalised quadrangles have been used extensively in extremal combinatorics, and particularly Ramsey theory (e.g. [7, 11, 17, 19, 25]), our result appears to be the first instance in Ramsey theory where the group theoretic structure of these geometries is exploited. We are hopeful that Kantor’s model of generalised quadrangles will lead to new results in other Ramsey problems as well.
In the probabilistic argument of Section 3, note that if we use and , then from equation (3.1) it follows that we can solve the following quadratic inequality in to ensure that the probability is :
One can check that this inequality is satisfied for all . Using that for any , , we obtained the following more refined upper bound.
Theorem 5.1.
For all ,
For further improvements to our upper bound we should perhaps explore triangle-free partial linear spaces that do not arise from generalised quadrangles. Moreover, if we could make the probabilistic argument of Section 3 deterministic, then this could also lead to an improvement in the bound.
We would like to make the following conjecture.
Conjecture 5.2.
For all
for some constant and a constant degree polynomial function .
Acknowledgements
We are grateful to Anita Liebenau for helpful discussions. We also thank Simona Boyadzhiyska for pointing out a significant gap in our first draft that was fixed later. We also want to thank the anonymous referee for a thorough proofreading and many helpful remarks.
References
- [1] S. A. Burr, P. Erdős, R. J. Faudree, C. C. Rousseau, and R. H. Schelp. Ramsey-minimal graphs for star-forests. Discrete Math., 33(3):227–237, 1981.
- [2] S. A. Burr, P. Erdős, and L. Lovász. On graphs of Ramsey type. Ars Combin., 1(1):167–190, 1976.
- [3] S. A. Burr, J. Nešetřil, and V. Rödl. On the use of senders in generalized Ramsey theory for graphs. Discrete Math., 54(1):1–13, 1985.
- [4] S. D. Cohen. Explicit theorems on generator polynomials. Finite Fields Appl., 11(3):337–357, 2005.
- [5] D. Conlon. A new upper bound for diagonal Ramsey numbers. Ann. of Math. (2), 170(2):941–960, 2009.
- [6] D. Conlon, J. Fox, and B. Sudakov. Recent developments in graph Ramsey theory. In Surveys in combinatorics 2015, volume 424 of London Math. Soc. Lecture Note Ser., pages 49–118. Cambridge Univ. Press, Cambridge, 2015.
- [7] A. Dudek and V. Rödl. On -free subgraphs in -free graphs and vertex Folkman numbers. Combinatorica, 31(1):39–53, 2011.
- [8] P. Erdős. Some remarks on the theory of graphs. Bull. Amer. Math. Soc., 53:292–294, 1947.
- [9] P. Erdős and G. Szekeres. A combinatorial problem in geometry. Compositio Math., 2:463–470, 1935.
- [10] J. Fox, A. Grinshpun, A. Liebenau, Y. Person, and T. Szabó. On the minimum degree of minimal Ramsey graphs for multiple colours. J. Combin. Theory Ser. B, 120:64–82, 2016.
- [11] Z. Füredi, A. Naor, and J. Verstraëte. On the Turán number for the hexagon. Advances in Mathematics, 203(2):476–496, July 2006.
- [12] H. Guo and L. Warnke. Packing nearly optimal Ramsey graphs. Combinatorica, 40(1):63–103, 2020.
- [13] H. Hàn, V. Rödl, and T. Szabó. Vertex Folkman numbers and the minimum degree of minimal Ramsey graphs. SIAM J. Discrete Math., 32(2):826–838, 2018.
- [14] W. M. Kantor. Generalized quadrangles associated with . J. Combin. Theory Ser. A, 29:212–219, 1980.
- [15] W. M. Kantor. Generalized polygons, SCABs and GABs. In Buildings and the geometry of diagrams (Como, 1984), volume 1181 of Lecture Notes in Math., pages 79–158. Springer, Berlin, 1986.
- [16] H. D. Kim, J. M. Kim, and I. Yie. Certain cubic polynomials over finite fields. J. Korean Math. Soc., 46(1):1–12, 2009.
- [17] A. Kostochka, D. Mubayi, and J. Verstraëte. Hypergraph Ramsey numbers: triangles versus cliques. J. Combin. Theory Ser. A, 120(7):1491–1507, 2013.
- [18] T. Łuczak. On Ramsey minimal graphs. Electron. J. Combin., 1:Research Paper 4, approx. 4, 1994.
- [19] D. Mubayi and J. Verstraëte. A note on pseudorandom Ramsey graphs. arXiv preprint, arXiv:/1909.01461., 2019.
- [20] S. E. Payne and J. A. Thas. Finite generalized quadrangles. EMS Series of Lectures in Mathematics. European Mathematical Society (EMS), Zürich, second edition, 2009.
- [21] F. P. Ramsey. On a Problem of Formal Logic. Proc. London Math. Soc. (2), 30(4):264–286, 1929.
- [22] V. Rödl and M. Siggers. On Ramsey minimal graphs. SIAM J. Discrete Math., 22(2):467–488, 2008.
- [23] A. Sah. Diagonal Ramsey via effective quasirandomnes. arXiv preprint, arXiv:2005.09251, 2020.
- [24] J. Spencer. Ramsey’s theorem—a new lower bound. J. Combinatorial Theory Ser. A, 18:108–115, 1975.
- [25] M. Tait. Degree Ramsey numbers for even cycles. Discrete Mathematics, 341(1):104–108, Jan. 2018.