Fact-checked by Grok 2 weeks ago

Ramsey theory

Ramsey theory is a branch of in that examines the conditions under which ordered structures inevitably emerge within sufficiently large or complex systems, even when those systems appear random or disordered. The theory asserts that complete disorder is impossible in large enough configurations, guaranteeing the presence of specific patterns or substructures. Named after the British mathematician Frank Plumpton Ramsey (1903–1930), the field originated from his seminal 1930 paper "On a Problem of Formal Logic", published in the Proceedings of the London Mathematical Society, where he proved the first general theorems on unavoidable sets in infinite structures motivated by Hilbert's in logic. Ramsey's work demonstrated that for any given number of colors and sizes, there exists a beyond which certain monochromatic or homogeneous subsets must appear, laying the groundwork for finite and infinite variants of the theory. At its core is , which states that in any sufficiently large structure—such as the edges of a colored with a fixed number of colors—there will be a monochromatic of arbitrary prescribed size, such as a . In terms, the Ramsey number R(s, t) is the smallest integer n such that any 2-coloring of the edges of the complete graph Kn contains either a monochromatic of size s or an independent set of size t. A well-known specific result is R(3, 3) = 6, implying that in any group of six people, where edges represent friendships (one color) or non-friendships (another color), there are either three mutual friends or three mutual strangers—often called the "party problem." Beyond graphs, Ramsey theory encompasses results in arithmetic progressions, hypergraphs, and infinite sets, including (1927), which guarantees monochromatic arithmetic progressions of any length in sufficiently long colorings of the integers, and Schur's theorem (1916), ensuring monochromatic solutions to x + y = z in colorings of natural numbers. These theorems highlight the theory's emphasis on the "preservation of properties under partitions." The field has profound implications in areas like , , and (e.g., and approximation algorithms), underscoring how randomness inevitably yields structure.

Historical Background

Origins in Early 20th Century

The intellectual foundations of Ramsey theory emerged in the early amid broader efforts in and logic to address decidability, infinite structures, and combinatorial order. A pivotal influence came from Hilbert's 1900 address at the in , where he outlined 23 unsolved problems to guide future research. The tenth problem specifically challenged mathematicians to devise a finite for determining whether any given —polynomials with coefficients—admits solutions, framing a core that intertwined , , and formal logic. This query underscored the tension between finite procedures and the complexity of mathematical truth, inspiring investigations into the boundaries of and the detection of patterns within arbitrary systems, themes that resonated with Ramsey theory's focus on inevitable order. In 1916, Issai Schur proved Schur's theorem, which states that for any positive integer r, there exists a number S(r) such that if the integers from 1 to S(r) are colored with r colors, there is a monochromatic solution to the equation x + y = z. This result established early principles of monochromatic solutions in colorings, prefiguring Ramsey-type guarantees. In the 1920s, advances in logical foundations further set the stage by grappling with undecidability and the nature of infinite sets. Thoralf Skolem's 1920 work, refining Leopold Löwenheim's earlier ideas, established the Löwenheim-Skolem theorem, proving that any first-order theory with an infinite model possesses a countable model of the same cardinality as the natural numbers. This result illuminated paradoxes in set theory, such as the coexistence of uncountable concepts within countable frameworks, and emphasized how chaotic infinite domains could harbor structured, enumerable substructures—a conceptual precursor to seeking guaranteed regularities in partitions. Complementing this, Emil Post's contributions in the early 1920s, including his development of truth-table methods for propositional logic and canonical normal forms in 1921, anticipated key insights into computability and undecidability. Post's explorations of systematic reductions in logical expressions highlighted the limitations of mechanical decision procedures for infinite structures, fostering a climate where the inescapability of order in disordered systems became a pressing motif. Combinatorial precursors also gained traction, notably through Bartel Leendert van der Waerden's 1927 on arithmetic progressions, which resolved a posed earlier in the decade. The asserts that for any positive integers k and r, there exists a number N(k, r) such that if the positive integers up to N(k, r) are colored with r colors, then at least one color class contains an of length k. Published in Nieuw Archief voor Wiskunde, this result demonstrated the persistence of monochromatic linear structures in finite partitions of the integers, bridging and partition calculus in a way that prefigured Ramsey-type guarantees of homogeneity. George Pólya's 1918 combinatorial investigations provided another indirect thread, particularly through his analysis of placement constraints akin to problems. In exploring variants of the n-queens puzzle—where queens on a must avoid mutual attacks—Pólya examined conditions under which partial configurations could be extended without conflicts, effectively modeling edge-avoidance in the queens graph. This work on discrete arrangements and their color-like partitions influenced subsequent studies of unavoidable patterns in finite structures, contributing to the evolving toolkit for analyzing divisions and symmetries in .

Frank Ramsey's 1930 Theorem

Frank Plumpton Ramsey (1903–1930) was a philosopher and known for his contributions to and . At the age of 25, he presented his seminal paper "On a Problem of Formal Logic" to the London Mathematical Society on December 13, 1928, which was subsequently published in 1930. The work addressed foundational issues in , including paradoxes arising from and the structure of formal systems. The core mathematical contribution of the paper is the proof of what is now known as the infinite Ramsey theorem. Specifically, it states that for any F, positive integer r, and finite number of classes s, if the r-element subsets of F are partitioned into s classes C_1, \dots, C_s, then there exists an infinite subset A \subseteq F such that all r-element subsets of A lie entirely in one C_i. The two-class case is a fundamental instance. In the context of , this theorem implies that in any finite-coloring of the edges of the complete infinite graph K_\infty, there is a monochromatic infinite . Ramsey's proof relies on the and proceeds by constructing homogeneous subsets through inductive selection. Philosophically, the theorem was motivated by efforts to resolve Hilbert's Entscheidungsproblem, which asked whether there exists an algorithm to determine the truth of any mathematical statement in first-order logic. Ramsey sought to demonstrate inherent order in large logical structures, showing decidability for certain classes of sentences involving existential quantifiers followed by universal ones in relational languages with equality. This approach highlighted the presence of uniform substructures in sufficiently complex systems, countering potential chaos in formal logics. Ramsey's insights were shaped by his interactions with , whom he had translated and critiqued, and his engagements with the during a 1924 visit to , where he met and Hans Hahn. These discussions on logic, decidability, and the nature of mathematical truth influenced the paper's focus on foundational problems. Tragically, Ramsey died on January 19, 1930, at the age of 26 from complications of , cutting short what promised to be a prolific career.

Post-Ramsey Developments

Following Frank Ramsey's foundational work on infinite structures, Richard Rado extended the theory to finite settings in his 1933 doctoral thesis, where he formalized generalizations of partition regularity for systems of linear equations, laying groundwork for finite analogs applicable to graph colorings and other combinatorial structures. Rado's results characterized the conditions under which finite colorings of integers guarantee monochromatic solutions, bridging Ramsey's infinite theorem to practical finite cases that influenced later graph-theoretic developments. In 1935, and George Szekeres advanced the field by proving the finite version of for graphs and introducing the notation for Ramsey numbers R(s,t), which denote the smallest number of vertices ensuring a monochromatic of size s or t in any 2-coloring of the edges of the . They specifically established that R(3,3)=6, demonstrating that in any group of 6 people, there are either 3 mutual friends or 3 mutual strangers, a result that popularized the "party problem" and spurred interest in explicit computations. In the same year, they also proved the , which guarantees a monotonic of length at least \sqrt{n} in any of n distinct real numbers, connecting Ramsey principles to order in permutations and sequences. The 1950s saw informal gatherings among combinatorialists, often facilitated by Erdős's extensive travels and collaborations, which accelerated research in Ramsey theory by fostering discussions on bounds and extensions. These interactions built momentum for computational approaches. A key milestone came in 1955 when Robert E. Greenwood and computed R(3,4)=9, showing that 9 vertices guarantee a monochromatic or independent set of size 4, while providing a for 8 vertices; this effort pioneered systematic techniques for small Ramsey numbers.

Fundamental Concepts

Graph Colorings and Partitions

In Ramsey theory, a foundational concept involves the k-coloring of the edges of a , which partitions the edge set into k subsets corresponding to the color classes. For a K_n, this assigns one of k colors to each of the \binom{n}{2} edges connecting pairs of vertices..pdf) This partitioning framework underpins the core question of Ramsey theory: whether any sufficiently large guarantees a monochromatic subset with desired structural properties. In the setting, it manifests as the search for subgraphs where all edges share the same color, ensuring order emerges from arbitrary divisions. A simple illustration arises in the 2-coloring of the edges of K_6, where any such coloring inevitably produces a monochromatic triangle—a complete subgraph K_3 with all edges the same color..pdf) The pigeonhole principle provides a trivial instance of this paradigm for 2 colors and clique size 2, as the single edge of K_2 must receive one color, yielding a monochromatic K_2. Such colorings extend briefly to hypergraphs in Ramsey theory, where edges are r-tuples of vertices for r \geq 3, and the coloring targets monochromatic complete r-uniform hypergraphs, as originally considered in the field's inception.

Definition of Ramsey Numbers

In Ramsey theory, the Ramsey number R(s, t) is defined as the smallest positive integer n such that every 2-coloring of the edges of the K_n on n vertices contains either a monochromatic of size s in the first color (say, ) or a monochromatic of size t in the second color (say, ). This definition quantifies the threshold at which unavoidable monochromatic structures emerge in edge colorings of complete graphs, serving as the foundational quantitative measure in the finite setting of the theory. Equivalently, R(s, t) > m if and only if there exists a 2-coloring of K_m with no K_s and no K_t, highlighting the existence of graphs just below the Ramsey threshold. The symmetric case, denoted R(k, k), arises when s = t = k and represents the smallest n guaranteeing a monochromatic of size k in any 2-edge-coloring of K_n. These diagonal Ramsey numbers are particularly challenging to determine, as they capture the balanced avoidance of equally sized cliques in both colors, often requiring more intricate constructions for lower bounds compared to off-diagonal cases where s \neq t. Ramsey numbers generalize naturally to multiple colors: for r \geq 3 colors, the multicolor Ramsey number R(s_1, s_2, \dots, s_r) is the smallest n such that any r-coloring of the edges of K_n contains a monochromatic K_{s_i} in the i-th color for some i. This extension broadens the theory to partitions into more than two sets, with early computations focusing on small uniform cases like R(3,3,3). No closed-form formula exists for computing general Ramsey numbers R(s, t), and their values grow exponentially with s and t. This rapid growth is evidenced by probabilistic lower bounds, which demonstrate that R(k, k) > (1 - o(1)) \frac{\sqrt{2}^k k}{e \sqrt{\ln k}} for large k, showing the of colorings avoiding large monochromatic cliques with high probability.

Infinite vs. Finite Settings

Ramsey theory encompasses both and finite settings, each addressing the inevitability of order in large structures under partitions, but differing fundamentally in , proof methods, and implications. In the setting, the focus is on colorings of structures, such as the edges of the complete graph on countably many vertices, guaranteeing the of monochromatic substructures without specifying their size. Ramsey's original in this context asserts that for any finite coloring of the edges, there exists an subset where all edges are the same color, proved using non-constructive techniques like the compactness principle or König's lemma, which rely on trees or limits to establish rather than explicit construction. In contrast, the finite setting examines colorings of finite structures and guarantees monochromatic subgraphs of fixed finite size, but requires determining explicit thresholds known as Ramsey numbers, which grow exponentially and are computationally challenging to compute even for small values. Proofs here are typically constructive, employing , the , or probabilistic methods to bound these numbers, though exact values remain elusive for most cases due to the rapid growth. Unlike the infinite case, finite Ramsey theory does not yield infinite cliques but ensures finite ones in sufficiently large graphs, highlighting the tension between guaranteed order and quantitative precision..pdf) A key distinction lies in the nature of the proofs: infinite versions are often non-constructive, deriving from abstract limits or topological , while finite proofs aim for explicit bounds but encounter immense computational hardness, as the search for optimal Ramsey numbers becomes intractable beyond modest sizes. This non-constructivity in the realm allows for elegant existential statements but offers little insight into finite approximations, whereas finite results provide tangible guarantees at the cost of complexity..pdf) The connection between infinite and finite paradigms is bridged by principles like and ultrafilters, which transfer infinite results to finite approximations by considering limits of finite configurations or "averaging" over infinite sets to select coherent subsets. For instance, the compactness principle implies that if every finite substructure satisfies a Ramsey property, then the infinite structure does as well, enabling derivations of weak finite versions from stronger infinite ones. However, while the infinite Ramsey theorem implies the existence of finite monochromatic cliques for sufficiently large graphs, it does not provide sharp bounds on Ramsey numbers, leaving the precise finite thresholds to separate combinatorial efforts.

Classical Results

Finite Graph Ramsey Theorem

The finite graph Ramsey theorem asserts that for any integers s, t \geq 2, there exists a smallest positive integer R(s,t), known as the Ramsey number, such that in any 2-coloring of the edges of the K_n on n = R(s,t) vertices with and , there is either a of size s or a of size t. This result guarantees the existence and finiteness of R(s,t), ensuring that sufficiently large graphs force monochromatic cliques under edge 2-colorings. Although Frank P. Ramsey's 1930 paper primarily established the infinite version of the theorem, the finite case for graphs follows implicitly from his arguments and was first explicitly proved and formalized by Richard Rado in his 1933 dissertation. Rado's work confirmed the theorem's applicability to finite structures, laying the groundwork for much of modern Ramsey theory in graph colorings. The standard proof proceeds by on s + t. The cases are straightforward: R(2,t) = t and R(s,2) = s, since a on t-1 vertices can be colored entirely to avoid a (clique of size 2) and a of size t, but any coloring of K_t forces one or the other; the symmetric case holds for s. Assume the result holds for all pairs with smaller sum. To establish it for (s,t), set n = R(s-1,t) + R(s,t-1). Consider an arbitrary 2-coloring of the edges of K_n. Fix a vertex v \in K_n. Let d_{\text{red}}(v) be the number of red edges incident to v, so the red neighborhood of v induces a complete subgraph on d_{\text{red}}(v) vertices, and the blue degree is d_{\text{blue}}(v) = n-1 - d_{\text{red}}(v). If d_{\text{red}}(v) \geq R(s-1,t), then by the induction hypothesis applied to the red neighborhood, there is either a red K_{s-1} (which together with v forms a red K_s) or a blue K_t. Similarly, if d_{\text{blue}}(v) \geq R(s,t-1), the blue neighborhood contains either a blue K_{t-1} (with v yielding a blue K_t) or a red K_s. Now suppose neither holds, so d_{\text{red}}(v) < R(s-1,t) and d_{\text{blue}}(v) < R(s,t-1). Then n-1 = d_{\text{red}}(v) + d_{\text{blue}}(v) < R(s-1,t) + R(s,t-1) - 1, implying n < R(s-1,t) + R(s,t-1), a contradiction. Thus, every 2-coloring of K_n contains the desired monochromatic clique, proving R(s,t) \leq R(s-1,t) + R(s,t-1). This recursive inequality yields an explicit upper bound: R(s,t) \leq \binom{s+t-2}{s-1}, which is at most $2^{s+t-2} and grows exponentially in s+t. Lower bounds show that R(s,t) grows at least exponentially in \min(s,t); for the diagonal case R(k,k), Joel Spencer improved the lower bound in 1975 to \Omega\left( \sqrt{2} \, k^{1/2} \, 2^{k/2} \right) using the . These bounds highlight the rapid growth of Ramsey numbers while confirming their finiteness.

Infinite Ramsey Theorem

The infinite Ramsey theorem asserts that for any 2-coloring of the edges of the complete graph K_{\aleph_0} on a countably infinite vertex set, there exists an infinite subset of vertices inducing a monochromatic complete subgraph, or infinite monochromatic clique. This result, proved by Frank P. Ramsey in his 1930 paper, guarantees the existence of such a homogeneous structure in any sufficiently large infinite graph under edge colorings. A standard proof constructs a tree where nodes at level k represent finite monochromatic cliques of size k, and branches extend these cliques using the finite : from a monochromatic K_k, embed it into a sufficiently large complete graph (of size at least R(k+1, k+1)) to force a monochromatic K_{k+1} containing the original K_k. This creates a finitely branching tree of finite monochromatic cliques. then guarantees an infinite path through the tree, corresponding to an infinite sequence of vertices forming an infinite monochromatic clique. The theorem generalizes to any finite number k of edge colors: in any k-coloring of K_{\aleph_0}, there exists an infinite monochromatic clique, proved by induction on k starting from the 2-color case. This infinite version is logically stronger than its finite counterpart and implies the finite Ramsey theorem via the compactness theorem of first-order logic, by encoding finite colorings into a theory whose models correspond to homogeneous sets. The proof requires the axiom of choice (explicitly assumed by Ramsey as the "axiom of selection"); without it, the theorem is independent of ZF set theory, and counterexamples exist in certain models.

Multicolor Extensions

Multicolor Ramsey theory extends the classical two-color results to edge colorings using an arbitrary finite number of colors, focusing on the guaranteed existence of monochromatic cliques in sufficiently large complete graphs. The multicolor Ramsey number R(s_1, s_2, \dots, s_r) is defined as the smallest integer n such that any r-coloring of the edges of the complete graph K_n contains a monochromatic copy of K_{s_i} in the i-th color for some i \in \{1, 2, \dots, r\}. For the symmetric case where all clique sizes are equal, this is denoted R_r(s) or R(s; r), representing the minimal n forcing a monochromatic K_s in any r-coloring of K_n. The finite multicolor Ramsey theorem states that for any fixed r \geq 2 and positive integers s_1, \dots, s_r, the number R(s_1, \dots, s_r) exists and is finite. This follows from the by induction on the number of colors r: assuming the result holds for r-1 colors, consider an r-coloring of K_n where n is large enough to apply the (r-1)-color theorem to the subgraph formed by ignoring the r-th color, ensuring a monochromatic clique in one of the first r-1 colors or a large enough structure in the r-th color to apply the two-color case. This inductive proof yields explicit but rapidly growing upper bounds on the multicolor numbers. In the infinite setting, the multicolor extension asserts that any finite r-coloring of the edges of the countably infinite complete graph K_{\aleph_0} contains an infinite monochromatic clique in at least one color. This generalizes the two-color infinite Ramsey theorem and holds for any finite r, as the proof relies on a compactness argument or iterative application of the finite case to extract increasingly large monochromatic cliques. A concrete example is the three-color Ramsey number R(3,3,3)=17, established in 1955 by constructing a three-coloring of K_{16} without monochromatic triangles and verifying that every three-coloring of K_{17} contains one. More generally, multicolor Ramsey numbers exhibit tower-like growth with respect to the number of colors r: for fixed s, R_r(s) grows at least exponentially and upper bounds from the inductive proof show R(r;s) \leq 2 \uparrow\uparrow (r-1) \cdot R(2;s), where $2 \uparrow\uparrow m denotes a power tower of $2's of height m. This rapid escalation underscores the theorem's strength, as even small increases in r$ lead to astronomically large values.

Specific Examples and Numbers

The Party Problem and R(3,3)

The party problem, popularized by Paul Erdős in his lectures on Ramsey theory, offers an accessible entry point to understanding Ramsey numbers through a social scenario. Consider a gathering of six people, where every pair of individuals is connected either by friendship (represented as a red edge) or by being strangers (a blue edge). In any such configuration, there is guaranteed to be either a group of three mutual friends or three mutual strangers, forming a monochromatic triangle in the underlying graph. This phenomenon illustrates the Ramsey number R(3,3) = 6, the smallest integer n such that every 2-edge-coloring of the complete graph K_n contains a monochromatic copy of K_3. The problem intuitively connects to social network analysis by highlighting inevitable structures—cliques of acquaintances or isolated subgroups—in sufficiently large interpersonal graphs. To establish that R(3,3) = 6, it must be shown both that no monochromatic triangle is forced in K_5 and that one is unavoidable in K_6. For the lower bound, a specific 2-coloring of K_5 evades monochromatic triangles: color the edges of a 5-cycle (a red pentagon) red, while the five diagonals form a blue pentagram (5-pointed star). This configuration, unique up to isomorphism and color reversal, contains no red K_3 (as the red graph is triangle-free) and no blue K_3 (as the blue graph is also a cycle of odd length without triangles). Thus, R(3,3) > 5. For the upper bound, consider an arbitrary 2-coloring of K_6. Select any vertex v, which connects to the other five vertices via five edges. By the pigeonhole principle, at least three of these edges share the same color—without loss of generality, assume they are red and incident to vertices a, b, c. Now examine the subgraph induced by \{a, b, c\}: if any edge among them is red (e.g., a-b), then \{v, a, b\} forms a red K_3; otherwise, all three edges are blue, yielding a blue K_3. In either case, a monochromatic triangle exists, so R(3,3) \leq 6. This proof can be formalized via degrees: in K_6, for vertex v, let d_r(v) and d_b(v) be its red and blue , with d_r(v) + d_b(v) = 5. The minority color has degree at most 2, so the majority color has degree at least 3, recursing to the K_3 case as above. Combining the bounds confirms R(3,3) = 6.

Known Small Ramsey Numbers

The determination of exact Ramsey numbers for small parameters relies heavily on computational methods due to the in graph sizes, making manual verification impractical beyond trivial cases. Known exact values are limited to a handful of small s and t, primarily obtained through exhaustive computer searches using algorithms that enumerate edge colorings or structures to identify critical configurations avoiding monochromatic cliques. These computations confirm the minimal n where any 2-coloring forces a monochromatic K_s or K_t. The following table summarizes the exact known small Ramsey numbers beyond R(3,3)=6, focusing on off-diagonal cases for R(3,t) and small diagonal or near-diagonal values:
Ramsey NumberExact ValueReference
R(3,4)9Greenwood and Gleason (1955)
R(3,5)14Graver and Yackel (1968)
R(3,6)18 and Radziszowski (1995)
R(3,7)23Graver and Yackel (1968)
R(3,8)28 and (1992)
R(3,9)36 and (1995)
R(4,4)18 and Radziszowski (1995)
R(4,5)25 and Radziszowski (1995)
For R(3,6), the value was confirmed via an exhaustive search on graphs of order 17 (which admit avoiding colorings) and 18 (which do not), a process that took significant computational resources in the . The diagonal Ramsey number R(5,5) remains unknown exactly, with current bounds of ≤ R(5,5) ≤ 46 as of 2024; the lower bound stems from constructive graphs avoiding K_5 in either color up to 42 vertices, while the upper bound was recently tightened through advanced computational verification of colorings on 46 vertices. Only the diagonal R(k,k) are exactly known for k ≤ 4 (including trivial cases for k=1,2). Off-diagonal numbers like R(3,t) for small t exhibit a pattern approximating R(3,t) \approx \frac{3t}{2} + O(1), as seen in the progression from R(3,4)=9 to R(3,9)=36, reflecting linear growth with fluctuations due to specific avoiding constructions. As of November 2025, no new exact values for these small classical Ramsey numbers have been established since R(3,9)=36 and R(4,5)=25 in , attributable to the exponential increase in search space—requiring checks of up to $2^{\binom{n}{2}} colorings—which renders brute-force methods infeasible for larger parameters without breakthroughs in algorithms or hardware.

Bounds for Larger Numbers

For larger Ramsey numbers R(k,k), where exact values are unknown beyond small k, researchers have developed asymptotic using probabilistic and combinatorial techniques. These bounds highlight the exponential growth of R(k,k) while revealing significant gaps between them, with the lower bounds growing roughly as (√2)^k and upper bounds as 4^k up to polylogarithmic factors. A seminal lower bound was established by in 1947 using the . By considering a random 2-coloring of the edges of the K_n with n ≈ (√2)^k / (e √k), Erdős showed that the probability of containing a monochromatic K_k is less than 1, implying the existence of a coloring without such a and thus R(k,k) > (1 + o(1)) (√2)^k / (e √k). This construction relies on random graphs avoiding both red and blue cliques of size k, marking one of the first applications of probability to . On the upper bound side, early recursive inequalities, such as those refined by Václav Chvátal, provide R(s,t) ≤ R(s-1,t) + R(s,t-1) with base cases leading to R(s,t) ≲ 2^{2(s+t)}, though more precise analysis yields tighter exponential estimates. A major improvement for diagonal cases came from Miklós Ajtai, János Komlós, and in 1980, who proved R(k,k) < (1 + o(1)) 4^k / √k by applying the Lovász Local Lemma to construct graphs with no clique or independent set of size k. This method avoids the cruder union bounds in probabilistic arguments, enabling better control over bad events like large monochromatic subgraphs. For asymmetric Ramsey numbers R(s,t), probabilistic techniques yield the general lower bound R(s,t) > (s-1)^{(t-1)/2} / \mathrm{poly}(\log t), obtained by random sampling to avoid blue K_t while controlling red cliques. The corresponding upper bound follows from recursive methods, giving R(s,t) < 2^{O(s+t)}. Additional techniques, such as Joel Spencer's entropy-based methods, have been used to derive refined lower bounds by analyzing the information-theoretic entropy of colorings to bound the probability of monochromatic structures. Despite these advances, gaps remain substantial even for moderately small k. For instance, computational efforts as of 2025, including semirandom constructions by Brendan McKay and Stanisław Radziszowski, have confirmed the lower bound R(5,5) > 42 (i.e., ≥ 43), while the upper bound stands at 46; for k > 5, the disparities grow exponentially, underscoring the challenge of closing these intervals.

Generalizations and Advanced Topics

Hypergraph Ramsey Theory

Hypergraph Ramsey theory generalizes the classical Ramsey theorem to structures where edges connect more than two vertices, specifically r-uniform hypergraphs in which every edge is an r-element subset for r ≥ 3. The r-uniform hypergraph Ramsey number R^{(r)}(s,t) is the smallest integer n such that any 2-coloring of the r-subsets of an n-element set forces either a monochromatic complete r-uniform hypergraph K_s^{(r)} (a set of s vertices with all its \binom{s}{r} r-subsets colored red) or K_t^{(r)} (all blue). The existence of finite Ramsey numbers R^{(r)}(s,t) for all finite r, s, t was established by Paul Erdős and Richard Rado in the 1950s through their work on canonical Ramsey theorems, which provide a stronger partitioning result for colorings. Unlike the graph case (r=2), where Ramsey numbers grow exponentially, hypergraph Ramsey numbers exhibit dramatically faster growth; for example, the diagonal numbers R^{(r)}(k,k) have lower bounds exponential in k^{r-1} and upper bounds involving towers of exponentials of height related to r. This rapid growth makes computation and tight bounds particularly challenging even for small parameters. The only known exact value for a non-trivial Ramsey number is R^{(3)}(4,4) = 13, determined computationally by exhaustive search showing that every 2-coloring of the triples on 13 vertices contains a monochromatic K_4^{(3)}, while a coloring on 12 vertices avoids it; no exact diagonal values are known for larger s or t. An infinite analog holds for hypergraphs: in any 2-coloring of the r-subsets of an , there exists an infinite monochromatic K_\infty^{(r)}, proved using compactness of the on the . Lower bounds for diagonal hypergraph Ramsey numbers are derived from probabilistic methods applied to random r-uniform hypergraphs. Specifically, a random 2-coloring of the r-subsets of has positive probability of avoiding monochromatic K_k^{(r)} when n is sufficiently small, yielding the bound R^{(r)}(k,k) > 2^{c k^{r-1}} for some constant c > 0 depending on r. For r=3, this gives double-exponential growth in k. Even R^{(3)}(5,5) remains unknown, with the current lower bound of 88 from constructive colorings avoiding monochromatic K_5^{(3)} on 87 vertices, while upper bounds from general theorems exceed 10^6, underscoring the vast gaps in our understanding as of 2025.

Arithmetic and Set-Theoretic Variants

Arithmetic Ramsey theory examines colorings of the integers and the emergence of monochromatic s, extending the combinatorial principles of Ramsey theory to additive structures. The van der Waerden number W(k, r) is defined as the smallest positive integer n such that any r-coloring of the set \{1, 2, \dots, n\} contains a monochromatic of length k. This concept was introduced by in 1927, building on earlier work by Schur on sum-free sets. Exact values for van der Waerden numbers are known only for small parameters; for instance, W(3, 2) = 9, meaning that in any 2-coloring of \{1, \dots, 9\}, there is a monochromatic 3-term arithmetic progression, but there exists a 2-coloring of \{1, \dots, 8\} avoiding such progressions. Similarly, W(3, 3) = 27, so any 3-coloring of \{1, \dots, 27\} forces a monochromatic 3-term progression, while a 3-coloring of \{1, \dots, 26\} without one is possible. For larger k or r, these numbers remain unknown, but upper bounds have been established showing super-exponential growth. Specifically, Timothy Gowers proved that W(3, k) \leq 2^{2^{2^{c \log k}}} for some constant c > 0, implying tower-type bounds that grow faster than any exponential function. In the set-theoretic variant, Ramsey theory addresses partitions of infinite sets, particularly focusing on the existence of homogeneous sets under certain measurability conditions. A key result is the Galvin–Prikry theorem, which states that every Borel subset of the space of infinite subsets of the natural numbers, [ \omega ]^\omega, is Ramsey: for any such Borel set B, there exists an infinite set H \subseteq \omega such that either [H]^\omega \subseteq B or [H]^\omega \cap B = \emptyset. This theorem, proved in the early 1970s, extends the infinite Ramsey theorem to analytic sets with Borel complexity, providing a partition calculus for descriptive set theory. A density-based extension of these ideas appears in from 1975, which asserts that any subset of the integers with positive upper asymptotic density contains of arbitrary finite length. Formally, for every \delta > 0 and integer k \geq 3, there exists N = N(k, \delta) such that any subset A \subseteq \{1, 2, \dots, N\} with |A| \geq \delta N contains a k-term . This result, often viewed as a density analogue of , has profound implications in additive combinatorics and was proved using uniformity norms on the integers. Recent advances connect these arithmetic Ramsey principles to , notably through the of 2004, which establishes that the primes contain arithmetic progressions of arbitrary length, despite having zero density. This builds directly on by adapting arguments to the primes. As of 2025, extended this framework to higher-degree polynomial progressions in dense sets, resolving long-standing conjectures on multidimensional patterns and linking to the structure of additive bases in the integers.

Structural Ramsey Theory

Structural Ramsey theory extends classical Ramsey theory to classes of finite relational structures, focusing on partition properties under rather than just edge colorings of graphs. In this framework, a class \mathcal{K} of finite structures has the Ramsey property if, for every A, B \in \mathcal{K} and every positive k, there exists C \in \mathcal{K} such that C \to (B)^k_A, meaning that in every k-coloring of the embeddings of A into C, there is a monochromatic embedding of B into C. This generalizes the notion of Ramsey numbers to arbitrary relational , such as tournaments or posets, where the "Ramsey number" for a fixed structure B in \mathcal{K} is the minimal size of C forcing a monochromatic copy of B in any coloring of the embeddings. A foundational result in this area is the Nešetřil–Rödl theorem from the , which establishes that every finite relational language admits a class of finite structures with the Ramsey property under embeddings. Specifically, for any finite structure A in a relational language and any number of colors k, there exists a finite structure C such that any k-coloring of the copies of A in C contains a monochromatic copy of A. This theorem, proved using partite constructions and inductive methods, confirms that structural Ramsey properties hold broadly for finite models, providing existence without explicit bounds. In the context of ordered graphs, where vertices are linearly ordered, the Ramsey property adapts Rado's selection to ensure that finite colorings force ordered substructures. This generalization implies that for ordered complete graphs, any coloring of ordered embeddings yields a monochromatic ordered , preserving the linear order in the homogeneous set. The stepping-up further extends these results to higher dimensions, lifting Ramsey properties from graphs to hypergraphs by constructing higher-uniformity structures from lower ones, thus proving existence of Ramsey witnesses in multidimensional relational settings without computing explicit sizes. Applications of structural Ramsey theory appear in topological dynamics, where Ramsey classes generate minimal dynamical systems with homogeneous actions, linking combinatorial partitions to recurrent behaviors in flows. Recent advances as of 2025 include developments in measurable Ramsey theory for spaces, showing that analytic sets in weak A_2 spaces are strategically Ramsey, enabling partition properties for continuous colorings on complete separable metric spaces.

Applications and Connections

In Combinatorial Optimization

Ramsey theory finds practical applications in combinatorial optimization, particularly in graph coloring problems aimed at conflict-free scheduling and frequency assignment. In these contexts, edges represent potential conflicts, such as interference between communication channels or scheduling overlaps, and coloring assigns resources like frequencies or time slots to avoid monochromatic cliques that could lead to unresolved conflicts. Ramsey numbers provide upper bounds on the minimal number of colors required to color a graph without creating a monochromatic clique of size k, ensuring that large-scale systems can be optimized with a bounded number of resources. For instance, in frequency assignment for wireless networks, the off-diagonal Ramsey number R(3,t) bounds the colors needed to prevent a monochromatic triangle while allowing larger independent sets, facilitating efficient spectrum allocation. A notable influence comes from Paul Erdős's , which establishes lower bounds on diagonal Ramsey numbers R(k,k) > (1+o(1)) \sqrt{2}^k / e, and has inspired approximation algorithms for variants of the traveling salesman problem (TSP). These lower bounds demonstrate the existence of graphs without large s or independent sets, informing hardness results and approximation ratios in metric TSP where edge colorings model tour constraints. By leveraging such constructions, algorithms achieve subexponential guarantees for TSP instances with bounded clique sizes, highlighting Erdős's foundational role in bridging Ramsey theory to optimization heuristics. Hedetniemi's , proposed in the 1960s, posited that the chromatic number of the of two s equals the minimum of their individual chromatic numbers. This was disproved by Shitov in 2019. A publication by Xiaoyu He and Yuval Wigderson showed that the conjecture is asymptotically false, with counterexamples where the chromatic number of the product is bounded away from the minimum by a positive fraction, impacting optimization in graph products for network design. In clique cover problems, Ramsey numbers enable approximation algorithms by relating the cover size to independent set approximations in the . Specifically, using the bound R(3,t) \sim t^2 / \log t, algorithms achieve approximations for covering graphs with cliques while avoiding large independent sets. Such methods are pivotal for optimizing clique-based partitions in . Recent advances in 2025 have extended Ramsey theory to quantum computing, where it bounds qubit entanglement in error-correcting codes. By analyzing Ramsey numbers for quantum graphs, researchers diagnose error thresholds and construct codes that minimize entanglement overhead while correcting phase flips, enhancing fault-tolerant quantum optimization solvers. For small Ramsey numbers like R(3,3)=6, these bounds inform hybrid classical-quantum algorithms for scheduling under noise. Ramsey theory establishes profound connections to , particularly through its infinite form and implications for foundational principles like the in . The infinite Ramsey theorem, which guarantees the existence of infinite homogeneous sets in colorings of infinite structures, combines with the —stating that a first-order theory has a model every finite does—to derive the finite Ramsey theorems from their infinite counterparts. This interplay highlights how infinite combinatorial guarantees underpin the finite cases central to logic, enabling constructions of models that satisfy infinite consistency conditions via finite approximations. Beyond this, Ramsey-theoretic principles are employed to reveal limitations in formal systems, such as proving undecidability or independence for fragments of Peano by constructing colorings that evade provable homogeneity in bounded . A seminal example is the Paris-Harrington theorem, developed in the 1970s, which strengthens the finite Ramsey theorem by incorporating a growth condition on the size of homogeneous sets. Specifically, for any positive integers k, r, and n, there exists a number N such that any k-coloring of the r-element subsets of an N-element set admits a homogeneous set H of size m > n where all r-subsets of H receive the same color, and m > r. Remarkably, while true, this theorem is independent of , meaning neither it nor its negation can be proved within , yet it is provable in stronger systems like the subsystem of known as ACA0. This result, the first natural mathematical statement shown independent of , underscores Ramsey theory's role in by exhibiting how combinatorial principles can outstrip the expressive power of basic arithmetic axioms. In , Ramsey theory intersects with , where problems like finding monochromatic s in edge-colored s are analyzed with respect to parameters such as subgraph size. The task of detecting a monochromatic of size k in a 2-edge-colored is W-hard when parameterized by k, implying that unless FPT = W, no fixed-parameter tractable algorithm exists, even though the underlying is W-complete. This hardness arises because Ramsey subgraphs inherit the intractability of clique detection across color classes, motivating approximation algorithms and kernelization techniques for practical analysis in large-scale data structures. Descriptive set theory leverages Ramsey principles to quantify homogeneity in analytic settings, particularly through the notion of Ramsey degrees for Borel colorings. For a finite relational structure F and a class K of structures (e.g., Borel definable graphs), the Ramsey degree T(F; K) measures the minimal complexity of homogeneous extensions under colorings. Formally, it is defined as T(F; K) = \sup\left\{ \min\left\{ |c''([X]^{|F|})| : X \in K \right\} : c : \bigcup_{X \in K} \binom{X}{|F|} \to \omega \right\}, where the supremum is over all finite-valued colorings c of the |F|-subsets, and c''([X]^{|F|}) denotes the image of the coloring restricted to homogeneous sets. This degree captures the inevitable "splitting" of colors in Borel contexts, with exact computations for simple structures like cliques revealing in spaces. Recent applications of finite Ramsey theory extend to , where it informs pattern avoidance in graph-based models to mitigate . In graph neural networks (GNNs), Ramsey guarantees ensure the emergence of monochromatic motifs like cliques or cycles in edge-colored representations of , guiding strategies that remove redundant to avoid dense "clique-like" overfits while preserving . For instance, recursive methods on Ramsey-constrained graphs have been proposed to optimize , balancing sparsity with homogeneity in learned embeddings.

Open Problems and Recent Advances

One of the most prominent open problems in Ramsey theory is determining the exact value of the diagonal Ramsey number R(5,5), the smallest integer such that any 2-coloring of the edges of the on that many vertices contains a monochromatic of size 5. The current known bounds are $43 \leq R(5,5) \leq 46 as of 2024. Another longstanding question concerns the precise asymptotic growth rate of the diagonal Ramsey numbers R(k,k). It is known that R(k,k) grows exponentially in k, with lower bounds of the form \Omega(2^{k/2}) and upper bounds of O(4^k / \sqrt{k}), but the exact base of the exponential remains unknown, with significant gaps persisting between these estimates. In the set-theoretic foundations of Ramsey theory, the extent to which infinite versions of Ramsey's theorem hold without the axiom of choice remains an active area of investigation. While the standard infinite Ramsey theorem for finite colorings is provable in ZF set theory alone, stronger partition properties and their implications for cardinal comparability often require choice principles, with partial results exploring models where certain Ramsey-like statements fail. A significant recent advance came in 2021 (published from earlier work), when Campos, Jenssen, Michelen, and Sahasrabudhe analyzed the triangle-free process—a randomized algorithm for building large triangle-free graphs—and used it to establish improved lower bounds for the off-diagonal Ramsey numbers R(3,t), showing R(3,t) > (1-o(1)) \frac{t^{5/2}}{(\log t)^{3/2} 2^{O(\sqrt{\log t})}} asymptotically. This approach not only refined previous estimates but also provided insights into the structural properties of graphs produced by such processes. In 2024, , He, and Wigderson extended the study of Ramsey numbers to sparse directed graphs, proving that for certain families of bounded-degree digraphs, the Ramsey numbers can exceed any in the number of vertices while remaining subexponential, and their methods yield algorithmic constructions for finding monochromatic copies in colorings of sparse structures. In 2025, , , and Xie provided an exponential improvement to lower bounds for diagonal Ramsey numbers using novel probabilistic constructions. Emerging connections to have introduced new tools for probing Ramsey numbers. For instance, techniques have been applied to compute bounds for small Ramsey numbers by embedding graph colorings into problems, demonstrating feasibility for R(3,3) and suggesting scalability for larger instances. Similarly, random-projector methods in quantum diagnostics offer statistical estimation frameworks for Ramsey thresholds using few qubits. Machine learning approaches are increasingly employed to search for Ramsey graph counterexamples and tighten bounds. Reinforcement learning frameworks, such as RamseyRL, have successfully identified new lower bounds for small off-diagonal numbers like R(K_{2,5}, K_{3,5}) by training agents to construct extremal graphs, outperforming traditional search heuristics in high-dimensional spaces. These techniques highlight a growing between and combinatorial search in Ramsey theory.

References

  1. [1]
    [PDF] Ramsey Theory
    May 13, 2016 · Ramsey theory is a branch of mathematics that focuses on the appearance of order in a substructure given a structure of a specific size.
  2. [2]
    [PDF] 12 Ramsey theory
    Ramsey theory is the study of structure and of disorder. The main message of Ramsey theory is that complete disorder is impossible—any sufficiently large ...
  3. [3]
    [PDF] ON A PBOBLEM OF FOKMAL LOGIC This paper is primarily ...
    ON A PROBLEM OF FORMAL LOGIC. 265. It -may happen that F contains a member Xi ... in the first part of the paper. Our line of argument is as follows ...
  4. [4]
    [PDF] An Introduction to Ramsey Theory on Graphs - VTechWorks
    May 2, 2011 · Ramsey theory deals with finding order amongst apparent chaos. Given a mathematical structure of interest and a setting where it may appear, ...
  5. [5]
    [PDF] RAMSEY THEORY Contents 1. Introduction 1 2. Arithmetic ...
    Aug 9, 2016 · This branch of mathematics concerns with “the preservation of properties under set partition”, and is applied in areas such as number theory, ...
  6. [6]
    Computational Ramsey Theory - cs.rit.edu
    Ramsey theory is often regarded as the study of how order emerges from randomness. It has applications in parallel programming, approximation algorithms, game ...
  7. [7]
    [PDF] HILBERT'S TENTH PROBLEM: What can we do with Diophantine ...
    For a given Diophantine equation the problem of deciding whether it has a solution in integers and the problem of deciding whether it has a solution in natural.
  8. [8]
    The Algebra of Logic Tradition - Stanford Encyclopedia of Philosophy
    Mar 2, 2009 · In 1920 Thoralf Skolem (1887–1963) simplified Löwenheim's proof by introducing Skolem normal forms, and in 1928 Skolem replaced his use of ...
  9. [9]
    [PDF] Emil Post and His Anticipation of Gödel and Turing
    Oct 3, 2024 · Emil Post is known to specialists in mathematical logic for several ideas in logic and computability theory: the structure theory of ...
  10. [10]
    [PDF] ramsey theory: van der waerden's theorem and the hales-jewett ...
    Actually, Van der Waerden's Theorem was originally stated in 1927 in the fol- lowing way: if the positive integers are partitioned into two classes, then at ...
  11. [11]
    A survey of known results and research areas for n-queens
    Jan 6, 2009 · In this paper we survey known results for the n -queens problem of placing n nonattacking queens on an n × n chessboard and consider ...<|control11|><|separator|>
  12. [12]
    Frank Ramsey (1903 - 1930) - Biography - MacTutor
    ... a meaningless game with marks on paper. His second paper on mathematics On a problem of formal logic was read to the London Mathematical Society on 13 ...Missing: details | Show results with:details
  13. [13]
  14. [14]
    Frank Ramsey - Stanford Encyclopedia of Philosophy
    Aug 14, 2019 · Ramsey also wrote his influential paper, 'On a Problem of Formal Logic' (1930), which contains the mathematical results now called 'Ramsey's ...Life and Works · The Foundations of Logic and... · Contributions to Mathematics
  15. [15]
    [PDF] Ramsey Theory Mathias Schacht - Fachbereich Mathematik
    In 1933 Rado (in his PhD-thesis [71] supervised by Schur) found a beautiful generalisation of Theorems 1.2-1.4. In fact he characterised all systems of lin- ear ...
  16. [16]
    [PDF] Ramsey Theory in the Work of Paul Erdős - UCSD Math
    Ramsey numbers are denoted as in most older Erdős papers by symbols of f(k), ƒ(3,k), g(k). He then defines analogously the function h(k, l) as "the least ...
  17. [17]
    Ramsey Theory in the Work of Paul Erdős - SpringerLink
    B. L. van der Waerden, Beweis einer Baudetschen Vermutung, Nieuw Arch. Wisk. 15 (1927), 212–216. B. L. van der Waerden, How the Proof of Baudet's Conjecture ...
  18. [18]
    [1111.3824] Higher-order Erdos--Szekeres theorems - arXiv
    Nov 16, 2011 · A famous 1935 Erdos--Szekeres theorem asserts that every such P contains a monotone subsequence S of \sqrt N points.
  19. [19]
    [PDF] Erd˝os–Szekeres theorem for multidimensional arrays - ETH Zürich
    Nov 5, 2019 · A classical paper of Erd˝os and Szekeres [14] from 1935 is one of the starting points of a very rich discipline within combinatorics: Ramsey ...
  20. [20]
    Combinatorial Relations and Chromatic Graphs | Canadian Journal ...
    Nov 20, 2018 · Information. Type: Research Article. Information. Canadian Journal of Mathematics , Volume 7 , 1955 ... R. E. Greenwood (a1) and A. M. Gleason (a2) ...
  21. [21]
    [PDF] Pigeonhole Principle and the Probabilistic Method - MIT Mathematics
    Feb 20, 2015 · The above proposition is a special case of Ramsey's theorem proved in 1930, which is a founda- tional result in Ramsey theory. This consists ...
  22. [22]
    [PDF] Walk through Combinatorics: Compactness principle - Boris Bukh
    The compactness principle states that if a hypergraph's finite induced hypergraphs are properly r-colorable, then the whole hypergraph is properly r-colorable.
  23. [23]
    [PDF] infinite ramsey theory - stefan geschke
    We will use some simple facts about compact right-topological semi- groups to prove highly non-trivial theorems in both finite and infinite. Ramsey theory.
  24. [24]
    [PDF] Using Ultrafilters to Prove Ramsey-type Theorems - arXiv
    Jan 4, 2022 · The purpose of this article is to introduce ul- trafilters in a friendly manner and present some applications to the branch of combinatorics.
  25. [25]
    Studien zur Kombinatorik | Mathematische Zeitschrift
    Download PDF · Mathematische Zeitschrift Aims and scope Submit manuscript ... Cite this article. Rado, R. Studien zur Kombinatorik. Math Z 36, 424–470 (1933).
  26. [26]
    [PDF] Ramsey Theory for Graphs
    It is customary in Ramsey theory to think of partitions as colourings: ... now define the edge set of Gk+1 so that the obvious extensions of 'p to all ...
  27. [27]
    [PDF] Ramsey Theory
    In 1928 the English mathematician Frank Plumpton Ramsey published his pa- per On a problem of formal logic [13] in which he proved what would become.<|control11|><|separator|>
  28. [28]
    [PDF] Completeness Compactness and an application to Ramsey's Theorem
    Apr 6, 2009 · We use the Compactness Theorem to prove the finitary version of Ramsey's Theorem from the infinitary version. First, some notation: If X is any ...Missing: principle | Show results with:principle
  29. [29]
    The independence of Ramsey's theorem | The Journal of Symbolic ...
    The independence of Ramsey's theorem. Published online by Cambridge University Press: 12 March 2014. E. M. Kleinberg. Show author details ...
  30. [30]
  31. [31]
    [PDF] Ramsey Numbers - MIT Mathematics
    May 13, 2018 · The main subject of the theory are complete graphs whose subgraphs can have some regular properties. Most commonly, we look for monochromatic ...
  32. [32]
    [PDF] The Study of Ramsey numbers r(C_k, C_k, C_k)
    K5 that avoids monochromatic triangles by coloring a pentagon in red and inscribing a star of blue within. Obviously, this is a 3-good 2-coloring of. K5.
  33. [33]
    Small Ramsey Numbers | The Electronic Journal of Combinatorics
    Aug 1, 1994 · We present data which, to the best of our knowledge, includes all known nontrivial values and bounds for specific graph, hypergraph and multicolor Ramsey ...Missing: revision | Show results with:revision
  34. [34]
    [2409.15709] $R(5,5)\le 46$ - arXiv
    We prove that the Ramsey number R(5,5) is less than or equal to~46. The proof uses a combination of linear programming and checking a large number of cases by ...Missing: Radziwiłł lower semirandom
  35. [35]
    [PDF] A survey of hypergraph Ramsey problems
    Just as for diagonal Ramsey numbers, a double exponential in nc lower bound for r4(5,n) and r4(6,n) would imply rk(k + 1,n) > twrk-1(nc ) and rk(k + 2,n) > twrk ...<|separator|>
  36. [36]
    [PDF] The First Classical Ramsey Number for Hypergraphs Is Computed
    Since R(13) turns out to be empty, we will infer that R(14) is too, and consequently R(4,4;3)=13. Let R(n,f ,t) be the set of all R(n) systems G with t ...
  37. [37]
    [PDF] Szemerédi's theorem and problems on arithmetic progressions
    Szemerédi's famous theorem on arithmetic progressions asserts that every subset of integers of positive asymptotic density contains arith- metic progressions of ...
  38. [38]
    [PDF] On the history of van der Waerden's theorem on arithmetic ...
    In this expository note, we discuss the celebrated theorem known as “van der Waerden's theo- rem on arithemetic progressions,” the history of work on upper and ...
  39. [39]
    [PDF] some results on a class of mixed van der waerden numbers
    Nov 2, 2016 · w(3; 2) = 9, w(3; 3) = 27, w(4; 2) = 35, w(3; 4) = 76,. 2010 AMS ... Snevily, On the van der Waerden numbers w(2; 3, t), Discr. Appl ...Missing: sources | Show results with:sources
  40. [40]
    Borel Sets and Ramsey's Theorem - jstor
    194 FRED GALVIN AND KAREL PRIKRY sets of w; X's, Y's, and Z's are finite subsets of w; small letters are elements of w. A < B means that a < b for all a e A ...
  41. [41]
    The primes contain arbitrarily long arithmetic progressions
    We prove that there are arbitrarily long arithmetic progressions of primes. There are three major ingredients. The first is Szemerédi's theorem, which asserts ...Missing: extensions | Show results with:extensions
  42. [42]
    Terence Tao's Breakthrough Extends Green-Tao Theorem to ...
    Jul 25, 2025 · Terence Tao announced a breakthrough extending the Green-Tao theorem to higher-degree polynomials, resolving a conjecture on patterns in dense ...
  43. [43]
    Twenty years of Nešetřil's classification programme of Ramsey classes
    Jan 28, 2025 · Following the influential Nešetřil-Rödl theorem, several Ramsey classes have been identified.
  44. [44]
    A New Proof of the Nešetřil–Rödl Theorem
    Jul 15, 2017 · In this paper we give a new proof of the Nešetřil–Rödl Theorem, a deep result of discrete mathematics which is one of the cornerstones of the structural Ramsey ...Missing: definition | Show results with:definition
  45. [45]
    Canonizing structural Ramsey theorems - ScienceDirect.com
    One of the cornerstones of the structural Ramsey theory is the famous Nešetřil–Rödl Theorem whose formulation requires some terminology. Let Θ = ( R i ) i ∈ I ...
  46. [46]
    [PDF] Ordered Ramsey numbers - MIT Mathematics
    ... Ajtai, Komlós and Szemerédi [2] improves this to r(Kn,K3) = O( n2 log ... Before tackling this theorem, we recall the famous. Lovász local lemma, which ...
  47. [47]
    [0907.0283] An improved bound for the stepping-up lemma - arXiv
    Jul 2, 2009 · An ingenious construction of Erdős and Hajnal known as the stepping-up lemma gives a negative partition relation for higher uniformity from one of lower ...Missing: dimensions | Show results with:dimensions
  48. [48]
    Ramsey theory and topological dynamics for first order theories - arXiv
    Dec 22, 2019 · We investigate interactions between Ramsey theory, topological dynamics, and model theory. We introduce various Ramsey-like properties for first order theories.Missing: applications | Show results with:applications
  49. [49]
    [PDF] CONFLICT-FREE COLORING - Corelab
    Conflict-free coloring hypergraphs induced by geometric shapes, like intervals on the line, or disks on the plane, has applica- tions in frequency assignment ...
  50. [50]
    [PDF] Frequency Planning and Ramifications of Coloring
    Dec 21, 2000 · Colors simply correspond to frequencies. Now, a conflict-free coloring of the vertices corresponds to an assignment of frequencies to the ...Missing: Ramsey scheduling
  51. [51]
    [PDF] Asy Lower Bounds on Ramsey Numbers
    Erdös developed it to get better lower bounds on R(k) as shown here ... Show that if you do a probabilistic experiment then you (a) always get the ...
  52. [52]
    [PDF] Hedetniemi's conjecture is asymptotically false - Yuval Wigderson
    Feb 28, 2020 · More precisely, we prove that there exists an absolute constant δ > 0 such that for all c sufficiently large, there exist graphs G and. H with ...
  53. [53]
    [PDF] Hedetniemi's conjecture is asymptotically false
    2021. TLDR. It is shown that if this "Hedetniemi-type" equality is not satisfied for some pair of graphs then the analogous equality is also not satisfaction ...
  54. [54]
    [PDF] Ramsey numbers and an approximation algorithm for the vertex ...
    Mar 12, 2015 · Summary. We show two results. First we derive an upper bound for the special Ramsey numbers rk(q), where rk(q) is the largest number of ...
  55. [55]
    [PDF] Ramsey Theory Applications - The Electronic Journal of Combinatorics
    Ramsey theory applications include number theory, algebra, geometry, topology, set theory, logic, ergodic theory, information theory, and theoretical computer ...Missing: combinatorialists | Show results with:combinatorialists
  56. [56]
    [PDF] A Mathematical Incompleteness in Peano Arithmetic
    Ramsey Theorem is provable in Peano Arithmetic. Our proof of 1.2 will use the Infinite Ramsey Theorem, and cannot be carried out in PA. 1.3. MAIN THEOREM.Missing: paper | Show results with:paper
  57. [57]
    Parameterized Complexity of Finding Subgraphs with Hereditary ...
    Jul 21, 2000 · We show that if II includes all independent sets but not all cliques or vice versa, then the problem is hard for the parameterized class W[1] ...
  58. [58]
    [PDF] Infinite-dimensional Ramsey theory for binary free amalgamation ...
    Dec 23, 2023 · This yields a direct recovery of exact big Ramsey degrees, namely one color per each finite diary coding a given finite structure, whereas ...
  59. [59]
    [2105.02383] Ramsey numbers of sparse digraphs - arXiv
    May 6, 2021 · Authors:Jacob Fox, Xiaoyu He, Yuval Wigderson. View a PDF of the paper titled Ramsey numbers of sparse digraphs, by Jacob Fox and 2 other ...
  60. [60]
    Toward computing bounds for Ramsey numbers using quantum ...
    Jul 11, 2025 · For example, R(5) is currently known to lie between 43 and 48. There are also less standard Ramsey numbers, some of which are much easier to ...
  61. [61]
    Random-projector quantum diagnostics of Ramsey numbers and a ...
    Aug 22, 2025 · Abstract:We introduce a statistical framework for estimating Ramsey numbers by embedding two-color Ramsey instances into a Z_2 \times ...
  62. [62]
    RamseyRL: A Framework for Intelligent Ramsey Number ... - arXiv
    Aug 23, 2023 · This paper explores the application of a best first search algorithm and reinforcement learning (RL) techniques to find counterexamples to specific Ramsey ...
  63. [63]
    Reinforcement learning for graph theory, II. Small Ramsey numbers
    Mar 29, 2024 · We illustrate this application by providing lower bounds for the small Ramsey numbers R(K_{2,5}, K_{3,5}), R(B_3, B_6) and R(B_4, B_5) and by ...<|control11|><|separator|>