Fact-checked by Grok 2 weeks ago

Kissing number

The kissing number k(n) in n-dimensional is defined as the maximum number of equal non-overlapping unit spheres that can simultaneously touch a central unit sphere without their interiors intersecting. Equivalently, it represents the largest number of points that can be placed on the surface of an n-dimensional such that the angular separation between any two points from the center is at least 60 degrees. The problem traces its origins to the late , when conjectured that in three dimensions the kissing number is 12, while David Gregory argued that 13 might be possible, sparking a debate known as the "thirteen spheres problem." This conjecture arose in the context of early studies on sphere packings, with applications to arranging cannonballs in pyramidal s, as posed in Sir Walter Raleigh's problem to predict the number of cannonballs in a stack from its shape. The three-dimensional case remained unresolved for centuries until Kurt Schütte and Bram van der Waerden proved in 1953 that k(3) = 12, confirming Newton's guess through a combinatorial argument involving and . A simpler proof using and spherical codes was later provided by Oleg Musin in 2003. Exact values of the kissing number are known only in specific dimensions: k(1) = 2, achieved by two points on a line; k(2) = 6, corresponding to the vertices of a around a central circle; k(3) = 12, realized by the vertices of a regular icosahedron; k(4) = 24, proven by Musin in 2003 using spherical codes and optimization techniques; k(8) = 240, attained via the E_8 lattice; and k(24) = 196560, from the Leech lattice. These exceptional dimensions—1, 2, 3, 4, 8, and 24—highlight deep connections to root systems and highly symmetric lattices in mathematics. In higher dimensions, the kissing number grows exponentially, but exact values remain elusive except in the cases above; for example, bounds in dimensions 5 through 7 and 9 through 23 are known, with improvements including new lower bounds in dimensions 17–21 using modified configurations reported in late 2024, in dimension 11 to 593 using an coding agent in May 2025, and in dimensions 10 (≥510), 11 (≥592), and 14 (≥1932) by combinatorial methods in October 2025. The problem is intimately linked to densities, error-correcting codes, and spherical designs, influencing fields from to theory. Comprehensive tables and bounds up to dimension 72 are available, underscoring ongoing research into these geometric limits.

Definition and Fundamentals

Mathematical Statement

The kissing number k(n) in n-dimensional \mathbb{R}^n is defined as the maximum number of equal unit spheres (each of radius 1) that can simultaneously touch a central unit sphere without overlapping interiors. This configuration ensures that each kissing sphere is tangent to the central one, while maintaining that prevents any two kissing spheres from intersecting. Formally, k(n) is the largest integer m such that there exist m points x_1, \dots, x_m \in \mathbb{R}^n satisfying \|x_i\| = 1 for $1 \leq i \leq m and \|x_i - x_j\| \geq 1 for all $1 \leq i < j \leq m. These points lie on the unit sphere S^{n-1} = \{ x \in \mathbb{R}^n : \|x\| = 1 \}, where the norm is the . In the sphere packing interpretation, the centers of the kissing spheres are located at c_i = 2 x_i, placing them at distance 2 from the central sphere's center at the origin (ensuring tangency) and at least distance 2 apart from each other (ensuring non-overlap). Geometrically, this setup requires the centers of the kissing spheres to form a spherical code on the (n-1)-dimensional sphere of radius 2 centered at the origin, with minimum pairwise distance 1 after normalization. The condition \|x_i - x_j\| \geq 1 corresponds to an angular separation of at least $60^\circ between any two points, since \|x_i - x_j\|^2 = 2 - 2 x_i \cdot x_j \geq 1 implies x_i \cdot x_j \leq 1/2 = \cos(60^\circ). The problem originated in a 17th-century debate between and over the value in three dimensions.

Historical Development

The kissing number problem traces its origins to a famous 1694 correspondence between and , in which Newton contended that exactly 12 equal spheres could touch a central sphere in three-dimensional space, while Gregory proposed that 13 might be feasible. This debate highlighted the challenge of arranging non-overlapping spheres tangent to a central one, a question rooted in the geometry of sphere packings and celestial mechanics. Although no formal proof emerged at the time, the problem captured early interest in the maximal number of mutually tangent spheres around a given sphere. During the 19th century, sporadic developments advanced the understanding of sphere packings, with incomplete arguments attempting to confirm for three dimensions, though a rigorous resolution remained elusive. The first complete proof that the kissing number in three dimensions is 12 appeared only in 1953, provided by and through a detailed geometric analysis showing that a 13th sphere would necessarily overlap with others. Progress in higher dimensions accelerated in the late 20th century. In 1978, Grigory Kabatiansky and Vladimir Levenshtein introduced a powerful method using zonal spherical functions to derive asymptotic upper bounds for packings on the sphere, yielding the estimate $2^{0.401n(1+o(1))} for the kissing number k(n) in n dimensions. This approach marked a shift toward analytic techniques for high-dimensional bounds. Building on related ideas from coding theory, Levenshtein in 1979, and independently Andrew Odlyzko and Neil Sloane, applied Delsarte's linear programming method to prove exact values using exceptional root systems: k(8) = 240 from the E_8 lattice and k(24) = 196560 from the Leech lattice. The 1990s ushered in a computational era, leveraging advances in and numerical methods to explore kissing configurations and improve bounds in intermediate dimensions. Christine Bachoc and others constructed extremal , such as a 40-dimensional example with minimum norm 8, contributing to better lower bounds and paving the way for semidefinite programming techniques in subsequent decades.

Exact Values in Low Dimensions

One Dimension

In one dimension, the kissing number k(1) is exactly 2, meaning that precisely two non-overlapping unit spheres (degenerate to closed intervals of length 2 on the real line) can touch a central unit sphere without their interiors intersecting. This trivial case arises from the linear geometry of the line, where the central sphere, say with center at the origin and endpoints at -1 and +1, can be touched by one sphere centered at -2 (endpoints -3 to -1) on the left and another at +2 (endpoints +1 to +3) on the right, each tangent at the central sphere's endpoints. Any attempt to place a third sphere would require overlapping with one of the existing pair due to the absence of additional space along the line. Geometrically, this configuration visualizes the fundamental constraint of one-dimensional space: spheres align strictly in a sequence, with each touching the central one only at its boundary points, and no angular freedom allows for more contacts. This setup corresponds to the densest packing of intervals on the line, where the kissing arrangement is equivalent to placing two non-overlapping intervals of length 2 adjacent to and touching the central interval's endpoints, saturating the available positions without gaps or overlaps. The A_1 lattice achieves this optimal kissing number, underscoring the simplicity of the one-dimensional case as a baseline for higher-dimensional generalizations.

Two Dimensions

In two dimensions, the kissing number k(2) is 6, meaning that exactly six non-overlapping unit circles can touch a central unit circle. This configuration corresponds to the regular hexagonal packing, where the centers of the surrounding circles lie on a circle of radius 2 centered at the origin and form a regular hexagon with side length 2. To see that 6 is the maximum, consider the centers of the kissing circles, each at distance 2 from the central center. The minimum angular separation \alpha between any two such centers must satisfy the condition that their mutual distance is at least 2 (to avoid overlap). This distance is given by the chord length formula: $4 \sin(\alpha/2) \geq 2, so \sin(\alpha/2) \geq 1/2, implying \alpha/2 \geq 30^\circ and thus \alpha \geq 60^\circ. Therefore, at most $360^\circ / 60^\circ = 6 circles can fit around the center. Equality holds in the hexagonal arrangement, where the angular separation is exactly $60^\circ, achieving tangency without overlap. This result has been known since ancient times through observations of hexagonal tilings in geometry, with the optimality of the hexagonal packing formalized in the late 19th century by Axel Thue in his work on circle arrangements.

Three Dimensions

In three dimensions, the kissing number k(3) is exactly 12, meaning that a maximum of twelve equal non-overlapping spheres can simultaneously touch a central sphere of the same radius. This configuration is achieved in arrangements such as the face-centered cubic (FCC) lattice or the hexagonal close-packed (HCP) structure, where each sphere contacts twelve neighbors. In these setups, the centers of the surrounding spheres lie at the vertices of a regular icosahedron appropriately scaled to ensure contact with the central sphere while maintaining non-overlap. The determination of this value resolved a long-standing debate originating from the correspondence between Isaac Newton and David Gregory in the late 17th century, where Newton argued for 12 and Gregory for 13. The impossibility of 13 was rigorously proved in 1953 by and , who demonstrated through geometric analysis that any attempt to place a thirteenth sphere would result in overlap with at least one of the existing twelve. This result has practical implications in materials science and engineering, notably in the stacking of cannonballs, which follows close-packed arrangements to maximize stability and density, and in atomic packings within crystalline structures of metals like copper and aluminum, where the FCC lattice enables efficient space filling and mechanical properties.

Four Dimensions

In four-dimensional Euclidean space, the kissing number k(4) is exactly 24, meaning that 24 equal non-overlapping unit spheres can touch a central unit sphere, and no more can do so. This value establishes an important benchmark in higher-dimensional geometry, shifting focus from the combinatorial challenges of three dimensions to algebraic structures like lattices. The configuration achieves the maximum through the centers of the kissing spheres positioned at the 24 shortest non-zero vectors of the D_4 lattice, a checkerboard lattice known for its high symmetry and packing efficiency. The lower bound of 24 was recognized early via this lattice construction, which demonstrates the feasibility of the arrangement without overlaps. The upper bound, confirming that 24 is maximal, was rigorously proven in 2003 by Oleg Musin using a computer-assisted optimization within for spherical codes. This method analyzes the possible angular separations on the 3-sphere, ensuring no additional sphere can fit without intersection, and builds on association schemes to bound code sizes. Although earlier bounds existed, such as those derived from spherical code techniques in the 1970s, Musin's proof definitively resolved the problem for dimension 4. Geometrically, the D_4 lattice and its kissing configuration can be interpreted in quaternionic space, where the minimal vectors correspond to the 24 units (elements of norm 1) in the ring of , a maximal order in the quaternions consisting of integer and half-integer coefficients. This quaternionic viewpoint highlights the algebraic underpinnings of the arrangement, linking it to the vertices of the regular 24-cell polytope.

Exact Values in Exceptional Dimensions

Eight Dimensions

The kissing number in eight dimensions, denoted k(8), is exactly 240. This value is achieved by the E_8 lattice, an even unimodular lattice in \mathbb{R}^8 whose 240 root vectors of minimal norm squared 2 correspond to the centers of the surrounding spheres in the optimal kissing configuration. The upper bound of 240 was established in 1979 by Andrew M. Odlyzko and Neil J. A. Sloane using a linear programming technique based on Delsarte's method, which leverages association schemes derived from Coxeter groups and expansions in spherical harmonics to constrain the possible number of equiangular lines on the sphere. This bound matches the lower bound from the E_8 lattice construction, confirming the exact value; the method exploits the high symmetry of the E_8 root system, whose Weyl group (a finite Coxeter group) acts transitively on the roots. Beyond geometry, the E_8 lattice plays a significant role in theoretical physics and coding theory. In heterotic string theory, the E_8 \times E_8 lattice serves as the momentum lattice for the internal degrees of freedom, enabling the unification of gauge interactions with minimal supersymmetry breaking. In error-correcting codes, the E_8 lattice modulo 2 yields the extended binary Hamming code of length 8, providing a precursor structure for higher-dimensional codes like those underlying the Leech lattice in 24 dimensions.

Twenty-Four Dimensions

In twenty-four dimensions, the kissing number k(24) is exactly 196560. This value is achieved by the configuration of minimal vectors in the \Lambda_{24}, a highly symmetric even unimodular lattice discovered by in 1967, where these 196560 vectors correspond to the directions in which equal spheres can touch a central sphere without overlapping. The provides the lower bound for k(24) through this explicit construction, demonstrating that at least 196560 spheres can kiss a central one. The upper bound matching this lower bound was established in 1979 by Andrew M. Odlyzko and Neil J. A. Sloane, who applied a discrete version of Michel Delsarte's linear programming method—originally developed for —to prove that no more than 196560 unit spheres can touch a central unit sphere in 24 dimensions. Independently, Vladimir I. Levenshtein arrived at the same result using similar techniques based on association schemes and spherical codes. Their proofs rely on optimizing polynomials over the sphere to bound the maximum size of spherical codes with minimum angular separation of 60 degrees, confirming the exactness of the Leech lattice configuration for the . This kissing configuration holds profound significance beyond geometry, as the Leech lattice underlies the densest known sphere packing in 24 dimensions, with its optimality among lattice packings proven by Henry Cohn and Abhinav Kumar in 2003 using theta series of modular forms to establish tight linear programming bounds. Furthermore, the lattice's structure connects to number theory and group theory through monstrous moonshine, a phenomenon linking the Monster simple group—whose smallest nontrivial representation has dimension 196883—with the j-invariant of elliptic modular functions, where the coefficient 196560 appears as the number of minimal vectors, reflecting deep symmetries in the moonshine module constructed by Igor Frenkel, James Lepowsky, and Arne Meurman in 1985. The Leech lattice's kissing arrangement thus exemplifies exceptional mathematical harmony in high dimensions, extending principles from the E8 lattice in eight dimensions as a foundational component in certain constructions of \Lambda_{24}.

Bounds in Higher Dimensions

Lower Bounds

Lower bounds on the kissing number τ(n) in dimensions n ≥ 5 are established through explicit geometric constructions that demonstrate achievable configurations of mutually tangent spheres without overlap. The most basic such bound is τ(n) ≥ 2n, obtained by centering spheres at the 2n points ±e_i (i = 1 to n), where e_i are the standard basis vectors in ℝ^n; these points lie on the unit sphere and have angular separation of at least 90 degrees, preventing intersections. In low dimensions, stronger explicit constructions yield improved bounds using lattice-based arrangements. For n=5, τ(5) ≥ 40 via a spherical code derived from a finite geometry or lattice packing. For n=6, the E_6 lattice provides τ(6) ≥ 72 minimal vectors, corresponding to a tight configuration. Similarly, in n=7, the Barnes-Wall lattice achieves τ(7) ≥ 126. Further progress in higher dimensions relies on advanced methods such as lattice packings, random coding techniques, and explicit spherical designs, leading to exponential growth in the bounds. For instance, recent work by Gil Fernández et al. establishes τ(10) ≥ 510 and τ(11) ≥ 592 through optimized spherical codes in high dimensions, building on semidefinite programming to refine configurations. These approaches demonstrate τ(n) growing at least as Ω(2^{c n}) for some c > 0, with c ≈ 0.207 from earlier random constructions, though explicit designs often yield tighter values up to n=48 and beyond.

Upper Bounds

Early upper bounds on the kissing number \tau(n) for dimensions n \geq 5 were polynomial in n, such as those derived from spherical code considerations by in the mid-20th century, which provided limits growing like O(n^2). These bounds, while significant for their time, were substantially improved upon in subsequent decades as researchers sought tighter constraints. A major breakthrough came in 1978 with the work of George A. Kabatiansky and Vladimir I. Levenshtein, who established the first exponential upper bound: \tau(n) \leq 2^{0.401n(1+o(1))}. This result, obtained using bounds on spherical codes via techniques over function spaces, demonstrated that \tau(n) grows at most exponentially with base approximately 1.32, far below the naive volume-based bound of $2^n. Modern improvements leverage the framework introduced by Henry Cohn and in 2003 for packings, which has been adapted to yield precise upper bounds for kissing numbers in low dimensions. For instance, relaxations of this method establish \tau(5) \leq 44, tightening previous estimates and suggesting the exact value may be 40. These techniques have been extended computationally to higher dimensions, with recent semidefinite bounds giving \tau(17) \leq 10{,}978, \tau(18) \leq 16{,}406, \tau(19) \leq 24{,}417, \tau(20) \leq 36{,}195, and \tau(21) \leq 53{,}524. Asymptotically, the Kabatiansky-Levenshtein bound remains the strongest known, confirming that \tau(n) grows exponentially with base less than 2; this contrasts with lower bounds from constructions, which achieve bases around 1.32 to 1.40, leaving an exponential gap that persists in high dimensions.

Recent Advances

In early 2025, mathematicians Henry Cohn and Anqi Li announced significant improvements to the lower bounds on kissing numbers in dimensions 17 through 21, marking the first substantial progress in these dimensions since the 1960s. Their work, leveraging novel asymmetric lattice modifications, established that the kissing number τ(17) is at least 5730, τ(18) ≥ 7654, τ(19) ≥ 11692, τ(20) ≥ 19448, and τ(21) ≥ 29768. Building on this, Irene Gil Fernández, Jaehoon Kim, Hong Liu, and Oleg Pikhurko published new asymptotic lower bounds for kissing numbers in high dimensions in August 2025, improving upon previous estimates by a factor of approximately 3.442 through advanced spherical code constructions. Their result provides K(d) ≥ (1 + o(1)) (√3 / π) · (1 / (4 √2 log(3/2))) · d^{3/2} · (2 √3 / e)^d as d → ∞, offering tighter guarantees for large d and influencing related . In May 2025, Google DeepMind's AlphaEvolve, a Gemini-powered agent, contributed novel algorithmic constructions for spherical codes, yielding minor enhancements to lower bounds in mid-range dimensions such as 11, where it achieved τ(11) ≥ 593, though these were soon outpaced by subsequent human efforts. By October 2025, researcher Mikhail Ganzhinov demonstrated the continued edge of human optimization over AI models in a direct comparison, establishing new lower bounds via highly symmetric line arrangements: τ(10) ≥ 510, τ(11) ≥ 592, and τ(14) ≥ 1932. These results surpassed prior AI-assisted attempts in dimensions 10 and 14 while nearly matching in dimension 11, highlighting manual methods' superiority for bound tightening in dimensions 10–14. Cohn and Li's advancements similarly exceeded DeepMind's explorations in higher dimensions 17–21, underscoring hybrid human-AI approaches' potential. These developments have narrowed the gaps between lower and upper bounds in dimensions 5–23, with updated tables showing ratios closer to unity in several cases, raising prospects for exact resolutions in low dimensions like 5–7 through refined computational searches.

Generalizations and Extensions

To Non-Euclidean Spaces

The kissing number concept extends naturally to non-Euclidean spaces, where the geometry alters the possible arrangements of spheres. In \mathbb{H}^n of constant negative , the kissing number \kappa(n, r) for congruent spheres of radius r > 0 is finite for fixed r but grows ly with r, allowing significantly more spheres to touch a central one than in the counterpart as r increases. Specifically, \log \kappa(n, r) \sim (n-1)r, reflecting the exponential growth due to negative . This enables arbitrarily large configurations for sufficiently large r, and in limiting cases involving horospheres—surfaces asymptotic to spheres centered at —infinitely many equal spheres can be arranged to the horosphere, mirroring the infinite packing of circles to a line in . In spherical space \mathbb{S}^n of constant positive , the compact geometry imposes stricter limitations, yielding finite kissing numbers that are typically smaller than Euclidean values. For instance, in two-dimensional , at most 5 equal circles can touch a central equal circle, in contrast to the 6 possible in the , due to the angular excess from positive . Higher-dimensional spherical kissing numbers \kappa_S(n, r) have been bounded using and spherical codes, confirming their finiteness across dimensions. Generalizations to other metrics, such as those induced by \ell_p-norms, further diversify the kissing number \tau(n,p). In spaces equipped with the \ell_p-norm, the unit ball's shape influences arrangements; for p = \infty, the kissing number achieves $3^n - 1, exponentially larger than the case for high n, as centers can be placed at all non-zero points in \{-2, 0, 2\}^n while maintaining tangency to the and non-overlap. This highlights how choice amplifies possible tangencies compared to the round ball. Discrete analogs arise on lattices or graphs, where the integer kissing number counts minimal-distance neighbors excluding the origin. In the standard \mathbb{Z}^n, this is $2n, corresponding to the coordinate axes directions. Advanced constructions yield lattices with exponentially large kissing numbers, approaching the continuous bounds in high dimensions and informing applications. The kissing number problem is intimately connected to the broader problem, which aims to determine the arrangement of non-overlapping equal that maximizes in . The kissing number establishes a local bound on this by specifying the maximum number of that can adjoin a central without overlap, influencing the overall configuration efficiency. In three dimensions, the kissing number of 12 corresponds to the local structure in the densest known packings, such as the face-centered cubic , which achieves the optimal asserted by the and proved by Hales. This local optimality implies that in three dimensions, any where no is touched by more than 12 others cannot exceed the density of the close packings, as part of the analysis in the proof of the . A closely related problem is the Tammes problem, which seeks the maximum number of points that can be placed on the surface of a such that the angular separation between any two is at least 60 degrees, equivalent to packing non-overlapping spherical caps of angular radius 30 degrees. In three dimensions, the Tammes problem for 12 points achieves the kissing configuration, linking it directly to the kissing number of 12, while for 13 points, it remains unsolved and relates to potential denser local arrangements in packings. This duality highlights how the kissing number addresses the three-dimensional case of the Tammes problem but generalizes to higher dimensions as a spherical code optimization. In , kissing configurations translate to constant-weight binary where each codeword has weight equal to half the dimension and the minimum is 2, corresponding to unit spheres touching the origin without overlapping interiors. These provide lower bounds on kissing numbers through constructions like those derived from Reed-Muller codes or other combinatorial designs, offering insights into asymptotic behavior and explicit bounds in dimensions up to 24. Such connections enable the use of coding-theoretic tools, like the Delsarte method, to derive upper bounds on kissing numbers via spherical code constraints. The center density \phi of a sphere packing is defined as the number of spheres per unit volume when the spheres have unit radius, quantifying how the local kissing arrangement contributes to global efficiency. High kissing numbers facilitate denser packings by allowing more minimal vectors in lattices; for instance, the in eight dimensions has a kissing number of 240 and center density \phi = 2^{-4}, while the in 24 dimensions achieves a kissing number of 196,560 with center density \phi = 1. Cohn and Elkies proved these lattices yield the densest packings in their dimensions using bounds that incorporate the kissing configurations. This demonstrates how maximal local kissing enhances overall packing performance in exceptional dimensions.

Proof Techniques and Algorithms

Classical Proof Methods

Classical proof methods for the kissing number problem primarily involve analytical and combinatorial techniques developed in the mid-20th century to establish exact values and tight bounds in low and specific higher dimensions. These methods focus on geometric constraints, algebraic structures, and on spheres, avoiding computational . deficit arguments exploit the of the unit to limit the number of points with minimum separation, corresponding to non-overlapping kissing spheres. In three dimensions, each kissing sphere's center projects to a point on the unit separated by at least 60 degrees from others. The subtended by a of half this angle provides a basic upper bound of approximately 14, but refined arguments using spherical excess—via the Gauss-Bonnet theorem relating area to angular excess in spherical triangles—tighten this to 12. Schütte and van der Waerden applied this approach in their seminal proof, demonstrating that 13 points with minimum angular distance 60 degrees cannot exist without violating area constraints derived from icosahedral arrangements, where the 12 vertices achieve the configuration with slightly larger separations of about 63.43 degrees. Lattice and root system analysis provide exact constructions and optimality proofs in dimensions tied to exceptional Lie algebras, leveraging the symmetry of positive definite quadratic forms and Weyl groups. In eight dimensions, the E₈ lattice, derived from the E₈ root system, admits 240 minimal vectors of equal length, yielding a kissing configuration of 240 spheres; the upper bound matching this is established by showing that any denser arrangement would require a quadratic form with more short vectors than possible under the lattice's integrality and even unimodularity constraints, as analyzed by Nebe and Sloane using the theta series of even unimodular lattices. Similarly, in 24 dimensions, the , constructed via the miracle octad generator and related to the Mathieu group M₂₄, achieves 196,560 minimal vectors, with the proof of optimality relying on bounds from the frame potential and root system multiplicities, confirming no superior lattice exists. These methods highlight the exceptional density in dimensions 8 and 24. Spherical harmonics and Fourier analysis offer a powerful framework for upper bounds by expanding the indicator function of the spherical code in an orthogonal basis of zonal harmonics, ensuring non-negativity to derive contradictions for excessive points. Introduced in the context of spherical designs, this technique was pivotal in Schütte and van der Waerden's three-dimensional proof, where the integral of the second-degree harmonic over the configuration leads to a positivity condition violated by 13 points. Generalized by Levenshtein in the using —the Fourier basis on the sphere—this method yields universal bounds, such as τ₈ ≤ 240 and τ₂₄ ≤ 196,560, matching the lattice constructions and proving exactness through the maximality of the Delsarte linear program over positive definite kernels. The approach underscores the role of harmonic positivity in capping code sizes. Association schemes and designs facilitate combinatorial impossibility proofs by modeling the adjacency relations among code points as a strongly regular graph or higher-order scheme, enabling double counting of triples to expose inconsistencies. In low dimensions, such as three and four, these tools count the possible inner products between vectors, showing that assuming more than the known kissing number leads to non-integer parameters in the or negative eigenvalues in the association . For instance, the three-dimensional case reduces to analyzing the spherical distance graph with connection numbers derived from icosahedral symmetry, proving no 13-vertex exists with the required parameters. This method, rooted in Bose-Mesner algebra, complements geometric arguments and was instrumental in early bounds before computational advances.

Computational Algorithms

Computational algorithms play a crucial role in approximating kissing numbers, particularly in dimensions where analytical proofs are infeasible, by leveraging numerical methods to search for optimal sphere configurations or derive bounds. techniques, such as a modification of Delsarte's method, have been used to prove exact kissing numbers in low dimensions. For instance, Musin proved the kissing number in four dimensions is 24 in 2003 using such an approach with computational verification to rule out higher configurations, building on earlier analytical insights. Linear and approaches provide powerful tools for establishing upper bounds on kissing numbers by formulating the problem as an optimization over auxiliary functions that enforce non-overlapping conditions. The seminal Cohn-Elkies method, introduced in 2003, adapts techniques from to , optimizing over radial kernels to derive tight upper bounds; extensions using relax the linear constraints to matrices, yielding even sharper bounds in higher dimensions by capturing more sophisticated inequalities. These methods prioritize computational efficiency in handling high-dimensional optimizations while ensuring rigorous guarantees. Recent advances in and have introduced neural network-based strategies for searching kissing configurations, aiming to discover new lower bounds through automated exploration of vast configuration spaces. In 2025, DeepMind's AlphaEvolve, powered by the , generated and refined algorithms to probe kissing numbers, achieving improvements such as a lower bound of 593 in eleven dimensions by evolving code for configuration generation. However, these AI-driven attempts often require human intervention for final refinements and validation, as demonstrated when mathematician Mikhail Ganzhinov achieved lower bounds of at least 510 in ten dimensions, 592 in eleven (compared to AlphaEvolve's 593), and 1932 in fourteen dimensions, surpassing the AI in dimensions 10 and 14. Such hybrid approaches highlight the potential of in heuristic search while underscoring the need for expert oversight in geometric optimization. Stochastic optimization methods, including , offer practical means for constructing lower bounds in high dimensions by iteratively perturbing initial arrangements and accepting suboptimal moves with decreasing probability to escape local minima. These techniques simulate to explore the energy landscape defined by inter- distances, enabling the discovery of large non-overlapping sets tangent to a central . For example, in dimensions 25 through 31, produced kissing configurations of size 488 in twenty-five dimensions, improving prior probabilistic lower bounds and providing scalable constructions for dimensions where exhaustive methods fail. By focusing on maximization of the number of kissing , these algorithms complement deterministic approaches in tackling unsolved cases.

References

  1. [1]
    Kissing numbers - Henry Cohn - MIT
    The kissing problem asks how many spheres can be arranged tangent to a given sphere, if they all have the same size and their interiors cannot overlap.Missing: history | Show results with:history
  2. [2]
    [PDF] Lecture 1 part 1: Kissing numbers and spherical codes
    Sir Walter Raleigh's problem: To develop a formula that would allow to know how many cannonballs can be in a given stack simply by looking at the shape of ...
  3. [3]
    Improved kissing numbers in seventeen through twenty-one ... - arXiv
    Nov 7, 2024 · We prove that the kissing numbers in 17, 18, 19, 20, and 21 dimensions are at least 5730, 7654, 11692, 19448, and 29768, respectively.Missing: upper bounds 2025
  4. [4]
    [PDF] The kissing number in four dimensions - Annals of Mathematics
    The kissing number problem asks for the maximal number k(n) of equal size nonoverlapping spheres in n-dimensional space that can touch another.
  5. [5]
    [PDF] Kissing numbers - arXiv
    Jul 13, 2015 · The maximum possible number of non-overlapping unit spheres that can touch a unit sphere in n dimensions is called kissing number. The problem ...
  6. [6]
    G. A. Kabatiansky, V. I. Levenshtein, “On Bounds for Packings on a ...
    Abstract: A method is proposed for obtaining bounds for packings in metric spaces, the method being based on the use of zonal spherical functions associated ...
  7. [7]
    Isodual Codes over ℤ2k and Isodual Lattices | Journal of Algebraic ...
    These codes give rise to isodual lattices; in particular, we construct a 22-dimensional isodual lattice with minimum norm 3 and kissing number 2464. ... Bachoc, C ...
  8. [8]
    Kissing Number -- from Wolfram MathWorld
    The kissing number is the number of hyperspheres in n dimensions that can touch an equivalent hypersphere without intersections. In 3D, it's 12.
  9. [9]
    Sphere Packings, Lattices and Groups - SpringerLink
    Sphere Packings and Kissing Numbers. J. H. Conway, N. J. A. ... Reviews. Third Edition. J.H. Conway and N.J.A. Sloane. Sphere Packings, Lattices and Groups.
  10. [10]
    Counterintuitive Properties of High Dimensional Space
    Figure 14: A proof that the kissing number is six in two dimensions. If more than six blue circles can kiss the red, then one of the angles must be less than 60 ...
  11. [11]
    Hexagon | Definition, Shape, Area, Angles, & Sides - Britannica
    Sep 12, 2025 · Beginning about the 6th century bce, the Greeks gathered and extended this practical knowledge and from it generalized the abstract subject now ...
  12. [12]
    [PDF] Circle Packing: A Mathematical Tale - UTK Math
    This is a main theme running through our story, that circle packing provides a bridge between the combinatoric on the one hand and the geometric on the other.
  13. [13]
    Sphere Packing -- from Wolfram MathWorld
    In three dimensions, there are three periodic packings for identical spheres: cubic lattice, face-centered cubic lattice, and hexagonal lattice.
  14. [14]
    The Science of Sticky Spheres | American Scientist
    Newton's kissing-number problem suggests a solar-system model of sphere packing, with a dozen planets all feeling the attraction of a central sun. The ...
  15. [15]
  16. [16]
    [math/0309430] The kissing number in four dimensions - arXiv
    Sep 26, 2003 · In this paper we present a solution of a long-standing problem about the kissing number in four dimensions. Namely, the equality k(4)=24 is proved.Missing: Levenshtein 1971
  17. [17]
    [2301.08272] A note on five dimensional kissing arrangements - arXiv
    Jan 19, 2023 · In this note we report on how we discovered a new, previously unknown arrangement of 40 unit spheres in dimension 5.
  18. [18]
    New lower bounds on kissing numbers and spherical codes in high ...
    Nov 1, 2021 · New lower bounds on kissing numbers and spherical codes in high dimensions. Authors:Irene Gil Fernández, Jaehoon Kim, Hong Liu, Oleg Pikhurko.Missing: 2025 | Show results with:2025
  19. [19]
    [PDF] New lower bounds on kissing numbers and spherical codes in high ...
    Abstract. Let the kissing number K(d) be the maximum number of non-overlapping unit balls in Rd that can touch a given unit ball.Missing: x_j | Show results with:x_j
  20. [20]
    (PDF) Some new bounds on the kissing numbers - ResearchGate
    General upper bounds for r(n, s) were given by Rankin [33] and Levenshtein [27] (see also [29]). Better upper bounds were found in some cases (s(W) = 1 7, ...
  21. [21]
    [PDF] New upper bounds on sphere packings I - Annals of Mathematics
    This paper develops new upper bounds for sphere packing density using linear programming, proving the best bounds known for dimensions 4 through 36.
  22. [22]
    [PDF] High-Accuracy Semidefinite Programming Bounds for Kissing ...
    Feb 10, 2009 · The kissing number in n-dimensional Euclidean space is the maximal number of nonoverlapping unit spheres that simulta-.Missing: 8D 24D 1980s
  23. [23]
  24. [24]
    Mathematicians Discover New Way for Spheres to 'Kiss'
    Jan 15, 2025 · It took mathematicians until 1952 to prove Newton right: In our familiar three dimensions, the maximum “kissing number” (opens a new tab) is 12.Missing: historical | Show results with:historical<|control11|><|separator|>
  25. [25]
    New lower bounds on kissing numbers and spherical codes in high ...
    Jul 23, 2025 · New lower bounds on kissing numbers and spherical codes in high dimensions. Irene Gil Fernández , Jaehoon Kim , Hong Liu , Oleg Pikhurko ...
  26. [26]
    A Gemini-powered coding agent for designing advanced algorithms
    May 14, 2025 · AlphaEvolve: A Gemini-powered coding agent for designing advanced algorithms ... For example, it advanced the kissing number problem. This ...
  27. [27]
    Human mathematician beats AI in the "kissing numbers" challenge
    Oct 30, 2025 · Human beats AI at “kissing numbers”. In May 2025, DeepMind announced AlphaEvolve, an agent that writes and refines code to search tough spaces.Missing: 2020-2025 | Show results with:2020-2025
  28. [28]
    Human ingenuity outpaces AI in finding new 'kissing number' bounds
    Oct 22, 2025 · Aalto University doctoral candidate Mikhail Ganzhinov recently established three new lower bounds for the kissing number: at least 510 in ...
  29. [29]
    Human outsmarts Google DeepMind AI, solving centuries-old ...
    Oct 23, 2025 · Like in many other fields, the news seemed to indicate the future of kissing problem investigations belonged to AI.Missing: 2020-2025 | Show results with:2020-2025
  30. [30]
    Public defence in Information Theory, M.Sc. Mikhail Ganzhinov
    Sep 2, 2025 · When. 3.10.2025 12:00 – 15:00 (UTC +3). Where. Maarintie 8. Lecture hall TU1 ... lower bounds for kissing numbers in dimensions d=10,11 and 14.
  31. [31]
    [PDF] table of kissing number bounds - DSpace@MIT
    Aug 16, 2025 · The kissing problem asks how many spheres can be arranged tangent to a given sphere, if they all have the same.Missing: 4D | Show results with:4D
  32. [32]
    [2003.05547] Kissing number in non-Euclidean spaces of constant ...
    Mar 11, 2020 · This paper provides upper and lower bounds on the kissing number of congruent radius r > 0 spheres in hyperbolic \mathbb{H}^n and spherical \mathbb{S}^n spaces.
  33. [33]
    [1802.00886] Lattices with exponentially large kissing numbers - arXiv
    Feb 3, 2018 · We construct a sequence of lattices \{L_{n_i}\subset \mathbb R^{n_i}\} for n_i\longrightarrow\infty, with exponentially large kissing numbers.Missing: integer | Show results with:integer
  34. [34]
    [PDF] A proof of the Kepler conjecture - Annals of Mathematics
    Up to Euclidean motion, U(D) is one of two arrangements: the kissing arrangement of the 12 balls around a central ball in the face-centered cubic packing or the ...
  35. [35]
    [PDF] A SURVEY ON THE KISSING NUMBERS Peter Boyvalenkov, Stefan ...
    In particular, only six kissing numbers are known: τ1 = 2, τ2 = 6 (these two are trivial), τ3 = 12 (some incomplete proofs appeared in 19th century and Schütte ...
  36. [36]
    Improved Lower Bounds for Kissing Numbers in Dimensions 25 ...
    Aug 25, 2016 · We achieve |S| = 488 by applying simulated annealing. We also improve the aforementioned probabilistic argument in the general case. Finally, ...