Fact-checked by Grok 2 weeks ago

Knot theory

Knot theory is a branch of mathematical topology that studies embeddings of circles (known as knots) in three-dimensional Euclidean space, focusing on their properties under continuous deformations that do not allow the curve to pass through itself. These deformations, called ambient isotopies, preserve the knot's topology while allowing stretching, twisting, and shrinking, with the central problem being to determine whether two given knots are equivalent under such transformations. Formally, a knot is defined as a smooth embedding of the circle S^1 into \mathbb{R}^3, or equivalently, a simple closed curve that is homeomorphic to S^1. The origins of knot theory trace back to the late 19th century, when Scottish mathematician Peter Guthrie Tait began systematic tabulations of knots in connection with early atomic theories proposed by physicist , who hypothesized that atoms were knotted vortices in the ether. Early work by Tait, Thomas Kirkman, and others focused on enumerating distinct knots, but progress stalled until the 1920s when J.W. Alexander introduced the first invariant, the , which assigns a Laurent to each knot to help distinguish them. A major breakthrough occurred in 1984 with ' discovery of the Jones polynomial, a new invariant that revealed unexpected connections to and revitalized the field. Key concepts in knot theory include knot invariants, quantities such as polynomials or numbers that remain unchanged under , enabling classification of knots; classical examples are the Jones, , and HOMFLY polynomials, alongside simpler invariants like the crossing number and unknotting number. The theory also extends to , which are collections of disjoint knots, and braids, whose closures yield knots, providing tools for computation via Reidemeister moves—three local transformations that generate all equivalent diagrams of a knot. Modern developments incorporate quantum invariants derived from representations of quantum groups and categorification, which upgrades invariants to richer structures like ; a notable achievement was the 2020 proof by that the is not slice, resolving a 50-year-old question using advanced invariants. In 2025, mathematicians at the University of Nebraska-Lincoln disproved the conjecture that the unknotting number is additive for connected sums of knots. Beyond , knot theory finds applications in , where it models recombination and supercoiling processes, as enzymes like topoisomerases resolve knotted strands without breakage. In physics, knots appear in quantum field theories, , and , with invariants like the linking to partition functions in models of three-dimensional . These interdisciplinary ties underscore knot theory's role in understanding complex structures in nature and technology, from to entanglement.

Basic Concepts

In knot theory, a is formally defined as a smooth of the circle S^1 into three-dimensional \mathbb{R}^3. Two such embeddings are considered the same knot if they are related by an , a continuous family of embeddings that deforms one into the other while keeping the image a closed without self-intersections. This definition captures the intuitive notion of a knotted loop of string, where the embedding ensures the curve is simple and closed, and ambient isotopy models flexible deformations in space without cutting or passing through itself. A generalizes the concept of a to multiple components and is defined as a smooth of a of finitely many s into \mathbb{R}^3. Each individual in the embedding forms a component, and the entire structure is studied up to . For instance, a two-component consists of two disjoint closed curves that may be linked together or separate, distinguishing links from single knots while sharing the same topological framework. Knots are classified as tame or wild based on their geometric regularity. A tame knot is one that is ambient isotopic to a polygonal knot, meaning it can be approximated by a finite number of straight-line segments forming a closed . The , which is the standard unknotted circle, and the , the simplest non-trivial with three crossings in its minimal diagram, are both examples of tame knots. In contrast, a lacks this polygonal and exhibits or irregularity at some point, such as a that spirals infinitely towards a point, making wild knots far more pathological and less studied in classical knot theory. Basic properties of knots include , , and decomposition into prime or composite forms. Since knots are embeddings of the orientable manifold S^1, they are inherently , but they are often equipped with a chosen —a consistent direction along the curve, represented by arrows in diagrams—to facilitate the study of invariants and linking. distinguishes knots that are not equivalent to their mirror images under orientation-preserving ambient isotopies; knots exist in distinct left-handed and right-handed enantiomers, as exemplified by the , whose mirror image cannot be deformed into the original without reflection. A is prime if it cannot be decomposed as the connected sum of two non-trivial knots, whereas a composite arises from such a sum, allowing a high-level into simpler prime components, though the full operation is explored elsewhere. The , or trivial knot, is the of S^1 into \mathbb{R}^3 that is ambient isotopic to the standard round lying in a plane. Despite its simplicity, recognizing whether a given is the —known as the unknot recognition problem—is computationally challenging; while decidable via algorithms like those developed by Haken in the using normal surface theory, the problem lies in and , with ongoing research into its exact complexity.

Knot Diagrams

A knot projection is a continuous map from a knot, embedded in , onto a plane, where the image consists of arcs that intersect only at double points, corresponding to locations where two parts of the knot overlap in the projection. This projection captures the spatial arrangement of the knot but loses the three-dimensional depth information at these intersection points. A regular projection is one in which exactly two arcs meet transversely at each double point, with no three or more arcs concurrent and no tangencies between arcs. To resolve the ambiguity in a regular projection, a knot diagram augments it by specifying at each double point which arc passes over the other, typically indicated by unbroken lines for overcrossings and breaks or gaps for undercrossings. Thus, a knot diagram represents an immersed in the plane with crossing data that encodes the knot's . Crossings in a knot diagram are classified by as over or under based on the relative positioning in the original . Additionally, each crossing is assigned a : a positive crossing (+1) occurs when the overpassing , when rotated counterclockwise to align with the underpassing , follows the , while a negative crossing (-1) follows the left-hand rule. This signing , illustrated in standard where the positive crossing has the over strand sloping upward to the right relative to the under strand, provides a consistent way to distinguish crossing types across different projections. Knot diagrams can be further characterized as alternating or non-alternating depending on the sequence of over and under crossings encountered when traversing the along the knot's orientation. In an alternating , crossings alternate strictly between over and under as the curve is followed; for example, the (denoted 4_1) admits an alternating with four crossings, where the pattern over-under-over-under repeats seamlessly. Non-alternating diagrams, by contrast, have sequences where the same type of crossing (over or under) occurs consecutively, often arising in more complex knots like the knot 8_19 (the simplest non-alternating ), though many knots possess both types of diagrams. To obtain a minimal or reduced representation, knot diagrams are simplified by eliminating unnecessary complexities, such as avoiding triple points in projections or ensuring no isolated loops at crossings, while preserving the knot type through careful manipulation. Basic rules for diagram construction include selecting projections that minimize the number of double points and consistently applying assignments based on the embedding's geometry, though further simplification typically requires equivalence-preserving transformations like Reidemeister moves, discussed later.

Equivalence and Invariants

Reidemeister Moves and Ambient Isotopy

provides the fundamental topological notion of for knots in . Two embeddings f, g: S^1 \to \mathbb{R}^3 of into are ambient isotopic if there exists a continuous family of homeomorphisms h_t: \mathbb{R}^3 \to \mathbb{R}^3, for t \in [0,1], such that h_0 is the , h_1 \circ f = g, and each h_t is orientation-preserving, ensuring the deformation avoids self-intersections by preserving the property. This deformation allows one to be continuously transformed into another without passing through itself, capturing the intuitive idea of "untangling" while maintaining the topological type. To determine combinatorially through knot diagrams, Reidemeister moves offer a set of local transformations that suffice to relate any two diagrams of the same . These moves, introduced by Kurt Reidemeister, consist of three types and generate all planar deformations corresponding to three-dimensional isotopies. Type I move involves adding or removing a single twist or untwist in a strand, creating or eliminating a small where the strand crosses itself once; this corresponds to rotating a portion of the knot around its axis. Type II move adds or removes a pair of nugatory crossings by overlapping two parallel strands, effectively introducing or eliminating a "bigon" without changing the linking structure. Type III move slides one strand over an existing crossing formed by two other strands, preserving the over-under information while rearranging the diagram locally. Each move can be performed in both directions and is reversible under . Reidemeister's theorem establishes that two knot diagrams represent ambient isotopic knots if and only if one can be transformed into the other via a finite sequence of these three Reidemeister moves combined with regular isotopies of the plane (continuous deformations without creating or destroying crossings). The proof involves showing that any ambient isotopy induces a sequence of such diagram changes during projection, and conversely, each move arises from a specific three-dimensional deformation: Type I from twisting, Type II from pulling strands apart or together, and Type III from sliding under or over. This equivalence reduces the continuous problem of isotopy to a discrete combinatorial one, allowing systematic manipulation of diagrams. For , which consist of multiple intertwined knots, the extension beyond single knots requires additional tools. Markov's provides a for link equivalence using representations: two braids have closures that are ambient isotopic if and only if they are related by a sequence of isotopies and Markov moves, including stabilization (adding trivial strands), conjugation, and destabilization. This , proved by Markov, bridges equivalence to the Artin , enabling a algebraic description of link while preserving the combinatorial spirit of Reidemeister moves for multi-component cases. Although Reidemeister moves theoretically determine knot equivalence, the general problem of verifying whether two diagrams are related by such a sequence remains computationally challenging, with algorithms relying on normal surface theory that are decidable but impractical for knots with more than a modest number of crossings due to in complexity. Knot invariants, such as polynomials or homology groups, serve as efficient checks for non-equivalence but cannot prove equivalence alone.

Classical Invariants

Classical invariants in knot theory are fundamental tools developed in the early to distinguish knots that are not ambient isotopic, relying on combinatorial and algebraic properties of knot diagrams and their complements. These invariants, such as the knot group and the crossing number, provide initial methods to detect non-triviality and differences between knots without requiring advanced geometric structures. Later developments like the build on these by incorporating homological information, while quadratic refinements such as the Arf invariant and offer further discrimination, particularly for concordance classes. The knot group, also known as the of the complement, is a key algebraic invariant that captures the topological complexity of a knot K \subset S^3. For a knot K, the complement is the S^3 \setminus K, and its \pi_1(S^3 \setminus K) is generated by loops around the knot strands. A concrete , called the Wirtinger presentation, arises from a with n crossings, where each arc between under- and overpasses provides a generator x_i for i = 1, \dots, n, and relations at each crossing enforce the local topology: if arc j passes under arcs i and k, the relation is x_j = x_i^{-1} x_k x_i or a conjugate depending on the crossing type. This , introduced by Wirtinger in , allows computation of the group from any and is invariant under Reidemeister moves, though it may not be minimal. The crossing number \operatorname{cr}(K) of a knot K is the minimal number of crossings over all possible diagrams of K, serving as a basic measure of knot complexity. It is a non-negative integer invariant, with the unknot having \operatorname{cr}(U) = 0, the \operatorname{cr}(3_1) = 3, and the \operatorname{cr}(4_1) = 4. Computing \operatorname{cr}(K) exactly is challenging, but for alternating knots, reduced alternating diagrams often achieve the minimum. This invariant, first systematically studied by Tait in the late , bounds other properties like the and helps in tabulation. The Alexander polynomial \Delta_K(t) is a Laurent polynomial invariant derived from the knot group or a Seifert surface. One definition abelianizes the knot group to obtain the infinite cyclic cover of the complement, whose first homology module over \mathbb{Z}[t, t^{-1}] is presented by the , a minor of which yields \Delta_K(t) up to units. Equivalently, for an oriented Seifert surface F bounding K with basis \{ \ell_1, \dots, \ell_{2g} \} for H_1(F; \mathbb{Z}), the Seifert matrix A has entries a_{ij} = \operatorname{lk}(\ell_i, \ell_j^+), where \ell_j^+ is the positive push-off and \operatorname{lk} is the ; then \Delta_K(t) = \det(t^{1/2} A - t^{-1/2} A^T), normalized to be positive at t=1 and symmetric \Delta_K(t^{-1}) = \Delta_K(t). This polynomial, introduced by Alexander in , detects the as non-trivial since \Delta_U(t) = 1. The Seifert matrix formulation, due to Seifert in , connects it to the of bounding surfaces. The Arf invariant and knot signature provide quadratic enhancements to the , acting on the of the complement or Seifert surface. The Arf \operatorname{Arf}(K) \in \mathbb{Z}/2\mathbb{Z} is defined via the mod-2 reduction of the Seifert form, measuring the quadratic refinement of the intersection form on H_1(F; \mathbb{Z}/2\mathbb{Z}); it vanishes for the and equals for the right-handed . The \sigma(K) is the of the Hermitian form t^{1/2} A + t^{-1/2} A^T on H_1(F; \mathbb{C}) at t=-[1](/page/1), or more generally the Tristram-Levine function, providing an integer related to the Seifert matrix eigenvalues. These , with the Arf due to Murasugi's 1969 interpretation for and the formalized by Levine in the 1960s, distinguish amphichiral and bound concordance obstructions. For the trefoil knot $3_1, the Alexander polynomial is \Delta_{3_1}(t) = t^{-1} - 1 + t, computed from its Wirtinger group presentation \langle x, y \mid x y x = y x y \rangle abelianized or from a Seifert \begin{pmatrix} -1 \end{pmatrix}. The figure-eight knot $4_1 has \Delta_{4_1}(t) = 3 - t - t^{-1}, arising from its group \langle a, b \mid a b a b^{-1} = b a^{-1} b a^{-1} \rangle or Seifert \begin{pmatrix} -1 & -1 \\ 0 & 1 \end{pmatrix}; both examples illustrate how non-trivial polynomials confirm these knots are distinct from the .

Algebraic Invariants

Algebraic invariants in knot theory provide powerful tools for distinguishing knots beyond basic topological properties, often through or homological structures derived from knot presentations or diagrams. These invariants capture algebraic features of the knot complement or associated groups, enabling rigorous . Seminal developments include surfaces bounding the knot and polynomials satisfying recursive relations, with later advancements incorporating categorification for enhanced discriminatory power. Seifert surfaces are orientable surfaces in three-dimensional space whose boundary is a given knot, providing a key algebraic invariant through their topological properties. Introduced by Herbert Seifert in 1934, these surfaces can be constructed algorithmically from any knot diagram by resolving crossings into Seifert circles—disjoint circles obtained by smoothing overcrossings—and connecting them with twisted bands at the original crossing sites. The Euler characteristic \chi(S) of a Seifert surface S is computed as \chi(S) = V - E + F, where V, E, and F are the numbers of vertices, edges, and faces in a cell decomposition derived from the diagram, offering a computable invariant related to the knot's complexity. The genus g(S) of the surface, defined as g(S) = \frac{2 - \chi(S) - b}{2} with b the number of boundary components (typically 1 for knots), measures the minimal "handles" needed, and the knot genus is the infimum over all such surfaces. The , an early algebraic , arises from the of the knot complement and can be computed using Fox calculus on group presentations. Developed by J.W. Alexander in 1928 as the first non-trivial , it is the of a minor of the Alexander matrix, obtained from the knot group's Wirtinger presentation. Ralph formalized its computation in the 1950s via free , introducing Fox free \partial/\partial x_i on the free group ring \mathbb{Z}[F], where for a word w in generators x_j, the derivative satisfies \partial x_k / \partial x_i = \delta_{ik} and the Leibniz rule \partial(uv)/\partial x_i = \partial u/\partial x_i \cdot v + u \cdot \partial v/\partial x_i. Applying the abelianization map to the matrix of relations yields the Alexander matrix, whose of (n-1) \times (n-1) minors gives the \Delta(t), normalized to be symmetric \Delta(t^{-1}) = \Delta(t) and \Delta(1) = 1. This refines classical linking numbers and detects amphichirality in some cases. The Jones polynomial, discovered by in 1984, marks a breakthrough in algebraic knot invariants, defined axiomatically through skein relations on oriented link diagrams. It satisfies V(L_+) - V(L_-) = (t^{1/2} - t^{-1/2}) V(L_0), where L_+, L_-, and L_0 are links differing only at a crossing (positive, negative, and smoothed, respectively), with normalization V(\bigcirc) = 1 for the . Louis Kauffman provided a combinatorial reformulation in 1987 using the Kauffman bracket \langle D \rangle, an unoriented state-sum invariant where \langle D \rangle = A \langle D_0 \rangle + A^{-1} \langle D_\infty \rangle at each crossing (smoothing to 0- or \infty-type), multiplied by (-A^3)^{-w(D)} for writhe w(D) and adjusted for orientation to yield the Jones polynomial via V(t) = f(A) with A = -t^{-3/4}. For the right-handed $3_1, the Jones polynomial is t + t^3 - t^4, distinguishing it from the and . The HOMFLY-PT generalizes the Jones polynomial to a two-variable Laurent P(L; v, z), capturing both Alexander-Conway and Jones behaviors as specializations. Introduced collaboratively in by Freyd, Yetter, Hoste, Lickorish, Millett, and Ocneanu, it obeys the skein relation v^{-1} P(L_+) - v P(L_-) = z P(L_0), with P(\bigcirc) = 1. Setting v = t^{-1}, z = t^{1/2} - t^{-1/2} recovers the Jones , while v=1, z=\sqrt{m} yields the Alexander-Conway form, making it a unifying for classical polynomials. Khovanov homology provides a categorification of the , replacing the with a bigraded whose recovers V(t). Developed by Mikhail Khovanov in , it assigns to a link diagram a complex built from the $2^n-cube of resolutions, where n is the number of crossings, with vertices as smoothings (0- or 1-resolution at each crossing) and edges connecting differing by one smoothing. The chain groups are direct sums of graded vector spaces from Frobenius algebras (e.g., polynomials modulo relations for the q-graded structure), with differentials increasing the homological grading by 1 and quantum grading by 1 or 3, yielding homology groups Kh^{i,j}(L) such that \sum (-1)^i q^j \dim Kh^{i,j} = V(q). This homological invariant detects mutations and provides finer distinctions than the alone.

Geometric and Quantum Invariants

A significant class of knots in three-dimensional space are hyperbolic knots, which constitute the majority of all knots excluding torus knots and satellite knots. For a hyperbolic knot K, its complement S^3 \setminus K admits a complete hyperbolic metric of finite volume, making it a hyperbolic 3-manifold. This structure is determined up to isometry by an ideal triangulation of the complement, where the hyperbolic volume serves as a key invariant, computable via algorithms that solve for the shapes of ideal tetrahedra satisfying gluing conditions. The geometrization theorem, conjectured by William Thurston and proved by Grigory Perelman, classifies all 3-manifolds and has profound implications for knot complements. Specifically, for a hyperbolic knot complement, Dehn filling along most slopes on the boundary torus yields another hyperbolic manifold, with only finitely many exceptional slopes producing non-hyperbolic geometries such as Seifert fibered spaces. This finiteness result, established by Thurston, underscores the rigidity of hyperbolic structures under Dehn surgery, enabling the construction of diverse 3-manifolds from knot exteriors. The Casson invariant provides another geometric measure for 3-manifolds arising from , counting irreducible representations of the into the SU(2) up to conjugation. For integral spheres obtained via Dehn on a in the , the Casson invariant \lambda(M) relates directly to the 's description, offering insights into the manifold's through formulas involving linking numbers and signatures. This invariant detects distinctions among manifolds that share the same group, highlighting geometric properties beyond algebraic ones. Quantum invariants extend classical knot polynomials into the realm of , providing non-trivial distinctions for s and 3-manifolds. The Reshetikhin-Turaev invariants, constructed from modular tensor categories associated to quantum groups, generalize the to colored representations and extend it to links in arbitrary 3-manifolds via presentations. In particular, the colored arise by assigning irreducible representations (colors) to strands, yielding a family of Laurent in a q that detect hyperbolic volumes in the large-N limit for SU(2) representations. Geometric measures of knot complexity include the stick number and ropelength, which quantify minimal realizations in . The stick number s(K) of a knot K is the smallest number of straight-line segments required to form a polygonal of K, providing a discrete bound on complexity that grows with crossing number. Ropelength, defined as the minimal length of a knotted of unit radius without self-intersection, combines length and thickness to model physical ropes and yields upper bounds scaling linearly with stick number for many knots. In a 2025 development, Mark Brittenham and Susan Hermiller disproved the long-standing conjecture that the unknotting number—a geometric complexity measure counting minimal crossing changes to unknot—is additive under connected sums, providing counterexamples where u(K_1 \# K_2) < u(K_1) + u(K_2) for specific knots with u(K_i) = 3. This result, resolving Kirby's Problem 1.69(B), reveals non-additive behavior in knot complexity and prompts reevaluation of related invariants.

Higher Dimensions and Generalizations

Higher-Dimensional Knots

In higher-dimensional knot theory, an n-knot is defined as the image of a smooth embedding of the n-sphere S^n into the (n+2)-sphere S^{n+2}, up to ambient isotopy. This generalizes the classical case of 1-knots in S^3. For example, a 2-knot is a smoothly embedded 2-sphere in S^4, often referred to as a knotted surface, whose complement exhibits non-trivial topology despite the embedding being locally unknotted. Such embeddings are studied in the smooth category, where the exterior (complement of a tubular neighborhood) provides the primary object of interest. One seminal construction for obtaining higher-dimensional knots from classical ones is Artin's spinning operation, introduced in 1925. Given a classical knot K \subset S^3, spinning rotates K around an axis in S^4 (viewed as S^3 \times I with endpoints identified appropriately), yielding a 2-knot whose exterior fibers over the circle with fiber the classical knot complement. This method produces non-trivial examples, such as the spun trefoil, and extends to higher dimensions via generalizations like twist-spinning. In dimensions where the codimension exceeds two, embeddings of spheres are unknotted. Specifically, all smooth embeddings of S^1 into S^4 (1-knots in 4D) are ambient isotopic to the standard unknot, as the high codimension allows general position arguments to resolve any intersections without obstruction. Unknotting spheres, which are embedded hyperspheres bounding regions containing the knot, play a role in these proofs by enabling isotopies through handle decompositions. In contrast, codimension-two embeddings, like n-knots in S^{n+2} for n \geq 2, resist simple unknotting due to the failure of the Whitney trick in low dimensions, leading to profound classification challenges that rely on algebraic topology. Key invariants for higher-dimensional knots arise from the topology of their exteriors. Alexander duality implies that the homology of the knot exterior E(K) satisfies \tilde{H}_i(E(K)) \cong \tilde{H}^{n+1-i}(K) for i \geq 0, providing cohomological information about the embedding. Seifert hypersurfaces, which are connected, oriented (n+1)-manifolds F^{n+1} \subset S^{n+2} with \partial F^{n+1} = K, always exist by general position and duality arguments; they generalize Seifert surfaces and allow definition of signatures and other bilinear forms. The Blanchfield form, a sesquilinear pairing on the torsion submodule of the Alexander module H_1(\tilde{E}(K); \mathbb{Z}[t^{\pm 1}]), serves as a complete concordance invariant for odd-dimensional knots, linking the Seifert pairing to the infinite cyclic cover. These tools, rooted in surgery theory, highlight the algebraic complexity of codimension-two classifications, where even simple homotopy equivalence of exteriors does not imply isotopy.

Virtual and Welded Knots

Virtual knots generalize classical knots by allowing diagrams that incorporate both classical crossings—where strands intersect in the plane—and virtual crossings, represented as intersections without over/under information, such as a small circle around the crossing point. These diagrams model embeddings of circles into thickened surfaces, where virtual crossings correspond to intersections that occur outside the surface when stabilized in three-dimensional space. Classical knot diagrams form a subset of virtual knot diagrams, with no virtual crossings. Two virtual knot diagrams are equivalent if one can be transformed into the other via a sequence of virtual Reidemeister moves, which extend the classical (RI, RII, RIII) to include operations involving virtual crossings: mixed virtual-classical moves (MV1, MV2, MV3) that adjust strands passing through both types of crossings, and purely virtual moves (VR2, VR3) that handle virtual crossings alone. Equivalence under these moves preserves the knot type, but three specific "forbidden moves" (F1, F2, F3)—which involve detours around virtual crossings—are not permitted, as they would alter the topological type. Any virtual knot can be unknoted by combining virtual Reidemeister moves with these forbidden moves, highlighting their role in distinguishing virtual from classical structures. Welded knots extend virtual knots further by relaxing the equivalence relation to include one of the forbidden moves, specifically the overcrossing commute (OC) move, where an overstrand can pass freely under or over a virtual crossing without changing the knot type. This relaxation embeds classical knot theory into welded knot theory while allowing more equivalences, and welded knots are closed under this operation. Welded knots have applications to the study of braid groups, particularly through welded braids, which generalize classical braids by incorporating virtual crossings and the OC move, providing a framework for analyzing permutation representations and group extensions in low-dimensional topology. Key invariants for virtual knots include the virtual Jones polynomial, obtained by extending the to account for virtual crossings via a modified skein relation that treats virtual crossings as non-interacting. This polynomial distinguishes many virtual knots not separable by classical invariants. The arrow polynomial, a multivariable refinement, incorporates arrow assignments on classical crossings to track oriented structures, providing stronger discrimination; for example, it detects non-triviality in virtual knots with the same . Virtual knot theory connects to Khovanov homology through extensions that define chain complexes over arbitrary coefficients, incorporating virtual crossings via diagrammatic enhancements and preserving homological grading. This categorification yields torsion information beyond the , with applications to detecting virtual unknotting.

Knot Operations and Constructions

Connected Sums and Satellite Knots

The connected sum of two oriented knots K_1 and K_2 in the 3-sphere S^3 is constructed by selecting an unknotted arc in the diagram of each knot, removing an open tubular neighborhood of that arc from S^3, and gluing the resulting boundary tori together via an orientation-preserving homeomorphism that matches the meridians and longitudes appropriately. This operation yields a new knot K_1 \# K_2 that is independent of the choice of arcs up to ambient isotopy. Many classical knot invariants exhibit additivity under connected sum; for instance, the satisfies \Delta_{K_1 \# K_2}(t) = \Delta_{K_1}(t) \Delta_{K_2}(t), up to units in the Laurent polynomial ring. A knot is termed prime if it cannot be expressed as the nontrivial connected sum of two knots, meaning any such decomposition involves the as one factor. The prime decomposition theorem asserts that every nontrivial knot in S^3 admits a unique decomposition as a connected sum of finitely many prime knots, unique up to reordering and choice of orientation. This result, originally established by , provides a fundamental factorization analogous to that of integers into primes and underpins much of classical knot classification. Satellite knots generalize connected sums by incorporating more complex embeddings. Formally, a satellite knot arises from a companion knot J in S^3 and a pattern knot P embedded in the interior of a solid torus V, via a homeomorphism f: V \to N(J) (where N(J) is a solid toroidal neighborhood of J) such that f(P) is the resulting knot and f restricted to the boundary of V is not isotopic to the core embedding of N(J). Here, J serves as the companion, dictating the overall scale, while P provides the intricate winding pattern. Cable knots exemplify satellites, obtained when the pattern P is a (p,q)-torus knot in V with \gcd(p,q)=1, yielding a knot that winds p times meridionally and q times longitudinally around J. Torus knots are a distinguished family that reside on the surface of a standard unknotted embedded in S^3. The (p,q)-torus knot, where p and q are coprime positive integers, traces a closed curve that wraps p times around the torus's longitudinal direction and q times around its meridional direction. A parametric representation in \mathbb{R}^3 (identifying with S^3 via stereographic projection if needed) is given by \begin{align*} x &= (R + r \cos(q t)) \cos(p t), \\ y &= (R + r \cos(q t)) \sin(p t), \\ z &= r \sin(q t), \end{align*} for t \in [0, 2\pi] and R > r > 0 controlling the torus geometry. These operations influence knot symmetries, including invertibility and amphichirality. A knot is invertible if it is ambient isotopic to its (with reversed ), and connected sums preserve invertibility when both summands do, though non-invertible knots exist and their connected sums can yield further non-invertible examples. Amphichirality, the property of being isotopic to one's , can also arise under these constructions; for example, certain connected sums of satellites around amphichiral companions like the produce amphichiral knots with specified properties.

Special Knot Families

Alternating knots are defined by diagrams in which the over and under crossings alternate as one traverses the knot, allowing for a coloring of the complementary regions. This property simplifies the study of their invariants and equivalence under Reidemeister moves. A key result concerning alternating knots is the resolution of , which states that any two reduced alternating diagrams of the same knot or are connected by a finite of flypes—specific moves that rotate portions of the diagram around a crossing. This was proved by Menasco and Thistlethwaite, establishing that flypes suffice for equivalence among reduced alternating diagrams and providing an to recognize them. Fibered knots are those whose complements in the admit a over , with the being an open surface (a punctured Seifert surface). The of this is the of the induced by traversing the base once, which encodes the twisting along the and determines properties like the via the Seifert form. The complement's Seifert structure implies that the bounds a surface, and the must be right-veering for certain fibered knots, as detected by . Pretzel knots are constructed by taking an even number of vertical strands, twisting them in columns according to parameters p_1, p_2, \dots, p_n (with p_i yielding knots), and connecting the top and bottom ends in a twisted manner to form a closed . Most pretzel knots are , meaning their complements admit a complete of finite volume, with the exceptions being certain knots like the . Montesinos knots generalize this by replacing the twisted columns with rational tangles, arranged in a and closed to form the ; they are denoted by parameters specifying each tangle's fraction. Like pretzel knots, most Montesinos knots are , though some admit Seifert fibered surgeries, and their hyperbolicity follows from the arborescent nature of their tangle decompositions. A slice knot bounds an disk in the 4-ball, while a ribbon knot is a special case where the disk arises from a collection of ribbon moves—pushing parts of the knot through disjoint disks in the 4-ball without creating new intersections. Ribbon knots are always slice, but the converse (the slice-ribbon conjecture) remains open. The Fox-Milnor condition provides an algebraic obstruction: for a slice knot, the Alexander polynomial satisfies \Delta_K(t) = f(t) f(t^{-1}) for some polynomial f(t) with integer coefficients. In 2025, advances in computational knot theory have explored the complexity of ribbon knots through studies of untying operations and ribbonlength, a measure of the minimal length-to-width ratio for folded ribbon presentations. Researchers improved upper bounds on ribbonlength for small-crossing ribbon knots using new folding constructions, revealing that certain complex ribbon knots admit more efficient untying sequences than previously expected. These findings, derived from combinatorial searches identifying over 1.6 million ribbon knots, highlight computational challenges in distinguishing ribbon from non-ribbon slice knots via untying complexity.

Tabulation and Notation

Knot Tables and Catalogs

Knot tables provide systematic enumerations of distinct knot types, typically ordered by crossing number, the minimal number of crossings in any diagram of the knot. The Rolfsen table, published in 1976, catalogs all 165 prime knots with up to 10 crossings, excluding the , and serves as a foundational reference for low-complexity knots. This table accounts for the Perko pair, where two entries in an initial count of 166 were later recognized as equivalent under . Building on this, the Hoste-Thistlethwaite tables extend the to all 1,701,936 prime knots up to 16 crossings, completed in 1998 through collaborative computational efforts. These tables include data on hyperbolic volumes for the majority of entries, which are knots, computed using normal surface theory in their complements to verify distinctness. For example, the hyperbolic volume provides a geometric that helps distinguish knots beyond diagrammatic equivalence. Link tables complement knot catalogs by enumerating prime links, with the Thistlethwaite link table listing all such links up to 13 crossings, though publicly available portions often focus on up to 11 crossings. The Knot Atlas , an ongoing online resource launched in the early , hosts these tables and associated data, including updates as recent as 2025 that incorporate enumerations of with 20 crossings. Enumerations up to 20 crossings total 2,199,471,680 distinct prime knot types. These resources use notation systems to label entries for easy reference across studies. Computational challenges in building these tables arise from the exponential growth in the number of possible diagrams as crossing number increases, necessitating exhaustive generation and reduction techniques. Algorithms typically generate all reduced alternating diagrams first, then introduce non-alternating ones via twist additions, applying Reidemeister moves to normalize representations and filter equivalents. Invariants such as the Jones polynomial and hyperbolic volume are employed to prune redundant candidates efficiently, ensuring computational feasibility up to higher crossings. Tabulations maintain uniqueness by presenting each knot or link exactly once, up to , with equivalence confirmed through a of diagrammatic and computations that detect all duplicates within the enumerated set. This approach guarantees comprehensive coverage without repetition for the specified complexity bounds.

Notation Systems

Notation systems in knot theory encode the structure of knot diagrams into sequences or symbols, enabling unique identification, systematic , and efficient computational manipulation without relying on graphical representations. These notations are crucial for constructing comprehensive knot tables and implementing algorithms in software for invariant calculations and equivalence checks. By representing crossings, orientations, and connectivity in a standardized form, they support the tabulation of millions of knots, as demonstrated in large-scale enumerations. The Alexander-Briggs notation, developed by J. W. Alexander and G. B. Briggs in 1927, assigns labels to prime knots based on their minimal crossing number, denoted as the leading integer, followed by a subscript indicating the sequential order within that crossing class. For instance, the right-handed , with three crossings, is labeled $3_1, while the is $4_1. This system orders knots by increasing crossing number and, within each class, by a ordering derived from their projections, facilitating early tabulations up to nine crossings using invariants. Its simplicity makes it ideal for manual and initial computational catalogs, though it requires supplementary data for higher crossings. Dowker-Thistlethwaite (DT) codes, introduced by C. H. Dowker and M. B. Thistlethwaite in the , provide a permutation-based encoding of knot diagrams by indexing crossings during traversal. To construct a DT code, orient the and traverse it from an arbitrary starting point, assigning consecutive odd numbers (1, 3, 5, ..., 2n-1) to the undercrossings encountered and recording the even number paired with each odd index based on the second visit to that crossing (the overcrossing). The result is a of even integers, with signs indicating information for non-alternating knots; for example, the has DT code [4, 6, 2] in its standard alternating . For prime alternating diagrams, the DT code uniquely determines the diagram up to flype , making it highly suitable for computational and in databases like KnotScape. Gauss codes, originating from C. F. Gauss's early work on linking numbers in the and formalized in modern knot theory, record the sequence of crossings encountered during an oriented traversal of the . Each crossing is labeled with a unique integer from 1 to n, and the code lists these labels in order, prefixed by 'O' for overcrossing or 'U' for undercrossing, along with the direction (e.g., entering or leaving the crossing arc). A standard example for the is the sequence O1 U2 O3 U1 O2 U3, capturing the cyclic passage through all three crossings. This notation preserves the immersion structure and is foundational for deriving other codes like , though it may not distinguish mirror images without additional signs; its sequential nature aids in algorithmic reconstruction of diagrams for software implementations. Conway notation, devised by J. H. Conway in 1967, employs a descriptive, tangle-based system where knots are built from rational tangles connected via algebraic operations like summation and multiplication. Tangles are 4-ended arcs, denoted by continued fractions (e.g., for a single twist), and knots are formed by closing them; the , for example, is simply "3," representing a 3-twist tangle closed appropriately. This hierarchical approach allows compact representation of complex knots, such as 5_2 as [3 2], and excels in enumerating prime knots up to 11 crossings by systematically combining tangles, reducing redundancy in tabulation efforts. Its symbolic form supports recursive computations and tangle calculus for derivations. Braid words represent knots as closures of braids in the Artin braid group, where every knot is isotopic to the closure of some braid, per Alexander's theorem from 1923. A braid on m strands is encoded as a word in generators \sigma_i (positive crossings) and \sigma_i^{-1} (negative), with the knot formed by connecting top to bottom endpoints; for the trefoil, a minimal representation is the closure of \sigma_1^3 on two strands. This notation leverages the braid group's presentation for computational advantages, such as Markov moves to check equivalence, and is particularly useful in quantum invariant calculations via representations like the Jones polynomial. The minimal braid index (smallest m) provides a measure of knot complexity, aiding software tools for generating and distinguishing knots in large tables. These notations offer key advantages in compactness, enabling the storage and comparison of vast knot collections—such as the 1,701,936 knots up to 16 crossings tabulated using codes—and facilitating software implementations for automated and computation. By converting diagrams to linear sequences or symbols, they minimize data requirements while preserving essential topological information, essential for advancing knot theory databases and algorithms.

History and Applications

Historical Development

The origins of knot theory in the 19th century were deeply rooted in physics, particularly Lord Kelvin's (William Thomson's) vortex atom hypothesis proposed in 1867, which posited that atoms were distinct knotted configurations of swirling vortices in the to explain chemical periodicity and stability. This idea motivated the Scottish physicist Peter Guthrie Tait to systematically enumerate and classify knots, beginning his work in the late 1860s as a means to catalog potential atomic structures; Tait published his first tables of prime knots up to seven crossings in 1876 and extended them to ten crossings by 1898, emphasizing amphichiral knots and resolving ambiguities in projections. Tait's enumerations, while containing some errors later corrected, laid the groundwork for distinguishing knotted from unknotted curves and influenced early combinatorial approaches. Complementing Tait's efforts, the Reverend Thomas Penyngton Kirkman independently compiled extensive lists of projections in the , sending Tait a catalog of 10-crossing knots in that helped refine the tables, though it included redundancies due to projection equivalences not fully understood at the time. In the early , mathematicians built on Tait and Kirkman's tabulations to seek rudimentary knot invariants and properties distinguishing knot types, amid the field's initial combinatorial focus. The 1920s and 1930s marked a shift toward rigorous mathematical foundations, with introducing the in 1923 as the first non-trivial derived from the of the knot complement, providing a means to detect unknotting and linking. developed his three moves in 1926, establishing equivalence classes for knot diagrams under and enabling algebraic manipulations without physical models. formalized the modern definition of a knot in 1928 as a continuous of into three-dimensional , shifting emphasis from physical vortices to pure and resolving ambiguities in earlier enumerations. Following , Ralph Fox advanced algebraic tools in knot theory during the 1950s and 1960s, developing Fox calculus for free groups and the free to compute Alexander invariants more systematically, which facilitated the study of knot groups and torsion. revitalized the field in the 1960s and 1970s through his combinatorial notation system introduced in 1970 for describing knots via tangles and his normalization of the into the Conway polynomial in 1968, emphasizing recursive definitions and enabling efficient tabulations up to higher crossings. A pivotal breakthrough occurred in when discovered the , a new Laurent that unexpectedly detects and distinguishes knots beyond the Alexander polynomial's capabilities, revolutionizing the field by linking knot theory to and statistical physics.

Modern Advances and Applications

A pivotal modern development in knot theory arose from its intersection with , particularly Edward Witten's 1989 formulation interpreting the Jones polynomial through the Chern-Simons action in three-dimensional quantum Yang-Mills theory. This approach not only provided a physical origin for the Jones invariant but also spawned the field of , enabling the computation of knot polynomials via path integrals and inspiring further links between topology and quantum invariants. Building on these foundations, categorification emerged in the early as a higher-level refinement of classical knot polynomials, with Heegaard Floer homology serving as a cornerstone. Introduced by Peter Ozsváth and Zoltán Szabó, this categorifies the and assigns graded homologies to that detect properties like fiberedness and L-space surgeries. The absolute grading in Heegaard Floer homology further connects knot invariants to and four-manifold topology, influencing ongoing research in . Recent advances in 2024 and 2025 have extended knot theory into condensed matter and . In disordered three-dimensional metals, chemical-potential and magnetic-type disorders induce transitions between distinct knot types in nodal lines, revealing emergent topological phases under symmetry constraints. In , knotted solitons have been identified as meta-stable configurations in extensions of the that incorporate the QCD , providing potential explanations for candidates. has also yielded data-driven knot invariants, such as the multiscale Gauss link integral, which quantifies knot in biological and physical datasets by integrating linking numbers across scales. Interdisciplinary applications abound in and . In , knots form during enzymatic processes, and topoisomerases resolve them to avoid replication stalling, with knot theory modeling the efficiency of these enzymes . Protein structures occasionally exhibit deep knots that stabilize folds, impacting enzymatic activity and disease-related misfolding. Chemically, the synthesis of molecular knots and higher-order topologies has progressed, exemplified by Sauvage's and work, which contributed to the 2016 for . In theoretical physics, knot theory underpins via Chern-Simons terms and models QCD axions as knotted field configurations in solutions. Prominent open problems continue to drive research. The Volume Conjecture, proposed by Rinat Kashaev, hypothesizes that the growth rate of the Jones polynomial at roots of unity asymptotically matches the hyperbolic volume of the knot complement, with partial verifications for specific families. Efficient algorithmic recognition of the remains unresolved, despite decidability proofs, as current methods scale poorly for complex diagrams.

References

  1. [1]
    [PDF] Introduction to Knot Theory
    Introduction. Knot theory is the study of closed curves suspended in three dimensional space and how they can be deformed and categorized without passing ...
  2. [2]
    Introduction to Knots - Knot Theory - Faculty Sites
    The central problem of Knot Theory is determining whether two knots can be rearranged (without cutting) to be exactly alike. A special case of this problem is ...
  3. [3]
    [PDF] Lectures notes on knot theory
    Definition 3 (Knot). A knot is a one-dimensional subset of R3 that is homeomorphic to S1. We can specify a knot K by specifying an embedding (smooth ...
  4. [4]
    [PDF] The history of Knot Theory - UCLA Mathematics
    Alexander's polynomial was the first discovered polynomial invariant in Knot Theory, and it remained the only polynomial invariant until the Jones polynomial ...
  5. [5]
    [PDF] A BRIEF INTRODUCTION TO KNOT THEORY Contents 1 ...
    Aug 14, 2017 · (Knot Invariant) A knot invariant is something (for example a number, matrix, or polynomial) that is associated with a knot. A link invariant is.
  6. [6]
    [PDF] An Introduction to the Theory of Knots
    Dec 11, 2002 · The simplest definitions in knot theory are based on the latter approach. Definition 1.1 (knot) A knot is a simple closed polygonal curve in R3.
  7. [7]
    Knot Theory and Physics 1 Introduction
    This article is an introduction to relationships between knot theory and theo- retical physics. We give an exposition of the theory of polynomial invariants.
  8. [8]
    [PDF] Modeling DNA Using Knot Theory: An Introduction
    In particular, knot theory gives a very nice way to model DNA recombination. The relationship between mathematics and DNA began in the 1950's with the discovery ...
  9. [9]
    Knots in math, science, and everyday life | News
    Dec 9, 2013 · Mathematically, however, a knot is defined as a single, closed curve in space that can move by stretching, shrinking and twisting, and it cannot ...
  10. [10]
    [PDF] Beginning Course: Lecture 1
    May 15, 2012 · Definition 1. A knot is a smooth embedding K : S1 → R3. In general, it's more interesting to treat knots as “floppy” objects: we want a ...
  11. [11]
    [PDF] An Overview of Knot Invariants - UChicago Math
    Formally: Definition 2.1. A knot is an embedding of the circle S1 into R3. Here, instead of a rope, we have the segment [0,1]. It is allowed to wrap around ...
  12. [12]
    [PDF] Knots, Polynomials, and Categorification - Jacob Rasmussen
    Definition 1.1.​​ An oriented knot in R3 is a smooth embedding K : S1 ,→ R3. We say that two knots K0, K1 : S1 ,→ R3 are isotopic if there is a smooth map Φ : S1 ...
  13. [13]
    [PDF] An introduction to knot theory and the knot group - UChicago Math
    A knot is an embedding of a circle in R3, forming a closed loop without loose ends. It's like a string knot, but fixed in space.Missing: formal | Show results with:formal
  14. [14]
    [PDF] Topology of knots. - UCSB Mathematics Department
    DEFINITION 9. A link is an embedded disjoint union of circles in 53. A link is split if it's the distant union of two proper sublinks. DEFINITION 10. The ...
  15. [15]
    [PDF] Knot theory and wild knots - CSUSB ScholarWorks
    A knot which is equivalent to a polygonal knot is. 6. Page 16. considered tame. The unknot and trefoil knots are depicted as polygonal knots in Figure 1.6.
  16. [16]
    [PDF] Chapter 2. What is a Knot?
    Jan 30, 2021 · Wild Knots and Unknottings. 2. (this is referred to as a “wild knot”; a knot with a finite number of crossings is a. “tame knot”). This is not ...
  17. [17]
    Logic-Topology Seminar: Complexity in knot theory
    Oct 3, 2022 · In 1962, Haken proved that the knot recognition problem is solvable (his method, with some gaps, worked for arbitrary compact 3-manifolds).
  18. [18]
    Knot Diagram -- from Wolfram MathWorld
    Rolfsen (1976) gives a table of knot diagrams for knots up to 10 ... "Table of Knots and Links." Appendix C in Knots and Links. Wilmington, DE ...
  19. [19]
    [PDF] KNOTS KNOTES Contents 1. Motivation, basic definitions and ...
    Jan 27, 2015 · People often ask whether knot theory is related to the physics of string theory. The answer is “not in the sense you might imagine”. The ...
  20. [20]
    [PDF] Knots and Reidemeister theorem - MIT Mathematics
    no vertex of K is mapped onto a crossing. Definition (Knot Diagram). A diagram of a knot is its regular projection with the additional information of which.
  21. [21]
    Linking Number -- from Wolfram MathWorld
    A link invariant defined for a two-component oriented link as the sum of +1 crossings and -1 crossing over all crossings between the two links divided by 2.
  22. [22]
    Alternating Knot -- from Wolfram MathWorld
    An alternating knot is a knot which possesses a knot diagram in which crossings alternate between under- and overpasses.
  23. [23]
    Reduced Knot Diagram -- from Wolfram MathWorld
    A knot diagram in which none of the crossings are reducible. See also Knot Diagram, Reducible Crossing Explore with Wolfram|Alpha
  24. [24]
    [PDF] Lecture 8
    Nov 3, 2022 · Two knots, links or tangles are equivalent by ambient isotopy if the ambient space R3 or (for tangles) R2 × [0,1] can be continuously deformed ...<|control11|><|separator|>
  25. [25]
    [PDF] Knot Diagrammatics - arXiv
    Nov 19, 2004 · Reidemeister [131] discovered a simple set of moves on link diagrams that captures the concept of ambient isotopy of knots in three-dimensional ...
  26. [26]
    [1911.03745] Markov's Theorem - arXiv
    Nov 9, 2019 · This survey consists of a detailed proof of Markov's Theorem based on Joan Birman's book "Braids, Links, and Mapping Class Groups" and Carlo ...Missing: seminal | Show results with:seminal
  27. [27]
    [1604.03778] Elementary knot theory - arXiv
    Apr 13, 2016 · The aim of this survey article is to highlight several notoriously intractable problems about knots and links, as well as to provide a brief discussion of what ...Missing: URL | Show results with:URL
  28. [28]
    [PDF] Wirtinger Presentation
    The first method for computing knot groups was introduced by Wirtinger around 1904 in his lectures in Vienna, but not given wide circulation until its ...
  29. [29]
    Knot theory IV. Knot invariants - Cornell Mathematics
    Minimum number of crossing points. A regular diagram of a knot K has at most a finite number of crossing points. However, this number c(D) is NOT a knot ...
  30. [30]
    [PDF] Topological Invariants of Knots and Links - JW Alexander
    Mar 31, 2003 · Knots and their diagrams. In order to avoid certain troublesome com- plications of a point-theoretical order we shall always think of a knot as.
  31. [31]
    [PDF] Alexander polynomial of knots - UC Berkeley math
    The Alexander polynomial is the very first polynomial knot invariant discovered. In this expository paper, we will discuss what they are, how to compute them, ...
  32. [32]
    Theorem 2. <f>ik)=0iff Ai-l) = ±1 (mod 8).
    THE ARF INVARIANT FOR KNOT TYPES. KUNIO MURASUGI. The purpose of this paper is to prove Theorem 1 below which gives a simple relation between the Arf ...
  33. [33]
    [PDF] THE ALEXANDER POLYNOMIAL | Nancy Scherich
    The format of this paper is first to describe Alexander's original defi- nition of the Alexander polynomial, a purely algebraic definition. Then a detailed ...
  34. [34]
    [PDF] Über das Geschlecht von Knoten
    Über das Geschlecht von Knoten. Seifert, H. pp. 571 - 592. Terms and Conditions. The Göttingen State and University Library provides access to digitized ...Missing: Herbert 1934 paper
  35. [35]
    [PDF] Topological invariants of knots and links
    TOPOLOGICAL INVARIANTS OF KNOTS AND LINKS*. BY. J. W. ALEXANDER. 1. Introduction. The problem of finding sufficient invariants to determine completely the knot ...Missing: 1923 | Show results with:1923
  36. [36]
    [PDF] Free Differential Calculus. I: Derivation in the Free Group Ring
    Jul 27, 2005 · The free differential calculus grew up naturally out of an analysis that I began in the years 1944-45 of the basic idea of Alexander's knot ...Missing: seminal | Show results with:seminal
  37. [37]
    a new polynomial invariant of knots and links1 - Project Euclid
    The purpose of this note is to announce a new isotopy invariant of oriented links of tamely embedded circles in 3-space. We represent links by plane ...
  38. [38]
    [PDF] STATE MODELS AND THE JONES POLYNOMIAL
    KAUFFMAN: Formal Knot Theory. Lecture Notes No. 30, Princeton University Press, 1983. 7. L. H. KAUFFMAN: Srarisrical Mechanics and the Jones Polynomial. (to ...
  39. [39]
    [PDF] Hyperbolic Knot Theory
    Jun 8, 2019 · If a knot K ⊂ S3 is neither a satellite nor a torus knot, then it is hyperbolic. After Thurston's theorem there is definitively a strong reason ...
  40. [40]
    [PDF] William P. Thurston The Geometry and Topology of Three-Manifolds
    The intent is to describe the very strong connection between geometry and low- dimensional topology in a way which will be useful and accessible (with some ...
  41. [41]
    Bounds on exceptional Dehn filling - MSP
    Nov 14, 2000 · Thurston demonstrated that if one has a hyperbolic knot complement, all but finitely many Dehn fillings give hyperbolic manifolds [14].
  42. [42]
    [PDF] Universal bounds for hyperbolic Dehn surgery - Annals of Mathematics
    This paper gives a quantitative version of Thurston's hyperbolic Dehn surgery theorem. Applications include the first universal bounds on the num-.
  43. [43]
    Casson's invariant and surgery on knots
    The aim of this paper is two-fold. Firstly, we give generalisations of Casson's invariant for a knot X'(K), by giving a count of representations of the ...
  44. [44]
    [PDF] Borromean surgery formula for the Casson invariant
    We give an explicit formula for the Casson invariant of an integral homology sphere given by such a surgery presentation. The formula involves simple classical ...Missing: descriptions | Show results with:descriptions
  45. [45]
    [PDF] Ribbon Graphs and Their Invariants Derived from Quantum Groups
    The present paper is intended to generalize the Jones polynomial of links and the related Jones-Conway and Kauffman polynomials to the case of graphs in R 3 ...
  46. [46]
    [PDF] arXiv:1310.2735v1 [math.GT] 10 Oct 2013
    Oct 10, 2013 · The Witten-Reshetikhin-Turaev invariants extend the Jones polynomials of links in S3 to invariants of links in 3-manifolds. Similarly, in [5] ...
  47. [47]
    On stick number of knots and links - ScienceDirect.com
    The stick number, s(K), of a knot type K is the minimum number of edges required to realize the knot as a polygon. Stick numbers, sometimes, called edge number, ...
  48. [48]
    [PDF] On the Minimum Ropelength of Knots and Links - Jason Cantarella
    How much rope does it take to tie a knot? We measure the ropelength of a knot as the quotient of its length and its thickness, the radius of the largest.
  49. [49]
    [2506.24088] Unknotting number is not additive under connected sum
    We give the first examples of a pair of knots K_1,K_2 in the 3-sphere for which their unknotting numbers satisfy u(K_1\#K_2)<u(K_1)+u(K_2).Missing: disproof conjecture complexity
  50. [50]
    Knot a problem: Husker mathematicians disprove decades-old theory
    Jul 25, 2025 · The unknotting number is how many changes you need to make to turn a knot into a simple loop, called the unknot. The fewer changes you need, ...
  51. [51]
    Higher-Dimensional Knots According to Michel Kervaire - EMS Press
    A knot Kn ⊂ Sn+2 is the image of a differentiable embedding of an n-dimensional homotopy sphere in Sn+2. Its exterior E(K) is the complement of an open tubular.
  52. [52]
    2-Knots - MSP
    An n-knot is a locally flat embedding K : Sn → Sn+2 . (We shall also use the terms “classical knot” when n = 1, “higher dimensional knot” when n ≥ 2 and.
  53. [53]
    [math/0410606] Knot spinning - arXiv
    Oct 28, 2004 · Knot spinning is a method for constructing higher-dimensional knots, introduced by E. Artin in 1926, with generalizations like twist and frame  ...
  54. [54]
    [PDF] Knot spinning
    Oct 28, 2004 · This exposition is intended to provide some introduction to higher-dimensional knots - embeddings of Sn−2 in Sn - through spinning ...
  55. [55]
    [PDF] Knots in Four Dimensions and the Fundamental Group
    By a classical knot we mean an embedding of S1 into S3. The triv- ial example of such an embedding is the unknot, which is shown in. Figure 1.Missing: formal | Show results with:formal<|control11|><|separator|>
  56. [56]
    [PDF] UNKNOTTING SPHERES IN CODIMENSION TWO
    We conclude by proving the promised unknotting theorem. THEOREM (3). Let M be a homotopy n-sphere imbedded in SnC2 such that Sn+' - M is homotopy equivalent to ...
  57. [57]
    B Existence of Seifert hypersurfaces - EMS Press
    A Seifert hypersurface of Ln is a differentiable, oriented and con- nected, (n + 1)-dimensional submanifold Fn+1 in Mn+2 that has Ln as boundary. The aim of ...
  58. [58]
    Blanchfield and Seifert algebra in high dimensional knot theory - arXiv
    Dec 13, 2002 · In this paper the analogy is applied to explain the relationship between the Seifert forms over a ring with involution and Blanchfield forms over the Laurent ...Missing: hypersurfaces | Show results with:hypersurfaces
  59. [59]
    [PDF] High-dimensional knot theory
    Page 1. Andrew Ranicki. High-dimensional knot theory. Algebraic surgery in codimension 2. Springer-Verlag. Berlin Heidelberg NewYork. London Paris Tokyo.<|control11|><|separator|>
  60. [60]
    [math/9811028] Virtual Knot Theory - arXiv
    Nov 5, 1998 · Virtual knot theory studies non-planar Gauss codes via knot diagrams with virtual crossings. This paper gives basic results and examples.
  61. [61]
    [0810.3858] Virtual Crossing Number and the Arrow Polynomial - arXiv
    Oct 21, 2008 · We introduce a new polynomial invariant of virtual knots and links and use this invariant to compute a lower bound on the virtual crossing number.
  62. [62]
  63. [63]
    [2503.15103] Data Driven Perspectives on Knot Theory - arXiv
    Mar 19, 2025 · This paper uses topological data analysis to extend knot theory, providing a way to understand knot invariants statistically and gain new ...Missing: virtual | Show results with:virtual
  64. [64]
    Knot Sum -- from Wolfram MathWorld
    Two oriented knots (or links) can be summed by placing them side by side and joining them by straight bars so that orientation is preserved in the sum.
  65. [65]
    Knots and Links
    Knots and links have an operation known as connected sum where two knots are joined into a single knot by cutting ... knot theory is how to tell when two knot di-.
  66. [66]
    [PDF] knot theory
    The Alexander polynomial of a connected sum of knots is the product of their individual polynomials (see. Chapter 6). Hence, the degree of the Alexander ...
  67. [67]
    Prime Knot -- from Wolfram MathWorld
    A knot is called prime if, for any decomposition as a connected sum, one of the factors is unknotted (Livingston 1993, pp. 5 and 78).
  68. [68]
    [PDF] Section 4.5. Connected Sums of Knots and Prime Decompositions
    Feb 16, 2021 · The Prime Decomposition Theorem was first proved by Horst Schubert. (June 11, 1919–2001). See: Horst Schubert, “Die eindeutige Zerlegbarkeit ...Missing: citation | Show results with:citation
  69. [69]
    [PDF] arXiv:1210.0885v3 [math.GT] 19 Mar 2013
    Mar 19, 2013 · In this paper we will primarily be concerned with satellite knots, which are defined in the following way. Definition 1.4. A knot K is a ...
  70. [70]
    [PDF] Satellite knots - UC Davis Math
    Mar 22, 2018 · More specifically, a satellite knot with pattern the (p, q)-torus knot and companion J is called a (p, q)-cable of J. See Figure 2.
  71. [71]
    Torus Knot -- from Wolfram MathWorld
    A (p,q) -torus knot is obtained by looping a string through the hole of a torus p times with q revolutions before joining its ends.Missing: parametrization | Show results with:parametrization
  72. [72]
    Torus Knot - Visions in Math
    Aug 15, 2024 · Recall that torus knots are knots (and links) that can be moved so that they are embedded on a torus (or doughnut shape surface). Kyle ...
  73. [73]
    [PDF] NON-INVERTIBLE KNOTS EXIST
    Two knots which differ only in their orientation are said to be inrerses of each other, and a knot is iwertible if it is equivalent to its inverse. The theorem.
  74. [74]
    [PDF] Amphichiral knots with large 4-genus
    Our examples are connected sums of certain satellites of the figure-eight knot. Example 1. Let J be a reversible knot and define K(J) to be as in Figure 1, ...
  75. [75]
    [PDF] arXiv:1901.00582v1 [math.GT] 3 Jan 2019
    Jan 3, 2019 · An alternating knot has a diagram where crossings alternate over-under-over-under, and can be colored in a checkerboard pattern with black and ...
  76. [76]
    [PDF] arXiv:math/0503270v2 [math.GN] 21 Nov 2007
    Nov 21, 2007 · b) (Tait's flyping Conjecture) Two reduced alternating diagrams of the same link, are related by a finite series of flypes. The first Tait's ...
  77. [77]
    SYMMETRIES OF FIBERED KNOTS - Project Euclid
    Specifically, we make use of the monodromy operator h , the intersection form Q, and the Seifert form B , all of which are defined on the homology of the fiber.
  78. [78]
    [PDF] arXiv:math/0012239v4 [math.GT] 24 Aug 2001
    Hence the monodromy of the fibration of the complement of a torus knot in S3 is a product of positive Dehn twists. These twists are nonseparating by our ...
  79. [79]
    [PDF] arXiv:2204.04093v2 [math.GT] 12 Sep 2024
    Sep 12, 2024 · Abstract. We prove that the knot Floer complex of a fibered knot detects whether the monodromy of its fibration is right-veering.
  80. [80]
    [PDF] arXiv:1904.12955v2 [math.GT] 7 Jun 2019
    Jun 7, 2019 · ... pretzel knots are odd pretzels whose parameters can be notated as a permutation of those of Ppa,´a, a,´a,...,aq (i.e. a pretzel with exactly ...
  81. [81]
    Some invariants of Pretzel links
    We show that nontrivial classical pretzel knots L(p,q,r) are hyperbolic with eight exceptions which are torus knots. We find Conway polynomials of n-pretzel ...
  82. [82]
    [PDF] arXiv:math/0502303v1 [math.GT] 15 Feb 2005
    Feb 15, 2005 · Example 1. Let K ∪ t1 be the two component link of Figure 1. Here K is the Montesinos knot given by the triple of rational tangles (1/3, −1/3, ...Missing: definition | Show results with:definition
  83. [83]
    [PDF] arXiv:1210.7547v2 [math.GT] 23 Jul 2013
    Jul 23, 2013 · According to Wu [Wu10, Wu11c], the only hyperbolic Montesinos knots of length three that are pret- zel knots and might admit Seifert fibered ...
  84. [84]
    [PDF] Alexander polynomials of equivariant slice and ribbon knots in S^ 3
    Definition 1.1. A knot in S3 is said to be slice if it bounds a slice disk,. i.e., a smooth 2-disk properly embedded in the 4-ball. It is called ribbon.
  85. [85]
    [PDF] SLICE-RIBBON CONJECTURE - McMaster University
    The definition of a slice knot first appeared in a paper by Fox and Milnor [6]. Slice knots are also intimately related with the failure of the Whitney.
  86. [86]
    [PDF] arXiv:2101.00865v1 [math.GT] 4 Jan 2021
    Jan 4, 2021 · Fox and Milnor observed in [FM66] that the Alexander polynomial of a slice knot can be factored in a special form. Theorem 2.1 ([FM66]). The ...
  87. [87]
    Ribbonlength upper bounds for small crossing knots and links - arXiv
    We study the folded ribbonlength of these folded ribbon knots, which is defined as the knot's length-to-width ratio. The {\em ribbonlength ...Missing: untying complexity
  88. [88]
    [PDF] Ribbon concordances and slice obstructions - Nathan Dunfield
    1.6M ribbon knots found by combinatorial band search and. “shaking the diagram” via exterior-to-link. 2.2M knots found to be not (top) slice via Herald-Kirk ...<|control11|><|separator|>
  89. [89]
    The Rolfsen Knot Table - Knot Atlas
    May 27, 2009 · In our table we removed Rolfsen's 10162 and renumbered the subsequent knots, so that our 10 crossings total is 165, one less than Rolfsen's 166.
  90. [90]
    [PDF] The First 1701936 Knots - Pitzer College
    Amphicheiral knots are important to chemists, who are often concerned with the right- or left-handedness of mole- cules. The figure-eight is amphicheiral, a ...
  91. [91]
    The Thistlethwaite Link Table - Knot Atlas
    Apr 13, 2009 · The Thistlethwaite Link Table is a list of all prime links with up to 13 crossings, though we present here only the links with up to 11 crossings.
  92. [92]
    Knot Atlas
    Jun 30, 2025 · The Knot Atlas is a user-editable, wiki-like site for knots, with tables of knots and links, and other resources like torus knots.The Rolfsen Knot Table · The Thistlethwaite Link Table · 36 Torus Knots · To DoMissing: 20 | Show results with:20
  93. [93]
    The enumeration and classification of prime 20-crossing knots - MSP
    Mar 5, 2025 · A fuller account of the history, up to the classification of 16-crossing knots, is given by Hoste, Thistlewaite and Weeks [14], and, to complete ...
  94. [94]
    Naming and Enumeration - Knot Atlas
    Feb 21, 2013 · KnotTheory` comes loaded with some knot tables; currently, the Rolfsen table of prime knots with up to 10 crossings [Rolfsen], the Hoste-Thistlethwaite tables.
  95. [95]
    DT (Dowker-Thistlethwaite) Codes - Knot Atlas
    Feb 21, 2013 · DTCode[i1, i2, ...] represents a knot via its DT (Dowker-Thistlethwaite) code, while DTCode[{i11,...}, {i21...}, ...] likewise represents a link. ...Missing: paper | Show results with:paper
  96. [96]
    Chapter 7. Knots and Links
    A link is stored using a link diagram, which consists of n crossings and 2 n strands that connect them, possibly with some additional unknot components that ...
  97. [97]
    Conway's Knot Notation -- from Wolfram MathWorld
    A concise notation based on the concept of the tangle used by Conway (1967) to enumerate prime knots up to 11 crossings.Missing: original paper
  98. [98]
    P G Tait (1831 - 1901) - Biography - MacTutor History of Mathematics
    The idea led Tait, Thomson and Maxwell to begin to work on knot theory since the basic building blocks, in Thomson's vortex atom theory, would be the rings ...
  99. [99]
    [PDF] Knots In Blue - Colby College
    collaborator of Lord Kelvin's, was much taken with the theory of vortex atoms and started a systematic listing, or “enumeration”, of knots. His goal was to ...
  100. [100]
    [PDF] arXiv:math/0312168v3 [math.HO] 15 Dec 2003
    This theory inspired Kelvin to enlist the aid of mathematicians Tait, Kirkman and Little to construct the first tables of knots. The vortex theory eventually ...Missing: Littlewood | Show results with:Littlewood
  101. [101]
    (PDF) History of Knot Theory - ResearchGate
    PDF | We present in this chapter (Chapter II) the history of ideas which lead up to the development of modern knot theory. We are more detailed when.
  102. [102]
    Kurt Reidemeister (1893 - 1971) - Biography - MacTutor
    Kurt Reidemeister was a pioneer of knot theory and his work had a great influence on group theory. ... It was Wirtinger who interested Reidemeister in knot theory ...
  103. [103]
    [PDF] History of Knot Theory | Semantic Scholar
    Mar 3, 2007 · We present in this chapter (Chapter II) the history of ideas which lead up to the development of modern knot theory.
  104. [104]
    Ralph Fox (1913 - 1973) - Biography - MacTutor
    Ralph Fox was an American mathematician who made important contributions to differential topology and knot theory. Thumbnail of Ralph Fox View one larger ...
  105. [105]
    Quantum field theory and the Jones polynomial
    Sep 27, 1988 · In this version, the Jones polynomial can be generalized fromS 3 to arbitrary three manifolds, giving invariants of three manifolds that are ...
  106. [106]
    Absolutely graded Floer homologies and intersection forms for four ...
    In Ozsváth and Szabó (Holomorphic triangles and invariants for smooth four-manifolds, math. SG/0110169, 2001), we introduced absolute gradings on the ...
  107. [107]
    Knot data analysis using multiscale Gauss link integral - PNAS
    We introduce a multiscale knot theory paradigm that extends its scope from qualitative to quantitative analysis, providing a cutting-edge computational biology ...
  108. [108]
    Molecular knots in biology and chemistry - IOPscience
    Aug 20, 2015 · In this review, we provide an overview on the various molecular knots found in naturally occurring biological systems (DNA, RNA and proteins), ...
  109. [109]
    A new algorithm for recognizing the unknot - MSP
    Jan 4, 1999 · The approach is to consider the knot as a closed braid, and to use the fact that a knot is un- knotted if and only if it is the boundary of a ...