Fact-checked by Grok 2 weeks ago

Alexander polynomial

The Alexander polynomial is a knot invariant in algebraic topology that assigns to each oriented knot or link in three-dimensional Euclidean space a Laurent polynomial with integer coefficients, unique up to multiplication by units in the ring \mathbb{Z}[t, t^{-1}] (specifically, \pm t^k for integer k). It was introduced by American mathematician J. W. Alexander II in his seminal 1928 paper, where it emerged from the study of the homology of the infinite cyclic cover of the knot complement as a module over \mathbb{Z}[t^{\pm 1}]. As the first polynomial invariant for knots, it provided a major advance in distinguishing knot types under ambient isotopy, remaining the primary such tool until the discovery of the Jones polynomial in 1984. The polynomial can be computed algebraically from the knot group's Wirtinger presentation via Fox free calculus, yielding the Alexander matrix whose (n-1)×(n-1) minor determinants generate the first elementary ideal, with the polynomial as its gcd, up to units. Geometrically, it arises from a Seifert surface spanning the knot, where a basis for the first homology yields a Seifert matrix V whose symmetrized form t^{1/2}V - t^{-1/2}V^T has determinant equal to the normalized Alexander polynomial \Delta_K(t). An equivalent recursive definition via the Conway skein relation, \Delta_{L_+} - \Delta_{L_-} = (t^{1/2} - t^{-1/2})\Delta_{L_0}, was later developed by John Horton Conway in 1968, facilitating computations from knot diagrams. Key properties include symmetry \Delta_K(t) = \Delta_K(t^{-1}) (up to units), evaluation \Delta_K(1) = 1, and the fact that its degree bounds the Seifert genus of the from below. However, it is not a complete , as it fails to distinguish a knot from its mirror image and many distinct knots share the same , for example, the Kinoshita–Terasaka knot and the both have the trivial Alexander . Despite these limitations, the Alexander remains foundational in , influencing developments in quantum invariants and .

Fundamentals

Definition

The Alexander polynomial is a fundamental invariant in , assigning to each oriented K embedded in the three-sphere S^3 a Laurent \Delta_K(t) \in \mathbb{Z}[t, t^{-1}] with coefficients, defined up to multiplication by units \pm t^k for some k \in \mathbb{Z}. This captures topological properties of the and serves as a distinguishing feature among different types. To define it formally, consider the knot complement X = S^3 \setminus K, an open 3-manifold whose fundamental group \pi_1(X) abelianizes to \mathbb{Z}, with the generator corresponding to the homotopy class of a meridian curve around K. The infinite cyclic cover \tilde{X} \to X is the unique connected covering space with deck transformation group isomorphic to \mathbb{Z}, induced by the kernel of the abelianization map \pi_1(X) \to \mathbb{Z}. The Alexander module is then the first homology group H_1(\tilde{X}; \mathbb{Z}) viewed as a module over the Laurent polynomial ring \Lambda = \mathbb{Z}[t, t^{-1}], where the action of t arises from the generator of the deck group. The Alexander polynomial \Delta_K(t) is derived from this module as a generator of its first Fitting ideal (up to units in \Lambda), which encodes the torsion structure of the module over the principal ideal domain \Lambda. For knots, the module is torsion, ensuring the polynomial is well-defined and non-trivial in general. Although initially formulated for knots, the construction extends naturally to links with \mu components by replacing the infinite cyclic cover with the corresponding \mathbb{Z}^\mu-cover of the link complement, yielding a multi-variable Alexander polynomial \Delta_L(t_1, \dots, t_\mu) \in \mathbb{Z}[t_1^{\pm 1}, \dots, t_\mu^{\pm 1}], again defined up to units.

Historical development

The Alexander polynomial emerged as a pivotal invariant in through the work of , who introduced it in his 1928 paper "Topological Invariants of Knots and Links," building on presentations given as early as 1927. This polynomial, derived from the of the infinite cyclic cover of the knot complement, marked the first algebraic tool capable of distinguishing many non-trivial s from the and from each other. The development occurred amid the rapid advancement of in the early 20th century, where Henri Poincaré's 1895 introduction of the in "Analysis Situs" provided essential machinery for analyzing the of spaces like knot complements. Kurt Reidemeister's 1926 formulation of equivalence moves for knot diagrams in "Knoten und Verkettungen" further enabled rigorous combinatorial studies of knot types, setting the stage for invariants like Alexander's. In , Alexander collaborated with G. B. Briggs on "On Types of Knotted Curves," establishing a systematic notation for tabulating prime knots up to eight crossings, which facilitated early computations and classifications using the . Among the key milestones, computed the polynomial for the (yielding t^{-1} - 1 + t) and the (yielding -t^{-1} + 3 - t) in his 1928 paper, illustrating its utility in distinguishing these knots. In the 1950s, Ralph Fox advanced the theory by developing the free differential calculus, a method for deriving Alexander invariants directly from group presentations via operations in the ring, as detailed in his series of papers beginning with "Free Differential Calculus. I" in 1954. This refinement simplified computations and extended the polynomial's applicability to links and groups. The Alexander polynomial's influence persisted into the 1980s, when Vaughan Jones introduced the Jones polynomial in 1984, revealing connections between classical invariants like the Alexander polynomial and quantum topology, thus inspiring further developments.

Computation

Presentation matrix method

The presentation matrix method computes the Alexander polynomial of a knot K \subset S^3 using a finite presentation of the knot group \pi_1(S^3 \setminus K). This approach begins with the Wirtinger presentation, derived from a diagram of the knot. In a knot diagram with n arcs, assign a generator x_i to each arc. At each crossing, the relation arises from the path along the under-arc being equal to the over-arc conjugating the incoming under-arc segment; specifically, if arcs a, b, and c meet at a crossing with b over and a incoming under to c outgoing under (oriented consistently), the relator is c = b a b^{-1}, or rearranged as b a b^{-1} c^{-1} = 1. This yields a presentation with n generators and n or more relators, though one relator is often redundant due to the group's structure. To obtain the Alexander matrix, apply Fox free differential calculus to the relators in the ring. The Fox derivative \partial / \partial x_j satisfies rules analogous to the : \partial / \partial x_j (1) = 0, \partial / \partial x_j (x_i) = \delta_{ij}, \partial / \partial x_j (w z) = \partial / \partial x_j (w) + w \cdot \partial / \partial x_j (z), and \partial / \partial x_j (w^{-1}) = - w^{-1} \cdot \partial / \partial x_j (w), extended linearly to the . For a \langle x_1, \dots, x_n \mid r_1, \dots, r_m \rangle with m \geq n, compute the Jacobian matrix A whose (i,j)-entry is the image of \partial r_i / \partial x_j under the abelianization map \phi: \mathbb{Z}[F] \to \mathbb{Z}[t, t^{-1}], where F is the on the x_k and \phi(x_k) = t for all k (since the abelianization of the group is \mathbb{Z}, generated by the ). This matrix A is m \times n. The Alexander polynomial \Delta_K(t) is then the determinant of any (n-1) \times (n-1) minor of A, obtained by deleting one row and one column; all such minors yield determinants differing by units \pm t^k in the Laurent polynomial ring \mathbb{Z}[t, t^{-1}]. The overall process involves: (1) abelianizing the knot group to \mathbb{Z} via the Hurewicz homomorphism, reflecting the infinite cyclic covering space of the knot complement; (2) lifting to the chain complex of the covering, where the Alexander module's presentation arises from the Fox derivatives; and (3) computing the torsion of the first homology of the infinite cyclic cover, encoded by \Delta_K(t) as the order ideal generator. Normalization conventions fix \Delta_K(1) = 1 and make the polynomial symmetric, \Delta_K(t) = \Delta_K(t^{-1}), up to units. For the trefoil knot, consider its standard diagram with three arcs and Wirtinger generators x, y. The single essential relator is r = x y x y^{-1} x^{-1} y^{-1}. The Fox derivatives are \partial r / \partial x = 1 + x y - x y x y^{-1} x^{-1} and \partial r / \partial y = x - x y x y^{-1} - x y x y^{-1} x^{-1} y^{-1}. Abelianizing with \phi(x) = \phi(y) = t yields the $1 \times 2 Alexander matrix entries $1 + t^2 - t and t - t^2 - 1. The 1×1 minor obtained by deleting the y-column is $1 - t + t^2, which is the Alexander polynomial up to units.

Seifert matrix approach

The Seifert matrix approach provides a geometric to compute the Alexander polynomial of an oriented in by leveraging a bounding orientable surface known as a Seifert surface. According to Seifert's theorem from 1934, every oriented bounds such a surface, which can be explicitly constructed from a knot diagram using Seifert's . This begins by resolving each crossing in the diagram: at an overcrossing followed by an undercrossing in the orientation direction, the strands are smoothed to form disjoint, oriented circles called Seifert circles. These circles are then filled with disks, and at each original crossing, a rectangular band (or ribbon) is attached between the corresponding disks, twisted according to the crossing sign—right-handed for positive crossings and left-handed for negative—to ensure the boundary of the resulting surface is exactly the . This disk-band decomposition yields an embedded orientable surface of minimal for many knots, facilitating the extraction of topological invariants. Given a Seifert surface F for the , with g, a basis \{\alpha_1, \dots, \alpha_{2g}\} is chosen for the first homology group H_1(F; \mathbb{Z}), consisting of simple closed curves on the surface. The Seifert matrix V is the $2g \times 2g integer matrix defined by V_{ij} = \mathrm{lk}(\alpha_i, \hat{\alpha}_j), where \hat{\alpha}_j is the positive push-off of \alpha_j (a parallel curve displaced along the positive normal direction to F in \mathbb{R}^3 \setminus K), and \mathrm{lk} denotes the in S^3. The linking number is computed as half the signed crossings between the two curves, ensuring V captures the intersection form on the surface relative to the ambient space. This matrix is well-defined up to S-equivalence (congruence by integer matrices P with P^T P = I), independent of the choice of surface or basis, as established by Seifert. The Alexander polynomial is then obtained from the Seifert matrix via the formula \Delta_K(t) = \det(V - t V^T), where V^T is the of V; this determinant is a Laurent in t with coefficients. Different choices of Seifert surface or basis yield polynomials differing by multiplication by units in the Laurent , specifically \pm t^k for some k. Normalization conventions fix a unique representative by requiring \Delta_K(1) = 1 (which holds for knots since \det(V - V^T) = \pm 1, reflecting the unimodular nature of the intersection form) and \Delta_K(t) = \Delta_K(t^{-1}), up to the unit factor; orientation reversal transposes V to V^T, negating the polynomial up to units. These conventions ensure the polynomial is a well-defined . For the figure-eight knot $4_1, a minimal Seifert surface of genus 1 is constructed via Seifert's algorithm from its standard four-crossing diagram, yielding two Seifert circles connected by two twisted bands. A basis for H_1(F; \mathbb{Z}) consists of two curves c_1 and c_2 along the cores of the bands. The corresponding Seifert matrix is V = \begin{pmatrix} -1 & 0 \\ 1 & 1 \end{pmatrix}, with entries computed from the linking numbers of the curves and their positive push-offs (e.g., V_{11} = -1 from one self-linking, V_{12} = 0 from no intersection in push-off). Then, V - t V^T = \begin{pmatrix} t-1 & -t \\ 1 & 1-t \end{pmatrix}, and \det(V - t V^T) = (t-1)(1-t) + t = -t^2 + 3t - 1. Normalizing by multiplying by t^{-1} gives \Delta_{4_1}(t) = -t^{-1} + 3 - t, satisfying \Delta(1) = 1 and the symmetry condition.

Properties

Algebraic characteristics

The Alexander polynomial \Delta_K(t) of a knot K resides in the \Lambda = \mathbb{Z}[t, t^{-1}] and exhibits a : \Delta_K(t) = \pm t^k \Delta_K(t^{-1}) for some k, which stems from the t \mapsto t^{-1} acting on the Alexander module of the complement. This ensures that the coefficients of \Delta_K(t) are palindromic when the polynomial is normalized appropriately. Conventionally, the Alexander polynomial is normalized to be monic (leading coefficient 1) with integer coefficients, and for knots, it satisfies \Delta_K(1) = 1, reflecting the fact that the knot complement has trivial in degree 1. Additionally, the degree of \Delta_K(t) is even under this , as the constant term and leading coefficient must align due to the . The degree of the Alexander polynomial provides bounds related to the knot's Seifert genus g(K): specifically, \deg \Delta_K(t) \leq 2g(K), with equality achieved when a minimal genus Seifert surface yields the presentation matrix. For alternating knots, this equality holds using the canonical Seifert surface derived from a reduced alternating diagram, highlighting the polynomial's role in measuring knot complexity. As an invariant of the H_1(\tilde{X}_K; \mathbb{Z})—the first of the infinite cyclic cover of the complement—the is determined up to units \pm t^k in \Lambda by the \Lambda-module structure: isomorphic Alexander modules yield identical Alexander polynomials. For links with \mu components, the Alexander polynomial generalizes to a multivariable form \Delta_L(t_1, \dots, t_\mu) \in \mathbb{Z}[t_1^{\pm 1}, \dots, t_\mu^{\pm 1}], satisfying analogous \Delta_L(t_1, \dots, t_\mu) = \pm t_1^{k_1} \cdots t_\mu^{k_\mu} \Delta_L(t_1^{-1}, \dots, t_\mu^{-1}) for integers k_i, with the total degree connected to the pairwise linking numbers among components.

Multiplicativity under knot operations

The Alexander polynomial exhibits multiplicativity under the connected sum of knots. Specifically, if K_1 and K_2 are knots in S^3, then the Alexander polynomial of their connected sum K = K_1 \# K_2 satisfies \Delta_K(t) = \Delta_{K_1}(t) \Delta_{K_2}(t), up to units in the Laurent polynomial ring \mathbb{Z}[t, t^{-1}]. This property arises from the direct sum decomposition of the groups of the infinite cyclic covers of the knot complements. Under the mirror image operation, the Alexander polynomial transforms in a specific way. For a knot K, the polynomial of its mirror mK is given by \Delta_{mK}(t) = \Delta_K(t^{-1}), again up to units. This relation implies that the Alexander polynomial cannot distinguish a knot from its mirror image, as the substitution t \mapsto t^{-1} yields an equivalent Laurent polynomial under normalization conventions. The Alexander polynomial remains invariant under certain mutation operations on knots. In particular, it is unchanged by genus 2 mutations, which preserve the structure of the knot complement relevant to the polynomial's definition via Seifert matrices or Fox calculus. Similarly, for knot diagrams, the polynomial is invariant under flype operations, as these are part of the equivalence relations that define ambient isotopy. This invariance holds in cases where the mutations or flypes do not alter the underlying Alexander module. For cable constructions, the Alexander polynomial admits an explicit formula. The (p, q)-cable knot C_{p,q}(K) of a knot K, where p and q are with p > 0, has Alexander polynomial \Delta_{C_{p,q}(K)}(t) = \Delta_K(t^p) \Delta_{T_{p,q}}(t), where T_{p,q} denotes the (p, q)- and \Delta_{T_{p,q}}(t) = \frac{(t^{pq} - 1)(t - 1)}{(t^p - 1)(t^q - 1)}. This cabling formula reflects the satellite nature of the construction, combining the knot's polynomial with that of the pattern . These operational properties highlight the Alexander polynomial's utility, yet it is not a complete . Counterexamples include the Kinoshita-Terasaka knot and the , which are distinct 11-crossing s related by but share the identical trivial Alexander polynomial \Delta(t) = 1. Such mutants demonstrate that the polynomial fails to detect all topological differences arising from local changes in knot diagrams.

Interpretations

Geometric and topological meaning

The Alexander polynomial \Delta_K(t) encodes significant topological information about a K \subset S^3 through the of its covering spaces. Specifically, the complement S^3 \setminus K admits an infinite cyclic cover \widetilde{M_K} corresponding to the kernel of the abelianization map \pi_1(S^3 \setminus K) \to [\mathbb{Z}](/page/Z), where the deck transformations are generated by a . The first group H_1(\widetilde{M_K}; \mathbb{Z}) forms a torsion over the \Lambda = \mathbb{Z}[t, t^{-1}], and \Delta_K(t) is a of the of this , up to units in \Lambda. This representation captures the torsion structure intrinsic to the 's topology, distinguishing it from free parts of the . Similarly, for the n-fold cyclic branched cover \Sigma_n(K) of S^3 branched along K, the order of the torsion subgroup of H_1(\Sigma_n(K); \mathbb{Z}) equals \left| \prod_{j=0}^{n-1} \Delta_K([\omega](/page/Omega)^j) \right|, where \omega is a primitive nth ; in particular, for the double branched cover (n=2), |H_1(\Sigma_2(K); \mathbb{Z})| = |\Delta_K(-1)|. A key geometric consequence arises in relation to the Seifert genus g(K), the minimal genus of an orientable surface in S^3 bounded by K. The degree of \Delta_K(t) provides a lower bound: g(K) \geq \frac{1}{2} \deg \Delta_K(t), reflecting the minimal dimension required for a Seifert matrix to produce a of that degree. The evaluation |\Delta_K(-1)|, known as the knot determinant, further ties into this via the double branched cover, where large values indicate substantial torsion that constrains minimal surface complexities in certain classes, such as alternating knots where it often aligns closely with $2g(K) + 1. S-equivalence offers another topological : two knots are S-equivalent if their Seifert matrices are related by integer congruence and stabilizations (adding trivial bands), and this preserves \Delta_K(t). The Blanchfield duality links this to the Alexander , endowing H_1(\widetilde{M_K}; \mathbb{Q}(t)/\mathbb{Z}[t, t^{-1}]) with a nondegenerate Hermitian form over \mathbb{Q}(t), ensuring the module's self-duality and tying S-equivalence classes to algebraic concordance invariants. Geometrically, \Delta_K(t) admits an interpretation as the Reidemeister torsion of the chain complex of the infinite cyclic cover \widetilde{M_K}, specifically a regularized \tau(\widetilde{M_K}) = \Delta_K(t) and units, the "size" of the acyclic complex after tensoring with the Novikov ring. This torsion invariant, originally developed by Reidemeister in and refined by and Milnor, highlights the polynomial's role in quantifying deviations from acyclicity in the knot complement's . Finally, evaluations of \Delta_K(t) at of unity connect to cyclotomic fields: for a primitive nth root \zeta, |\Delta_K(\zeta)| divides the order of H_1(\Sigma_n(K); \mathbb{Z}), mirroring how Dedekind zeta values at of unity relate to class numbers in cyclotomic extensions. In arithmetic , this analogy portrays knots as "primes" in the "3-manifold world," with \Delta_K(t) akin to an Iwasawa polynomial governing growth in infinite towers of covers, akin to \mathbb{Z}_p-extensions of number fields.

Connections to modern homology theories

The Alexander polynomial of a knot K in S^3 serves as the graded of the knot groups \widehat{\HFK}(K), a bigraded invariant introduced by Ozsváth and Szabó in their development of Heegaard . Specifically, if \widehat{\HFK}_{i,j}(K) denotes the bigraded groups with Maslov grading i and Alexander grading j, then \Delta_K(t) = \sum_{i,j} (-1)^i t^j \rank \widehat{\HFK}_{i,j}(K), where \Delta_K(t) is the symmetrized Alexander polynomial. This relation positions the Alexander polynomial as the decategorification of \widehat{\HFK}(K), with the full providing a refinement that detects additional topological features, such as concordance obstructions beyond those from \Delta_K(t). In the Ozsváth-Szabó framework, the knot Floer complex \CFK^-(K) further refines this connection, where the Alexander polynomial equals the Euler characteristic \sum_i (-1)^i t^{\gr(i)} \dim \HFK^i(K, t) in the associated graded theory, with \gr denoting the Alexander grading. This categorification has been instrumental in modern low-dimensional topology, enabling computations of Seiberg-Witten invariants and Dehn surgery effects through the surgery exact triangle. Connections also extend to Khovanov homology, a categorification of the Jones polynomial, though it does not directly lift the Alexander polynomial; instead, related constructions in Khovanov-Rozansky homology yield Euler characteristics that specialize to quantum invariants encompassing the Alexander in certain limits. For instance, the sl(n) Khovanov-Rozansky theories provide a framework where the Poincaré polynomial in quantum and homological gradings relates to the Alexander polynomial via specialization at q = t and a = 1 in the HOMFLY-PT polynomial. In gauge-theoretic approaches, Floer homology for knots, developed by Kronheimer and Mrowka, recovers the Alexander polynomial from the torsion of the homology groups \KHI(K), mirroring the Heegaard Floer relation and providing evidence for the conjectural equivalence between these theories. Recent post-2000 advancements, particularly through bordered Heegaard Floer homology, refine these links via pairing theorems and surgery formulas; for example, the bordered invariants categorify the Alexander polynomial's behavior under satellite operations and Dehn fillings, as shown in the link surgery formula.

Extensions

Behavior under satellite constructions

Satellite knots are constructed by embedding a pattern knot P in a V and then mapping V homeomorphically onto a of a knot C in S^3, with the image of P yielding the satellite knot S. The w of P around the core of V measures how many times P wraps around C. The Alexander polynomial of a knot satisfies the formula \Delta_S(t) = \Delta_P(t) \cdot \Delta_C(t^w), where \Delta_P(t) is the Alexander polynomial of the pattern P viewed in the (normalized such that the pattern yields \Delta_{\text{unknot}}(t^w) = 1), and \Delta_C(t) is that of the . This multiplicative structure arises from the decomposition of the infinite cyclic cover of the satellite complement into covers associated to P and C. Cables represent a special class of satellites where the pattern is a T(m,n) on the boundary torus of the , with m. For the (m,n)- of K, the Alexander polynomial is \Delta_{m,n}(K)(t) = \Delta_K(t^m) \cdot \Delta_{T(m,n)}(t), where \Delta_{T(m,n)}(t) = \frac{(t^{mn}-1)(t-1)}{(t^m-1)(t^n-1)} (up to units t^k). This reflects the pattern's contribution from the torus knot polynomial, scaled by the companion's polynomial evaluated at higher powers. Whitehead doubles are satellites with a specific in the that clasps around the core with w=0. Untwisted Whitehead doubles of any K have trivial Alexander \Delta_W(t) = 1, independent of \Delta_K(t), since the \Delta_P(t) = 1 and t^0 = 1. This triviality implies that untwisted Whitehead doubles are non-fibered for non-trivial companions, as fibered knots have monic Alexander polynomials of even matching twice the , but here the is 0 while the exceeds 0. Satellite constructions preserve topological concordance properties linked to the ; for instance, satellites with trivial Alexander polynomial, such as untwisted doubles, are topologically slice, meaning concordant to the in the topological category. This follows from the general result that knots with \Delta_K(t) = 1 bound locally flat disks in the 4-ball topologically.

Alexander–Conway variant

The Alexander–Conway polynomial, introduced by in 1969 and published in 1970, reparametrizes the original Alexander by substituting z = t^{1/2} - t^{-1/2}, yielding \nabla(z) = \Delta(t), a in z with coefficients and only non-negative powers. This form addresses limitations of the Laurent polynomial structure in the t-variable by providing a cleaner, ordinary representation that simplifies computations. A key advantage of the Alexander–Conway variant is its presentation via Conway coefficients, which enable an arc index formulation suitable for open arcs and link diagrams, facilitating recursive calculations without fractional issues inherent in the original Laurent form. The polynomial is computed using the skein \nabla(L_+) - \nabla(L_-) = z \nabla(L_0), where L_+, L_-, and L_0 denote link diagrams differing locally at a crossing (positive, negative, and smoothed, respectively), with the normalization \nabla(\text{[unknot](/page/Unknot)}) = 1. For links, the Alexander–Conway polynomial extends naturally to a multi-variable version \nabla(z_1, \dots, z_\mu), where \mu is the number of components, related to the multi-variable Alexander polynomial through symmetrization over the variables. This multi-variable form preserves the skein relation in each variable while accommodating oriented components. The Alexander–Conway variant proves especially useful in , where its coefficients generate Vassiliev invariants; for instance, the coefficient of z^2 yields the simplest nontrivial finite type invariant of order 2.

References

  1. [1]
    None
    ### Summary of the Alexander Polynomial in Knot Theory
  2. [2]
  3. [3]
    [PDF] Knot Theory and the Alexander Polynomial - Elizabeth Denne
    Apr 15, 2008 · A familiar invariant of surfaces is the genus, which measures the number of holes in a surface. Make this a knot invariant as follows.
  4. [4]
    Alexander Polynomial -- from Wolfram MathWorld
    The Alexander polynomial is a knot invariant discovered in 1923 by J. W. Alexander (Alexander 1928) ... "Topological Invariants of Knots and Links." Trans. Amer.
  5. [5]
    [PDF] arXiv:2403.15732v1 [math.GT] 23 Mar 2024
    Mar 23, 2024 · ∆(t) is realized by a knot in the 3–sphere as its Alexander polynomial. (Here, . = shows the equality up to units ±ti in the Laurent polynomial ...
  6. [6]
    None
    ### Summary of Alexander Polynomial Definition from J.W. Alexander's 1928 Paper
  7. [7]
  8. [8]
    [PDF] a short introduction to the alexander polynomial
    The Alexander polynomial of a link is defined up to some indeterminacy. ... Knots and links. Publish or Perish Inc., Berkeley, Calif., 1976. Mathematics ...
  9. [9]
    Topological Invariants of Knots and Links - jstor
    TOPOLOGICAL INVARIANTS OF KNOTS AND LINKS* ... * Presented to the Society, May 7, 1927; received by the editors, October 13, 1927. 275. Page 2. 276 J. W. ...
  10. [10]
    On Types of Knotted Curves - jstor
    ALEXANDER AND G. B. BRIGGS. 1. The problem of determining the various possible types of closed, knotted curves in 3-space was originally studied ...
  11. [11]
    Free Differential Calculus. I: Derivation in the Free Group Ring - jstor
    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 polynomial [1].<|control11|><|separator|>
  12. [12]
    Free Differential Calculus, V. The Alexander Matrices Re-Examined
    In FDC II, I defined the Alexander polynomial of a group G (having a finite presentation in which there are more generators than relations).
  13. [13]
    [PDF] Section 6.4. Knot Groups and the Alexander Polynomial
    Mar 6, 2021 · The algorithm we present was developed by Ralph Fox in five papers in the. 1950s: 1. Fox, R., “Free Differential Calculus, I: Derivation in the ...
  14. [14]
    Über das Geschlecht von Knoten - EuDML
    Seifert, H.. "Über das Geschlecht von Knoten." Mathematische Annalen 110 (1935): 571-592. <http://eudml.org/doc/159739>. @article{Seifert1935,Missing: Herbert 1934 PDF
  15. [15]
    [PDF] MATH 309 - Solution to Homework 4
    Apr 3, 2019 · The Alexander polynomial of a projection of the figure 8 via the linking matrix. ... Figure 5: A Seifert surface for the figure-eight knot.
  16. [16]
    4 1 - Knot Atlas
    Jul 14, 2007 · 4_1 is also known as "the Figure Eight knot", as some people think it looks like a figure `8' in one of its common projections.
  17. [17]
    [PDF] Three flavors of twisted invariants of knots - Stefan Friedl's homepage
    Jul 8, 2014 · example, Seifert [Se34] showed that the Alexander polynomial can be normalized such that ∆K(t−1)=∆K(t) and ∆K(1) = 1, and that any polynomial ...
  18. [18]
    [PDF] THE ALEXANDER POLYNOMIAL | Nancy Scherich
    Grid diagrams are a representation of knots and links that are used to describe the Minesweeper Matrix. This section develops some basic theory of grid diagrams ...
  19. [19]
    [PDF] The Alexander Polynomial - Yuqing Shi
    A knot K ≈ S1 in S3 has genus zero if and only if it is a unknot. Proof. If K is an unknot, K bounds a disk in S3, which has genus zero. Thus g(K) = 0.
  20. [20]
    On the genus of the alternating knot, I. - Project Euclid
    THEORFM 1.1. For any alternating knot zvitha constant incidence number, the genus is exactly equal to one half of the degree of its Alexander polynomial.
  21. [21]
    [PDF] On the Alexander Polynomial
    We have included, in Chapter III, a proof of a theorem (Theorem 5), which is a generalization, for A > 1, of a theorem proved by Seifert [11] for A = 1. ... 10 ( ...
  22. [22]
    [PDF] Explicit Formulas for the Alexander Polynomial of Pretzel Knots - arXiv
    Recall that the Alexander polynomial of a connected sum is the product of the Alexander polynomials of its summands. Finally, we use the skein relation to ...
  23. [23]
    [PDF] Alexander polynomial of knots - Berkeley Math
    First discovered by J.W. Alexander in 1928, the Alexander polyno- mial was the only known polynomial invariant of knot types for over 50 years, until Jones ...
  24. [24]
    [PDF] behavior of knot invariants under genus 2 mutation
    Genus 2 mutation preserves Alexander and Jones polynomials, but not HOMFLY-PT. It can also change sl2-Khovanov homology.
  25. [25]
    [PDF] square numbers and polynomial invariants of achiral knots
    Also, the skein and Alexander polynomial are invariant under mutation, so they are well-defined on a mutation equivalence class.
  26. [26]
    [PDF] A Cable Knot and BPS-Series
    Jan 13, 2023 · ... knot [13]. 3.2 The Alexander polynomial. The cabling formula for the Alexander polynomial of a knot K is [18]. ∆C(p,q)(K)(t)=∆K(tp)∆T(p,q) ...
  27. [27]
    Computing knot Floer homology in cyclic branched covers - MSP
    Jul 25, 2008 · The order of H1.†m.K// is equal to. Qm1. jD0 БK .!j /, where БK is the Alexander polynomial of K, and ! is a primitive mth root of unity (Fox ...
  28. [28]
    [PDF] Knots, Polynomials, and Categorification - Jacob Rasmussen
    These lectures give an introduction to knot polynomials and their cat- egorifications. Topics covered include the Jones and Alexander polynomials,. Khovanov ...
  29. [29]
    [PDF] S-EQUIVALENCE OF KNOTS 1. Introduction An oriented knot k is a ...
    Every oriented knot is spanned by an oriented surface, a Seifert surface, and this gives rise to a matrix of linking numbers called a Seifert matrix. Any two ...
  30. [30]
    [PDF] Reidemeister torsion, peripheral complex, and Alexander ...
    A different approach to the study of Alexander polynomials relies on the use of Reidemeister torsion. Milnor [Mi62, Mi66] showed that the Alexander polynomial ...
  31. [31]
    [PDF] Knots in Number Theory - Universiteit Leiden
    We build the theory to compute the Alexander polynomial from the ground up, after which we will show that the same construction very much applies to the Iwasawa ...
  32. [32]
    [PDF] Abelian invariants of satellite knots - Paul Melvin
    Since the Alexander polynomial of a knot K is just detAK(t) , we have. Corollary (Seifert [S]). As(t) = AE(t)Ac(t w). Remarks. (i). If w = 0 , then the theorem ...
  33. [33]
    [PDF] arXiv:0806.2172v2 [math.GT] 15 Jun 2008
    Jun 15, 2008 · Our intuition comes from the fact that the Alexander polynomial of cable knots is determined by the formula. (1). ∆Kp,q (t)=∆Tp,q (t) · ∆K ...
  34. [34]
    Knot Floer homology of Whitehead doubles - MSP
    Dec 17, 2007 · In particular, the Alexander polynomial of the 0–twisted Whitehead double of K is trivial. It is thus an interesting question to ask how, if ...
  35. [35]
    New topologically slice knots - Project Euclid
    In the early 1980's Mike Freedman showed that all knots with trivial Alexander polynomial are topologically slice (with fundamental group Z ℤ ).
  36. [36]
    [PDF] John Horton Conway: The Man and His Knot Theory
    May 27, 2022 · An Alexander polynomial of a link can be calculated from its Conway polynomial by substituting z = t1/2−t-1/2. Laurent polynomials are not ...<|control11|><|separator|>
  37. [37]
    [PDF] Section 10.1. The Conway Polynomial of a Knot
    Feb 6, 2021 · −1 and in this polynomial we have (t − 1 + t. −1. )|t=1 = 1 so this is the normalized version of the Alexander polynomial of the trefoil knot.
  38. [38]
    Vassiliev Invariant -- from Wolfram MathWorld
    Vassiliev invariants, discovered around 1989, provided a radically new way of looking at knots. The notion of finite type (a.k.a. Vassiliev) knot invariants ...