Torus knot

Last updated
A (3,-7)-3D torus knot. TorusKnot3D.png
A (3,−7)-3D torus knot.
EureleA Award showing a (2,3)-torus knot. Eurelea.png
EureleA Award showing a (2,3)-torus knot.
(2,8) torus link (2,8)-Torus Link.svg
(2,8) torus link

In knot theory, a torus knot is a special kind of knot that lies on the surface of an unknotted torus in R3. Similarly, a torus link is a link which lies on the surface of a torus in the same way. Each torus knot is specified by a pair of coprime integers p and q. A torus link arises if p and q are not coprime (in which case the number of components is gcd(p, q)). A torus knot is trivial (equivalent to the unknot) if and only if either p or q is equal to 1 or −1. The simplest nontrivial example is the (2,3)-torus knot, also known as the trefoil knot.

Contents

the (2,-3)-torus knot, also known as the left-handed trefoil knot Trefoil knot left.svg
the (2,−3)-torus knot, also known as the left-handed trefoil knot

Geometrical representation

A torus knot can be rendered geometrically in multiple ways which are topologically equivalent (see Properties below) but geometrically distinct. The convention used in this article and its figures is the following.

The (p,q)-torus knot winds q times around a circle in the interior of the torus, and p times around its axis of rotational symmetry. [note 1] . If p and q are not relatively prime, then we have a torus link with more than one component.

The direction in which the strands of the knot wrap around the torus is also subject to differing conventions. The most common is to have the strands form a right-handed screw for p q > 0. [3] [4] [5]

The (p,q)-torus knot can be given by the parametrization

where and . This lies on the surface of the torus given by (in cylindrical coordinates).

Other parameterizations are also possible, because knots are defined up to continuous deformation. The illustrations for the (2,3)- and (3,8)-torus knots can be obtained by taking , and in the case of the (2,3)-torus knot by furthermore subtracting respectively and from the above parameterizations of x and y. The latter generalizes smoothly to any coprime p,q satisfying .

Properties

Diagram of a (3,-8)-torus knot. TorusKnot-3-8.png
Diagram of a (3,−8)-torus knot.

A torus knot is trivial iff either p or q is equal to 1 or −1. [4] [5]

Each nontrivial torus knot is prime [6] and chiral. [4]

The (p,q) torus knot is equivalent to the (q,p) torus knot. [3] [5] This can be proved by moving the strands on the surface of the torus. [7] The (p,−q) torus knot is the obverse (mirror image) of the (p,q) torus knot. [5] The (−p,−q) torus knot is equivalent to the (p,q) torus knot except for the reversed orientation.

The (3, 4) torus knot on the unwrapped torus surface, and its braid word (3, 4) torus knot.svg
The (3, 4) torus knot on the unwrapped torus surface, and its braid word

Any (p,q)-torus knot can be made from a closed braid with p strands. The appropriate braid word is [8]

(This formula assumes the common convention that braid generators are right twists, [4] [8] [9] [10] which is not followed by the Wikipedia page on braids.)

The crossing number of a (p,q) torus knot with p,q > 0 is given by

c = min((p1)q, (q1)p).

The genus of a torus knot with p,q > 0 is

The Alexander polynomial of a torus knot is [3] [8]

where

The Jones polynomial of a (right-handed) torus knot is given by

The complement of a torus knot in the 3-sphere is a Seifert-fibered manifold, fibred over the disc with two singular fibres.

Let Y be the p-fold dunce cap with a disk removed from the interior, Z be the q-fold dunce cap with a disk removed from its interior, and X be the quotient space obtained by identifying Y and Z along their boundary circle. The knot complement of the (p, q) -torus knot deformation retracts to the space X. Therefore, the knot group of a torus knot has the presentation

Torus knots are the only knots whose knot groups have nontrivial center (which is infinite cyclic, generated by the element in the presentation above).

The stretch factor of the (p,q) torus knot, as a curve in Euclidean space, is Ω(min(p,q)), so torus knots have unbounded stretch factors. Undergraduate researcher John Pardon won the 2012 Morgan Prize for his research proving this result, which solved a problem originally posed by Mikhail Gromov. [11] [12]

Connection to complex hypersurfaces

The (p,q)torus knots arise when considering the link of an isolated complex hypersurface singularity. One intersects the complex hypersurface with a hypersphere, centred at the isolated singular point, and with sufficiently small radius so that it does not enclose, nor encounter, any other singular points. The intersection gives a submanifold of the hypersphere.

Let p and q be coprime integers, greater than or equal to two. Consider the holomorphic function given by Let be the set of such that Given a real number we define the real three-sphere as given by The function has an isolated critical point at since if and only if Thus, we consider the structure of close to In order to do this, we consider the intersection This intersection is the so-called link of the singularity The link of , where p and q are coprime, and both greater than or equal to two, is exactly the (p,q)torus knot. [13]

List

(36,3) torus link Torus link (36,3).png
(36,3) torus link

The figure on the right is torus link (72,4) .

Table
#
A-B ImagePQ Cross
#
001 Blue Unknot.png 0
3a131 (3-2) torus knot.png 233
5a251 A (5,2)-torus knot.png 255
7a771 A (7,2)-torus knot.png 277
8n3819 A (4,3)-torus knot.png 348
9a4191 A (9,2)-torus knot.png 299
10n2110124 A (5,3)-torus knot.png 3510
11a367 Simple-knotwork-cross-12crossings.svg 21111
13a4878 (2,13)-Torus Knot.png 21313
14n21881 A (7,3)-torus knot.png 3714
15n41185 A (5,4)-torus knot.png 4515
15a8526321515
16n783154 A (8,3)-torus knot.png 3816
21717
21919
31020
A (7,4)-torus knot.png 4721
22121
31122
22323
A (6,5)-torus knot.png 5624
22525
31326
A (9,4)-torus knot.png 4927
22727
A (7,5)-torus knot.png 5728
31428
22929
23131
A (8,5)-torus knot.png 5832
31632
41133
23333
31734
A (7,6)-torus knot.png 6735
23535
A (9,5)-torus knot.png 5936
A (8,7)-torus knot.png 7848
A (9,7)-torus knot.png 7954
A (9,8)-torus knot.png 8963

g-torus knot

A g-torus knot is a closed curve drawn on a g-torus. More technically, it is the homeomorphic image of a circle in which can be realized as a subset of a genus g handlebody in (whose complement is also a genus g handlebody). If a link is a subset of a genus two handlebody, it is a double torus link. [14]

For genus two, the simplest example of a double torus knot that is not a torus knot is the figure-eight knot. [15] [16]

Notes

  1. Note that this use of the roles of p and q is contrary to what appears on. [1] It is also inconsistent with the pictures that appear in: [2]

See also

Related Research Articles

In mathematics, de Moivre's formula states that for any real number x and integer n it holds that

<span class="mw-page-title-main">Genus (mathematics)</span> Number of "holes" of a surface

In mathematics, genus has a few different, but closely related, meanings. Intuitively, the genus is the number of "holes" of a surface. A sphere has genus 0, while a torus has genus 1.

In mathematics, a quadric or quadric surface (quadric hypersurface in higher dimensions), is a generalization of conic sections (ellipses, parabolas, and hyperbolas). It is a hypersurface (of dimension D) in a (D + 1)-dimensional space, and it is defined as the zero set of an irreducible polynomial of degree two in D + 1 variables; for example, D = 1 in the case of conic sections. When the defining polynomial is not absolutely irreducible, the zero set is generally not considered a quadric, although it is often called a degenerate quadric or a reducible quadric.

In mechanics and geometry, the 3D rotation group, often denoted SO(3), is the group of all rotations about the origin of three-dimensional Euclidean space under the operation of composition.

<span class="mw-page-title-main">Vapnik–Chervonenkis theory</span> Branch of statistical computational learning theory

Vapnik–Chervonenkis theory was developed during 1960–1990 by Vladimir Vapnik and Alexey Chervonenkis. The theory is a form of computational learning theory, which attempts to explain the learning process from a statistical point of view.

<span class="mw-page-title-main">Trefoil knot</span> Simplest non-trivial closed knot with three crossings

In knot theory, a branch of mathematics, the trefoil knot is the simplest example of a nontrivial knot. The trefoil can be obtained by joining the two loose ends of a common overhand knot, resulting in a knotted loop. As the simplest knot, the trefoil is fundamental to the study of mathematical knot theory.

In probability theory, the Borel–Kolmogorov paradox is a paradox relating to conditional probability with respect to an event of probability zero. It is named after Émile Borel and Andrey Kolmogorov.

In mathematics, Milnor maps are named in honor of John Milnor, who introduced them to topology and algebraic geometry in his book Singular Points of Complex Hypersurfaces and earlier lectures. The most studied Milnor maps are actually fibrations, and the phrase Milnor fibration is more commonly encountered in the mathematical literature. These were introduced to study isolated singularities by constructing numerical invariants related to the topology of a smooth deformation of the singular space.

In mathematics, more specifically in dynamical systems, the method of averaging exploits systems containing time-scales separation: a fast oscillationversus a slow drift. It suggests that we perform an averaging over a given amount of time in order to iron out the fast oscillations and observe the qualitative behavior from the resulting dynamics. The approximated solution holds under finite time inversely proportional to the parameter denoting the slow time scale. It turns out to be a customary problem where there exists the trade off between how good is the approximated solution balanced by how much time it holds to be close to the original solution.

In physics, spherical multipole moments are the coefficients in a series expansion of a potential that varies inversely with the distance R to a source, i.e., as  Examples of such potentials are the electric potential, the magnetic potential and the gravitational potential.

A quasi-Hopf algebra is a generalization of a Hopf algebra, which was defined by the Russian mathematician Vladimir Drinfeld in 1989.

<span class="mw-page-title-main">Voigt effect</span>

The Voigt effect is a magneto-optical phenomenon which rotates and elliptizes linearly polarised light sent into an optically active medium. The effect is named after the German scientist Woldemar Voigt who discovered it in vapors. Unlike many other magneto-optical effects such as the Kerr or Faraday effect which are linearly proportional to the magnetization, the Voigt effect is proportional to the square of the magnetization and can be seen experimentally at normal incidence. There are also other denominations for this effect, used interchangeably in the modern scientific literature: the Cotton–Mouton effect and magnetic-linear birefringence, with the latter reflecting the physical meaning of the effect.

In physics, the Green's function for Laplace's equation in three variables is used to describe the response of a particular type of physical system to a point source. In particular, this Green's function arises in systems that can be described by Poisson's equation, a partial differential equation (PDE) of the form

In mathematics, the spin representations are particular projective representations of the orthogonal or special orthogonal groups in arbitrary dimension and signature. More precisely, they are two equivalent representations of the spin groups, which are double covers of the special orthogonal groups. They are usually studied over the real or complex numbers, but they can be defined over other fields.

<span class="mw-page-title-main">Near sets</span> Concept in mathematical set theory

In mathematics, near sets are either spatially close or descriptively close. Spatially close sets have nonempty intersection. In other words, spatially close sets are not disjoint sets, since they always have at least one element in common. Descriptively close sets contain elements that have matching descriptions. Such sets can be either disjoint or non-disjoint sets. Spatially near sets are also descriptively near sets.

The Bowring series of the transverse mercator published in 1989 by Bernard Russel Bowring gave formulas for the Transverse Mercator that are simpler to program but retain millimeter accuracy.

In mathematics, singular integral operators of convolution type are the singular integral operators that arise on Rn and Tn through convolution by distributions; equivalently they are the singular integral operators that commute with translations. The classical examples in harmonic analysis are the harmonic conjugation operator on the circle, the Hilbert transform on the circle and the real line, the Beurling transform in the complex plane and the Riesz transforms in Euclidean space. The continuity of these operators on L2 is evident because the Fourier transform converts them into multiplication operators. Continuity on Lp spaces was first established by Marcel Riesz. The classical techniques include the use of Poisson integrals, interpolation theory and the Hardy–Littlewood maximal function. For more general operators, fundamental new techniques, introduced by Alberto Calderón and Antoni Zygmund in 1952, were developed by a number of authors to give general criteria for continuity on Lp spaces. This article explains the theory for the classical operators and sketches the subsequent general theory.

Functional regression is a version of regression analysis when responses or covariates include functional data. Functional regression models can be classified into four types depending on whether the responses or covariates are functional or scalar: (i) scalar responses with functional covariates, (ii) functional responses with scalar covariates, (iii) functional responses with functional covariates, and (iv) scalar or functional responses with functional and scalar covariates. In addition, functional regression models can be linear, partially linear, or nonlinear. In particular, functional polynomial models, functional single and multiple index models and functional additive models are three special cases of functional nonlinear models.

In representation theory of mathematics, the Waldspurger formula relates the special values of two L-functions of two related admissible irreducible representations. Let k be the base field, f be an automorphic form over k, π be the representation associated via the Jacquet–Langlands correspondence with f. Goro Shimura (1976) proved this formula, when and f is a cusp form; Günter Harder made the same discovery at the same time in an unpublished paper. Marie-France Vignéras (1980) proved this formula, when and f is a newform. Jean-Loup Waldspurger, for whom the formula is named, reproved and generalized the result of Vignéras in 1985 via a totally different method which was widely used thereafter by mathematicians to prove similar formulas.

<span class="mw-page-title-main">Lissajous-toric knot</span>

In knot theory, a Lissajous-toric knot is a knot defined by parametric equations of the form

References

  1. Torus Knot on Wolfram Mathworld .
  2. "36 Torus Knots", The Knot Atlas. .
  3. 1 2 3 Livingston, Charles (1993). Knot Theory. Mathematical Association of America. p. [ page needed ]. ISBN   0-88385-027-3.
  4. 1 2 3 4 Murasugi, Kunio (1996). Knot Theory and its Applications. Birkhäuser. p. [ page needed ]. ISBN   3-7643-3817-2.
  5. 1 2 3 4 Kawauchi, Akio (1996). A Survey of Knot Theory. Birkhäuser. p. [ page needed ]. ISBN   3-7643-5124-1.
  6. Norwood, F. H. (1982-01-01). "Every two-generator knot is prime". Proceedings of the American Mathematical Society. 86 (1): 143–147. doi: 10.1090/S0002-9939-1982-0663884-7 . ISSN   0002-9939. JSTOR   2044414.
  7. Baker, Kenneth (2011-03-28). "p q is q p". Sketches of Topology. Retrieved 2020-11-09.
  8. 1 2 3 Lickorish, W. B. R. (1997). An Introduction to Knot Theory. Springer. p. [ page needed ]. ISBN   0-387-98254-X.
  9. Dehornoy, P.; Dynnikov, Ivan; Rolfsen, Dale; Wiest, Bert (2000). Why are Braids Orderable? (PDF). p. [ page needed ]. Archived from the original (PDF) on 2012-04-15. Retrieved 2011-11-12.
  10. Birman, J. S.; Brendle, T. E. (2005). "Braids: a Survey". In Menasco, W.; Thistlethwaite, M. (eds.). Handbook of Knot Theory. Elsevier. p. [ page needed ]. ISBN   0-444-51452-X.
  11. Kehoe, Elaine (April 2012), "2012 Morgan Prize", Notices of the American Mathematical Society , vol. 59, no. 4, pp. 569–571, doi: 10.1090/noti825 .
  12. Pardon, John (2011), "On the distortion of knots on embedded surfaces", Annals of Mathematics , Second Series, 174 (1): 637–646, arXiv: 1010.1972 , doi:10.4007/annals.2011.174.1.21, MR   2811613, S2CID   55567836
  13. Milnor, J. (1968). Singular Points of Complex Hypersurfaces. Princeton University Press. p. [ page needed ]. ISBN   0-691-08065-8.
  14. Rolfsen, Dale (1976). Knots and Links. Publish or Perish, Inc. p. [ page needed ]. ISBN   0-914098-16-0.
  15. Hill, Peter (December 1999). "On Double-Torus Knots (I)". Journal of Knot Theory and Its Ramifications. 08 (8): 1009–1048. doi:10.1142/S0218216599000651. ISSN   0218-2165.
  16. Norwood, Frederick (November 1989). "Curves on surfaces". Topology and Its Applications. 33 (3): 241–246. doi: 10.1016/0166-8641(89)90105-3 .