Writhe

Last updated

In knot theory, there are several competing notions of the quantity writhe, or . In one sense, it is purely a property of an oriented link diagram and assumes integer values. In another sense, it is a quantity that describes the amount of "coiling" of a mathematical knot (or any closed simple curve) in three-dimensional space and assumes real numbers as values. In both cases, writhe is a geometric quantity, meaning that while deforming a curve (or diagram) in such a way that does not change its topology, one may still change its writhe. [1]

Contents

In knot theory, the writhe is a property of an oriented link diagram. The writhe is the total number of positive crossings minus the total number of negative crossings.

A direction is assigned to the link at a point in each component and this direction is followed all the way around each component. For each crossing one comes across while traveling in this direction, if the strand underneath goes from right to left, the crossing is positive; if the lower strand goes from left to right, the crossing is negative. One way of remembering this is to use a variation of the right-hand rule.

Knot-crossing-plus.svg Knot-crossing-minus.svg
Positive
crossing
Negative
crossing

For a knot diagram, using the right-hand rule with either orientation gives the same result, so the writhe is well-defined on unoriented knot diagrams.

A Type I Reidemeister move changes the writhe by 1 Reidemeister move 1.svg
A Type I Reidemeister move changes the writhe by 1

The writhe of a knot is unaffected by two of the three Reidemeister moves: moves of Type II and Type III do not affect the writhe. Reidemeister move Type I, however, increases or decreases the writhe by 1. This implies that the writhe of a knot is not an isotopy invariant of the knot itself — only the diagram. By a series of Type I moves one can set the writhe of a diagram for a given knot to be any integer at all.

Writhe of a closed curve

Writhe is also a property of a knot represented as a curve in three-dimensional space. Strictly speaking, a knot is such a curve, defined mathematically as an embedding of a circle in three-dimensional Euclidean space, . By viewing the curve from different vantage points, one can obtain different projections and draw the corresponding knot diagrams. Its writhe (in the space curve sense) is equal to the average of the integral writhe values obtained from the projections from all vantage points. [2] Hence, writhe in this situation can take on any real number as a possible value. [1]

In a paper from 1961, [3] Gheorghe Călugăreanu proved the following theorem: take a ribbon in , let be the linking number of its border components, and let be its total twist. Then the difference depends only on the core curve of the ribbon, [2] and

.

In a paper from 1959, [4] Călugăreanu also showed how to calculate the writhe Wr with an integral. Let be a smooth, simple, closed curve and let and be points on . Then the writhe is equal to the Gauss integral

.

Numerically approximating the Gauss integral for writhe of a curve in space

Since writhe for a curve in space is defined as a double integral, we can approximate its value numerically by first representing our curve as a finite chain of line segments. A procedure that was first derived by Michael Levitt [5] for the description of protein folding and later used for supercoiled DNA by Konstantin Klenin and Jörg Langowski [6] is to compute

,

where is the exact evaluation of the double integral over line segments and ; note that and . [6]

To evaluate for given segments numbered and , number the endpoints of the two segments 1, 2, 3, and 4. Let be the vector that begins at endpoint and ends at endpoint . Define the following quantities: [6]

Then we calculate [6]

Finally, we compensate for the possible sign difference and divide by to obtain [6]

In addition, other methods to calculate writhe can be fully described mathematically and algorithmically, some of them outperform method above (which has quadratic computational complexity, by having linear complexity). [6]

A simulation of an elastic rod relieving torsional stress by forming coils

Applications in DNA topology

DNA will coil when twisted, just like a rubber hose or a rope will, and that is why biomathematicians use the quantity of writhe to describe the amount a piece of DNA is deformed as a result of this torsional stress. In general, this phenomenon of forming coils due to writhe is referred to as DNA supercoiling and is quite commonplace, and in fact in most organisms DNA is negatively supercoiled. [1]

Any elastic rod, not just DNA, relieves torsional stress by coiling, an action which simultaneously untwists and bends the rod. F. Brock Fuller shows mathematically [7] how the “elastic energy due to local twisting of the rod may be reduced if the central curve of the rod forms coils that increase its writhing number”.

See also

Related Research Articles

In physics, the cross section is a measure of the probability that a specific process will take place in a collision of two particles. For example, the Rutherford cross-section is a measure of probability that an alpha particle will be deflected by a given angle during an interaction with an atomic nucleus. Cross section is typically denoted σ (sigma) and is expressed in units of area, more specifically in barns. In a way, it can be thought of as the size of the object that the excitation must hit in order for the process to occur, but more exactly, it is a parameter of a stochastic process.

<span class="mw-page-title-main">Dipole</span> Electromagnetic phenomenon

In physics, a dipole is an electromagnetic phenomenon which occurs in two ways:

<span class="mw-page-title-main">Discrete Fourier transform</span> Type of Fourier transform in discrete mathematics

In mathematics, the discrete Fourier transform (DFT) converts a finite sequence of equally-spaced samples of a function into a same-length sequence of equally-spaced samples of the discrete-time Fourier transform (DTFT), which is a complex-valued function of frequency. The interval at which the DTFT is sampled is the reciprocal of the duration of the input sequence. An inverse DFT (IDFT) is a Fourier series, using the DTFT samples as coefficients of complex sinusoids at the corresponding DTFT frequencies. It has the same sample-values as the original input sequence. The DFT is therefore said to be a frequency domain representation of the original input sequence. If the original sequence spans all the non-zero values of a function, its DTFT is continuous, and the DFT provides discrete samples of one cycle. If the original sequence is one cycle of a periodic function, the DFT provides all the non-zero values of one DTFT cycle.

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances. They were named after French engineer and physicist Claude-Louis Navier and the Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

In vector calculus, the divergence theorem, also known as Gauss's theorem or Ostrogradsky's theorem, is a theorem relating the flux of a vector field through a closed surface to the divergence of the field in the volume enclosed.

<span class="mw-page-title-main">Linking number</span> Numerical invariant that describes the linking of two closed curves in three-dimensional space

In mathematics, the linking number is a numerical invariant that describes the linking of two closed curves in three-dimensional space. Intuitively, the linking number represents the number of times that each curve winds around the other. In Euclidean space, the linking number is always an integer, but may be positive or negative depending on the orientation of the two curves.

<span class="mw-page-title-main">Holonomy</span> Concept in differential geometry

In differential geometry, the holonomy of a connection on a smooth manifold is the extent to which parallel transport around closed loops fails to preserve the geometrical data being transported. Holonomy is a general geometrical consequence of the curvature of the connection. For flat connections, the associated holonomy is a type of monodromy and is an inherently global notion. For curved connections, holonomy has nontrivial local and global features.

In mathematics, the total variation identifies several slightly different concepts, related to the (local or global) structure of the codomain of a function or a measure. For a real-valued continuous function f, defined on an interval [a, b] ⊂ R, its total variation on the interval of definition is a measure of the one-dimensional arclength of the curve with parametric equation xf(x), for x ∈ [a, b]. Functions whose total variation is finite are called functions of bounded variation.

<span class="mw-page-title-main">Phasor</span> Complex number representing a particular sine wave

In physics and engineering, a phasor is a complex number representing a sinusoidal function whose amplitude, and initial phase are time-invariant and whose angular frequency is fixed. It is related to a more general concept called analytic representation, which decomposes a sinusoid into the product of a complex constant and a factor depending on time and frequency. The complex constant, which depends on amplitude and phase, is known as a phasor, or complex amplitude, and sinor or even complexor.

In calculus, the Leibniz integral rule for differentiation under the integral sign, named after Gottfried Wilhelm Leibniz, states that for an integral of the form where and the integrands are functions dependent on the derivative of this integral is expressible as where the partial derivative indicates that inside the integral, only the variation of with is considered in taking the derivative.

<span class="mw-page-title-main">Weierstrass–Enneper parameterization</span> Construction for minimal surfaces

In mathematics, the Weierstrass–Enneper parameterization of minimal surfaces is a classical piece of differential geometry.

The gradient theorem, also known as the fundamental theorem of calculus for line integrals, says that a line integral through a gradient field can be evaluated by evaluating the original scalar field at the endpoints of the curve. The theorem is a generalization of the second fundamental theorem of calculus to any curve in a plane or space rather than just the real line.

<span class="mw-page-title-main">Pendulum (mechanics)</span> Free swinging suspended body

A pendulum is a body suspended from a fixed support such that it freely swings back and forth under the influence of gravity. When a pendulum is displaced sideways from its resting, equilibrium position, it is subject to a restoring force due to gravity that will accelerate it back towards the equilibrium position. When released, the restoring force acting on the pendulum's mass causes it to oscillate about the equilibrium position, swinging it back and forth. The mathematics of pendulums are in general quite complicated. Simplifying assumptions can be made, which in the case of a simple pendulum allow the equations of motion to be solved analytically for small-angle oscillations.

In many-body theory, the term Green's function is sometimes used interchangeably with correlation function, but refers specifically to correlators of field operators or creation and annihilation operators.

<span class="mw-page-title-main">Axis–angle representation</span> Parameterization of a rotation into a unit vector and angle

In mathematics, the axis–angle representation parameterizes a rotation in a three-dimensional Euclidean space by two quantities: a unit vector e indicating the direction of an axis of rotation, and an angle of rotation θ describing the magnitude and sense of the rotation about the axis. Only two numbers, not three, are needed to define the direction of a unit vector e rooted at the origin because the magnitude of e is constrained. For example, the elevation and azimuth angles of e suffice to locate it in any particular Cartesian coordinate frame.

<span class="mw-page-title-main">Fisher's noncentral hypergeometric distribution</span>

In probability theory and statistics, Fisher's noncentral hypergeometric distribution is a generalization of the hypergeometric distribution where sampling probabilities are modified by weight factors. It can also be defined as the conditional distribution of two or more binomially distributed variables dependent upon their fixed sum.

In geometry, the polar sine generalizes the sine function of angle to the vertex angle of a polytope. It is denoted by psin.

In mathematics, the Möbius energy of a knot is a particular knot energy, i.e., a functional on the space of knots. It was discovered by Jun O'Hara, who demonstrated that the energy blows up as the knot's strands get close to one another. This is a useful property because it prevents self-intersection and ensures the result under gradient descent is of the same knot type.

In optics, the Fraunhofer diffraction equation is used to model the diffraction of waves when the diffraction pattern is viewed at a long distance from the diffracting object, and also when it is viewed at the focal plane of an imaging lens.

An impulse vector, also known as Kang vector, is a mathematical tool used to graphically design and analyze input shapers that can suppress residual vibration. The impulse vector can be applied to both undamped and underdamped systems, as well as to both positive and negative impulses in a unified manner. The impulse vector makes it easy to obtain impulse time and magnitude of the input shaper graphically. A vector concept for an input shaper was first introduced by W. Singhose for undamped systems with positive impulses. Building on this idea, C.-G. Kang introduced the impulse vector to generalize Singhose's idea to undamped and underdamped systems with positive and negative impulses.

References

  1. 1 2 3 Bates, Andrew (2005). DNA Topology. Oxford University Press. pp. 36–37. ISBN   978-0-19-850655-3.
  2. 1 2 Cimasoni, David (2001). "Computing the writhe of a knot". Journal of Knot Theory and Its Ramifications . 10 (387): 387–395. arXiv: math/0406148 . doi:10.1142/S0218216501000913. MR   1825964. S2CID   15850269.
  3. Călugăreanu, Gheorghe (1961). "Sur les classes d'isotopie des nœuds tridimensionnels et leurs invariants". Czechoslovak Mathematical Journal (in French). 11 (4): 588–625. doi: 10.21136/CMJ.1961.100486 . MR   0149378.
  4. Călugăreanu, Gheorghe (1959). "L'intégrale de Gauss et l'analyse des nœuds tridimensionnels" (PDF). Revue de Mathématiques Pure et Appliquées (in French). 4: 5–20. MR   0131846.
  5. Levitt, Michael (1986). "Protein Folding by Restrained Energy Minimization and Molecular Dynamics". Journal of Molecular Biology . 170 (3): 723–764. CiteSeerX   10.1.1.26.3656 . doi:10.1016/s0022-2836(83)80129-6. PMID   6195346.
  6. 1 2 3 4 5 6 Klenin, Konstantin; Langowski, Jörg (2000). "Computation of writhe in modeling of supercoiled DNA". Biopolymers . 54 (5): 307–317. doi:10.1002/1097-0282(20001015)54:5<307::aid-bip20>3.0.co;2-y. PMID   10935971.
  7. Fuller, F. Brock (1971). "The writhing number of a space curve". Proceedings of the National Academy of Sciences of the United States of America . 68 (4): 815–819. Bibcode:1971PNAS...68..815B. doi: 10.1073/pnas.68.4.815 . MR   0278197. PMC   389050 . PMID   5279522.

Further reading