Worm-like chain

Last updated

The worm-like chain (WLC) model in polymer physics is used to describe the behavior of polymers that are semi-flexible: fairly stiff with successive segments pointing in roughly the same direction, and with persistence length within a few orders of magnitude of the polymer length. The WLC model is the continuous version of the KratkyPorod model.

Contents

Model elements

An illustration of the WLC model, with position and unit tangent vectors as shown. Worm-like Chain.png
An illustration of the WLC model, with position and unit tangent vectors as shown.

The WLC model envisions a continuously flexible isotropic rod. [1] [2] [3] This is in contrast to the freely-jointed chain model, which is only flexible between discrete freely hinged segments. The model is particularly suited for describing stiffer polymers, with successive segments displaying a sort of cooperativity: nearby segments are roughly aligned. At room temperature, the polymer adopts a smoothly curved conformation; at K, the polymer adopts a rigid rod conformation. [1]

For a polymer of maximum length , parametrize the path of the polymer as . Allow to be the unit tangent vector to the chain at point , and to be the position vector along the chain, as shown to the right. Then:

and the end-to-end distance . [1]

The energy associated with the bending of the polymer can be written as:

where is the polymer's characteristic persistence length, is the Boltzmann constant, and is the absolute temperature. At finite temperatures, the end-to end distance of the polymer will be significantly shorter than the maximum length . This is caused by thermal fluctuations, which result in a coiled, random configuration of the undisturbed polymer.

The polymer's orientation correlation function can then be solved for, and it follows an exponential decay with decay constant 1/P: [1] [3]

Mean squared end-to-end distance as a function of persistence length. Crossover from flexible to stiff behavior in the worm-like chain polymer model.svg
Mean squared end-to-end distance as a function of persistence length.

A useful value is the mean square end-to-end distance of the polymer: [1] [3]

Note that in the limit of , then . This can be used to show that a Kuhn segment is equal to twice the persistence length of a worm-like chain. In the limit of , then , and the polymer displays rigid rod behavior. [2] The figure to the right shows the crossover from flexible to stiff behavior as the persistence length increases.

Comparison between the worm-like chain model and experimental data from the stretching of l-DNA. Worm-like Chain Model vs. DNA.svg
Comparison between the worm-like chain model and experimental data from the stretching of λ-DNA.

Biological relevance

Experimental data from the stretching of Lambda phage DNA is shown to the right, with force measurements determined by analysis of Brownian fluctuations of a bead attached to the DNA. A persistence length of 51.35 nm and a contour length of 1560.9 nm were used for the model, which is depicted by the solid line. [4]

Other biologically important polymers that can be effectively modeled as worm-like chains include:

Stretching worm-like chain polymers

Upon stretching, the accessible spectrum of thermal fluctuations reduces, which causes an entropic force acting against the external elongation. This entropic force can be estimated from considering the total energy of the polymer:

.

Here, the contour length is represented by , the persistence length by , the extension is represented by , and external force is represented by .

Laboratory tools such as atomic force microscopy (AFM) and optical tweezers have been used to characterize the force-dependent stretching behavior of biological polymers. An interpolation formula that approximates the force-extension behavior with about 15% relative error is: [11]

A more accurate approximation for the force-extension behavior with about 0.01% relative error is: [4]

,

with , , , , , .

A simple and accurate approximation for the force-extension behavior with about 1% relative error is: [12]

Approximation for the extension-force behavior with about 1% relative error was also reported: [12]

Extensible worm-like chain model

The elastic response from extension cannot be neglected: polymers elongate due to external forces. This enthalpic compliance is accounted for the material parameter , and the system yields the following Hamiltonian for significantly extended polymers:

,

This expression contains both the entropic term, which describes changes in the polymer conformation, and the enthalpic term, which describes the stretching of the polymer due to the external force. Several approximations for the force-extension behavior have been put forward, depending on the applied external force. These approximations are made for stretching DNA in physiological conditions (near neutral pH, ionic strength approximately 100 mM, room temperature), with stretch modulus around 1000 pN. [13] [14]

For the low-force regime (F < about 10 pN), the following interpolation formula was derived: [15]

.

For the higher-force regime, where the polymer is significantly extended, the following approximation is valid: [16]

.

As for the case without extension, a more accurate formula was derived: [4]

,

with . The coefficients are the same as the one of the above described formula for the WLC model without elasticity.

Accurate and simple interpolation formulas for the force-extension and extension-force behaviors for the extensible worm-like chain model are: [12]

See also

Related Research Articles

<span class="mw-page-title-main">Inner product space</span> Generalization of the dot product; used to define Hilbert spaces

In mathematics, an inner product space is a real vector space or a complex vector space with an operation called an inner product. The inner product of two vectors in the space is a scalar, often denoted with angle brackets such as in . Inner products allow formal definitions of intuitive geometric notions, such as lengths, angles, and orthogonality of vectors. Inner product spaces generalize Euclidean vector spaces, in which the inner product is the dot product or scalar product of Cartesian coordinates. Inner product spaces of infinite dimension are widely used in functional analysis. Inner product spaces over the field of complex numbers are sometimes referred to as unitary spaces. The first usage of the concept of a vector space with an inner product is due to Giuseppe Peano, in 1898.

<span class="mw-page-title-main">Uncertainty principle</span> Foundational principle in quantum physics

In quantum mechanics, the uncertainty principle is any of a variety of mathematical inequalities asserting a fundamental limit to the accuracy with which the values for certain pairs of physical quantities of a particle, such as position, x, and momentum, p, can be predicted from initial conditions.

In mechanics, the virial theorem provides a general equation that relates the average over time of the total kinetic energy of a stable system of discrete particles, bound by a conservative force, with that of the total potential energy of the system. Mathematically, the theorem states

The Cauchy–Schwarz inequality is considered one of the most important and widely used inequalities in mathematics.

In physics, a Langevin equation is a stochastic differential equation describing how a system evolves when subjected to a combination of deterministic and fluctuating ("random") forces. The dependent variables in a Langevin equation typically are collective (macroscopic) variables changing only slowly in comparison to the other (microscopic) variables of the system. The fast (microscopic) variables are responsible for the stochastic nature of the Langevin equation. One application is to Brownian motion, which models the fluctuating motion of a small particle in a fluid.

<span class="mw-page-title-main">Polymer physics</span>

Polymer physics is the field of physics that studies polymers, their fluctuations, mechanical properties, as well as the kinetics of reactions involving degradation and polymerisation of polymers and monomers respectively.

<span class="mw-page-title-main">Helmholtz free energy</span> Thermodynamic potential

In thermodynamics, the Helmholtz free energy is a thermodynamic potential that measures the useful work obtainable from a closed thermodynamic system at a constant temperature (isothermal). The change in the Helmholtz energy during a process is equal to the maximum amount of work that the system can perform in a thermodynamic process in which temperature is held constant. At constant temperature, the Helmholtz free energy is minimized at equilibrium.

In quantum mechanics, perturbation theory is a set of approximation schemes directly related to mathematical perturbation for describing a complicated quantum system in terms of a simpler one. The idea is to start with a simple system for which a mathematical solution is known, and add an additional "perturbing" Hamiltonian representing a weak disturbance to the system. If the disturbance is not too large, the various physical quantities associated with the perturbed system can be expressed as "corrections" to those of the simple system. These corrections, being small compared to the size of the quantities themselves, can be calculated using approximate methods such as asymptotic series. The complicated system can therefore be studied based on knowledge of the simpler one. In effect, it is describing a complicated unsolved system using a simple, solvable system.

In mathematics, the Hodge star operator or Hodge star is a linear map defined on the exterior algebra of a finite-dimensional oriented vector space endowed with a nondegenerate symmetric bilinear form. Applying the operator to an element of the algebra produces the Hodge dual of the element. This map was introduced by W. V. D. Hodge.

In mathematics, spectral theory is an inclusive term for theories extending the eigenvector and eigenvalue theory of a single square matrix to a much broader theory of the structure of operators in a variety of mathematical spaces. It is a result of studies of linear algebra and the solutions of systems of linear equations and their generalizations. The theory is connected to that of analytic functions because the spectral properties of an operator are related to analytic functions of the spectral parameter.

<span class="mw-page-title-main">Equipartition theorem</span> Theorem in classical statistical mechanics

In classical statistical mechanics, the equipartition theorem relates the temperature of a system to its average energies. The equipartition theorem is also known as the law of equipartition, equipartition of energy, or simply equipartition. The original idea of equipartition was that, in thermal equilibrium, energy is shared equally among all of its various forms; for example, the average kinetic energy per degree of freedom in translational motion of a molecule should equal that in rotational motion.

The fluctuation–dissipation theorem (FDT) or fluctuation–dissipation relation (FDR) is a powerful tool in statistical physics for predicting the behavior of systems that obey detailed balance. Given that a system obeys detailed balance, the theorem is a proof that thermodynamic fluctuations in a physical variable predict the response quantified by the admittance or impedance of the same physical variable, and vice versa. The fluctuation–dissipation theorem applies both to classical and quantum mechanical systems.

In physics, the S-matrix or scattering matrix relates the initial state and the final state of a physical system undergoing a scattering process. It is used in quantum mechanics, scattering theory and quantum field theory (QFT).

<span class="mw-page-title-main">LSZ reduction formula</span> Connection between correlation functions and the S-matrix

In quantum field theory, the Lehmann–Symanzik–Zimmerman (LSZ) reduction formula is a method to calculate S-matrix elements from the time-ordered correlation functions of a quantum field theory. It is a step of the path that starts from the Lagrangian of some quantum field theory and leads to prediction of measurable quantities. It is named after the three German physicists Harry Lehmann, Kurt Symanzik and Wolfhart Zimmermann.

In polymer chemistry, an ideal chain is the simplest model to describe polymers, such as nucleic acids and proteins. It assumes that the monomers in a polymer are located at the steps of a hypothetical random walker that does not remember its previous steps. By neglecting interactions among monomers, this model assumes that two monomers can occupy the same location. Although it is simple, its generality gives insight about the physics of polymers.

The Debye–Waller factor (DWF), named after Peter Debye and Ivar Waller, is used in condensed matter physics to describe the attenuation of x-ray scattering or coherent neutron scattering caused by thermal motion. It is also called the B factor, atomic B factor, or temperature factor. Often, "Debye–Waller factor" is used as a generic term that comprises the Lamb–Mössbauer factor of incoherent neutron scattering and Mössbauer spectroscopy.

<span class="mw-page-title-main">Kuhn length</span>

The Kuhn length is a theoretical treatment, developed by Hans Kuhn, in which a real polymer chain is considered as a collection of Kuhn segments each with a Kuhn length . Each Kuhn segment can be thought of as if they are freely jointed with each other. Each segment in a freely jointed chain can randomly orient in any direction without the influence of any forces, independent of the directions taken by other segments. Instead of considering a real chain consisting of bonds and with fixed bond angles, torsion angles, and bond lengths, Kuhn considered an equivalent ideal chain with connected segments, now called Kuhn segments, that can orient in any random direction.

Photon polarization is the quantum mechanical description of the classical polarized sinusoidal plane electromagnetic wave. An individual photon can be described as having right or left circular polarization, or a superposition of the two. Equivalently, a photon can be described as having horizontal or vertical linear polarization, or a superposition of the two.

Rubber elasticity refers to a property of crosslinked rubber: it can be stretched by up to a factor of 10 from its original length and, when released, returns very nearly to its original length. This can be repeated many times with no apparent degradation to the rubber. Rubber is a member of a larger class of materials called elastomers and it is difficult to overestimate their economic and technological importance. Elastomers have played a key role in the development of new technologies in the 20th century and make a substantial contribution to the global economy. Rubber elasticity is produced by several complex molecular processes and its explanation requires a knowledge of advanced mathematics, chemistry and statistical physics, particularly the concept of entropy. Entropy may be thought of as a measure of the thermal energy that is stored in a molecule. Common rubbers, such as polybutadiene and polyisoprene, are produced by a process called polymerization. Very long molecules (polymers) are built up sequentially by adding short molecular backbone units through chemical reactions. A rubber polymer follows a random, zigzag path in three dimensions, intermingling with many other rubber molecules. An elastomer is created by the addition of a few percent of a cross linking molecule such as sulfur. When heated, the crosslinking molecule causes a reaction that chemically joins (bonds) two of the rubber molecules together at some point. Because the rubber molecules are so long, each one participates in many crosslinks with many other rubber molecules forming a continuous molecular network. As a rubber band is stretched, some of the network chains are forced to become straight and this causes a decrease in their entropy. It is this decrease in entropy that gives rise to the elastic force in the network chains.

<span class="mw-page-title-main">Kicked rotator</span>

The kicked rotator, also spelled as kicked rotor, is a paradigmatic model for both Hamiltonian chaos and quantum chaos. It describes a free rotating stick in an inhomogeneous "gravitation like" field that is periodically switched on in short pulses. The model is described by the Hamiltonian

References

  1. 1 2 3 4 5 Doi and Edwards (1988). The Theory of Polymer Dynamics.
  2. 1 2 Rubinstein and Colby (2003). Polymer Physics.
  3. 1 2 3 4 Kirby, B.J. Micro- and Nanoscale Fluid Mechanics: Transport in Microfluidic Devices.
  4. 1 2 3 4 Bouchiat, C (1999). "Estimating the Persistence Length of a Worm-Like Chain Molecule from Force-Extension Measurements". Biophysical Journal. 76 (1): 409–413. Bibcode:1999BpJ....76..409B. doi: 10.1016/S0006-3495(99)77207-3 . PMC   1302529 . PMID   9876152.
  5. J. A. Abels and F. Moreno-Herrero and T. van der Heijden and C. Dekker and N. H. Dekker (2005). "Single-Molecule Measurements of the Persistence Length of Double-Stranded RNA". Biophysical Journal. 88 (4): 2737–2744. Bibcode:2005BpJ....88.2737A. doi:10.1529/biophysj.104.052811. PMC   1305369 . PMID   15653727.
  6. Bernard, Tinland (1997). "Persistence Length of Single-Stranded DNA". Macromolecules. 30 (19): 5763. Bibcode:1997MaMol..30.5763T. doi:10.1021/ma970381+.
  7. Chen, Huimin; Meisburger, Steve P. (2011). "Ionic strength-dependent persistence lengths of single-stranded RNA and DNA". PNAS. 109 (3): 799–804. Bibcode:2012PNAS..109..799C. doi: 10.1073/pnas.1119057109 . PMC   3271905 . PMID   22203973.
  8. L. J. Lapidus and P. J. Steinbach and W. A. Eaton and A. Szabo and J. Hofrichter (2002). "Single-Molecule Effects of Chain Stiffness on the Dynamics of Loop Formation in Polypeptides. Appendix: Testing a 1-Dimensional Diffusion Model for Peptide Dynamics". Journal of Physical Chemistry B. 106: 11628–11640. doi:10.1021/jp020829v.
  9. Gittes, F (1993). "Flexural rigidity of microtubules and actin filaments measured from thermal fluctuations in shape". Journal of Cell Biology. 120 (4): 923–934. doi:10.1083/jcb.120.4.923. PMC   2200075 . PMID   8432732.
  10. Khalil, A. S.; Ferrer, J. M.; Brau, R. R.; Kottmann, S. T.; Noren, C. J.; Lang, M. J.; Belcher, A. M. (2007). "Single M13 bacteriophage tethering and stretching". Proceedings of the National Academy of Sciences. 104 (12): 4892–4897. doi: 10.1073/pnas.0605727104 . ISSN   0027-8424. PMC   1829235 . PMID   17360403.
  11. Marko, J.F.; Siggia, E.D. (1995). "Statistical mechanics of supercoiled DNA". Physical Review E. 52 (3): 2912–2938. Bibcode:1995PhRvE..52.2912M. doi:10.1103/PhysRevE.52.2912. PMID   9963738.
  12. 1 2 3 Petrosyan, R. (2016). "Improved approximations for some polymer extension models". Rehol Acta. 56: 21–26. arXiv: 1606.02519 . doi:10.1007/s00397-016-0977-9. S2CID   100350117.
  13. Wang, Michelle D.; Hong Yin; Robert Landick; Jeff Gelles; Steven M. Block (1997). "Stretching DNA with Optical Tweezers". Biophysical Journal. 72 (3): 1335–1346. Bibcode:1997BpJ....72.1335W. doi:10.1016/S0006-3495(97)78780-0. PMC   1184516 . PMID   9138579.
  14. Murugesapillai, Divakaran; McCauley, Micah J.; Maher, L. James; Williams, Mark C. (2017). "Single-molecule studies of high-mobility group B architectural DNA bending proteins". Biophysical Reviews. 9 (1): 17–40. doi:10.1007/s12551-016-0236-4. PMC   5331113 . PMID   28303166.
  15. Marko, J.F.; Eric D. Siggia (1995). "Stretching DNA". Macromolecules. 28 (26): 8759–8770. Bibcode:1995MaMol..28.8759M. doi:10.1021/ma00130a008.
  16. Odijk, Theo (1995). "Stiff Chains and Filaments under Tension". Macromolecules. 28 (20): 7016–7018. Bibcode:1995MaMol..28.7016O. doi:10.1021/ma00124a044.

Further reading