Fact-checked by Grok 2 weeks ago

Persistence length

The persistence length is a key parameter in polymer physics that characterizes the stiffness of semiflexible polymer chains, defined as the distance along the chain over which the orientation of the local tangent vector decorrelates exponentially due to thermal fluctuations. In the worm-like chain (WLC) model, a continuum description of such polymers, it is mathematically expressed as l_p = \frac{\kappa}{k_B T}, where \kappa is the bending rigidity, k_B is Boltzmann's constant, and T is the absolute temperature. Physically, it represents the effective length of a rigid rod segment: chains behave as straight rods on scales shorter than l_p, while exhibiting random coil-like flexibility on longer scales. The concept originates from the Kratky-Porod model of , which introduced the WLC as an interpolation between rigid rods and freely jointed chains, capturing the continuous bending energy E = \frac{\kappa}{2} \int_0^L \left( \frac{\partial \mathbf{t}(s)}{\partial s} \right)^2 ds, where \mathbf{t}(s) is the unit at s and L is the contour length. The persistence length emerges from the of the \langle \mathbf{t}(s) \cdot \mathbf{t}(0) \rangle = e^{-s / l_p}, quantifying how competes with elastic resistance to bending. In , it influences key observables like the end-to-end distance and ; for instance, in the high-force stretching regime of the WLC, the fractional extension is approximated as \frac{\langle R \rangle}{L} \approx 1 - \frac{1}{4} \sqrt{\frac{k_B T}{f l_p}}, where f is the applied force. Persistence length plays a crucial role in biophysics, particularly for biopolymers like DNA, where l_p \approx 50 nm (about 150 base pairs) in physiological conditions, enabling the modeling of chromatin packaging and supercoiling. It is also essential for cytoskeletal filaments, such as actin (l_p \approx 18 μm), which determines their buckling resistance and role in cell motility, and microtubules (l_p \approx 5.2 mm), critical for intracellular transport. Experimental measurements often involve atomic force microscopy, optical tweezers, or fluorescence microscopy to fit force-extension curves or analyze fluctuation spectra, providing insights into material properties under biological conditions.

Introduction

Definition

The persistence length, denoted as l_p, is defined as the characteristic length scale over which correlations in the tangent vectors of a polymer chain decay exponentially. Physically, this means that on length scales shorter than l_p, the chain behaves like a due to its inherent , whereas on longer scales exceeding l_p, it transitions to behaving like a flexible , allowing significant and . The persistence length is directly related to the chain's bending rigidity \kappa through the relation l_p = \frac{\kappa}{k_B T}, where k_B is Boltzmann's constant and T is the absolute temperature; this expression arises in the worm-like chain model as a measure of thermal fluctuations opposing bending. It is typically expressed in nanometers for biopolymers (e.g., approximately 50 nm for double-stranded DNA) or in angstroms for synthetic polymers (e.g., around 6-7 Å for polyethylene glycol). In contrast to the contour length L, which is the total end-to-end length of the fully extended , l_p specifically characterizes the local independent of overall size.

Historical Context

Early statistical theories of chains, such as Werner Kuhn's model for flexible molecules in , provided foundational ideas on conformation, local , and elasticity in the context of , where chains were modeled as random coils with restricted rotations around bonds. This work laid the groundwork for quantifying how chain rigidity influences macroscopic properties like and elasticity, transitioning from rigid models to more realistic flexible structures. The persistence length was explicitly defined and refined in 1949 by Otto Kratky and Giovanni Porod, who applied it to describe the semi-flexible nature of polymer chains through analysis of X-ray scattering in solutions. Their model distinguished scattering regions corresponding to rigid and flexible behaviors, with the persistence length emerging as the characteristic distance over which the chain direction remains correlated, directly linking microscopic bending rigidity to observable scattering laws. This refinement marked a key milestone, enabling quantitative assessment of chain stiffness in experimental data from synthetic polymers. The adoption of the persistence length extended to biophysics during the 1960s and 1970s, as researchers applied it to model the semi-flexible conformations of biological macromolecules like DNA and proteins, building on hydrodynamic and light-scattering measurements to characterize their bending properties. This period saw the evolution from the fully flexible freely jointed chain (FJC) models—pioneered by Kuhn for highly compliant chains—to semi-flexible frameworks that incorporated persistence length as a parameter for local bending rigidity, better capturing the intermediate stiffness of biopolymers. Key contributions included hydrodynamic studies that validated the concept for DNA, emphasizing its role in solution behavior. In the post-1990s era, the persistence length gained prominence in single-molecule biophysics through techniques like optical tweezers, which directly probed chain mechanics and integrated the concept into force-extension analyses of DNA, confirming its utility in describing semi-flexible polymer dynamics under tension.

Theoretical Foundations

Worm-Like Chain Model

The worm-like chain (WLC) model, introduced by Kratky and Porod in 1949, serves as the foundational theoretical framework for describing the conformational statistics of semi-flexible polymers, bridging the behaviors of rigid rods and flexible Gaussian chains. This model treats the polymer as a continuous, inextensible space curve parameterized by arc length s along its contour, with uniform bending rigidity and neglect of torsional effects. Thermal fluctuations in the chain configuration arise from the competition between this rigidity and entropic tendencies, governed by the bending energy in the Hamiltonian: H = \frac{\kappa}{2} \int_0^L \left( \frac{d \mathbf{t}}{ds} \right)^2 \, ds, where \mathbf{t}(s) is the unit tangent vector to the chain at position s, L is the contour length, and \kappa is the bending rigidity. The persistence length l_p = \kappa / k_B T, with k_B T the thermal energy, quantifies the chain's stiffness, representing the characteristic distance over which tangent orientations remain correlated. The statistical mechanics of the WLC is based on the chain as a space curve in thermal equilibrium, where the key parameters—contour length L, persistence length l_p, and effective Kuhn segment length b = 2 l_p—determine its overall properties. In the limit l_p \gg L, the chain behaves as a rigid rod with nearly straight configuration and end-to-end distance approaching L. Conversely, when l_p \ll L, it transitions to Gaussian chain behavior, resembling a freely jointed chain of segments of length b = 2 l_p. The orientation correlations decay exponentially as \langle \mathbf{t}(0) \cdot \mathbf{t}(s) \rangle = \exp(-s / l_p), reflecting the progressive loss of directional memory along the chain due to thermal bending. Despite its utility, the WLC model has limitations, as it neglects excluded volume interactions (self-avoidance) and treats the chain as continuous rather than discrete. This continuous approximation is valid primarily when l_p exceeds the size of individual monomers, ensuring that bending occurs smoothly over scales much larger than molecular subunits.

Mathematical Derivation

The persistence length l_p in the worm-like chain (WLC) model arises from the bending energy of the polymer, given by the Hamiltonian H_b = \frac{\kappa}{2} \int_0^L \left( \frac{\partial \mathbf{t}(s)}{\partial s} \right)^2 ds, where \mathbf{t}(s) is the unit tangent vector along the chain contour at arc length s, L is the contour length, and \kappa is the bending rigidity. Applying the to the Fourier modes of the vector fluctuations yields an average energy of \frac{1}{2} k_B [T](/page/Temperature) per mode, leading to the definition l_p = \frac{\kappa}{k_B [T](/page/Temperature)}, where k_B is Boltzmann's constant and T is the . The orientational of the vectors then follows from the Gaussian statistics of the chain configuration, resulting in the \langle \mathbf{t}(0) \cdot \mathbf{t}(s) \rangle = e^{-s / l_p}. This characterizes the decay of directional memory along the chain, with l_p setting the scale over which correlations persist. The mean-squared end-to-end distance \langle R^2 \rangle is obtained by integrating the tangent correlations: \langle \mathbf{R}^2 \rangle = \left\langle \left( \int_0^L \mathbf{t}(s) \, ds \right)^2 \right\rangle = 2 \int_0^L (L - s) \langle \mathbf{t}(0) \cdot \mathbf{t}(s) \rangle \, ds = 2 l_p L \left[ 1 - \frac{l_p}{L} \left( 1 - e^{-L / l_p} \right) \right]. This exact expression for the WLC model interpolates between limiting cases. In the flexible limit L \gg l_p, it reduces to \langle R^2 \rangle \approx 2 l_p L, resembling a with effective segment length $2 l_p. In the rigid rod limit L \ll l_p, it approaches \langle R^2 \rangle \approx L^2, as expected for a straight chain. For stretched chains under tension, the force-extension relation is approximated by the Marko-Siggia interpolation formula, derived from a saddle-point approximation to the partition function in the strong-stretching regime combined with low-force Gaussian behavior: F(x) \approx \frac{k_B T}{l_p} \left[ \frac{1}{4} \left(1 - \frac{x}{L}\right)^{-2} - \frac{1}{4} + \frac{x}{L} \right], where x is the end-to-end extension and F is the applied force. This formula accurately captures the nonlinear response for semi-flexible polymers like DNA over a wide range of extensions. The mean-squared radius of gyration \langle R_g^2 \rangle, which measures the spatial extent of the chain, is derived similarly by averaging over all pairwise distances along the contour: \langle R_g^2 \rangle = \frac{L l_p}{3} - l_p^2 + 2 \frac{l_p^3}{L} \left( 1 - e^{-L / l_p} \right). This expression, known as the Benoit-Doty formula, follows from double integration of the tangent correlations in the definition R_g^2 = \frac{1}{L^2} \int_0^L \int_0^L \frac{|\mathbf{r}(s) - \mathbf{r}(s')|^2}{2} \, ds \, ds'. In the flexible limit L \gg l_p, it yields \langle R_g^2 \rangle \approx \frac{L l_p}{3}, consistent with a Gaussian chain where R_g^2 = \langle R^2 \rangle / 6. In the rigid limit L \ll l_p, it approaches \langle R_g^2 \rangle \approx \frac{L^2}{12}, the value for a uniform rod.

Measurement Methods

Experimental Techniques

Atomic force microscopy (AFM) enables direct visualization of contours adsorbed onto a substrate, allowing researchers to fit the observed curvatures to the (WLC) model for persistence length extraction. In this approach, molecules are typically deposited on surfaces treated with divalent cations to promote adhesion, and the end-to-end distance distributions or tangent-tangent correlations are analyzed to yield persistence lengths around 50 nm for double-stranded under physiological conditions. Complementary methods within AFM imaging, such as polynomial fitting of traces, minimize measurement errors to below 0.4% for contour lengths, enhancing accuracy in curvature-based fits. Optical tweezers and magnetic tweezers provide single-molecule force-extension measurements, where polymer chains are stretched and the data fitted to the Marko-Siggia interpolation formula derived from the WLC model.78780-0.pdf) For double-stranded DNA, these techniques consistently report a persistence length of approximately 50 nm, reflecting its bending rigidity under low forces (below 1 pN). Optical tweezers use laser traps to manipulate micron-sized beads attached to chain ends, while magnetic tweezers employ superparamagnetic beads and external fields for gentler manipulations, both enabling torsional as well as extensional probing without significant hydrodynamic interference at dilute concentrations. Small-angle X-ray scattering (SAXS) and small-angle neutron scattering (SANS) characterize ensemble-averaged chain conformations in solution through analysis of scattering profiles in the low-q regime, where the Debye function modified for the WLC model extracts the persistence length from the apparent radius of gyration. SAXS has been applied to RNA and DNA, revealing persistence lengths that vary with folding states, such as a dramatic reduction from unfolded to folded conformations in group I ribozymes. SANS complements this for deuterated or contrast-matched samples, particularly in polymer brushes or micelles modeled as worm-like chains, yielding persistence lengths on the order of 10-20 nm for semi-flexible systems like cellulose derivatives. These scattering methods avoid surface effects but require careful correction for inter-chain interactions in semi-dilute solutions. Fluorescence microscopy tracks the or thermal fluctuations of labeled chains in solution, using end-to-end distance correlations or mean-squared displacements to infer persistence length via WLC statistics. For cytoskeletal filaments like , this technique measures rigidity by analyzing correlations from video images of fluorescently stained filaments, reporting persistence lengths of 10-17 μm. The method excels for studies of dynamic assemblies but demands high temporal resolution to capture fluctuations accurately.00319-8) Cyclization assays quantify the efficiency of loop formation in short chains, where the probability of intramolecular ligation inversely relates to persistence length through WLC cyclization j-factors. For DNA fragments of 200-350 base pairs, ligation yields fit to WLC models yield persistence lengths that decrease with temperature, from 53 nm at 5°C to 36 nm at 60°C, highlighting sequence and environmental influences. This approach is particularly useful for probing local stiffness in oligonucleotides but assumes negligible electrostatic effects at controlled ionic strengths. Key challenges in these experimental techniques include achieving dilute solutions to prevent aggregation and inter-chain interactions, which can artificially stiffen apparent persistence lengths. Sample preparation often involves surface adsorption artifacts in AFM or labeling-induced perturbations in fluorescence, while hydrodynamic interactions in scattering and tweezers methods require modeling corrections for accurate low-force regimes. For fragile or short chains, such as single-stranded DNA, stem-loop formations and structural heterogeneity further complicate measurements, necessitating multi-technique validation.

Computational Approaches

Molecular dynamics (MD) simulations provide a powerful computational framework for determining the persistence length of polymers by modeling chain conformations using all-atom or coarse-grained representations that incorporate bending potentials. In these simulations, the tangent-tangent correlation function \langle \mathbf{t}(0) \cdot \mathbf{t}(s) \rangle = e^{-s/l_p} is computed from equilibrated trajectories, allowing l_p to be fitted via exponential decay analysis. Coarse-grained MD models, which reduce computational cost while retaining essential stiffness parameters, have been used to predict l_p values for DNA on the order of 50 nm under varying ionic conditions. Monte Carlo (MC) methods offer an alternative stochastic approach to sample worm-like chain (WLC) configurations under the WLC Hamiltonian, generating ensemble statistics for validation of persistence length estimates. These simulations discretize the chain into segments with bond angles governed by a bending energy U(\theta) = -\kappa \cos\theta, enabling efficient exploration of conformational space to compute correlation functions and fit l_p. For semiflexible polymers, lattice-based MC simulations have demonstrated that l_p scales linearly with the bending rigidity parameter in the strong-stiffness limit, providing benchmarks for continuum models. Analytical approximations within the Kratky-Porod framework yield exact solutions for the WLC in two dimensions, where the end-to-end distance distribution can be derived closed-form, facilitating direct computation of l_p. In three dimensions, numerical integration techniques, such as path-integral methods or series expansions, are employed to approximate correlation functions and higher-order statistics due to the increased orientational complexity. Ab initio calculations using quantum mechanics derive the bending modulus \kappa for small oligomers, enabling prediction of l_p = \kappa / k_B T from first principles without empirical force fields. Density functional theory (DFT) applied to short peptide or DNA segments has yielded \kappa values consistent with experimental l_p for biological polymers, bridging quantum-scale interactions to mesoscopic stiffness. Software tools like LAMMPS facilitate MD simulations of polymer chains with customizable bending potentials to compute l_p from trajectory analysis. Custom WLC simulators, such as PolymerCpp, implement discrete WLC models for generating 3D chain ensembles and directly estimating persistence lengths via correlation fits. Computational approaches surpass experimental methods by providing access to extreme regimes, such as ultra-high forces or precise variations, where direct measurements are challenging or infeasible.

Applications

Synthetic Polymers

The persistence length plays a crucial role in classifying synthetic polymers as rod-like or coil-like, which directly influences their processing and performance characteristics such as melt viscosity and chain alignment under flow. Rod-like polymers, characterized by high persistence lengths on the order of tens of nanometers, exhibit limited flexibility and tend to align more readily, leading to anisotropic properties. For instance, poly(p-phenylene terephthalamide) (PPTA), the polymer in fibers, has a persistence length of approximately 29 nm, enabling efficient orientation during processing and contributing to high tensile strength. In contrast, coil-like polymers with low persistence lengths below 1 nm behave as flexible random walks, promoting higher melt viscosities due to entanglements but allowing easier flow in thermoplastics. exemplifies this with a persistence length of about 0.57 nm, resulting in significant chain coiling that affects its rheological behavior in extrusion processes. High persistence lengths in synthetic polymers significantly impact mechanical properties by favoring ordered structures over random entanglement networks. In polyaramids like PPTA, the elevated persistence length promotes the formation of liquid crystalline phases in concentrated solutions, which facilitates self-alignment and enhances and strength upon solidification into fibers. Conversely, low persistence lengths in flexible thermoplastics such as enable dense entanglement networks, which provide through energy dissipation but limit inherent . (PVC), with a persistence length of 1-2 nm determined via light scattering, illustrates an intermediate case where moderate supports applications in rigid plastics while allowing sufficient flexibility for processing. In polyelectrolytes, a subclass of synthetic polymers, the persistence length is modulated by electrostatic interactions, decreasing with increased charge screening from added salts that reduce repulsion between charged groups along the chain. This tunability arises from the additive nature of intrinsic and electrostatic contributions to the total persistence length, allowing control over chain extension in solution. Persistence length can be further tuned through chemical modifications, such as incorporating rigid side chains to increase backbone stiffness or copolymerization to adjust sequence and flexibility, as demonstrated in polyimides where varying hard segment rigidity improves optical and dielectric properties. Industrially, persistence length is pivotal in fiber spinning, where high values in rod-like polymers like PPTA enable gel spinning techniques to achieve ultra-high molecular orientation and tensile strengths exceeding 3 GPa in fibers. In nanocomposites, incorporating synthetic polymers with high persistence lengths, such as , as reinforcing agents enhances matrix tensile strength by promoting load transfer and alignment at the nanoscale, critical for applications in composites and protective materials.

Biological Systems

In biological systems, the persistence length characterizes the stiffness of such as nucleic acids and cytoskeletal filaments, influencing their conformational behavior and roles in cellular processes. For double-stranded B-DNA, the persistence length is approximately 50 nm under physiological conditions, which sets the scale for its semi-rigid structure and enables tight packaging into nucleosomes where about 147 base pairs wrap around cores. This stiffness also facilitates DNA looping over hundreds of base pairs, crucial for gene regulation by bringing distant regulatory elements into proximity. Cytoskeletal proteins exhibit much longer persistence lengths, reflecting their role in maintaining cellular architecture and enabling force transmission. filaments, key components of the , have a persistence length of about 17 μm in their ATP-bound state, providing the rigidity needed for , shape maintenance, and formation. , composed of dimers, possess an even greater persistence length of 1–5 mm, allowing them to act as tracks for intracellular transport via motor proteins like and while withstanding compressive forces during . Single-stranded , in contrast, is highly flexible with a persistence length of 1–2 nm, promoting compact folding into functional structures like ribozymes or ribosomal components through base-pairing interactions. Polysaccharides such as in the display a persistence length around 6–7 nm, contributing to and elasticity by forming extended, entangled networks. The persistence length of these biopolymers is modulated by environmental factors, particularly ionic strength, which alters electrostatic interactions. For DNA, low ionic strength increases the persistence length to over 100 nm due to enhanced repulsion between negatively charged phosphate backbones, while high salt screens these charges and reduces it toward the intrinsic value; similar effects occur with multivalent ions like Mg²⁺, which can decrease it to 25–30 nm. Protein binding, such as linker histones to DNA or regulatory proteins to actin, can further tune stiffness, adapting biopolymer conformations to cellular needs. These persistence lengths underpin critical biological functions by dictating how biopolymers assemble and respond to mechanical cues. In viral capsid assembly, the semi-rigid nature of DNA or RNA genomes (with persistence lengths enabling spooling without excessive bending energy) drives efficient packaging into protein shells, as seen in bacteriophages where electrostatics and polymer stiffness balance to form stable virions. For chromatin folding, DNA's 50 nm persistence length influences higher-order compaction into 30-nm fibers and loops, facilitating epigenetic regulation and genome organization within the nucleus. In cellular mechanotransduction, the high persistence lengths of actin (∼17 μm) and microtubules (1–5 mm) allow them to transmit forces from the extracellular matrix to the nucleus, integrating mechanical signals with signaling pathways for processes like cell migration and differentiation.

References

  1. [1]
    [PDF] Mechanics and statistics of the worm-like chain - Soft Math Lab
    The worm-like chain model is a continuum model for a flexible polymer under external force, used to calculate the average polymer extension.
  2. [2]
    The Persistence Length of Semiflexible Polymers in Lattice Monte ...
    Feb 10, 2019 · Usually, the stiffness of polymers is characterized by persistence length that can be understood as the distance over which the orientation of ...
  3. [3]
    The Persistence Length of DNA Is Reached from ... - PubMed Central
    The buckling persistence length P of a polymer is defined to be the Euler length corresponding to an RMS compression force acting at the ends of a rodlike ...
  4. [4]
    Measurement of the persistence length of cytoskeletal filaments ...
    Persistence length of a biological filament is an important quantitative descriptor of its resistance to bending, key to understanding the forces exerted on or ...
  5. [5]
    Über die Gestalt fadenförmiger Moleküle in Lösungen - Colloid and Polymer Science
    **Title:** Über die Gestalt fadenförmiger Moleküle in Lösungen
  6. [6]
    Röntgenuntersuchung gelöster Fadenmoleküle - Wiley Online Library
    Röntgenuntersuchung gelöster Fadenmoleküle. O. Kratky,. O. Kratky. (Institut ... First published: 1949. https://doi.org/10.1002/recl.19490681203. Citations ...
  7. [7]
  8. [8]
    Atomic Force Microscopy of Long and Short Double-Stranded ...
    Persistence length is a measure of the flexibility of DNA. Persistence length is proportional to the rigidity of a polymer chain and is a measure of the ...
  9. [9]
    Approaches for Determining DNA Persistence Length Using Atomic ...
    We describe ten complementary approaches for determining DNA persistence length by AFM imaging. The combination of different approaches provides increased ...
  10. [10]
    Accurate length determination of DNA molecules visualized by ...
    The best method for DNA length determination uses a polynomial fitting of degree 3 over a 5-point window, achieving less than 0.4% error.
  11. [11]
    [PDF] Stretching DNA - The Rockefeller University
    ABSTRACT: A statistical mechanical treatment of the wormlike chain model (WLC) is used to analyze experiments in which double-stranded DNA, tethered at one ...
  12. [12]
    Magnetic tweezers measurements of the nanomechanical properties ...
    We have used magnetic tweezers manipulation techniques, which allow us to measure the contour and persistence lengths together with the bending and torsional ...
  13. [13]
    Persistence Length Changes Dramatically as RNA Folds
    Dec 29, 2005 · In this Letter, we use SAXS data and theoretical results for the wormlike chain (WLC) to obtain l p for a 195 nucleotide group I ribozyme ...
  14. [14]
    Wormlike Chain Parameters of Cellulose and Cellulose Derivatives
    ... wormlike chain model. The persistence length q was calculated by four methods. The q values at the unperturbed state agreed closely with those derived using ...Missing: seminal | Show results with:seminal
  15. [15]
    Incorporating Intermicellar Interactions in the Fitting of SANS Data ...
    Determination of the Local Structure and the Persistence Length of Wormlike Micelles of Potassium Oleate by Small-Angle Neutron Scattering. Journal of ...
  16. [16]
    Measurement of the persistence length of polymerized actin using ...
    Sep 1, 1993 · The rigidity of actin filaments was measured by observing their Brownian movement using fluorescence microscopy and extracting the correlation ...
  17. [17]
    Temperature dependence of DNA persistence length
    The first approach was based on measuring the j-factors of short DNA fragments at various temperatures. Fitting the measured j-factors by the theoretical ...
  18. [18]
    Cyclization of short DNA fragments and bending fluctuations ... - PNAS
    During this analysis we assumed, regardless of the microscopic mechanism of bending, that DNA persistence length always equals 48 nm (1, 22, 24). In the ...
  19. [19]
    Measuring the Conformation and Persistence Length of Single ...
    Oct 23, 2018 · results show a larger persistence length for ssDNA than ordinary synthetic polymers, and the obtained length per base is larger than the ...Missing: units | Show results with:units<|control11|><|separator|>
  20. [20]
    A systematically coarse-grained model for DNA and its predictions ...
    Jan 21, 2010 · Persistence length, degree of stacking, and twist are studied by molecular dynamics simulation as functions of temperature, salt ...
  21. [21]
    A Molecular Dynamics Study of Mechanical and Conformational ...
    We investigate the mechanical and conformational behavior of P3AT thin films during deformation. The density profiles and measures of local mobility identify a ...
  22. [22]
    The Persistence Length of Semiflexible Polymers in Lattice Monte ...
    In this paper, theoretical persistence lengths of polymers with two different bending potentials were analyzed and examined by using lattice Monte Carlo ...Missing: history | Show results with:history
  23. [23]
    Mechanics and statistics of the worm-like chain - AIP Publishing
    Feb 1, 2018 · When a worm-like chain that is much longer than its persistence length is extended weakly by an external force, it behaves like a Gaussian chain ...
  24. [24]
    Sequence-Dependent Persistence Lengths of DNA - ACS Publications
    Thus, for the WLC, or any other isotropic model, the Flory persistence length (5) coincides with the original notion of Kratky–Porod persistence length, namely ...
  25. [25]
    LAMMPS Molecular Dynamics Simulator
    LAMMPS is a classical molecular dynamics code with a focus on materials modeling. It's an acronym for Large-scale Atomic/Molecular Massively Parallel Simulator.Download LAMMPS · LAMMPS documentation · LAMMPS Tutorials · Picture galleryMissing: persistence length
  26. [26]
    kmdouglass/PolymerCpp: 2D and 3D wormlike chain ... - GitHub
    PolymerCpp is a small program for generating three-dimensional wormlike chains (WLC), a common and relatively simple model in polymer physics.
  27. [27]
    Effects of tube persistence length on dynamics of mildly entangled ...
    Jul 1, 2012 · (For PE, the persistence lengths of polymer and tube are 5.7 and 16.4 Å at 140 °C [Fetters et al. (1994)], respectively.) The time dependent ...
  28. [28]
    Chapter 4: Rigid-chain polymers: Aromatic polyamides, heterocyclic ...
    Polymer conformation and persistence length. In the polymer chain of PPTA the benzene units are connected with amide bonds which prefer a trans conformation ...
  29. [29]
    Plot for determining persistence length of polystyrene (solid green...
    Plot for determining persistence length of polystyrene (solid green circles), poly(p-vinylbenzyl chloride) (open red triangles), and poly(vinyl chloride) (solid ...
  30. [30]
    Persistence length of polyelectrolytes with precisely located charges
    At high ionic strength, the total persistence length decreases for both macromolecules because the electrostatic repulsions between ionized groups are screened.
  31. [31]
    Tuning the persistence lengths of main chain towards colorless and ...
    Apr 9, 2024 · It was found that tuning the persistence lengths of hard segments of liquid-crystal-like PI can significantly improve the general properties of CPI films.
  32. [32]
    High Performance Fibers Based on Rigid and Flexible Polymers
    Structural Transformations in the Course of Gel Spinning of High-Strength Polymer Fibers ... Persistence length of the "rodlike" molecule poly(p-phenylene-trans ...<|control11|><|separator|>
  33. [33]
    Persistence Length and Nanomechanics of Random Bundles of ...
    Aug 6, 2025 · In addition, several synthetic polymers can be modeled as semiflexible WLCs [4]. Single-walled carbon nanotubes have a persistence length ...
  34. [34]
    Ionic effects on the elasticity of single DNA molecules - PNAS
    Multivalent ions lead to persistence lengths as low as 250–300 Å, well below the high-salt “fully neutralized” value of 450–500 Å characteristic of DNA in ...
  35. [35]
    Nucleosome packaging and nucleosome positioning of genomic DNA
    Taking the persistence length of DNA to be 50 nm and assuming that DNA in a nucleosome is uniformly bent in a circular trajectory at a radius of curvature ...
  36. [36]
    Allostery of actin filaments: Molecular dynamics simulations ... - PNAS
    The computed persistence lengths of F-actin composed of G-ATP (16 μm) and of G-ADP (8.5 μm) agree well with the experimental values for the two states.
  37. [37]
    Ionic strength-dependent persistence lengths of single-stranded ...
    Dec 27, 2011 · Published literature on the flexibilities of ssDNA and ssRNA reveals a rather large range of persistence lengths that span 10–60 Å, measured ...
  38. [38]
    Mechanics of DNA packaging in viruses - PNAS
    The objective of the current article is to respond to such measurements on viruses and to use the theory of elasticity and a simple model of charge and ...
  39. [39]
    Long-range compaction and flexibility of interphase chromatin in ...
    Nov 23, 2004 · Fiber flexibility can be described mathematically by polymer chain models, where its stiffness is quantified as the persistence length (Lp; refs ...