Fact-checked by Grok 2 weeks ago

Polymer physics

Polymer physics is the branch of physics that examines the structure, conformations, dynamics, and physical properties of polymers, which are long-chain macromolecules consisting of many repeating monomeric units connected by covalent bonds. These materials, ranging from synthetic plastics like polyethylene to biological molecules such as DNA, exhibit unique behaviors due to their large size and flexibility, with the degree of polymerization N often spanning from 10 to $10^9. The field integrates statistical mechanics, thermodynamics, and soft matter physics to understand how polymer chains interact in various states, including solutions, melts, and solids. At the core of polymer physics lies the study of single-chain conformations and dimensions, often modeled using the or approximation, where non-adjacent monomers have no interactions beyond . In this model, the mean squared end-to-end distance scales as \langle R^2 \rangle = N b^2, with b as the length and scaling exponent \nu = 1/2, reflecting a of 2. Real chains deviate from ideality due to effects or attractions, leading to swollen (\nu \approx 3/5) or collapsed configurations in good or poor solvents, respectively, as described by Flory theory. These scaling relations govern the R \sim N^\nu and critical concentrations for entanglement, such as c^* \sim N^{1-3\nu}. Polymer physics also addresses collective behaviors in multi-chain systems, including thermodynamics of solutions, blends, and separations. In melts and concentrated solutions, chains form entangled networks exhibiting , where and dynamic responses arise from or Rouse modes. Solid-state properties encompass glassy, crystalline, and rubbery s, with entropic elasticity dominating in rubbers and microphase separation in copolymers enabling into nanostructures. Applications span , from designing conductive polymers to biomolecular simulations, highlighting structure-function relationships.

Fundamentals

Definition and Scope

Polymer physics is the branch of physics that investigates the physical properties, structure, and dynamic behavior of polymers—long-chain macromolecules composed of repeating monomeric units covalently linked together—primarily through the application of and continuum theories. This field addresses phenomena at multiple scales, from the conformational statistics of individual chains to the collective behaviors of polymer ensembles in solutions and melts, including phase transitions such as glass transitions and . Its scope extends to practical applications in , enabling the design of s with specific mechanical, thermal, and optical properties for uses in plastics, elastomers, and biomedical devices. Key historical milestones include Paul J. Flory's foundational work in 1941–1942, where he developed a statistical mechanical framework for the of high polymer solutions, deriving expressions for and chain configurations that explained non-ideal behaviors in dilute and concentrated regimes. In the 1970s, advanced the field by introducing concepts to describe polymer dynamics and conformations, linking polymer statistics to and establishing simple scaling laws for chain behavior in solutions and melts, as elaborated in his 1979 monograph. Polymer physics is crucial for elucidating in polymer melts and solutions, where time-dependent responses arise from chain entanglements and relaxations, as well as , which stems from entropic restoring forces in crosslinked networks. These insights underpin the understanding of polymer solutions, where chain-solvent interactions govern and , and melts, where topological constraints dictate flow properties.

Basic Polymer Architecture

Polymers are composed of repeating structural units known as monomers, which are linked together by strong covalent bonds to form a continuous backbone . These monomers typically consist of a central backbone atom or sequence, often carbon-based in synthetic polymers, with pendant side groups that can vary in size and chemistry, influencing the polymer's overall . The covalent bonding along the backbone provides mechanical strength and flexibility, distinguishing polymers from smaller molecules. Polymer chains exhibit diverse topologies, including linear structures where monomers form a single continuous backbone without branches, branched architectures featuring side chains emanating from the main chain at various points, and cross-linked networks where multiple chains are interconnected via covalent bonds between side groups or backbones. The degree of polymerization, denoted as N, quantifies the average number of repeating units in a chain and serves as a primary metric for chain length. The contour length L of a polymer chain represents its maximum end-to-end extension when fully stretched, given by L = N l, where l is the length of each segment along the backbone. In polymer physics, the statistical segment length, often termed the b, provides a conceptual for describing chain flexibility on larger scales, equivalent to an effective rigid whose length accounts for local conformational freedom beyond the monomer scale. Synthetic polymers, such as polyethylene, typically feature simple linear or branched carbon backbones with alkyl side groups, enabling scalable production and tunable properties through polymerization control. In contrast, biopolymers like DNA exhibit more complex architectures, including double-helical linear chains with nucleotide monomers featuring sugar-phosphate backbones and nitrogenous base side groups that enable specific pairing and folding. These architectural differences arise from biological synthesis versus chemical manufacturing, affecting rigidity and functionality in their respective environments.

Static Chain Models

Ideal Chains

In polymer physics, the ideal chain model provides the foundational statistical description for the conformational behavior of a in a dilute under conditions, where monomer-monomer interactions effectively cancel with solvent-monomer interactions, leading to no net long-range forces. This model treats the polymer as a chain of N rigid segments, each of length l, connected by freely rotating joints that allow independent orientation of successive segments without any restrictions from or torsional potentials. Crucially, it assumes the absence of volume exclusion effects between non-adjacent segments and neglects all long-range interactions, simplifying the chain to a non-interacting entity whose statistics are governed solely by . The probability distribution of the end-to-end vector \mathbf{R} for an ideal chain of large N follows a Gaussian form, derived from the central limit theorem applied to the summation of independent random bond vectors: P(\mathbf{R}) = \left( \frac{3}{2\pi N l^2} \right)^{3/2} \exp\left( -\frac{3 R^2}{2 N l^2} \right), where R = |\mathbf{R}|. This distribution implies that the mean-squared end-to-end distance is \langle R^2 \rangle = N l^2, reflecting the diffusive nature of the chain's configuration space, with the root-mean-square end-to-end distance scaling as R_{\text{rms}} = l \sqrt{N}. The Gaussian statistics emerge from the random flight model, originally formulated for polymer chains by considering the chain as a three-dimensional random walk of N steps, each of length l, where the position after N steps is the vector sum of uncorrelated displacements. A key structural parameter for the ideal chain is the radius of gyration R_g, which quantifies the spatial extent of the chain's mass distribution relative to its center of mass. For large N, it is given by R_g = \sqrt{ \frac{N l^2}{6} }, obtained by averaging the squared distances of all segments from the center of mass over the Gaussian ensemble. This yields R_g^2 = \langle R^2 \rangle / 6, establishing a direct relation between the overall chain size and its end-to-end metric. Under theta conditions, the characteristic size of the ideal chain—whether measured by R, R_{\text{rms}}, or R_g—exhibits Gaussian scaling R \sim N^{1/2}, a hallmark of unperturbed chain dimensions that serves as the baseline for understanding more complex polymer behaviors. Here, the segment length l corresponds to the effective Kuhn length derived from the underlying polymer architecture of covalently linked monomers.

Real Chains

Real polymer chains exhibit deviations from the fully flexible models due to local arising from angles and short-range interactions, leading to semi-flexible behaviors that influence their static configurations. The (WLC) model provides a foundational description for such semi-flexible polymers, treating the chain as a continuous, smooth curve with intrinsic bending rigidity rather than discrete rigid segments. This model, originally formulated as the Kratky-Porod model, incorporates an energy cost for deviations in the chain's local tangent direction, parameterized by a bending modulus B. A key quantity is the persistence length \xi = B / k_B T, which quantifies the chain's : it represents the distance along the contour over which the direction of the chain correlates exponentially with the k_B T. For a chain of contour length L, the mean-square end-to-end distance in the WLC model is given by \langle R^2 \rangle = 2 \xi L \left[ 1 - \frac{\xi}{L} \left( 1 - e^{-L/\xi} \right) \right], which captures the transition from rigid to flexible limits without invoking long-range interactions. In the rod-like regime, where L \ll \xi, the chain remains nearly straight, yielding R \approx L and Gaussian statistics break down due to dominant stiffness. As L increases beyond \xi, a crossover occurs to the coil-like regime (L \gg \xi), where \langle R^2 \rangle \approx 2 \xi L and the chain behaves akin to an ideal flexible chain with effective segment length proportional to \xi. This persistence length thus serves as a bridge between microscopic bond rigidity and macroscopic chain dimensions, essential for polymers like DNA or actin filaments where stiffness measurably perturbs ideal configurations. Even with stiffness, real chains in solution experience balanced short-range attractions and repulsions at Flory's theta point, a temperature \Theta where the second virial coefficient vanishes, allowing the ideal chain approximation to hold despite local effects. At \Theta, polymer-solvent interactions are balanced enthalpically such that the net second virial coefficient vanishes, with attractions countering excluded volume repulsions on short scales, preserving Gaussian statistics for sufficiently long chains. This condition, derived from mean-field considerations of pairwise interactions, marks the boundary between collapsed and expanded states influenced by temperature, though local stiffness modulates the precise location of \Theta.

Interaction Effects

Excluded Volume

In polymer physics, the excluded volume effect arises from short-range repulsive interactions between non-adjacent units along the chain, which prohibit segmental overlap and cause the polymer coil to swell relative to its configuration. This phenomenon is quantified by the excluded volume parameter v, defined as the over the pairwise potential v = \int \left(1 - e^{-U(\mathbf{r})/kT}\right) d\mathbf{r}, where U(\mathbf{r}) is the potential between two s separated by \mathbf{r}; for hard-core repulsions, v approximates the effective volume excluded by each pair. This parameter corresponds directly to the second virial coefficient in the describing the of dilute solutions. The of chains with is captured by the (SAW) model, in which the chain trajectory avoids revisiting lattice sites or overlapping in continuous space, leading to long-range correlations that alter the scaling of chain dimensions. In three dimensions, the root-mean-square end-to-end distance scales as R \sim N^\nu, where N is the number of monomers and the Flory exponent \nu \approx 0.588 characterizes the swollen coil; this value emerges from both high-precision simulations and field-theoretic calculations, exceeding the exponent of \nu = 1/2. A mean-field for SAW statistics was introduced by Flory, balancing the entropic elasticity of the chain against repulsive interactions in a variational expression F \approx \frac{R^2}{N l^2} + \frac{v N^2}{R^3}, where l is the segment length and the first term approximates the Gaussian chain while the second accounts for binary collisions proportional to density. Minimizing F with respect to R yields the scaling R \sim N^{3/5}, or \nu = 3/5 = 0.6, providing a simple estimate of the swelling effect that qualitatively captures experimental observations for flexible polymers in good solvents. Subsequent refinements using renormalization group (RG) techniques, treating the excluded volume as a perturbation in a field-theoretic framework, yield more precise exponents by resumming higher-order corrections and accounting for the fixed-point behavior of the interaction strength. These methods confirm \nu \approx 0.588 in three dimensions and establish universal scaling laws, including des Cloizeaux corrections to the leading asymptotic behavior, such as R \sim N^\nu (1 + c N^{-\omega} + \cdots), where \omega is the correction-to-scaling exponent approximately 0.68, enabling better fits to finite-chain data.

Solvent and Temperature Influences

The behavior of polymer chains in solution is profoundly influenced by the quality of the solvent and the temperature, which modulate the interactions between polymer segments and solvent molecules. The Flory-Huggins theory provides a foundational mean-field model for the of polymer-solvent mixtures, describing the of mixing through the Flory interaction parameter \chi, which quantifies the energetic preference between unlike contacts (polymer-solvent) versus like contacts (polymer-polymer or solvent-solvent). In this framework, \chi determines the solubility and conformational preferences of the polymer; specifically, the theta condition occurs when \chi = 0.5 at the theta temperature T_\theta, where polymer-solvent interactions balance the repulsive tendencies, leading to behavior akin to a with no net expansion or contraction. In good solvents, where \chi < 0.5, favorable polymer-solvent interactions dominate, causing the polymer chains to swell and adopt expanded conformations to maximize solvation, as the effective becomes positive and promotes chain repulsion. Conversely, in poor solvents with \chi > 0.5, unfavorable interactions lead to chain collapse into compact globules or, at higher concentrations, into polymer-rich and solvent-rich domains, minimizing contact between polymer and solvent. These solvent effects effectively tune the excluded volume parameter, shifting the balance between attractive and repulsive forces on the chain. The temperature dependence of \chi, often approximated as \chi \approx A + B/T where A and B are constants, introduces critical solution temperatures in polymer mixtures: the (UCST) marks the boundary below which mixing is favored for systems with exothermic interactions, while the (LCST) indicates immiscibility above it for endothermic cases, as predicted by the Flory-Huggins . In semi-dilute solutions, where polymer volume fraction \phi exceeds the overlap concentration, the solvent quality further manifests through the blob picture introduced by de Gennes, conceptualizing the solution as a network of correlation blobs of size \xi. Within each blob, chains behave as in dilute solution, but inter-blob interactions screen long-range effects; in good solvents, the correlation length scales as \xi \sim \phi^{-3/4}, reflecting the swollen nature of the blobs and leading to a concentration-dependent screening that alters overall solution properties without phase separation at moderate \phi. This scaling arises from the Flory exponent \nu \approx 3/5 in good solvents, ensuring that the blob size decreases with increasing \phi to maintain local ideality on larger scales.

Dynamic Behaviors

Chain Flexibility

Chain flexibility refers to the capacity of polymer segments to bend and reorient locally on characteristic time scales, governing the viscoelastic response of dilute polymer solutions where chains move independently without topological constraints. This local dynamics underpins phenomena such as and flow properties, distinguishing it from global chain motions in denser systems. In unentangled regimes, flexibility manifests through cooperative motions of segments, enabling the chain to explore conformational space while responding to and external forces. The model provides a foundational description of chain flexibility for unentangled polymers in dilute solutions, treating the chain as a composed of beads linked by entropic springs. modes of characterize the relaxation, with the relaxation time for the p-th mode scaling as \tau_p \sim p^{-2}, reflecting faster decay for higher-frequency (shorter wavelength) bends. The longest relaxation time, known as the time \tau_R, scales with the as \tau_R \sim N^2, determining the overall time for the chain to equilibrate its end-to-end vector. These scalings arise from balancing frictional and entropic restoring forces, as derived in the original formulation. Local bending modes are quantified using the b as the effective rigid segment over which induce without significant . The energetic penalty for is captured by the continuum expression E = \frac{[\kappa](/page/Kappa)}{2} \int_0^L \left( \frac{d^2 \mathbf{r}}{ds^2} \right)^2 ds, where [\kappa](/page/Kappa) is the rigidity, \mathbf{r}(s) is the chain position along length s, and L is the total length; this form highlights the quadratic cost of local changes. The relates to [\kappa](/page/Kappa) via b = 2[\kappa](/page/Kappa) / k_B T, setting the scale for flexible versus semiflexible behavior, with influencing these modes in real chains by modulating the onset of . In dilute limits, chain flexibility directly impacts hydrodynamic properties, as seen in the intrinsic viscosity [\eta] \sim R_g^3 / M, where R_g is the radius of gyration and M is the molecular weight; this scaling reflects the pervaded volume of the coiled chain, with the constant of proportionality \Phi \approx 2.9 \times 10^{23} mol^{-1} (in cgs units) from hydrodynamic theories. For ideal chains, this yields [\eta] \sim M^{1/2}, emphasizing how flexibility enlarges the effective size without excluded volume effects. As molecular weight increases beyond a critical value M_c \approx 10^4 - 10^5 g/mol (depending on polymer type), flexibility transitions to entangled dynamics, where interchain overlaps hinder independent bending and introduce longer relaxation times.

Reptation Dynamics

Reptation dynamics describes the constrained motion of chains in dense, entangled melts, where chains cannot move freely due to topological constraints from surrounding molecules. In the Doi-Edwards theory, a chain is envisioned as being confined within a virtual tube formed by the entanglements with neighboring chains, with the tube diameter d scaling as d \sim N_e^{1/2} l, where N_e is the number of segments between entanglement points and l is the length of a statistical segment. This confinement restricts lateral motion, forcing the chain to reptate—slithering curvilinearly along the tube's contour like a snake—until it disengages and renews the tube. The theory builds on earlier ideas by de Gennes, extending them to account for flexible chains in melts. The curvilinear diffusion coefficient along the is D_c \sim 1/N, where N is the total number of segments in the , reflecting the proportional to chain length. The time, \tau_\mathrm{rep}, required for complete tube disengagement, scales as \tau_\mathrm{rep} \sim N^3 / N_e, leading to a center-of-mass coefficient D \sim 1/N^2. Tube disengagement involves the chain's primitive path unraveling, while constraint release—arising from the motion of surrounding chains—allows partial tube renewal, accelerating relaxation. Local chain flexibility contributes to early-time by enabling small-scale curvilinear displacements within the . In the Doi-Edwards model with constraint release, the modulus in the plateau exhibits G(t) \sim 1/t^{1/2}, bridging the entanglement plateau to terminal flow. Experimental validation of dynamics is evident in rheological measurements of entangled linear polymers, where the zero-shear scales as \eta_0 \sim N^{3.4}, slightly higher than the predicted N^3 due to subtle effects like release and tube length fluctuations. This scaling holds for a wide range of molecular weights above the entanglement threshold, as observed in and melts. The plateau G_N^0 \sim \rho kT / N_e further confirms the entanglement density's role in dictating the viscoelastic response.

Model Examples

Freely Jointed Chain

The freely jointed (FJC) model describes a as a sequence of N rigid bonds, each of fixed \ell, connected at joints where the bond directions are completely independent and randomly oriented without any restrictions. This idealization assumes no energetic or steric correlations between consecutive bonds, capturing the essence of a fully flexible in the absence of long-range interactions. The model, originally developed to understand the elasticity of networks, provides a foundational framework for calculations in polymer physics. The end-to-end vector \mathbf{R} of the chain is the vector sum \mathbf{R} = \sum_{i=1}^N \mathbf{r}_i, where each \mathbf{r}_i is a bond vector of magnitude \ell. Due to the independence of the bond vectors, the mean-square end-to-end distance is exactly \langle R^2 \rangle = N \ell^2, as the cross terms \langle \mathbf{r}_i \cdot \mathbf{r}_j \rangle = 0 for i \neq j and \langle r_i^2 \rangle = \ell^2. This result highlights the random walk nature of the chain, where the squared displacement scales linearly with the number of steps. The characteristic ratio C_\infty, defined as the limiting value of \langle R^2 \rangle / (N \ell^2) for large N, equals 1 for the FJC, signifying no local stiffness or persistence beyond a single bond. Each bond vector \mathbf{r}_i has an orientation uniformly distributed over the surface of a of \ell, corresponding to equal probability for all directions in . The statistical properties of the chain, such as the distribution of \mathbf{R}, are derived using generating functions that account for the isotropic nature of these orientations, often through transforms or characteristic functions in the formulation. For large N, the end-to-end distance distribution approaches a Gaussian form, serving as a concrete realization of statistics. Despite its simplicity, the FJC model overlooks bond angle restrictions and torsional barriers inherent in real polymer backbones, leading to an underestimation of chain stiffness that is compensated in more advanced models incorporating local constraints.

Representations

In polymer physics, the spatial configuration of an ideal polymer chain is often represented as a three-dimensional on a , where each step corresponds to a segment of the chain with fixed length l. This analogy captures the statistical ensemble of possible conformations assumed by the chain under , treating successive bonds as uncorrelated vectors. The variance of each step is \sigma^2 = l^2, leading to the mean-squared end-to-end distance \langle R^2 \rangle = N \sigma^2 = N l^2 for a chain comprising N segments. This spatial model provides a foundational description for the size scaling of coils, emphasizing how the chain's overall extension grows with the of its length, R \sim \sqrt{N}. In the continuous limit, it aligns with the freely jointed chain model as a Gaussian distribution of end-to-end vectors. The approach originates from early statistical treatments that likened molecular bonds to independent displacements, enabling analytical tractability for chain statistics. Temporally, the motion of a in solution evokes , where the center-of-mass diffusion traces a over time. The mean-squared displacement of the center of mass satisfies \langle r^2(t) \rangle = 6 D t in three dimensions, with the self-diffusion constant D inversely proportional to the chain length, D \sim 1/N, reflecting the cumulative hydrodynamic drag on the segments. This scaling arises in the Rouse model, which idealizes the as harmonically linked beads undergoing uncorrelated forces from the . The properties of random walks in polymer representations are profoundly influenced by spatial dimensionality. In one and two dimensions, simple symmetric random walks are recurrent, returning to the (or any ) with probability one, which implies a high likelihood of self-intersections in low-dimensional confinements. In contrast, three-dimensional walks are transient, with return probability less than one, allowing chains to explore space more freely without inevitable recrossings. These dimensionality-dependent behaviors underpin variations in polymer configurational and stability across different spaces. Random walk models also connect directly to experimental probes like small-angle scattering, where the static structure factor for an ideal chain is described by the Debye function. This form factor, P(q) = \frac{2}{u^2} (e^{-u} + u - 1) with u = q^2 R_g^2 and radius of gyration R_g^2 = \langle R^2 \rangle / 6, quantifies the angular distribution of scattered intensity, decaying as q^{-2} at intermediate wavevectors q to reflect the Gaussian coil statistics. The function enables extraction of chain dimensions from , , or scattering data on dilute solutions.

References

  1. [1]
    [PDF] Introduction to polymer physics Lecture 1
    A polymer molecule is a chain of many elementary units called monomers, connected by covalent bonds. An ideal chain has no interactions between monomers except ...
  2. [2]
    Polymer Physics | Materials Science and Engineering
    This course presents the mechanical, optical, and transport properties of polymers with respect to the underlying physics and physical chemistry of polymers.
  3. [3]
    [PDF] Lecture-polymers.pdf - Physics
    Mar 8, 2021 · A polymer molecule is a chain consisting of many elementary units called monomers. These monomers are attached to each other by covalent bonds.
  4. [4]
    Polymer Physics | Stanford Online
    This course teaches the fundamentals of the equilibrium properties and dynamic behaviors of polymers.
  5. [5]
    Polymer physics across scales: Modeling the multiscale behavior of ...
    Dec 18, 2019 · Many materials applications involve physical processes that occur at length scales where the polymer chains exhibit conformation correlations ...
  6. [6]
    Thermodynamics of High Polymer Solutions - AIP Publishing
    A statistical mechanical treatment of high polymer solutions has been carried out on the basis of an idealized model, originally proposed by Meyer.
  7. [7]
    Scaling Concepts in Polymer Physics by Pierre-Gilles Gennes
    In this book, a noted physicist explains the radical changes that have taken place in this exciting and rapidly developing field.
  8. [8]
    [PDF] The Challenges Facing the Current Paradigm Describing ...
    Nov 2, 2023 · The main question raised in polymer physics remains: how do these long chains interact and move as a group when submitted to shear deformation ...
  9. [9]
    Molecular Theory of Rubber Elasticity | Polymer Journal - Nature
    Jan 1, 1985 · The theory establishes connections between network structure and elastic properties. All essential parameters are determined by the connectivity ...
  10. [10]
    [PDF] chapters 14/15: polymer structures, applications, & processing
    A polymer is a macromolecule (long molecules) built of small covalently bonded units called monomers (“mer” from the Greek word meros meaning part).
  11. [11]
    Basic Polymer Structure | MATSE 81: Materials In Today's World
    The four basic polymer structures are linear, branched, crosslinked, and networked. Linear polymers have long chains, while crosslinked polymers have chains  ...Missing: degree contour length statistical segment
  12. [12]
    [PDF] Ideal chains
    Ideal chains have no interactions between monomers far apart, even if they approach each other in space, and are a starting point in polymer physics.Missing: topologies cross-
  13. [13]
    [PDF] Polymers
    Consider that all linear polymer chains can be reduced to a step length and a free, universal joint. This is the Kuhn model, and the step length is called ...
  14. [14]
    2. Advanced Technology Applications | Polymer Science and ...
    Biopolymers Versus Synthetic Polymers. The volume of biopolymers in the world far exceeds that of synthetic macromolecules. Biological polymers include DNA, ...
  15. [15]
    [PDF] 04. Polymers I - DigitalCommons@URI
    May 5, 2017 · Biopolymers include nucleic acids, proteins, ad sugars. They are as- sembled organically. • Most synthetic polymers are carbon based. • ...Missing: polyethylene | Show results with:polyethylene
  16. [16]
    Principles of Polymer Chemistry by Paul J. Flory Jr. | Hardcover
    by Paul J. Flory Jr. Guest Lecturer in: Format. Hardcover$113.95. Add to cart. Open Access. Open Access. This work can be downloaded for non-commercial ...<|control11|><|separator|>
  17. [17]
    [PDF] Ideal Chain Statistics, Free Energy and Chain Deformation
    We write equations for the size of the pore in terms of the number of segments in the blob, and then estimate the length that the polymer occupies in the tube, ...
  18. [18]
    [PDF] Lecture 3: Ideal polymer chains - ETH Zürich
    We will use the FJC model to approximate the mean-square end-to-end distance for a polymer chain. First, consider a random walk on a one-dimensional (1D) ...
  19. [19]
    [PDF] 1 Single Polymer Chain - Wiley-VCH
    Section 1.2 shows that the. Gaussian distribution successfully describes the end-to-end distance of an ideal chain. Page 5. 1.2 Statistics of a Single Polymer ...
  20. [20]
  21. [21]
    [PDF] Mechanics and statistics of the worm-like chain - Soft Math Lab
    This tutorial utilizes a variety of mathematical and physi- cal approaches in an effort to explain the mechanical and probabilistic aspects of the worm-like ...Missing: seminal | Show results with:seminal
  22. [22]
    Breakdown of the Kratky-Porod wormlike chain model ... - IOP Science
    Sep 8, 2011 · This model is widely used as “standard model” for describing conformations and force vs. extension curves of stiff polymers.
  23. [23]
    [PDF] Nobel Lecture, December 11, 1974 - PAUL J. FLORY
    The concept of a chain molecule consisting of atoms covalently linked is as old as modern chemistry. It dates from the origins of the graphic formula introduced ...
  24. [24]
  25. [25]
    Statistical physics of the freely jointed chain | Phys. Rev. E
    Jun 1, 1996 · Using the method of constraints proposed by S. F. Edwards and A. G. Goodyear [J. Phys. A 5, 965 (1972); 5, 1188 (1972)], we do a complete ...Missing: seminal paper
  26. [26]
    [PDF] Chapter 1 Polymer Physics The Isolated Polymer Chain Random ...
    A random walk can be constructed more easily on a lattice where the number of choices in direction for each step are limited, and the step length is fixed, b ( ...
  27. [27]
    [PDF] Spatial Configuration of Macromolecular Chains
    The skeletal bonds of the molecular chain were thus likened to the steps in a random walk in three dimensions, the steps being uncorrelated one to another.
  28. [28]
    [PDF] RANDOM WALKS - Probability Theory in Leiden
    The random walk (Sn)n∈N0 is called recurrent when P(σ0 < ∞) = 1. Otherwise, it is called transient. We have seen in (1.56) that one-dimensional simple random ...
  29. [29]
    [PDF] FORM FACTORS FOR POLYMER SYSTEMS 1. THE DEBYE ...
    THE DEBYE FUNCTION FOR GAUSSIAN CHAINS. Polymer coils in theta solvents or in the melt state follow Gaussian chain statistics whereby the inter-monomer ...Missing: ideal | Show results with:ideal