Fact-checked by Grok 2 weeks ago

Excluded volume

Excluded volume is a fundamental concept in and that describes the finite spatial extent of particles or molecular segments, which prevents their centers from approaching closer than a certain due to steric repulsion, effectively excluding other particles from that volume. In the context of chains, this leads to long-range interactions between non-adjacent monomers, causing the chain to swell and adopt an expanded conformation in , deviating from the ideal Gaussian model where such exclusions are ignored. The idea of excluded volume was originally introduced by Werner Kuhn in 1934 in the context of chain molecule configurations in solution, and it was soon applied to macromolecules by Paul J. Flory, who in the 1940s and 1950s developed seminal theories incorporating these effects into polymer conformation. Flory's mean-field approach balances the entropic elasticity of the against the repulsive excluded volume interactions, predicting that the R_g scales as R_g \sim N^{\nu}, where N is the number of monomers and the Flory exponent \nu \approx 3/5 (or 0.6) in three dimensions for good solvents, compared to \nu = 1/2 for chains. This scaling reflects the self-avoiding nature of the polymer, where distant segments repel each other, reducing the probability of chain overlaps and resulting in a of approximately 5/3 for the . The magnitude of excluded volume effects depends critically on solvent quality, parameterized by the Flory-Huggins interaction parameter . In good solvents (), repulsive monomer-solvent interactions enhance chain swelling, while in poor solvents (), attractive forces lead to collapse into a compact globule with \nu \approx 1/3. At the theta point (, e.g., 34.6°C for in ), excluded volume repulsions are perfectly balanced by attractions, yielding behavior with no net swelling. In semidilute solutions, screening of long-range interactions by overlapping chains restores Gaussian scaling, forming a network of blobs where excluded volume is locally relevant but globally shielded. Beyond polymers, excluded volume influences phenomena such as , where it drives compact native structures, and the phase behavior of colloidal suspensions, but its most profound impact remains in describing the dilute and semidilute regimes of polymer solutions, underpinning applications in , , and . Experimental validation through scattering and neutron scattering confirms the predicted exponents, with \nu \approx 0.588 in three dimensions from theory and simulations refining Flory's approximation.

Fundamental Concepts

Definition and Physical Meaning

Excluded volume refers to the region of space surrounding the center of a particle within which the centers of other particles are forbidden from entering due to the finite physical size of the molecules, effectively reducing the available volume for particle motion in a system. This concept arises solely from the geometric constraints imposed by the non-zero extent of particles, preventing their overlap without invoking any energetic interactions. In dense systems, such as liquids or solutions, this inaccessible volume significantly influences the overall behavior by limiting the number of possible configurations. The idea of accounting for molecular volume in thermodynamic models traces back to , who in 1873 introduced a correction (the parameter b) in his to represent the effective volume occupied by gas molecules, laying the groundwork for understanding finite-size effects in fluids. The specific "excluded volume" was later formalized by Werner Kuhn in 1934, who applied it to describe the steric hindrance in solutions of chain-like molecules, marking a key development in . Physically, excluded volume can be analogized to billiard balls on a table, where each ball requires a "personal space" equivalent to its diameter around its center, such that the effective occupied volume exceeds the actual material volume of the balls and restricts their free movement. Unlike attractive forces in van der Waals interactions, which are energy-driven, excluded volume is a purely entropic effect stemming from geometric restrictions that reduce the system's configurational entropy without altering potential energy. In dilute gases, where particle density is low, excluded volume effects are negligible, allowing behavior to approximate reality closely. However, in liquids and dense phases, these effects dominate, enforcing efficient packing and contributing to phenomena like phase transitions and .

Mathematical Formulation

The excluded volume for two spherical particles with radii r_1 and r_2 interacting via a hard-core potential is the volume inaccessible to the center of one particle due to the presence of the other, given by v_{\text{ex}} = \frac{4}{3} \pi (r_1 + r_2)^3. This quantity arises from the geometric constraint that the centers of the particles cannot approach closer than r_1 + r_2. A more general formulation for arbitrary particle shapes and potentials employs the Mayer f-function, defined for the pair interaction as f_{ij}(\mathbf{r}) = e^{-\beta u_{ij}(\mathbf{r})} - 1, where \beta = 1/(k_B T), k_B is Boltzmann's constant, T is , and u_{ij}(\mathbf{r}) is the pairwise . For hard-core repulsions, u_{ij} \to \infty at short separations (specifically, for |\mathbf{r}| < r_1 + r_2), yielding f_{ij} \approx -1 within the excluded region and f_{ij} = 0 otherwise. The excluded volume then follows as the integral v_{\text{ex}} = -\int f_{ij}(\mathbf{r}) \, d\mathbf{r}, which reduces to the spherical case above for hard spheres. This formulation connects directly to the virial expansion of the equation of state for a classical fluid, where the pressure P is expressed as P/(k_B T) = \rho + B_2 \rho^2 + \cdots, with \rho the number density. The second virial coefficient is B_2 = \frac{1}{2} \int \left(1 - e^{-\beta u(r)}\right) d\mathbf{r}, equivalent to -\frac{1}{2} \int f_{ij}(\mathbf{r}) \, d\mathbf{r}. For hard-core potentials, $1 - e^{-\beta u(r)} = 1 inside the excluded region, leading to B_2 \approx \frac{1}{2} v_{\text{ex}}. In d-dimensional space, the excluded volume for hard spheres of diameter \sigma scales as v_{\text{ex}} \sim \sigma^d, reflecting the volume of the hypersphere defined by the closest-approach distance; the precise prefactor involves the gamma function via the d-dimensional ball volume formula. For simple liquids such as argon, empirical estimates from the van der Waals constant b \approx 0.032 L/mol yield an excluded volume per molecule approximately 4 times the actual molecular volume v_m = \frac{4}{3} \pi r^3, consistent with the hard-sphere approximation where b = 4 N_A v_m and N_A is Avogadro's number.

Applications in Fluid Systems

Hard Sphere Model

The hard sphere model serves as the paradigmatic system for studying excluded volume effects in dense fluids, where particles are idealized as smooth, impenetrable spheres of diameter \sigma. The interparticle potential is infinite for center-to-center separations less than \sigma (preventing overlaps) and zero otherwise, ensuring that interactions arise solely from geometric constraints rather than energetic attractions. This setup isolates the entropic contributions from excluded volume, making the model analytically tractable at low densities and amenable to simulations at higher ones. The key dimensionless is the packing fraction \eta = \frac{\pi}{6} \rho \sigma^3, where \rho is the number density, representing the volume fraction occupied by the spheres. Exact analytical treatments of the hard sphere model are limited, but Onsager's 1949 theory provides a seminal result for elongated hard bodies, such as rods, where excluded volume drives a first-order isotropic-to-nematic liquid crystalline phase transition. In this approach, the free energy is dominated by orientational entropy losses due to pairwise excluded volumes, predicting the transition at a critical density where nematic ordering maximizes configurational freedom despite the geometric constraints. Although originally formulated for rod-like particles, the theory highlights how shape anisotropy amplifies excluded volume effects beyond spherical symmetry, influencing applications in colloidal suspensions. For spherical particles, insights into the model's behavior at moderate to high densities come from simulations and approximate theories. Molecular dynamics simulations reveal a kinetic glass transition around \eta \approx 0.58, where dynamical arrest occurs due to caging by neighboring spheres, preventing long-time diffusion without crystallization. The Percus-Yevick approximation, an integral equation closure for the pair correlation function, yields an analytical equation of state via the compressibility route: P = \rho kT \frac{1 + \eta + \eta^2 - \eta^3}{(1 - \eta)^3}, which accurately captures pressure increases near contact but underestimates at higher \eta. These results underscore the model's utility in benchmarking numerical methods for dense packings. The phase diagram of the hard sphere fluid features a first-order fluid-solid transition driven entirely by excluded volume entropy maximization in the ordered phase. Freezing occurs at \eta \approx 0.494, where the fluid becomes unstable to crystallization into a face-centered cubic lattice, while melting happens at \eta \approx 0.545 upon compression of the solid. These coexistence points, determined from early molecular dynamics simulations, demonstrate how geometric packing efficiency governs the transition without thermal or energetic inputs. Beyond the solid phase, the model exhibits no further equilibria up to random close packing at \eta \approx 0.64. Despite its foundational role, the hard sphere model overlooks attractive interactions present in real fluids, necessitating perturbative additions like van der Waals terms for quantitative predictions in atomic or molecular systems. This limitation confines its exact applicability to idealized colloids or short-range repulsions, though it remains a cornerstone for understanding entropy-dominated phenomena.

Van der Waals and Mean-Field Approximations

In 1873, Johannes Diderik van der Waals introduced the concept of excluded volume in his doctoral thesis on the continuity of the gaseous and liquid states, quantifying it as the co-volume parameter b to account for the finite size of molecules in real gases and liquids. This parameter represents the volume excluded by intermolecular repulsions, modifying the ideal gas law to better describe deviations at high densities. Van der Waals' work laid the foundation for mean-field treatments of non-ideal fluids, emphasizing the role of molecular volume in preventing complete compression. The van der Waals equation of state incorporates excluded volume through a correction to the available volume: \left(P + \frac{a}{V_m^2}\right)(V_m - b) = RT, where V_m is the molar volume, a accounts for attractive forces, and b is the excluded volume per mole. For spherical molecules modeled as hard spheres, b = 4 N_A v_{m}, with v_{m} the actual volume of one molecule and N_A Avogadro's number; this factor of 4 arises empirically from the pairwise excluded volume between two touching spheres, which is eight times the molecular volume, but halved per molecule in the mean-field limit. The free volume correction appears as (V - N \beta)^N in the configurational partition function, where \beta = b / N_A is the excluded volume per molecule, approximating the system as one where molecules occupy a reduced space without overlap. From free volume theory, the statistical mechanical derivation treats the insertion probability of a test particle as approximately the unoccupied fraction, p \approx 1 - \beta \rho, where \rho = N/V is the number density and \beta = 4 v_m. The full excess chemical potential due to excluded volume is then \mu_{ex} = kT \left[ -\ln(1 - \beta \rho) + \frac{\beta \rho}{1 - \beta \rho} \right], obtained from the excess Helmholtz free energy A_{ex} = -N kT \ln(1 - \beta \rho), with the partition function Q incorporating the reduced volume (V - N \beta)^N / N!. This expression captures the entropic penalty of crowding at higher densities, bridging microscopic repulsions to macroscopic thermodynamics in the low-to-moderate density regime. In mean-field approximations, such as the random phase approximation (RPA), excluded volume effects are integrated into liquid structure via the , yielding the structure factor: S(k) = \frac{1}{1 - \rho \hat{c}(k)}, where \hat{c}(k) is the Fourier transform of the direct correlation function c(r), approximated as c(r) \approx -f(r) with f(r) = e^{-u(r)/kT} - 1 the . For pure excluded volume (hard-core repulsions), f(r) = -1 for r < \sigma (molecule diameter) and 0 otherwise, modeling pairwise exclusion without correlations beyond mean field. This framework extends the hard-sphere baseline to compressible fluids by combining repulsive contributions with averaged attractions. These approximations elucidate liquid compressibility, as the long-wavelength limit S(0) = \rho kT \kappa_T shows how excluded volume reduces \kappa_T (isothermal compressibility) near close packing, preventing divergence seen in ideal gases. Additionally, excluded volume shifts the liquid-gas critical point, with the van der Waals model predicting a critical molar density \rho_{m,c} = 1/(3b) and temperature T_c = 8a/(27 R b), where finite molecular size increases \rho_{m,c} relative to the point-particle limit and stabilizes the transition against thermal fluctuations. Such shifts are essential for understanding real fluid phase behavior, including supercritical compressibility anomalies.

Applications in Polymer Physics

Single-Chain Conformations

In the absence of excluded volume interactions, a single polymer chain adopts configurations resembling a Gaussian random walk, where the mean-square end-to-end distance scales as R^2 = N l^2, with N the number of segments and l the segment length; equivalently, the radius of gyration follows R_g^2 = N l^2 / 6. This ideal chain model, valid at the theta point where attractive and repulsive interactions balance, ignores the physical constraint that chain segments cannot occupy the same space, leading to unphysically compact statistics for long chains. Excluded volume effects, arising from the finite size of monomers, cause chain segments to repel each other, resulting in swollen conformations that deviate from Gaussian statistics. In three dimensions, the radius of gyration scales as R_g \sim N^\nu with Flory exponent \nu \approx 0.588, rather than the ideal \nu = 0.5; this value, refined from Flory's mean-field estimate of \nu = 3/5 = 0.60 via renormalization group theory, reflects the balance between entropic elasticity and repulsive interactions. The self-avoiding walk (SAW) provides a lattice-based model for these swollen chains, enforcing strict non-intersection of segments to mimic excluded volume. In two dimensions, exact enumeration yields \nu = 3/4, while in three dimensions, numerical simulations and series expansions confirm \nu \approx 0.588, consistent with continuum limits. This model captures the universal scaling behavior of isolated chains in good solvents, where long-range correlations prevent collapse. Perturbation theory addresses excluded volume through the Edwards equation, a diffusion-like equation for the chain's Green function \psi(\mathbf{r}, N), the probability density of finding the chain end at position \mathbf{r} after N steps: \frac{\partial \psi}{\partial N} = \frac{l^2}{6} \nabla^2 \psi - v \phi(\mathbf{r}) \psi(\mathbf{r}, N), where v > 0 is the excluded volume parameter quantifying pairwise repulsions and \phi(\mathbf{r}) is the self-consistent local monomer density field. For small v, this yields perturbative corrections to ideal statistics, such as a first-order swelling R_g^2 \approx (N l^2 / 6) (1 + (4/3) z), with z \propto v N^{1/2} / l^3 the crossover variable; larger v drives the system to the asymptotic swollen regime. Crossover behavior occurs as a function of or quality, transitioning from the theta point (v = 0, Gaussian chains) to good conditions (v > 0, swollen SAW-like statistics). At the theta point, second virial coefficients vanish, but higher-order attractions can induce collapse below it; above, the chain expands continuously, with scaling functions describing the smooth interpolation between regimes. Experimental verification comes from light scattering on dilute solutions, where the exponent deviates from 0.5 toward 0.588 in good solvents like , confirming excluded volume swelling for chains up to N \approx 10^4 monomers. These measurements, corrected for polydispersity, align closely with predictions rather than Flory's mean-field value.

Flory-Huggins Theory

The Flory-Huggins theory describes the of solutions using a model that inherently incorporates excluded volume effects by assigning one site per unit or molecule, thereby restricting site occupancy and preventing molecular overlap. Developed independently by and Maurice Huggins in 1942, the model treats polymers as chains of N segments on a regular with z, where molecules and polymer segments compete for sites, leading to reduced configurational freedom for longer chains. The core of the model lies in the Flory-Huggins interaction parameter \chi = \frac{z \Delta \epsilon}{[kT](/page/KT)}, where \Delta \epsilon represents the change in interaction energy for a polymer-solvent contact relative to solvent-solvent or polymer-polymer contacts, k is Boltzmann's constant, and T is ; this parameter quantifies energetic penalties or attractions beyond the pure volume exclusion enforced by the , with \chi > 0.5 favoring demixing. Excluded volume per , effectively the volume v, manifests in the entropy term by limiting the number of available configurations, particularly penalizing the placement of entire chains. Flory's 1942 contribution emphasized how this volume exclusion, combined with \chi, sets limits, explaining why polymers exhibit conditions or poor at high concentrations. The dimensionless Helmholtz free energy of mixing per lattice site is expressed as \frac{F}{kT} = \frac{\phi}{N} \ln \phi + (1 - \phi) \ln (1 - \phi) + \chi \phi (1 - \phi), where \phi is the polymer volume fraction; the logarithmic terms capture the mixing entropy diminished by excluded volume constraints on chain conformations, while the \chi term adds the mean-field enthalpy of interactions. This formulation predicts phase behavior through the common tangent construction on the free energy curve, yielding the binodal line that separates stable single-phase regions from two-phase coexistence. For typical \chi > 0, the theory exhibits upper critical solution temperature (UCST) behavior, with phase separation occurring below a critical temperature where \chi_c = \frac{1}{2} \left(1 + \frac{1}{\sqrt{N}}\right)^2; the critical polymer concentration is \phi_c = \frac{1}{1 + \sqrt{N}}, which approaches zero for large N, indicating that long polymers phase separate at arbitrarily low concentrations. This critical point marks the onset of instability, derived from the conditions where the second and third derivatives of the free energy with respect to \phi vanish. As a mean-field approach, Flory-Huggins neglects spatial correlations and concentration fluctuations, overestimating near the critical point and failing for short chains where excluded volume induces stronger swelling effects. These limitations are partially addressed by the (RPA), which incorporates Gaussian chain statistics to better capture long-wavelength fluctuations in long-chain systems, improving predictions of the spinodal and .

Extensions and Modern Developments

In Biological Macromolecules

In biological macromolecules, excluded volume effects play a crucial role in shaping the conformations and interactions of polymers such as DNA and proteins within the crowded cellular environment. For DNA, which can be modeled as a wormlike chain with persistence length l_p \approx 50 nm, excluded volume interactions cause chain swelling, preventing collapse and promoting an extended configuration essential for packaging in the nucleus. This swelling arises from repulsive interactions that limit chain self-overlap, leading to a characteristic deflection length \lambda \sim (v / l_p)^{1/3}, where v is the excluded volume parameter representing the effective volume per unit length excluded due to chain thickness and stiffness. In , excluded volume restricts the conformational space available to the polypeptide chain, providing a partial resolution to the Levinthal paradox by limiting the number of accessible states and guiding the chain toward the native structure through steric constraints rather than exhaustive . Models like the Gō model incorporate hard-core repulsions to simulate these excluded volume effects, ensuring that non-native contacts are penalized by repulsive potentials, which smooths the energy landscape and accelerates folding kinetics. Macromolecular crowding in cells, where macromolecules occupy a volume fraction \phi \approx 0.3, amplifies excluded volume effects by depleting available space, which hinders of unfolded proteins but enhances folding rates by favoring compact states through entropic stabilization. This crowding also induces , as demonstrated in 1990s studies showing that crowders promote the segregation of macromolecules into dense phases, altering and promoting aggregation or formation . Experimental evidence from (SAXS) on nucleosomes reveals that compaction is limited by excluded volume v_{ex}, as the core particles and impose steric barriers that prevent excessive folding, maintaining a between packaging efficiency and accessibility in fibers. Additionally, the in biological shapes—such as rigid rods (e.g., ) versus flexible coils (e.g., actin-associated polymers)—enhances orientational ordering under excluded volume interactions, leading to nematic-like alignment that facilitates cytoskeletal and .

In Soft Matter and Colloids

In physics, excluded volume effects play a central role in the behavior of colloidal suspensions, where micron-sized particles dispersed in a interact primarily through steric repulsion due to their finite size, preventing overlap and leading to purely forces. This interaction is often idealized as the model, in which particles behave as impenetrable spheres with no attractive potentials, making the volume fraction \phi the sole control parameter for thermodynamic properties. Such systems exhibit rich phase behavior driven by the maximization of configurational , as the excluded volume restricts accessible , favoring ordered structures at higher densities. Colloidal serve as a quintessential model for studying fundamental phenomena like , , and interfacial properties in , with experimental realizations using sterically stabilized polymethylmethacrylate (PMMA) particles in solvents. The of hard-sphere colloidal suspensions features a at low \phi, transitioning to a coexistence region of and face-centered cubic (FCC) at intermediate densities, and a fully crystalline at higher \phi. Experimental measurements by Pusey and Megen using on PMMA suspensions revealed the fluid- freezing transition at \phi_f \approx 0.494 and the - melting transition at \phi_m \approx 0.545, closely matching theoretical predictions from Monte Carlo simulations (freezing at ≈0.494 and melting at ≈0.545). At even higher densities, around \phi \approx 0.58, a colloidal glass forms, where particles become kinetically arrested despite remaining in a -like thermodynamic state, as predicted by mode-coupling theory. These transitions are governed exclusively by , which increases osmotic pressure and drives entropy gain through ordering, without reliance on temperature-dependent energetics. In binary or multicomponent colloidal mixtures, excluded volume induces effective attractions via the Asakura-Oosawa (AO) depletion mechanism, particularly when smaller depletant particles or polymers are present alongside larger colloids. The AO model treats depletants as ideal gas particles excluded from a thin shell (of thickness equal to the depletant radius R_d) around each colloid, leading to an osmotic pressure imbalance that attracts colloids when their surfaces approach within $2R_d. The resulting square-well-like potential has a depth of approximately -\frac{3}{2} \phi_d k_B T (where \phi_d is the depletant volume fraction) and range $2R_d, promoting phase separation, gelation, or clustering depending on \phi_d and the size ratio. This entropic attraction has been verified in experiments with sterically stabilized colloids and non-adsorbing polymers, influencing applications like colloidal stabilization and self-assembly in soft materials. Extensions of the AO model account for many-body effects and polydispersity, enhancing predictions for realistic soft matter systems such as emulsions or microgel suspensions. Beyond equilibrium phases, excluded volume governs dynamical properties in sheared colloidal suspensions, such as shear thickening, where hydrodynamic interactions couple with steric repulsion to increase viscosity at high \phi. In confined geometries, like slits or pores, excluded volume amplifies entropic forces, leading to and enhanced . These effects underscore the versatility of excluded volume as a design principle in , enabling tunable structures from photonic crystals to responsive materials without chemical bonds.

References

  1. [1]
    [PDF] 1 Introduction - MIT
    2.4 Excluded Volume Interaction. If a chain is in solution, because the monomers have a finite volume we expect. that chain will be self-repelling.
  2. [2]
    3/30/98 Chapter 1 Polymer Physics
    The expanded coil is larger than the Gaussian coil since excluded volume is a type of self-repulsion for the coil. In order to obtain a scaling law which ...
  3. [3]
    [PDF] Non-Ideal Chains: Size, Statistics, Free-Energy
    Poor solvents Excluded volume is negative due to large attractive interactions between the monomers, which prefer monomer-monomer contact strongly over monomer- ...
  4. [4]
    Generalized equation of state and exclusion factor ... - AIP Publishing
    If an approaching particle has volume v0 and radius r, the excluded volume is defined as the volume of a sphere of radius R⫹r. Expressing the radii through ...
  5. [5]
  6. [6]
    [PDF] Thermodynamics of d-dimensional hard sphere fluids confined to ...
    Mar 15, 2011 · We derive an analytical expression of the second virial coefficient of d-dimensional hard sphere fluids confined to slit pores by applying ...Missing: sigma^ | Show results with:sigma^
  7. [7]
    Phase diagram of the hard-sphere potential model in three and four ...
    Apr 22, 2025 · The HS and pHS models have only one relevant parameter, which is the filling or packing fraction, defined for any spatial dimension as.
  8. [8]
    [PDF] of colloidal particles - by lars onsager
    Introdzution. The shapes of colloidal particles are often reasonably com- pact, so that no diameter greatly exceeds the cube root of the volume of.Missing: original | Show results with:original
  9. [9]
    Hard spheres: crystallization and glass formation - Journals
    Dec 28, 2009 · This concentration (by weight) is then identified with the freezing volume fraction ϕF=0.494 found in accurate computer simulations of ...
  10. [10]
    Equation of State for Hard Spheres | The Journal of Chemical Physics
    Simple and exact expressions have been found for the compressibility and pressure equations of state predicted by the Percus—Yevick equation for hard spheres. ...
  11. [11]
    Phase Transition for a Hard Sphere System - AIP Publishing
    B. J. Alder, T. E. Wainwright; Phase Transition for a Hard Sphere System ... https://doi.org/10.1063/1.1743957. Article history. Received: August 12 1957.
  12. [12]
    [PDF] Johannes D. van der Waals - Nobel Lecture
    I intend to discuss in sequence: (1) the broad outlines of my equation of state and how I arrived at it;. (2) what my attitude was and still is to that equation ...
  13. [13]
    Taking Another Look at the van der Waals Equation of State–Almost ...
    Aug 2, 2019 · Repulsive part of vdW EoS, with the original covolume and an “effective co-volume” (half of the original covolume) compared to the Carnahan– ...
  14. [14]
    [PDF] 1 The Statistical Mechanical Derivation of the van der Waals ...
    Feb 23, 2005 · The first thermodynamic property that we will calculate is the pressure because that is what everyone associates with the van der Waals equation ...
  15. [15]
    [PDF] Nobel Lecture, December 11, 1974 - PAUL J. FLORY
    At the asymptote for chains of great length and large excluded volume ẞ, the exponent a should reach 0.80 according to the treatment given above. Although this ...<|separator|>
  16. [16]
    The statistical mechanics of polymers with excluded volume
    The probability distribution of the configurations of a polymer consisting of freely hinged links of length l and excluded volume v is studied.
  17. [17]
    [PDF] Polymer excluded volume exponent v: An experimental verification ...
    Jan 1, 1980 · la théorie critique (v = 0,588) que celle de Flory (v = 0,60). Ceci est confirmé par les mesures de A2. Abstract. 2014 Three light scattering ...
  18. [18]
    Self-avoiding walks - ScienceDirect.com
    The “true” self-avoiding walk is defined as the statistical problem of a traveller who steps randomly, but tries to avoid places he has already visited.
  19. [19]
    Crossover from the theta to the Excluded Volume Single Chain ...
    Hydrodynamic radius and intrinsic viscosity of polystyrene in the crossover region from θ to good‐solvent conditions. Journal of Polymer Science: Polymer ...
  20. [20]
    Thermodynamics of High Polymer Solutions - AIP Publishing
    Paper presented by Dr. Huggins before the Wilder D. Bancroft Colloid ... This content is only available via PDF. Open the PDF for in another window. © 1942 ...
  21. [21]
    Flory-Huggins Equation - an overview | ScienceDirect Topics
    The Flory-Huggins equation obtained by Flory (Flory, 1942) and Huggins (Huggins, 1941) is for the thermodynamics of polymer solutions.
  22. [22]
    Analytical Solution to the Flory–Huggins Model - ACS Publications
    Aug 17, 2022 · The Ginzburg–Landau solution describes the binodal near the critical point (Figure 1B,C). We now aim to extend this solution to cover χ far ...
  23. [23]
    [PDF] Chapter 2 Polymer Physics Concentrated Solutions and Melts
    For a dilute chain in a good solvent the chain experiences repulsive interactions between chain elements which lead to excluded volume and good solvent scaling.<|control11|><|separator|>
  24. [24]
    [PDF] Mean Field Flory Huggins Lattice Theory
    The excluded volume effect is precisely compensated by the solvent quality effect. Previously we called this the θ condition, so χ = 1/2 is also the theta ...
  25. [25]
    Electro-entropic excluded volume effects on DNA in nanochannels
    Oct 22, 2013 · The model bead excluded volume diameters correspond to the change of polymer width due to the electrostatic screening length (high ionic ...
  26. [26]
    [PDF] Scaling regimes for wormlike chains confined to cylinders
    Nov 25, 2023 · Consequently, the Odijk length or the deflection length, ld = (R2lp)1/3, emerges [23] coupling the chain stiffness and radius of con ...
  27. [27]
    A backbone-based theory of protein folding | PNAS
    The Levinthal paradox adds a temporal dimension to the basic conundrum ... excluded volume restrictions in the marginally compact phase (47, 59). Such ...
  28. [28]
    Simple two‐state protein folding kinetics requires near‐levinthal ...
    Jul 31, 2003 · Even for Gō and Gō-like models, energy landscapes cannot be completely smooth because of repulsive interactions (including excluded volume ...<|separator|>
  29. [29]
    An upper limit for macromolecular crowding effects - BMC Biophysics
    May 31, 2011 · In cells, these macromolecular crowders have a large range of sizes and can occupy 30% or more of the available volume. We chose to study the ...Discussion · Dynamics And Stability · Methods
  30. [30]
    Minireview - The Influence of Macromolecular Crowding and ...
    Excluded volume theory predicts that at the high level of macromolecular fractional volume occupancy characteristic of all living cells (i.e. ⬎0.20 – 0.30), the.
  31. [31]
    Phase separation in cytoplasm, due to macromolecular crowding, is ...
    The macromolecular diversity and concentrations in the fluid phase of cytoplasm constitute conditions necessary and sufficient for aqueous phase separation.
  32. [32]
    Biophysical analysis and small-angle X-ray scattering-derived ...
    The last column in the table lists the expected particle exclusion volumes under the assumption that 1:1 complexes are formed between the nucleosomes and MeCP2.
  33. [33]
    Effect of Geometrical Asymmetry on the Phase Behavior of Rod-Coil ...
    The orientational interaction is generally described as the excluded volume interaction between the rod-like units that can be defined as the Maier–Saupe ...
  34. [34]
  35. [35]