Fact-checked by Grok 2 weeks ago

Sedimentation coefficient

The sedimentation coefficient, denoted as s, is a fundamental biophysical parameter that quantifies the rate at which a or particle sediments under the influence of a centrifugal field in (AUC). It is defined as the ratio of the sedimentation velocity (v) to the centrifugal acceleration (ω²r), where ω is the and r is the radial distance from the axis of rotation, yielding the equation s = v / (ω²r). Expressed in units (S), with 1 S equivalent to 10⁻¹³ seconds, the sedimentation coefficient serves as a characteristic property of biological macromolecules, reflecting their , , and hydrodynamic in . Named after the Swedish chemist Theodor Svedberg, who invented the in the 1920s and received the in 1926 for this work, the sedimentation coefficient enables precise characterization of biomolecular properties without requiring chemical modification or immobilization. Its value is influenced by several factors, including the macromolecule's (M), partial specific volume (), the solvent's (ρ) and (η), and the translational friction coefficient (f), as described by the relation s = M(1 - ρ) / (NAf), where NA is Avogadro's number. To standardize measurements across varying experimental conditions, the observed sexp is corrected to s20,w at 20°C in , using the formula s20,w = sexp × (η / η20,w) × ((1 - ρ20,w) / (1 - ρ)). In practice, sedimentation coefficients are determined through sedimentation velocity experiments in , where the movement of concentration boundaries is monitored optically (e.g., via or ) and fitted to the Lamm equation, which models the radial distribution of solute concentration over time. This approach allows for the resolution of heterogeneous samples into distributions of s values, providing insights into molecular weight, oligomeric state, and conformational changes. For instance, ribosomal subunits are classically identified by their s values, such as the 30S and 50S prokaryotic subunits, while proteins like exhibit s ≈ 4.5 S. The sedimentation coefficient's significance extends to studying biomolecular interactions, including self-association, hetero- formation, and binding, under near-native conditions, making it indispensable in biochemistry for purity assessment, aggregation detection, and therapeutic protein development. Advanced computational tools, such as SEDFIT and SEDPHAT, further enhance analysis by accounting for and non-ideality, enabling high-resolution c(s) distributions even for complex systems like proteins or fibrils. Despite its sensitivity to experimental variables, the sedimentation coefficient remains a cornerstone of hydrodynamic studies, complemented by techniques like for comprehensive characterization.

Definition and Fundamentals

Definition

The sedimentation coefficient, denoted as s, is defined as the ratio of a particle's terminal sedimentation velocity v_t to the applied centrifugal acceleration a, expressed as s = \frac{v_t}{a}, where a = \omega^2 r, with \omega representing the angular velocity of the rotor and r the radial distance from the axis of rotation. This parameter characterizes the sedimentation behavior of macromolecules or particles in a centrifugal field, reflecting their intrinsic properties such as size, shape, and density relative to the solvent. Physically, s quantifies the rate at which particles migrate toward the bottom of a tube under the influence of the , normalized to the strength of the field, which makes it independent of the specific rotor speed or instrument settings. The term "Svedberg," abbreviated as S, serves as the unit for this coefficient, where 1 S equals $10^{-13} seconds, honoring the pioneering work of Theodor , who developed in the 1920s to study colloidal particles and macromolecules. A representative example is the prokaryotic , which sediments at 70 S, indicating its relatively large size and compact structure compared to smaller macromolecules like proteins that typically exhibit coefficients in the range of 1–10 S.

Units and Notation

The sedimentation coefficient, denoted as s, is expressed in units of time, specifically seconds, as it represents the ratio of a particle's sedimentation to the applied centrifugal . This dimensional form arises historically from adaptations of , which originally described under gravitational acceleration where the coefficient s = v / g yields units of time, with v as (length per time) and g as (length per time squared); in ultracentrifugation, g is replaced by \omega^2 r ( squared times radial distance), preserving the time dimension. The standard unit for reporting sedimentation coefficients is the (S), named after Theodor , where 1 S = $10^{-13} seconds; this scale was chosen because typical coefficients for macromolecules fall in the range of 1 to 100 S, making the numerical values convenient for experimental reporting. In SI units, the sedimentation coefficient thus equates to $10^{-13} s per , emphasizing its temporal nature rather than a direct measure of size or mass. This unit is standardized for aqueous solutions at 20°C in , ensuring comparability across experiments by normalizing for solvent and effects. Common notations distinguish between observed and corrected values: s denotes the experimentally observed sedimentation coefficient under specific conditions, while s_{20,w} refers to the value corrected to standard conditions of 20°C in to account for and variations. Additionally, s^\circ (or s^0) specifically indicates the sedimentation coefficient extrapolated to infinite dilution, eliminating concentration-dependent interactions between solute molecules. These notations facilitate precise communication in biophysical analyses, such as those involving proteins or nucleic acids.

Theoretical Principles

Sedimentation Velocity

The terminal sedimentation velocity, v_t, represents the constant speed attained by a particle sedimenting through a under the influence of an applied , once the driving balances the opposing frictional drag.28003-1) This steady-state motion occurs after an initial acceleration phase, where the particle reaches a uniform independent of time.28003-1) In , the sedimentation process is driven by a centrifugal field rather than Earth's gravity, providing much higher accelerations for studying macromolecules. The centrifugal a = \omega^2 r acts on the particle, where \omega is the angular velocity of the rotor and r is the radial distance from the axis of rotation.28003-1) This contrasts with gravitational , where acceleration is limited to g \approx 9.8 \, \mathrm{m/s^2}, as the centrifugal approach enables precise measurement of sedimentation rates for large biomolecules that would otherwise sediment too slowly.28003-1) The frictional drag opposing sedimentation is quantified by for spherical particles, giving the frictional coefficient f = 6 \pi \eta r_0, where \eta is the viscosity and r_0 is the particle's . At , this drag force f v_t equals the net driving force. Additionally, reduces the effective of the particle due to displacement, expressed as m(1 - \bar{v} \rho), where m is the particle , \bar{v} is the partial , and \rho is the density.28003-1) This correction accounts for the buoyant force, ensuring the velocity reflects the particle's intrinsic properties rather than interactions alone. The sedimentation coefficient s normalizes v_t by the a, providing a characteristic parameter for the particle.28003-1)

Derivation of the Coefficient

The derivation of the sedimentation coefficient begins with the force balance on a sedimenting particle in a centrifugal field, where the particle achieves a constant v_t when the net driving equals the opposing frictional . The driving arises from the centrifugal minus the buoyant , given by m (1 - \bar{v} \rho) \omega^2 r, where m is the of the particle, \bar{v} is the partial specific volume, \rho is the solvent , \omega is the , and r is the radial from the of . The opposing frictional is f v_t, with f as the frictional coefficient. At equilibrium, m (1 - \bar{v} \rho) \omega^2 r = f v_t, yielding the terminal velocity v_t = \frac{m (1 - \bar{v} \rho) \omega^2 r}{f}. The sedimentation coefficient s is defined as the ratio of the terminal velocity to the , s = \frac{v_t}{\omega^2 r}. Substituting the expression for v_t gives the form s = \frac{m (1 - \bar{v} \rho)}{f}. This equation, originally developed by in the context of ultracentrifugation, normalizes the sedimentation behavior independent of the instrument's speed and position. For spherical particles under low conditions, provides the frictional coefficient as f = 6 \pi \eta r_0, where \eta is the solvent and r_0 is the . Substituting this into the expression for s results in s = \frac{m (1 - \bar{v} \rho)}{6 \pi \eta r_0}. This form assumes spherical geometry, (low ), negligible diffusion, and absence of particle interactions. To relate the sedimentation coefficient to molecular properties, express the particle mass as m = \frac{M}{N_A}, where M is the and N_A is Avogadro's number. The resulting equation is s \approx \frac{M (1 - \bar{v} \rho)}{N_A 6 \pi \eta r_0}, which links sedimentation behavior directly to the molar mass of the . This approximation holds under the stated assumptions and facilitates the characterization of molecular size and shape.

Experimental Methods

Ultracentrifugation Techniques

() is a primary technique for measuring sedimentation coefficients, employing high-speed rotors that generate centrifugal forces up to approximately 300,000 times at speeds of 60,000 rpm to induce of macromolecules in . Samples are placed in specialized sector-shaped cells within the rotor to minimize and wall effects, ensuring that sedimenting boundaries move radially outward without mixing due to density differences. Optical detection systems, such as at UV-visible wavelengths or interference optics, monitor the migration of these boundaries in , allowing direct observation of sedimentation behavior without disrupting the experiment. In AUC protocols, rotors are accelerated stepwise to the target speed to prevent initial and ensure stable , with typical run times ranging from several hours to multiple days depending on the and desired boundary resolution. This setup enables the determination of coefficients by tracking the velocity of boundaries under controlled centrifugal fields, providing insights into macromolecular size, shape, and interactions. Preparative ultracentrifugation complements by focusing on large-scale isolation of particles based on their sedimentation coefficients, often using fixed-angle or swinging-bucket rotors to fractionate samples into distinct layers. A common method involves density , where a preformed of increasing concentration separates components by their sedimentation rates, with faster-sedimenting (higher s values) penetrating deeper into the before is reached. Fractions are collected post-run for further analysis, making this technique valuable for purifying biomolecules like proteins and viruses. The foundational development of ultracentrifugation traces back to Theodor Svedberg, who constructed the first oil-turbine-driven in 1924, capable of 12,000 rpm, and refined it in the 1920s to study colloidal particles, earning the 1926 for this work on disperse systems. Modern advancements since the early 2000s include fluorescence detection systems integrated into , which use excitation and photomultiplier tubes to label and track specific macromolecules at lower concentrations with higher sensitivity than traditional optics.

Data Acquisition and Analysis

In for determining sedimentation coefficients, raw data are acquired through optical scanning systems that monitor the movement of solute in the sample cell. These systems produce boundary patterns revealing a sedimenting boundary, where macromolecules migrate toward the cell bottom under , and a non-sedimenting plateau representing the depleted region near the . Traditional detection methods include , which visualize gradients as deflections in a light beam, and Rayleigh interference , which quantify concentration via phase shifts in interference fringes corresponding to solute displacement. The sedimentation coefficient s is derived from the radial position x of the boundary as a function of time t, using the relation s = \frac{1}{\omega^2 x} \frac{dx}{dt}, where \omega denotes the angular velocity of the rotor. This expression arises from the balance of centrifugal and frictional forces on the sedimenting species. For cases where boundary motion is non-linear—due to initial acceleration or varying field strength—numerical integration techniques approximate the path by solving the underlying Lamm equation, which governs the spatiotemporal concentration profile. Analysis of these datasets relies on specialized software to fit theoretical models to the observed scans and extract s. SEDFIT, developed for sedimentation velocity experiments, employs finite-element solutions to the Lamm equation for generating sedimentation coefficient distributions c(s), enabling of sample heterogeneity and correction for diffusion-induced broadening. SEDPHAT extends this capability with global nonlinear least-squares fitting across multiple datasets or detection wavelengths, facilitating multi-signal integration for complex systems like interacting proteins. These tools typically achieve sedimentation coefficient precisions of ±1-5% through robust statistical optimization. Significant error sources in s determination include , which progressively broadens the and biases estimates if unaccounted for, and inaccuracies in the partial \bar{v}, which influences the buoyant correction in the effective mass. Least-squares fitting in SEDFIT and SEDPHAT mitigates these by simultaneously optimizing for and parameters, though experimental signal-to-noise ratios remain critical for precision.

Influencing Factors

Concentration Dependence

In non-ideal solutions, the observed sedimentation coefficient s decreases with increasing solute concentration c primarily due to hydrodynamic interactions among macromolecules, which hinder their motion, and backflow effects where the solvent displaced by sedimenting particles creates opposing flows. These non-ideality effects become significant at concentrations above approximately 1 mg/mL for typical proteins, complicating the interpretation of sedimentation velocity data without corrections. A common empirical approach to quantify this dependence is the linear approximation s = s^\circ (1 - k_s c), where s^\circ is the sedimentation coefficient at infinite dilution and k_s is the , a measure of non-ideality typically ranging from 0.007 to 0.008 L/g for globular proteins. This model, first established in the through studies on derivatives and viruses, provides a practical first-order correction for moderate concentrations. For higher concentrations, advanced models incorporate non-linear terms via virial expansions, such as s = s^\circ (1 - k_s c + \beta c^2 + \cdots), where higher-order coefficients account for multi-body interactions and are linked to the osmotic second virial coefficient B_2 through relations like k_s \approx 2 B_2 M - k_D, with k_D being the corresponding non-ideality coefficient. These expansions, building on theoretical frameworks including hydrodynamic corrections for particle interactions, enable more accurate modeling in concentrated solutions. To obtain s^\circ, experimental data from multiple concentrations are analyzed using extrapolation techniques, such as plotting $1/s versus c, which yields a straight line with intercept $1/s^\circ and slope related to k_s. This correction is essential for reliable molecular weight estimation via the Svedberg equation, as uncorrected s values can lead to systematic errors in macromolecular characterization.

Solvent and Temperature Effects

The sedimentation coefficient s is inversely proportional to the solvent viscosity \eta, as frictional drag on the sedimenting particle increases with higher \eta, slowing the sedimentation rate. Additionally, solvent density \rho modulates the buoyancy term (1 - \bar{v} \rho), where \bar{v} is the partial specific volume of the solute; higher \rho enhances , reducing s. These effects are accounted for in the standard relation s = \frac{m (1 - \bar{v} \rho)}{f}, with the frictional coefficient f proportional to \eta. Temperature influences s primarily through its impact on solvent viscosity and density, both of which vary significantly. Viscosity decreases with rising temperature, accelerating sedimentation; for water, \eta \approx 1.00 cP at 20°C but drops to \approx 0.65 cP at 40°C. Density also declines slightly (e.g., from 0.998 g/cm³ at 20°C to 0.992 g/cm³ at 40°C), further increasing the buoyancy term. Corrections for temperature are exponential due to the nonlinear viscosity-temperature relationship, typically transforming observed s_{T,B} (at temperature T in buffer B) to standard conditions via s_{20,w} = s_{T,B} \frac{(1 - \bar{v} \rho)_{20,w} \, \eta_{T,B}}{(1 - \bar{v} \rho)_{T,B} \, \eta_{20,w}}, where subscripts denote water (w) at 20°C. Solvent composition alters s beyond pure water, particularly in non-aqueous or mixed media used for specialized studies. In deuterated water (D₂O), higher density (≈1.105 g/cm³ vs. 0.998 g/cm³ for H₂O at 20°C) reduces s via enhanced buoyancy, enabling precise \bar{v} measurements through comparative sedimentation equilibrium. This isotope substitution is valuable for neutron scattering contrast but requires corrections for the ≈10% slower sedimentation in D₂O. Cosolvents like urea modify both \eta and \rho, with 8 M urea increasing \eta by ≈1.45-fold and \rho by ≈8% relative to water, thus decreasing s. Urea also perturbs \bar{v} slightly (e.g., by 0.01–0.02 cm³/g for proteins), reflecting hydration changes during denaturation. Standardization to s_{20,w} ensures comparability across experiments, using literature-derived factors for \eta and \rho. Seminal tables from Kawahara and Tanford (1966) provide empirical polynomials, such as \eta/\eta_0 = 1 + 3.75 \times 10^{-2}C + 3.15 \times 10^{-3}C^2 + 3.10 \times 10^{-4}C^3 for urea (C in M), enabling precise corrections in denaturant solutions. Modern computational tools, like SEDNTERP software, automate these using updated databases for diverse solvents, incorporating compressibility and cosolvent interactions for accuracy within 1–2%.

Applications

Macromolecular Characterization

The sedimentation coefficient, s, serves as a key parameter in characterizing the physicochemical properties of synthetic and natural , such as proteins, nucleic acids, and polymers, by providing insights into their size, shape, and interactions in solution. Measured via (AUC), s reflects the rate at which a sediments under , influenced by its mass, , and frictional drag. This enables precise determination of molecular attributes without relying on matrix-based separation techniques, making it particularly valuable for studying isolated in their native or near-native states. One primary application is the estimation of molecular weight (M) through the Svedberg equation: M = \frac{s R T}{D (1 - \bar{v} \rho)} where R is the , T is the absolute temperature, D is the diffusion coefficient, \bar{v} is the , and \rho is the solvent density. This relationship combines sedimentation velocity data (s) with diffusion measurements to yield absolute , assuming knowledge of \bar{v} and \rho, which are often determined experimentally or estimated from composition for proteins. For instance, this approach has been instrumental in confirming the molecular weights of globular proteins like at approximately 64,500 Da, using s values around 4.5 S combined with corresponding D. The sedimentation coefficient also reveals macromolecular shape and conformation via the frictional ratio, f/f_0, defined as the ratio of the actual frictional coefficient (f) to that of a hypothetical anhydrous sphere of equivalent mass (f_0). A value of f/f_0 = 1 corresponds to a perfect sphere, while f/f_0 > 1 indicates deviations due to non-sphericity, hydration, or asymmetry; typical globular proteins exhibit f/f_0 \approx 1.2, elongated proteins or DNA up to 2.0–3.0, and flexible polymers even higher. This parameter is derived from s and M using f/f_0 = \frac{R T}{M (1 - \bar{v} \rho) s} / f_0, allowing assessment of conformational changes or compactness in proteins, rigid rod-like DNA structures, and coiled synthetic polymers like polystyrene. Oligomerization states can be detected through shifts in s, as association increases effective mass while frictional effects moderate the change; for example, dimerization typically elevates s by a factor of 1.4–1.6 relative to the , rather than the ideal factor of 2, due to alterations and hydrodynamic interactions. This concentration-dependent variation in s enables monitoring of self-association equilibria in proteins and polymers, distinguishing from dimers or higher oligomers without assuming ideality. Representative examples illustrate these applications: globular proteins like sediment at 4.5 S, reflecting their compact tetrahedral structure with f/f_0 \approx 1.25, while viral capsids, such as (AAV) particles, show s \approx 95 S for filled capsids versus 65 S for empty ones, highlighting and differences. Integration of s distributions from further assesses polydispersity in heterogeneous samples, such as mixtures or protein aggregates, by resolving multiple and quantifying their relative abundances to evaluate sample homogeneity.

Biological and Biomedical Uses

In biological research, sedimentation coefficients are essential for analyzing ribosomal assemblies and their associated polysomes, which are key to protein synthesis. Bacterial ribosomes sediment at 70S, consisting of 50S and 30S subunits, while eukaryotic cytoplasmic ribosomes sediment at 80S, with 60S and 40S subunits. Sucrose density gradient centrifugation exploits these distinct sedimentation profiles to isolate polysomes—multiple ribosomes bound to mRNA—from cellular extracts, enabling studies of translation regulation and ribosome biogenesis in organisms like Synechocystis sp. Sedimentation analysis also characterizes viral particles and biomimetic nanoparticles for biomedical applications. For instance, HIV-1 virus-like particles, modeling mature structures, exhibit sedimentation coefficients around 200–300 S in , aiding in the assessment of assembly and maturation states critical for development. Density gradient methods, such as gradients, size nanoparticle-based vehicles like silver or nanoparticles by their rates, which depend on particle mass and shape, facilitating optimization for targeted therapies. In studies of protein complexes, sedimentation coefficients reveal assembly dynamics relevant to disease. The 26S proteasome, named for its sedimentation coefficient, degrades ubiquitinated proteins and its dysfunction contributes to aggregate accumulation; its ~26 S value confirms the integration of a core with 19S regulatory caps. Spliceosomes, dynamic RNA-protein machines, form complexes sedimenting at 40–60 S depending on the assembly stage, such as the 35S tri-snRNP intermediate or the 60S active . In Alzheimer's disease, sedimentation velocity ultracentrifugation profiles amyloid-β aggregates, distinguishing oligomeric species (e.g., ~10–20 S for small oligomers) from , linking their size distribution to . Recent advancements integrate sedimentation analysis with cryo-electron microscopy (cryo-EM) and for comprehensive of biological assemblies. Sedimentation velocity provides purity and heterogeneity data to select optimal samples for cryo-EM, as in characterizing or viral complexes before high-resolution imaging. Hybrid approaches combine sedimentation with to map protein stoichiometries in complexes, such as detecting co-sedimenting partners in proteomic workflows. In therapeutics, post-2010 applications include assessing purity; for example, sedimentation velocity quantifies empty versus full capsids in A71 vaccines (e.g., ~160 S for full particles), ensuring potency and in formulations. Similar methods evaluate vectors for , distinguishing genomic-filled particles (~120–140 S) from empty ones.

References

  1. [1]
    Analytical Ultracentrifugation: Sedimentation Velocity and ... - NIH
    The sedimentation coefficient depends on the total macromolecular concentration. In the simplest analysis, the viscosity of solutions increases with increasing ...
  2. [2]
    Current Methods in Sedimentation Velocity and ... - PubMed Central
    We can define the sedimentation coefficient s as the linear velocity u of sedimentation a protein exhibits per unit gravitational field ω2r (with rotor ...
  3. [3]
    [PDF] Sedimentation for Analysis and Separation of Macromolecules
    Since it is the molecular properties that most interest us, we can factor out the latter terms by defining a sedimentation coefficient s = v/rω2, which has cgs ...
  4. [4]
    [PDF] Sedimentation
    The sedimentation coefficient is a function of the size and shape of the sedimenting particle. At the boundary x, we have a concentration c. 2 of solute. These ...
  5. [5]
    Analytical Ultracentrifugation: A Versatile and Valuable Technique ...
    The technique was pioneered by Swedish chemist Theodor Svedberg, who was awarded the Nobel Prize in Chemistry in 1926 for his innovative work on dispersed ...
  6. [6]
    Sedimentation coefficients of RNA from 70S and 80S ribosomes
    Sedimentation coefficients of RNA from 70S and 80S ribosomes. Proc Natl Acad Sci U S A. 1967 Jan;57(1):164-9. doi: 10.1073/pnas.57.1.164.
  7. [7]
    [PDF] Biomolecular hydrodynamics - Casegroup
    Measurements are reported in Svedberg units, where 1S=10-13 s. Hence the. Svedberg coefficient is proportional to the mass of the particle. ... sedimentation ...
  8. [8]
    Determination of Sedimentation Coefficients for Small Peptides
    Sedimentation coefficients are given in Svedberg units (10−13 s−1). Sedimentation coefficients were determined from fits of the experimental data taking into ...
  9. [9]
    [PDF] AUC - sedimentation velocity experiment
    s. = 1 s0. +. 1 s0 ks c. (3). Here, s0 is the sedimentation coefficient corrected to infinite dilution and c is the solute concentration. ks is a measure for ...
  10. [10]
    [PDF] THE SVEDBERG - The ultracentrifuge - Nobel Lecture, May 19, 1927
    The ultracentrifuge uses high-speed centrifugation to study colloids, measuring sedimentation speed or equilibrium to determine particle size and molecular ...<|control11|><|separator|>
  11. [11]
    Analytical Ultracentrifugation Experiments - Beckman Coulter
    These experiments measure the sedimentation rate of molecules in solution in response to the centrifugal force, providing high-resolution information on the ...
  12. [12]
    svprotocols - Sedfit
    For larger complexes, use a lower rotor speed (30,000 to 40,000 rpm), and for smaller proteins a higher rotor speed (60,000 rpm). Use a rotor temperature of 20 ...
  13. [13]
    Analytical Ultracentrifugation (AUC): An Overview of the Application ...
    Dec 22, 2020 · Analytical ultracentrifugation (AUC) represents a broadly applicable and information-rich method for investigating macromolecular characteristics.Abstract · INTRODUCTION · GENERAL AND... · BASICS OF SV-AUC DATA...
  14. [14]
    Ultracentrifugation: Types, Techniques & Applications
    Density gradient ultracentrifugation ... It separates particles based on their buoyant density using a gradient medium like sucrose or cesium chloride.
  15. [15]
    Sucrose Density Gradient Centrifugation - ScienceDirect.com
    Sucrose density gradient centrifugation is a method used to purify and identify virus particles by separating them based on their sedimentation coefficients ...
  16. [16]
    Sucrose Density Gradient Centrifugation Protocols and Methods
    Sucrose Density Gradient Centrifugation is a technique used for fractionation of macromolecules like DNA, RNA, and proteins.
  17. [17]
    Accounting for Photophysical Processes and Specific Signal ...
    A fluorescence detection system for the analytical ultracentrifuge and its application to proteins, nucleic acids, viroids and viruses. In Analytical ...Missing: post- | Show results with:post-
  18. [18]
    Optical Systems Used for AUC - Beckman Coulter
    Optical systems used for AUC include absorbance spectrophotometers, which are the most frequently used, and Rayleigh interferometers.
  19. [19]
  20. [20]
  21. [21]
    Quantifying the concentration dependence of sedimentation ...
    Apr 10, 2021 · Time-dependence of g∗(s) is removed by employing Eq. (5) with r substituted for rb to generate a distribution with sedimentation coefficient s ...
  22. [22]
    A Reappraisal of Sedimentation Nonideality Coefficients for ... - NIH
    Feb 27, 2019 · The choice of techniques for measuring the second virial coefficient depends on many factors, including purity, experimental time, concentration ...
  23. [23]
    [PDF] Presenter: Borries Demeler Topic: Sedimentation Transport I
    Feb 7, 2019 · Transport Processes – Fundamental Equations: We have: Svedberg equation: Stokes-Einstein: s depends on and ρ s and D are inversely ...
  24. [24]
    [PDF] Dynamic viscosity of liquid water from 0 °C to 100 °C - VaxaSoftware
    Dynamic viscosity kg / (m·s). 0. 0.001792. 34. 0.000734. 68. 0.000416. 1. 0.001731. 35. 0.000720. 69. 0.000410. 2. 0.001674. 36. 0.000705. 70. 0.000404.
  25. [25]
    Water - Dynamic and Kinematic Viscosity at Various Temperatures ...
    Free online calculator - figures and tables with viscosity of water at temperatures ranging 0 to 360°C (32 to 675°F) - Imperial and SI Units.
  26. [26]
    PARTIAL SPECIFIC VOLUME MEASUREMENTS BY ...
    The partial specific volumes of several macromolecules were determined from their sedimentation rates in aqueous and heavy water solutions with results ...
  27. [27]
    [PDF] Partial specific volumes and interactions with solvent components of α
    The partial specific volume, V of a protein in a particular solvent is a parameter which is necessary for the determination of the molecular weight of the ...
  28. [28]
    [PDF] Determination of Membrane Protein Molecular Weights and ...
    The monomeric mass and partial specific volume of each of the proteins were calculated using the program. SEDNTERP (Laue et al., 1992) and the amino acid ...
  29. [29]
    Analytical Ultracentrifugation (AUC): An Overview of the Application ...
    Dec 22, 2020 · Analytical ultracentrifugation (AUC) represents a broadly applicable and information‐rich method for investigating macromolecular characteristics.
  30. [30]
    Size and Shape of Protein Molecules at the Nanometer Level ...
    The relationship of S to size and shape of the protein is given by the Svedberg formula: ... Thus, typical proteins have sedimentation coefficients of 2–20 S.
  31. [31]
    [PDF] 361847: Introduction to Analytical Ultracentrifugation
    differences in sedimentation coefficient, and of diffusion coefficients. *A ... sedimentation and chemical equilibrium. The three sets of Ωr versus c ...
  32. [32]
    Size-Distribution Analysis of Proteins by Analytical Ultracentrifugation
    A method for directly fitting the time derivative of sedimentation velocity data and an alternative algorithm for calculating sedimentation coefficient ...<|control11|><|separator|>
  33. [33]
    heart cytochrome <i>c</i> by hexametaphosphate anions
    Using this simple approximation, most models predict a sedimentation coefficient 1.4- 1.6 times (as opposed to 2 times) greater than the monomeric protein.
  34. [34]
    Characterization and quantification of adeno-associated virus ... - NIH
    Oct 25, 2023 · SV analysis of sample AAV8 BCM #1 revealed filled capsids sedimenting at 95 S (Figure 7A), along with a nucleic acid signal sedimenting below 20 ...
  35. [35]
    Measuring macromolecular size distributions and interactions at ...
    Oct 24, 2018 · Here we present a sedimentation velocity technique that, for the first time, can resolve macromolecular size distributions at high ...
  36. [36]
    Sedimentation coefficients of RNA from 70S and 80S ribosomes.
    In the case of Neurospora crassa, a purified ribosome pellet' was suspended in 2 M LiCl and incubated at 4°C for 16 hours. '5 The RNA precipitate was washed ...Missing: polysome | Show results with:polysome
  37. [37]
    [PDF] Purification, identification, and functional analysis of polysomes from ...
    Sep 21, 2017 · Sucrose gradient centrifugation is used to separate polysomes, 70S ribosomes, and the two ribosomal subunits according to their sedimentation.
  38. [38]
    Purification, identification, and functional analysis of polysomes from ...
    Mar 15, 2017 · The technique presented here will be applied to the purification and use of polysomes, 70S ribosomes, and 50S and 30S subunits from different S.
  39. [39]
    Size-Separation of Silver Nanoparticles Using Sucrose Gradient Ce
    This technique makes use of the sedimentation coefficients of nanoparticles in the surrounding medium, which can vary with nanoparticles form and mass [5], and ...Abstract · Materials And Methods · Results And DiscussionMissing: vehicles | Show results with:vehicles<|separator|>
  40. [40]
    Leveraging Analytical Ultracentrifugation for Comprehensive ...
    Lipid nanoparticles (LNPs) are an advanced drug delivery system that encapsulates therapeutic molecules, such as nucleic acids, proteins, or small molecules ...Missing: vehicles | Show results with:vehicles
  41. [41]
    Protein and gene structures of 20S and 26S proteasomes - PubMed
    The two types of proteasomes with apparent sedimentation coefficients of 20S and 26S consist of a number of heterogeneous polypeptides and are unusually ...
  42. [42]
    (PDF) Identification of a 35S U4/U6.U5 Tri-snRNP Complex ...
    Sep 6, 2017 · Remarkably, ZIP-containing tri-snRNP has a sedimentation coefficient ~35S, a tri-snRNP that has not been described before. We showed that the ...
  43. [43]
    Three-Dimensional Structure of the Native Spliceosome by Cryo ...
    In fact, each of the individual subcomplexes resembles the 60S in vitro-assembled spliceosome also with respect to sedimentation coefficient (60S–70S), ...
  44. [44]
    Sedimentation Velocity Analysis of the Size Distribution of Amyloid ...
    In this chapter, we describe preparative ultracentrifuge methods to isolate specific amyloid aggregates, as well as analytical ultracentrifuge methods for ...
  45. [45]
    Amyloid β Oligomeric Species Present in the Lag Phase of Amyloid ...
    May 29, 2015 · Our study focused on Aβ42 and its oligomeric assemblies in the lag phase of amyloid formation, as studied by sedimentation velocity (SV) ...
  46. [46]
    Integral approach to biomacromolecular structure by analytical ...
    Jun 8, 2020 · Sedimentation velocity-AUC (SV-AUC) is conducted for the same solution subjected to the SAS measurement. The measured sedimentation coefficient ...
  47. [47]
    Proteomic complex detection using sedimentation - PubMed
    Mar 1, 2007 · Proteins in a putative complex can be identified since they sediment faster than predicted from their monomer molecular weight. Using ...Missing: coefficient | Show results with:coefficient
  48. [48]
    Quantitation of Enterovirus A71 Empty and Full Particles by ...
    Apr 8, 2024 · In this study, we analyzed the application of sedimentation velocity analytical ultracentrifugation (SV-AUC) in characterizing the EPs and FPs ...
  49. [49]
    Methodological Validation of Sedimentation Velocity Analytical ...
    May 8, 2024 · The recovery rate of partial particles and full particles were >110% or <80% since their percentages were <10% in the mix empty capsid samples: ...<|control11|><|separator|>