Mason–Weaver equation

Mason–Weaver equation

The Mason–Weaver equation (named after Max Mason and Warren Weaver) describes the sedimentation and diffusion of solutes under a uniform force, usually a gravitational field.[1] Assuming that the gravitational field is aligned in the z direction (Fig. 1), the Mason–Weaver equation may be written


\frac{\partial c}{\partial t} = 
D \frac{\partial^{2}c}{\partial z^{2}} + 
sg \frac{\partial c}{\partial z}

where t is the time, c is the solute concentration (moles per unit length in the z-direction), and the parameters D, s, and g represent the solute diffusion constant, sedimentation coefficient and the (presumed constant) acceleration of gravity, respectively.

The Mason–Weaver equation is complemented by the boundary conditions


D \frac{\partial c}{\partial z} + s g c = 0

at the top and bottom of the cell, denoted as za and zb, respectively (Fig. 1). These boundary conditions correspond to the physical requirement that no solute pass through the top and bottom of the cell, i.e., that the flux there be zero. The cell is assumed to be rectangular and aligned with the Cartesian axes (Fig. 1), so that the net flux through the side walls is likewise zero. Hence, the total amount of solute in the cell


N_{tot} = \int_{z_{b}}^{z_{a}} dz \ c(z, t)

is conserved, i.e., dNtot / dt = 0.

Figure 1: Diagram of Mason–Weaver cell and Forces on Solute

Contents

Derivation of the Mason–Weaver equation

A typical particle of mass m moving with vertical velocity v is acted upon by three forces (Fig. 1): the drag force fv, the force of gravity mg and the buoyant force ρVg, where g is the acceleration of gravity, V is the solute particle volume and ρ is the solvent density. At equilibrium (typically reached in roughly 10 ns for molecular solutes), the particle attains a terminal velocity vterm where the three forces are balanced. Since V equals the particle mass m times its partial specific volume \bar{\nu}, the equilibrium condition may be written as


f v_{term} = m (1 - \bar{\nu} \rho) g \ \stackrel{\mathrm{def}}{=}\  m_{b} g

where mb is the buoyant mass.

We define the Mason–Weaver sedimentation coefficient s \ \stackrel{\mathrm{def}}{=}\  m_{b} / f = v_{term}/g. Since the drag coefficient f is related to the diffusion constant D by the Einstein relation


D = \frac{k_{B} T}{f}
,

the ratio of s and D equals


\frac{s}{D} = \frac{m_{b}}{k_{B} T}

where kB is the Boltzmann constant and T is the temperature in kelvins.

The flux J at any point is given by


J = -D \frac{\partial c}{\partial z} - v_{term} c 
  = -D \frac{\partial c}{\partial z} - s g c.

The first term describes the flux due to diffusion down a concentration gradient, whereas the second term describes the convective flux due to the average velocity vterm of the particles. A positive net flux out of a small volume produces a negative change in the local concentration within that volume


\frac{\partial c}{\partial t} = -\frac{\partial J}{\partial z}.

Substituting the equation for the flux J produces the Mason–Weaver equation


\frac{\partial c}{\partial t} = 
D \frac{\partial^{2}c}{\partial z^{2}} + 
sg \frac{\partial c}{\partial z}.

The dimensionless Mason–Weaver equation

The parameters D, s and g determine a length scale z0


z_{0} \ \stackrel{\mathrm{def}}{=}\  \frac{D}{sg}

and a time scale t0


t_{0} \ \stackrel{\mathrm{def}}{=}\  \frac{D}{s^{2}g^{2}}

Defining the dimensionless variables \zeta \ \stackrel{\mathrm{def}}{=}\  z/z_{0} and \tau \ \stackrel{\mathrm{def}}{=}\  t/t_{0}, the Mason–Weaver equation becomes


\frac{\partial c}{\partial \tau} =
\frac{\partial^{2} c}{\partial \zeta^{2}} + 
\frac{\partial c}{\partial \zeta}

subject to the boundary conditions


\frac{\partial c}{\partial \zeta} + c = 0

at the top and bottom of the cell, ζa and ζb, respectively.

Solution of the Mason–Weaver equation

This equation may be solved by separation of variables. Defining c(\zeta,\tau) \ \stackrel{\mathrm{def}}{=}\  e^{-\zeta/2} T(\tau) P(\zeta), we obtain the two equations coupled by a constant β


\frac{\partial T}{\partial \tau} + \beta T = 0

\frac{\partial^{2} P}{\partial \zeta^{2}} + 
 \left[ \beta - \frac{1}{4} \right] P = 0

where acceptable values of β are defined by the boundary conditions


\frac{dP}{d\zeta} + \frac{1}{2} P = 0

at the upper and lower boundaries, ζa and ζb, respectively. Since the T equation has the solution T(τ) = T0e − βτ, where T0 is a constant, the Mason–Weaver equation is reduced to solving for the function P(ζ).

The ordinary differential equation for P and its boundary conditions satisfy the criteria for a Sturm–Liouville problem, from which several conclusions follow. First, there is a discrete set of orthonormal eigenfunctions Pk(ζ) that satisfy the ordinary differential equation and boundary conditions. Second, the corresponding eigenvalues βk are real, bounded below by a lowest eigenvalue β0 and grow asymptotically like k2 where the nonnegative integer k is the rank of the eigenvalue. (In our case, the lowest eigenvalue is zero, corresponding to the equilibrium solution.) Third, the eigenfunctions form a complete set; any solution for c(ζ,τ) can be expressed as a weighted sum of the eigenfunctions


c(\zeta, \tau) = 
\sum_{k=0}^{\infty} c_{k} P_{k}(\zeta) e^{-\beta_{k}\tau}

where ck are constant coefficients determined from the initial distribution c(ζ,τ = 0)


c_{k} = 
\int_{\zeta_{a}}^{\zeta_{b}} d\zeta \ 
c(\zeta, \tau=0) e^{\zeta/2} P_{k}(\zeta)

At equilibrium, β = 0 (by definition) and the equilibrium concentration distribution is


e^{-\zeta/2} P_{0}(\zeta) = B e^{-\zeta} = B e^{-m_{b}gz/k_{B}T}

which agrees with the Boltzmann distribution. The P0(ζ) function satisfies the ordinary differential equation and boundary conditions at all values of ζ (as may be verified by substitution), and the constant B may be determined from the total amount of solute


B = N_{tot} \left( \frac{sg}{D} \right) 
\left( \frac{1}{e^{-\zeta_{b}} - e^{-\zeta_{a}}} \right)

To find the non-equilibrium values of the eigenvalues βk, we proceed as follows. The P equation has the form of a simple harmonic oscillator with solutions P(\zeta) = e^{i\omega_{k}\zeta} where


\omega_{k} = \pm \sqrt{\beta_{k} - \frac{1}{4}}

Depending on the value of βk, ωk is either purely real (\beta_{k}\geq\frac{1}{4}) or purely imaginary (\beta_{k} < \frac{1}{4}). Only one purely imaginary solution can satisfy the boundary conditions, namely, the equilibrium solution. Hence, the non-equilibrium eigenfunctions can be written as

P(ζ) = Acos ωkζ + Bsin ωkζ

where A and B are constants and ω is real and strictly positive.

By introducing the oscillator amplitude ρ and phase ϕ as new variables,


u \ \stackrel{\mathrm{def}}{=}\  \rho \sin(\phi) \ \stackrel{\mathrm{def}}{=}\  P

v \ \stackrel{\mathrm{def}}{=}\  \rho \cos(\phi) \ \stackrel{\mathrm{def}}{=}\  - \frac{1}{\omega} 
\left( \frac{dP}{d\zeta} \right)

\rho \ \stackrel{\mathrm{def}}{=}\  u^{2} + v^{2}

\tan(\phi) \ \stackrel{\mathrm{def}}{=}\  v / u

the second-order equation for P is factored into two simple first-order equations


\frac{d\rho}{d\zeta} = 0

\frac{d\phi}{d\zeta} = \omega

Remarkably, the transformed boundary conditions are independent of ρ and the endpoints ζa and ζb


\tan(\phi_{a}) = 
\tan(\phi_{b}) = \frac{1}{2\omega_{k}}

Therefore, we obtain an equation


\phi_{a} - \phi_{b} + k\pi = k\pi = 
\int_{\zeta_{b}}^{\zeta_{a}} d\zeta \ \frac{d\phi}{d\zeta} = 
\omega_{k} (\zeta_{a} - \zeta_{b})

giving an exact solution for the frequencies ωk


\omega_{k} = \frac{k\pi}{\zeta_{a} - \zeta_{b}}

The eigenfrequencies ωk are positive as required, since ζa > ζb, and comprise the set of harmonics of the fundamental frequency \omega_{1} \ \stackrel{\mathrm{def}}{=}\  \pi/(\zeta_{a} - \zeta_{b}). Finally, the eigenvalues βk can be derived from ωk


\beta_{k} = \omega_{k}^{2} + \frac{1}{4}

Taken together, the non-equilibrium components of the solution correspond to a Fourier series decomposition of the initial concentration distribution c(ζ,τ = 0) multiplied by the weighting function eζ / 2. Each Fourier component decays independently as e^{-\beta_{k}\tau}, where βk is given above in terms of the Fourier series frequencies ωk.

See also

References

  1. ^ Mason, M; Weaver W (1924). "The Settling of Small Particles in a Fluid". Physical Review 23: 412–426. Bibcode 1924PhRv...23..412M. doi:10.1103/PhysRev.23.412. 

Wikimedia Foundation. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • Mason-Weaver equation — The Mason Weaver equation describes the sedimentation and diffusion of solutes under a uniform force, usually a gravitational field.cite journal | last = Mason | first = M | coauthors = Weaver W | year = 1924 | title = The Settling of Small… …   Wikipedia

  • Équation de Mason-Weaver — L équation de Mason Weaver est une équation décrivant la sédimentation et la diffusion de solutés sous l action d une force uniforme, typiquement un champ de pesanteur[1]. Sommaire 1 Expression mathématique …   Wikipédia en Français

  • Weaver — Cette page d’homonymie répertorie les différents sujets et articles partageant un même nom. Sur les autres projets Wikimedia : « Weaver », sur le Wiktionnaire (dictionnaire universel) Weaver est un mot anglais qui signifie… …   Wikipédia en Français

  • Sedimentation — describes the motion of molecules in solutions or particles in suspensions in response to an external force such as gravity, centrifugal force or electric force. Sedimentation may pertain to objects of various sizes, ranging from suspensions of… …   Wikipedia

  • List of mathematics articles (M) — NOTOC M M estimator M group M matrix M separation M set M. C. Escher s legacy M. Riesz extension theorem M/M/1 model Maass wave form Mac Lane s planarity criterion Macaulay brackets Macbeath surface MacCormack method Macdonald polynomial Machin… …   Wikipedia

  • Equipartition de l'énergie — Équipartition de l énergie Agitation thermique d’un peptide avec une structure en hélice alpha. Les mouvements sont aléatoires et complexes, l’énergie d’un atome peut fluctuer énormément. Néanmoins, le théorème d’équipartition permet de calculer… …   Wikipédia en Français

  • Theoreme d'equipartition — Équipartition de l énergie Agitation thermique d’un peptide avec une structure en hélice alpha. Les mouvements sont aléatoires et complexes, l’énergie d’un atome peut fluctuer énormément. Néanmoins, le théorème d’équipartition permet de calculer… …   Wikipédia en Français

  • Théorème d'équipartition — Équipartition de l énergie Agitation thermique d’un peptide avec une structure en hélice alpha. Les mouvements sont aléatoires et complexes, l’énergie d’un atome peut fluctuer énormément. Néanmoins, le théorème d’équipartition permet de calculer… …   Wikipédia en Français

  • Théorème d’équipartition — Équipartition de l énergie Agitation thermique d’un peptide avec une structure en hélice alpha. Les mouvements sont aléatoires et complexes, l’énergie d’un atome peut fluctuer énormément. Néanmoins, le théorème d’équipartition permet de calculer… …   Wikipédia en Français

  • Équipartition de l'énergie — Agitation thermique d’un peptide avec une structure en hélice alpha. Les mouvements sont aléatoires et complexes, l’énergie d’un atome peut fluctuer énormément. Néanmoins, le théorème d’équipartition permet de calculer l’énergie cinétique moyenne …   Wikipédia en Français

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”