20.11.2013 Views

Thermophoresis as persistent random walk - Saint Anselm College

Thermophoresis as persistent random walk - Saint Anselm College

Thermophoresis as persistent random walk - Saint Anselm College

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Physics Letters A 373 (2009) 2122–2124<br />

Contents lists available at ScienceDirect<br />

Physics Letters A<br />

www.elsevier.com/locate/pla<br />

<strong>Thermophoresis</strong> <strong>as</strong> <strong>persistent</strong> <strong>random</strong> <strong>walk</strong><br />

A.V. Plyukhin<br />

Department of Mathematics, <strong>Saint</strong> <strong>Anselm</strong> <strong>College</strong>, Manchester, NH 03102, USA<br />

article info abstract<br />

Article history:<br />

Received 16 January 2009<br />

Received in revised form 20 March 2009<br />

Accepted 11 April 2009<br />

Availableonline17April2009<br />

Communicated by C.R. Doering<br />

In a simple model of a continuous <strong>random</strong> <strong>walk</strong> a particle moves in one dimension with the velocity<br />

fluctuating between +v and −v. If v is <strong>as</strong>sociated with the thermal velocity of a Brownian particle<br />

and allowed to be position dependent, the model accounts readily for the particle’s drift along the<br />

temperature gradient and recovers b<strong>as</strong>ic results of the conventional thermophoresis theory.<br />

© 2009 Elsevier B.V. All rights reserved.<br />

PACS:<br />

05.40.-a<br />

66.10.C-<br />

82.70.Dd<br />

<strong>Thermophoresis</strong> [1–4] is the systematic drift of a Brownian particle<br />

caused by a temperature gradient ∇T . The steady-state thermophoretic<br />

velocity acquired by the particle is given by<br />

V T =−D T ∇T , (1)<br />

where D T is generally referred to <strong>as</strong> thermal diffusion coefficient.<br />

A similar phenomenon exists on the molecular level in g<strong>as</strong> mixtures<br />

and is known <strong>as</strong> the Soret effect. In most c<strong>as</strong>es D T is positive,<br />

i.e. the particle moves toward a colder region. However, thermophilic<br />

behavior, D T < 0, h<strong>as</strong> also been observed experimentally<br />

[5–7]. Among other well-known field-driven transport effects, thermophoresis<br />

is perhaps the most subtle: it takes place in a stationary<br />

state when the pressure is uniform, and therefore the average<br />

force on a particle which is fixed in space is strictly zero. Interesting<br />

in itself, thermophoresis h<strong>as</strong> many applications, in particular in<br />

the context of molecular Brownian motors driven by thermal fluctuations<br />

[8–10].<br />

Despite many efforts, a realistic theoretical model of thermophoresis,<br />

valid over a wide region of densities of the host fluid,<br />

seems to be lacking, and even the physical origin of the effect is<br />

still under debate [1]. Many authors argue that the effect cannot<br />

be described by a simple modification of the over-damped diffusion<br />

equation in the configurational space only (Smoluchowski<br />

equation), but requires a kinetic description. Well-known example<br />

of such approach is that by van Kampen [11], which takes <strong>as</strong> a<br />

starting point the Kramers equation with position-dependent temperature<br />

∂ f<br />

∂t =−v ∂ f<br />

∂x − F ∂ f<br />

m ∂ v + γ ∂ (<br />

vf + kT(x) )<br />

∂ f<br />

. (2)<br />

∂ v m ∂ v<br />

E-mail address: aplyukhin@anselm.edu.<br />

Here f (x, v, t) is the joint probability density of position and velocity,<br />

F is an external force, and γ is the friction coefficient which<br />

may depend on x and is <strong>as</strong>sumed to be large. Using the expansion<br />

in powers of γ −1 , one can obtain to the lowest order the<br />

corresponding overdamped Smoluchowski equation for the spatial<br />

density f (x, t) in the form<br />

∂ f<br />

∂t + ∂ (<br />

∂x J = 0, J = μF − D ∂ )<br />

∂x − D ∂ T<br />

T f (x, t), (3)<br />

∂x<br />

where the mobility μ = 1/mγ , the diffusion coefficient D = kT μ,<br />

and the thermal diffusion coefficient is<br />

D T =<br />

k<br />

mγ = D = kμ. (4)<br />

T<br />

Therefore, the thermophoretic force F T on the particle, defined<br />

through the relation μF T =−D T ∇T ,equals<br />

F T =−k∇T . (5)<br />

The result (4) often underestimates the value of D T by up to one<br />

order of magnitude [1], and cannot, of course, account for thermophilic<br />

behavior corresponding to a negative D T .Anotherrestriction<br />

of the approach is that the validity of the Kramers equation<br />

for the c<strong>as</strong>e of nonuniform temperature is postulated rather than<br />

proved. More “microscopic” approaches lead to the expression for<br />

D T with additional terms involving rather complicated correlation<br />

functions [12–14]. However, conceptually the result (4) is important<br />

<strong>as</strong> a demonstration that the expression for the current J<br />

in the Smoluchowski equation (3) cannot be obtained by simple<br />

insertion of the position dependence into the conventional driftdiffusion<br />

expression J = (μF − D∂/∂x) f for an isothermal system.<br />

The simplicity of the result (4) suggests that it might have<br />

a transparent physical interpretation. Yet the transition from the<br />

0375-9601/$ – see front matter © 2009 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.physleta.2009.04.027


A.V. Plyukhin / Physics Letters A 373 (2009) 2122–2124 2123<br />

Kramers equation (2) in ph<strong>as</strong>e space to Eq. (3) for the spatial distribution<br />

function f (x, t) = ∫ dv f (x, v, t) (the elimination of the<br />

f<strong>as</strong>t variable v) isnotatrivialstep[15,16], which makes the origin<br />

of the result (4) somewhat obscure. The same perhaps may<br />

be said of the Luttinger’s method of fictitious external fields [17],<br />

which also leads to the result (4) [18,19]. The aim of this Letter is<br />

to formulate a minimal qualitative model of thermophoresis, which<br />

leads to the expression (4) elementary and directly. Besides pedagogical<br />

merits, such model might be useful for numerical modeling<br />

of stoch<strong>as</strong>tic processes at nonuniform temperature.<br />

The model is a slightly generalized version of the well-known<br />

stoch<strong>as</strong>tic process of the continuous <strong>persistent</strong> <strong>random</strong> <strong>walk</strong><br />

[20–22]. In its simplest setting, the process describes a particle<br />

moving in one dimension with fixed speed v suffering occ<strong>as</strong>ionally<br />

a complete reversal of direction. Let f + (x, t) and f − (x, t) be<br />

the probability density for the particle moving to the right and to<br />

the left, respectively. Reversals of velocity are Poisson distributed,<br />

i.e. occurring with a constant rate 1/2τ , so that the probability for<br />

reversal in a time interval dt is dt/2τ . For an infinitesimal time<br />

step one can write<br />

f + (x, t + dt) = f + (x − vdt, t) − f + (x − vdt, t) dt/2τ<br />

+ f − (x + vdt, t) dt/2τ , (6)<br />

and a similar equation for f − (x, t). The corresponding differential<br />

equations read<br />

∂ f +<br />

∂t<br />

=−v ∂ f +<br />

∂x<br />

− f + − f −<br />

,<br />

2τ<br />

∂ f −<br />

∂t<br />

= v ∂ f −<br />

∂x<br />

+ f + − f −<br />

, (7)<br />

2τ<br />

and lead to the telegrapher’s equation for the total density f =<br />

f + + f − ,<br />

∂ 2 f<br />

∂t + 1 2 τ<br />

∂ f<br />

∂t = v2 ∂2 f<br />

∂x 2 . (8)<br />

The same equation holds also for the difference Δ = f + − f − , and<br />

therefore for the components f + and f − separately.<br />

Suppose the particle at t = 0 is at the origin with equal probability<br />

to be in each of the two velocity states,<br />

f ± (x, 0) = 1 2 δ(x), ∂ f ± (x, 0)<br />

=∓ v ∂<br />

δ(x). (9)<br />

∂t 2 ∂x<br />

Here the second initial condition follows from the first one and<br />

Eq. (7). Respectively, the initial conditions for the total density f =<br />

f + + f − are<br />

∂ f (x, 0)<br />

f (x, 0) = δ(x), = 0. (10)<br />

∂t<br />

The corresponding solution of the telegrapher’s equation is well<br />

known [26]. In the long-time limit t ≫ τ , vt ≫ x it coincides exactly<br />

with solution of the Smoluchowski equation<br />

)<br />

f (x, t) ≈ (4π Dt) −1/2 exp<br />

(− x2<br />

(11)<br />

4Dt<br />

with the diffusion coefficient D = τ v 2 . However, unlike the overdamped<br />

Smoluchowski equation (3), Eq.(7) incorporates effects of<br />

inertia of the particle. This advantage, which w<strong>as</strong> recognized and<br />

used beneficially in many previous works (see [22] and references<br />

therein), allows to account for thermophoresis in a particularly<br />

simple way.<br />

To link the model to the problem of thermophoresis, it is natural<br />

to identify v with a typical thermal speed of a Brownian<br />

particle. The choice is ambiguous. For instance, one can set v equal<br />

to the root mean square velocity v rms =〈v 2 〉 1/2 = √ 3kT/m. However,<br />

one can show that in this c<strong>as</strong>e the final result for the thermophoretic<br />

force would differ from Eq. (5) by the factor 3/2 (see<br />

Eq. (16) below). As will be shown, a perfect agreement with the<br />

standard results D T = D/T and F T =−k∇T can be achieved if v<br />

is identified not with v rms but with the most probable speed of a<br />

Brownian particle v mp (for which the Maxwell speed distribution<br />

h<strong>as</strong> a maximum):<br />

v = v mp = √ 2kT/m. (12)<br />

The second parameter of the model 1/τ should be identified with<br />

the friction coefficient γ which appears in the Kramers equation<br />

(2) and in the corresponding Langevin equation ˙v =−γ v + ξ(t).<br />

For nonuniform temperature, the velocity in Eq. (7) is position<br />

dependent, v(x) = √ 2kT(x)/m. In this c<strong>as</strong>e instead of the telegrapher’s<br />

equation (8) one obtains [23]<br />

∂ 2 f<br />

∂t + 1 (<br />

∂ f<br />

2 τ ∂t = ∂2 f<br />

v2 ∂x + 2 v dv ) ∂ f<br />

dx ∂x . (13)<br />

Suppose the temperature gradient is constant, so that T (x) = T +<br />

x ∇T and<br />

v(x) = √ (<br />

2kT(x)/m = v<br />

1 + ∇T<br />

T x ) 1/2<br />

, (14)<br />

where v is the thermal velocity corresponding to the temperature<br />

T , v = √ 2kT/m. ThenEq.(13) reads <strong>as</strong><br />

∂ 2 f<br />

∂t + 1 2 τ<br />

(<br />

∂ f<br />

∂t = v2 1 + ∇T<br />

T<br />

) ∂ 2 x f ∇T<br />

+ v2<br />

∂x2 2T<br />

∂ f<br />

∂x .<br />

For a small gradient (∇T /T )x ≪ 1, the equation is simplified to<br />

the form<br />

∂ 2 f<br />

∂t + 1 ∂ f<br />

2 τ ∂t = ∂2 f<br />

v2 ∂x − F T ∂ f<br />

2 m ∂x , (15)<br />

where the thermophoretic force F T coincides (thanks to the setting<br />

v = √ 2kT/m) with the expression (5) of the standard theory:<br />

F T =− 1 ∇T<br />

mv2 =−k∇T . (16)<br />

2 T<br />

It can be shown that in the long-time limit the solution of Eq. (15)<br />

with the boundary conditions (10) coincides with the solution of<br />

the overdamped equation (3) and describes the drift along the<br />

temperature gradient, superimposed on the diffusion<br />

f (x, t) ≈<br />

1<br />

√ 4π Dt<br />

exp<br />

(− (x − V T t) 2 )<br />

4Dt<br />

(17)<br />

with the drift velocity V T = F T τ /m. Using(16) and recalling D =<br />

τ v 2 ,onegetsV T =−D T ∇T with D T = D/T , thus recovering the<br />

result (4) of the conventional theory.<br />

The <strong>as</strong>ymptotic solution (17) can be obtained <strong>as</strong> follows. After<br />

applying the transformation f (x, t) = φ(x, t) exp(−t/2τ +<br />

F T x/2mv 2 ),Eq.(15) takes the form<br />

∂ 2 φ<br />

∂t 2<br />

= ∂2 φ<br />

v2 ∂x + 1 φ, 2 τ∗<br />

2 (18)<br />

with 1/τ 2 ∗ = 1/4τ 2 − F T /4m 2 v 2 , while the boundary conditions<br />

corresponding to (10) read<br />

φ(x, 0) = δ(x),<br />

∂φ(x, 0)<br />

∂t<br />

= 1<br />

2τ δ(x).<br />

Note that the model makes sense only under the <strong>as</strong>sumption τ∗ 2 ><br />

0, which guarantees the positiveness of f + and f − [25]. Eq.(18) is<br />

the modified telegrapher’s equation whose solution is well known<br />

[26]. Transforming back from φ to f ,theresultcanbewrittenin<br />

the following form<br />

(<br />

f (x, t) = exp<br />

− t<br />

2τ +<br />

F T x<br />

2mv 2 )<br />

(φ 1 + φ 2 + φ 3 ), (19)


2124 A.V. Plyukhin / Physics Letters A 373 (2009) 2122–2124<br />

where the functions φ i (x, t) are<br />

φ 1 (x, t) = 1 [ ]<br />

δ(x − vt) + δ(x + vt) ,<br />

2<br />

φ 2 (x, t) = 1 ( 1 √ )<br />

4vτ I 0 v<br />

vτ 2 t 2 − x 2 θ ( vt −|x| ) ,<br />

∗<br />

φ 3 (x, t) =<br />

t<br />

2τ ∗<br />

√<br />

v 2 t 2 − x 2 I 1<br />

( 1<br />

vτ ∗<br />

√<br />

v 2 t 2 − x 2 )<br />

θ ( vt −|x| ) , (20)<br />

and θ(x) is the unit step function. Using the <strong>as</strong>ymptotic form of<br />

the modified Bessel functions for large argument I α (x) ≈ 1 √ 2π x<br />

e x ,<br />

it is an e<strong>as</strong>y matter to prove that in the limit of strong damping<br />

F T τ /mv ≪ 1 and of long time t ≫ τ , vt ≫ x, the exact solution<br />

(19) is reduced to the simple form (17).<br />

One interesting problem related to thermophoresis is that of<br />

Brownian motors driven by position-dependent temperature. In<br />

particular, the Büttiker–Landauer [27,28] motor is essentially a<br />

Brownian particle diffusing in the periodic potential field subject<br />

to a spatially inhomogeneous temperature. In this context it is of<br />

interest to generalize the model to the c<strong>as</strong>e of the presence of<br />

an external force F . As discussed in [25], in this c<strong>as</strong>e Eq. (7) for<br />

f ± (x, t) should be extended <strong>as</strong> follows<br />

∂ f ±<br />

∂t<br />

∂ 2 f<br />

∂t + 1 2 τ<br />

=∓v ∂ f ±<br />

∂x<br />

∓ f + − f −<br />

2τ<br />

∂ f<br />

∂t = ∂2 f<br />

v2 ∂x − F T ∂ f<br />

2 m ∂x − ∂ ∂x<br />

± F<br />

2mv ( f + + f − ). (21)<br />

Respectively, Eq. (15) for the total density f = f + + f − is generalized<br />

to the form<br />

( ) F<br />

m f . (22)<br />

This equation differs from the overdamped equation (3) of the<br />

standard theory by the presence of the term f tt . This term, related<br />

to inertial effects, is unimportant in the long time limit, but<br />

may be responsible for wave-like behavior at short times [22].<br />

Summarizing, in this Letter we discussed a stoch<strong>as</strong>tic process<br />

which underlies thermophoresis in a way similar to that <strong>as</strong> the discrete<br />

<strong>random</strong> <strong>walk</strong> underlies isothermal diffusion. The model leads<br />

to the telegrapher’s equation whose <strong>as</strong>ymptotic solution coincides<br />

with the solution of the overdamped Smoluchowski equation (3).<br />

This is consistent with an observation that the telegrapher’s equation<br />

is reduced to the Smoluchowski equation under conditions<br />

v →∞, τ → 0, v 2 τ = const. Since the model takes into account<br />

inertial effects, one might hope that it can resolve difficulties of the<br />

overdamped theory of thermally driven Brownian motors [8–10].<br />

However, similar to the Kramers equation (2), the model is b<strong>as</strong>ed<br />

on the <strong>as</strong>sumption that the thermalization of the particle to a local<br />

temperature is f<strong>as</strong>ter than any other process involved. Since the<br />

characteristic time scales for inertial and thermalization effects are<br />

typically the same, the application of the model beyond the overdamped<br />

regime, strictly speaking, is not justified, and should be<br />

undertaken with caution. Despite of this limitation, the mapping<br />

of thermophoresis onto a <strong>random</strong> <strong>walk</strong> problem may offer some<br />

advantage for both analytical and numerical modeling, in particular<br />

for problems with complicated or velocity-dependent [24,25]<br />

boundary conditions.<br />

Acknowledgements<br />

I thank Richard Bowles, Stephen Shea, and the referees for stimulating<br />

comments.<br />

References<br />

[1] R. Piazza, A. Parola, J. Phys.: Cond. Matter 20 (2008) 153102.<br />

[2] M. Braibanti, D. Vigolo, R. Piazza, Phys. Rev. Lett 100 (2008) 108303.<br />

[3] S. Duhr, D. Braun, Phys. Rev. Lett. 96 (2006) 168301.<br />

[4] E. Bringuier, A. Bourdon, Phys. Rev. E 67 (2003) 011404;<br />

E. Bringuier, A. Bourdon, Physica A 385 (2007) 9.<br />

[5] M. Giglio, A. Vendramini, Phys. Rev. Lett. 38 (1977) 26.<br />

[6] B.-J. de Gans, R. Kita, B. Müller, S. Wiegand, J. Chem. Phys. 118 (2003) 8073.<br />

[7] S. Iacopini, R. Piazza, Europhys. Lett. 63 (2003) 247.<br />

[8] I. Derényi, R.D. Astumian, Phys. Rev. E 59 (1999) R6219.<br />

[9] T. Hondou, K. Sekimoto, Phys. Rev. E 62 (2000) 6021.<br />

[10] R. Benjamin, R. Kawai, Phys. Rev. E 77 (2008) 051132.<br />

[11] N.G. van Kampen, IBM J. Res. Develop. 32 (1988) 107;<br />

N.G. van Kampen, J. Phys. Chem. Solids 49 (1988) 673.<br />

[12] G. Nicolis, J. Chem. Phys. 43 (1965) 1110.<br />

[13] D.N. Zubarev, A.G. B<strong>as</strong>hkirov, Physica 39 (1968) 334.<br />

[14] J.-E. Shea, I. Oppenheim, J. Phys. Chem. 100 (1996) 19035.<br />

[15] M.E. Widder, U.M. Titulaer, Physica A 154 (1989) 452.<br />

[16] J.M. Sancho, M. San Miguel, D. Dürr, J. Stat. Phys. 28 (1982) 291.<br />

[17] J.M. Luttinger, Phys. Rev. 135 (1964) A1505.<br />

[18] A.L. Efros, Sov. Phys. JETP 23 (1966) 536.<br />

[19] A.G. B<strong>as</strong>hkirov, Theor. Math. Phys. 1 (1969) 213.<br />

[20] S. Goldstein, Quart. J. Mech. Appl. Math. 4 (1951) 129.<br />

[21] M. Kac, Rocky Mountain J. Math. 4 (1974) 497.<br />

[22] G.H. Weiss, Physica A 311 (2002) 381.<br />

[23] J. M<strong>as</strong>oliver, G.H. Weiss, Phys. Rev. E 49 (1994) 3852.<br />

[24] J. M<strong>as</strong>oliver, J.M. Porra, G.H. Weiss, Phys. Rev. E 45 (1992) 2222.<br />

[25] A.V. Plyukhin, K.S. Kim, Phys. Rev. E 61 (2000) 3207.<br />

[26] E. Zauderer, Partial Differential Equations of Applied Mathematics, Wiley, New<br />

York, 1983.<br />

[27] M. Büttiker, Z. Phys. B 68 (1987) 161.<br />

[28] R. Landauer, J. Stat. Phys. 53 (1988) 233.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!