+ All Categories
Home > Documents > Mode-coupling glass transition in a fluid confined by a periodic potential

Mode-coupling glass transition in a fluid confined by a periodic potential

Date post: 10-Oct-2016
Category:
Upload: sriram
View: 214 times
Download: 2 times
Share this document with a friend
8
PHYSICAL REVIEW E 84, 061501 (2011) Mode-coupling glass transition in a fluid confined by a periodic potential Saroj Kumar Nandi, 1,* Sarika Maitra Bhattacharyya, 2,and Sriram Ramaswamy 1,1 Centre for Condensed Matter Theory, Department of Physics, Indian Institute of Science, Bangalore 560 012, India 2 Complex Fluids & Polymer Engineering Group, National Chemical Laboratory, Dr. Homi Bhabha Road, Pune 411 008, India (Received 15 April 2011; revised manuscript received 29 October 2011; published 2 December 2011) We show that a fluid under strong spatially periodic confinement displays a glass transition within mode- coupling theory at a much lower density than the corresponding bulk system. We use fluctuating hydrodynamics, with confinement imposed through a periodic potential whose wavelength plays an important role in our treatment. To make the calculation tractable we implement a detailed calculation in one dimension. Although we do not expect simple 1d fluids to show a glass transition, our results are indicative of the behavior expected in higher dimensions. In a certain region of parameter space we observe a three-step relaxation reported recently in computer simulations [S. H. Krishnan, Ph.D. thesis, Indian Institute of Science (2005); Kim et al., Eur. Phys. J. Special Topics 189, 135 (2010)] and a glass-glass transition. We compare our results to those of Krakoviack [Phys. Rev. E 75, 031503 (2007)] and Lang et al. [Phys. Rev. Lett. 105, 125701 (2010)]. DOI: 10.1103/PhysRevE.84.061501 PACS number(s): 64.70.P, 64.70.QI. INTRODUCTION AND RESULTS Confined between two atomically smooth surfaces a few molecular diameters apart, fluids depart markedly from their bulk behavior, displaying a prodigious increase in viscosity and structural relaxation times, and shear thinning and vis- coelasticity at remarkably low shear rates and frequencies. The consensus from experiments [13] in a Surface Force Apparatus adapted for shear studies [4] and simulations [513] is that strong confinement moves the system into the regime of the glass transition. Such a trend is indeed found theoretically [14,15] by adapting mode-coupling theory (MCT) [1618] to the case of a fluid confined in a porous medium modeled as a random aggregate of hard spheres. MCT for a fluid between two smooth planar walls shows similar slowing down [19]. Strictly planar walls, however, exert no force parallel to themselves, so that the dynamics remains momentum-conserving, and long-wavelength density fluctuations travel as sound waves in directions parallel to the walls, unless no-slip is imposed by hand. Questions about the nature of finite-size scaling [20,21] at the glass transition, and the related issue of cooperatively rearranging regions [7], provide further motivation for studying the effect of confinement on the slowing down of the dynamics of a liquid. The extraction of viscosity and friction parame- ters from confined fluid flow experiments is discussed in Refs. [22,23]. In this paper we study the dynamics of a confined dense fluid, in a coarse-grained approach. Confinement enters the theory through an external potential; the resulting static inhomogeneous density background is a surrogate for the potential in our calculation. The mean density and temperature of the system are encoded in the static structure factor of the fluid in the absence of confinement. Our work differs in detail from those of Refs. [14,19] in our use of the fluctuating * [email protected] [email protected] [email protected]; also at JNCASR, Bangalore 560 064, India. hydrodynamic approach (see, e.g., Ref. [18]), encoding the interactions between particles and confining medium in the free-energy functional, and examining the problem in detail in one space dimension. We summarize our main results below and in Figs. 15: (1) We present a particularly transparent derivation, from the equations of fluctuating hydrodynamics, of the mode-coupling equations for the memory function and time-correlation function of the density field of a confined fluid. (2) In the fluid phase, confinement renders the density dynamics diffusive at long wavelengths, with a diffusivity calculable, via mode coupling, in terms of properties of the inhomogeneous background density field. (3) We show in detail (see Fig. 2) that strong enough confinement can drive the system through a glass transition, in conditions under which the system in bulk would be a fluid. To make the calculation tractable we work in one dimension, where a glass transition is unlikely for a fluid without confinement. However, the structure of the calculation makes it clear that similar behavior is expected in realistic higher-dimensional systems. (4) The strength of potential required to produce the transition is lowest when the wavelength of the potential matches the length scale corresponding to the structure factor peak of the fluid (Fig. 3). (5) For densities ρ 0 below a value ρ c that depends on ,a continuous onset of the glassy state is observed as the potential strength is increased (Fig. 4). (6) In the glassy state for ρ 0 ρ c , we predict a three-step relaxation of the density correlation function (Fig. 5), as seen in recent molecular dynamics studies [24,25] of confined fluids. (7) Correspondingly, for a certain range of densities, two thresholds of confinement strength are seen for the onset of the nonergodicity parameter: first a continuous onset, then a discontinuous jump. The prediction of a transition in one dimension, the finding that a periodic background suffices to enhance the transition, the emergence of diffusive dynamics for the density, and the multistep relaxation of the collective density correlator in the continuous transition regime are all features that distinguish our work from existing theoretical studies [14,19] of the glass transition under confinement. In particular, our approach offers a natural MCT-based explanation of the multiple plateaux in the intermediate scattering function seen in the simulations 061501-1 1539-3755/2011/84(6)/061501(8) ©2011 American Physical Society
Transcript
Page 1: Mode-coupling glass transition in a fluid confined by a periodic potential

PHYSICAL REVIEW E 84, 061501 (2011)

Mode-coupling glass transition in a fluid confined by a periodic potential

Saroj Kumar Nandi,1,* Sarika Maitra Bhattacharyya,2,† and Sriram Ramaswamy1,‡1Centre for Condensed Matter Theory, Department of Physics, Indian Institute of Science, Bangalore 560 012, India

2Complex Fluids & Polymer Engineering Group, National Chemical Laboratory, Dr. Homi Bhabha Road, Pune 411 008, India(Received 15 April 2011; revised manuscript received 29 October 2011; published 2 December 2011)

We show that a fluid under strong spatially periodic confinement displays a glass transition within mode-coupling theory at a much lower density than the corresponding bulk system. We use fluctuating hydrodynamics,with confinement imposed through a periodic potential whose wavelength plays an important role in our treatment.To make the calculation tractable we implement a detailed calculation in one dimension. Although we do notexpect simple 1d fluids to show a glass transition, our results are indicative of the behavior expected in higherdimensions. In a certain region of parameter space we observe a three-step relaxation reported recently incomputer simulations [S. H. Krishnan, Ph.D. thesis, Indian Institute of Science (2005); Kim et al., Eur. Phys.J. Special Topics 189, 135 (2010)] and a glass-glass transition. We compare our results to those of Krakoviack[Phys. Rev. E 75, 031503 (2007)] and Lang et al. [Phys. Rev. Lett. 105, 125701 (2010)].

DOI: 10.1103/PhysRevE.84.061501 PACS number(s): 64.70.P−, 64.70.Q−

I. INTRODUCTION AND RESULTS

Confined between two atomically smooth surfaces a fewmolecular diameters apart, fluids depart markedly from theirbulk behavior, displaying a prodigious increase in viscosityand structural relaxation times, and shear thinning and vis-coelasticity at remarkably low shear rates and frequencies.The consensus from experiments [1–3] in a Surface ForceApparatus adapted for shear studies [4] and simulations[5–13] is that strong confinement moves the system intothe regime of the glass transition. Such a trend is indeedfound theoretically [14,15] by adapting mode-coupling theory(MCT) [16–18] to the case of a fluid confined in a porousmedium modeled as a random aggregate of hard spheres.MCT for a fluid between two smooth planar walls showssimilar slowing down [19]. Strictly planar walls, however,exert no force parallel to themselves, so that the dynamicsremains momentum-conserving, and long-wavelength densityfluctuations travel as sound waves in directions parallel tothe walls, unless no-slip is imposed by hand. Questionsabout the nature of finite-size scaling [20,21] at the glasstransition, and the related issue of cooperatively rearrangingregions [7], provide further motivation for studying the effectof confinement on the slowing down of the dynamics ofa liquid. The extraction of viscosity and friction parame-ters from confined fluid flow experiments is discussed inRefs. [22,23].

In this paper we study the dynamics of a confined densefluid, in a coarse-grained approach. Confinement enters thetheory through an external potential; the resulting staticinhomogeneous density background is a surrogate for thepotential in our calculation. The mean density and temperatureof the system are encoded in the static structure factor ofthe fluid in the absence of confinement. Our work differs indetail from those of Refs. [14,19] in our use of the fluctuating

*[email protected][email protected][email protected]; also at JNCASR, Bangalore 560 064,

India.

hydrodynamic approach (see, e.g., Ref. [18]), encoding theinteractions between particles and confining medium in thefree-energy functional, and examining the problem in detailin one space dimension.

We summarize our main results below and in Figs. 1–5:(1) We present a particularly transparent derivation, from theequations of fluctuating hydrodynamics, of the mode-couplingequations for the memory function and time-correlationfunction of the density field of a confined fluid. (2) Inthe fluid phase, confinement renders the density dynamicsdiffusive at long wavelengths, with a diffusivity calculable, viamode coupling, in terms of properties of the inhomogeneousbackground density field. (3) We show in detail (see Fig. 2)that strong enough confinement can drive the system througha glass transition, in conditions under which the system inbulk would be a fluid. To make the calculation tractable wework in one dimension, where a glass transition is unlikelyfor a fluid without confinement. However, the structure of thecalculation makes it clear that similar behavior is expectedin realistic higher-dimensional systems. (4) The strength ofpotential required to produce the transition is lowest whenthe wavelength � of the potential matches the length scalecorresponding to the structure factor peak of the fluid (Fig. 3).(5) For densities ρ0 below a value ρc that depends on �, acontinuous onset of the glassy state is observed as the potentialstrength is increased (Fig. 4). (6) In the glassy state for ρ0 � ρc,we predict a three-step relaxation of the density correlationfunction (Fig. 5), as seen in recent molecular dynamics studies[24,25] of confined fluids. (7) Correspondingly, for a certainrange of densities, two thresholds of confinement strengthare seen for the onset of the nonergodicity parameter: firsta continuous onset, then a discontinuous jump.

The prediction of a transition in one dimension, the findingthat a periodic background suffices to enhance the transition,the emergence of diffusive dynamics for the density, and themultistep relaxation of the collective density correlator in thecontinuous transition regime are all features that distinguishour work from existing theoretical studies [14,19] of the glasstransition under confinement. In particular, our approach offersa natural MCT-based explanation of the multiple plateaux inthe intermediate scattering function seen in the simulations

061501-11539-3755/2011/84(6)/061501(8) ©2011 American Physical Society

Page 2: Mode-coupling glass transition in a fluid confined by a periodic potential

NANDI, BHATTACHARYYA, AND RAMASWAMY PHYSICAL REVIEW E 84, 061501 (2011)

−20−10

010

c k0 10 20 30 40

0123

kS

k(0)

0 10 20 300

1

2

2.7

k

δmk

0 10 20 300

1

2

3

4

k

S

k(0)

Sk

δm(x)

(d)

(b)

(a)

(c)

FIG. 1. (Color online) (a) A schematic presentation of a one-dimensional fluid confined by periodic external walls, which wereexpress in terms of a periodic mean density profile. (b) The directcorrelation function ck and the static structure factor S

(0)k for a

one-dimensional bulk fluid. (c) The Fourier transformed backgrounddensity δmk . (d) The bulk structure factor S

(0)k and its modification Sk

due to the external potential.

of Refs. [24,25]. The double onset of the nonergodicityparameter seen in our calculations provides a much strongerexample of the confinement-induced glass-glass transitionthan that seen in Ref. [14] and is similar to the behavior foundin Ref. [15] for the tagged-particle density, for confinement ina disordered medium.

The rest of this paper is organized as follows. Section IIbegins with a discussion of possible dynamical effects ofconfinement in Sec. II A and continues in Sec. II B with generalconclusions from MCT for the fluctuating hydrodynamics of

confined fluids. Section III presents our calculation in detailfor a model one-dimensional fluid, leading to a phase diagramas a function of density and confinement strength. We close inSec. IV with a discussion and summary.

II. MODE-COUPLING THEORY FOR A FLUID IN ACONFINING POTENTIAL

A. Remarks on the dynamics of confined fluids

Before entering into the calculations that led to the aboveresults, some general remarks are necessary. The viscositymeasured in the surface force apparatus (SFA) experimentscorresponds to gradients along the z direction normal tothe plates and velocity in the xy =⊥ plane. The densitymodulations contributing to these components of the viscositymust have a wave-vector component in the ⊥ plane, becauseshear with gradient along z does not directly affect structuresvarying only along z. In fact, the relevant fourier modes ofthe density must have wave vectors with nonzero z and ⊥components. Layering alone, with wave vector only along z,will not suffice. Next, what features of confinement are re-sponsible for the dramatic slowing down? It seems reasonablyclear that the pressure applied in the SFA experiments is notdirectly responsible, as it does not amount to a bulk hydrostaticcompression, but simply goes into determining the thickness ofthe confined fluid layer. The confining walls rule out motionnormal to their surface; they limit motion parallel to theirsurface through no-slip; and they alter the static structure of thefluid through steric or potential interactions. Can no-slip aloneproduce a significant slowing down of the density? To makethis question concrete, consider a fluid with viscosity η andvelocity field v, with density field ρ governed by a free-energyfunctional F [ρ] (for example, a Ramakrishnan-Yussouff [26]functional), lying between a pair of walls parallel to the plane,separated by distance w along the z direction. If we imposeno-slip at the walls, the dynamics of the in-plane velocityfield, on length scales in the ⊥ plane large compared to w

is governed by lubrication: (η/w2)v⊥ � −ρ∇⊥δF/δρ. The

−5 −2 1 4 7 100

0.2

0.4

0.6

0.8

1

log10

(time)

φ k=5.

0(t)

ρ0=0.7715

ρ0=0.7720

ρ0=0.7722

ρ0=0.7724

ρ0=0.7725

ρ0=0.7726

−4 −2 0 2 4 60

0.2

0.4

0.6

0.8

1

log10

(time)

φ k=5.

0(t)

M=10.5× 10−4

M=11.0× 10−4

M=11.5× 10−4

M=11.8× 10−4

M=12.0× 10−4

M=12.2× 10−4

(a) (b)

FIG. 2. (Color online) (a) The relaxation of the normalized coherent intermediate scattering function φk(t) for k = 5.0 for an unconfinedone-dimensional fluid shows the mode-coupling transition at density ρ0 = 0.7726. (b) We set the density ρ0 = 0.70 so that the system is ina fluid state [φk(t) relaxes to zero] in the absence of an external potential. For an external potential with period � = 1.3, corresponding tofirst maximum of the structure factor of the fluid without confinement, we increase M from a small value and observe the MCT transition atM = 12.2 × 10−4. The plot shows the relaxation of φk(t) for k = 5.0.

061501-2

Page 3: Mode-coupling glass transition in a fluid confined by a periodic potential

MODE-COUPLING GLASS TRANSITION IN A FLUID . . . PHYSICAL REVIEW E 84, 061501 (2011)

continuity equation ∂tρ = −∇ · (ρv) then implies an effectivetwo-dimensional dynamics:

∂tρ = ∇⊥ ·(

w2ρ

ηρ∇⊥

δFeff

δρ

)+ ∇⊥ · f⊥. (1)

In writing (1) we have implicitly averaged over z, so that ρ

depends only on x and y, and we have written the z-averagedforce density on the right-hand side in a form similar tothe bulk, with an effective free-energy density Feff. Equation(1) is of the type analyzed in Refs. [27,28], and f is amultiplicative noise whose form [27,29,30] guarantees thatequal-time correlations of the two-dimensional density fieldare governed by Feff. To linear order in density fluctuations, atlong wavelength, (1) is a diffusion equation, with diffusivitykBT w2η−1ρ0(1 − ρ0cq=0), where cq is the direct pair correla-tion function of the confined fluid, and kB and T , respectively,are Boltzmann’s constant and absolute temperature. Nonlineareffects enter via the interactions embodied in Feff and can leadto a slowing down with increasing density and decreasingtemperature [27,31]. The confinement scale w in Eq. (1)appears in two different ways: one, explicitly, in determiningthe diffusivity ∼w2/η, and two, implicitly, within Feff. Thevanishing of the bare diffusivity ∝w2, a consequence of no-slipalone, can readily be absorbed into a rescaling of the time inEq. (1) and cannot be further enhanced in an MCT feedbackmechanism [32]. A decrease in w can lead to a glass transitiononly through structural inputs via Feff. We are thus obliged toconsider the effect of the walls on the structure of the fluid andthence on its dynamics [33].

B. Fluctuating hydrodynamics of confined fluids

As we remarked in the Sec. I, our treatment of the effectof confinement will replace walls by an external potential.Consider a fluid with velocity field v and density [34] fieldρ with mean ρ0, in the presence of an externally imposedpotential U (x). The equations of fluctuating hydrodynamicsfor an isothermal fluid, extended down to the length scalesrelevant to the slow dynamics of a fluid near structural arrest[18], are the continuity equation

∂tρ + ∇ · (ρv) = 0 (2)

and the generalized Navier-Stokes equation

ρ(∂t + v · ∇)v = η∇2v + (ζ + η/3)∇∇ · v − ρ∇ δFU

δρ+ f,

(3)

where ζ and η are the bare bulk and shear viscosities, andthermal fluctuations enter through the Gaussian white noise fwith 〈f(0,0)f(r,t)〉 = −2kBT [ηI∇2 + (ζ + η/3)∇∇]δ(r)δ(t).

Let us linearize the equations of motion by ignoring δρv in(2) and (3), and replace velocity in the divergence of Eq. (3)using Eq. (2) to obtain the equation of motion for the densityfluctuation alone:

∂2δρ(r,t)∂t2

= DL 2 ∂δρ(r,t)∂t

+ ∇ ·(

ρ∇ δFU

δρ

)+ ∇ · f,

(4)

where DL = (ζ + 4η/3)/ρ0. We take the space-Fourier trans-form of the above equation and obtain

∂2δρk(t)

∂t2+ k

∂δρk(t)

∂t=

[∇ ·

(ρ∇ δFU

δρ

) ]k

− ik · fLk ,

(5)

where k = DLk2 is the bare longitudinal dampingcoefficient.

The confining potential U has been incorporated into thedensity-wave free-energy functional [26]:

βFU =∫

r

{ρ(r) ln

ρ(r)

ρ0− [ρ(r) − ρ0]

}

− 1

2

∫r,r′

c(r − r′)[ρ(r) − ρ0][ρ(r′) − ρ0]

∫rU (r)ρ(r), (6)

where∫

r ≡ ∫dr, β = 1/kBT , and c(r − r′), the direct pair

correlation function in the absence of U , is a coarse-grainedexpression of the intermolecular interactions in the fluid. Ourreference state is m(r), the inhomogeneous equilibrium densityfield in the presence of U , which satisfies δFU [m]/δm(r) = 0,i.e.,

lnm(r)

ρ0= −βU (r) +

∫r′

c(r − r′)δm(r′), (7)

where δm(r) ≡ m(r) − ρ0. Writing ρ(r,t) = ρ0 + δm(r) +δρ(r,t), the force density from (6), after replacing U (r) interms of m(r) using Eq. (7), takes the form

ρ∇ δβFU

δρ(r,t)= ∇

∫r′

[δ(r − r′) − ρ0c(r − r′)]δρ(r′,t)

−∇∫

r′δm(r)c(r − r′)δρ(r′,t)

− ∇m(r)

m(r)

∫r′

[δ(r − r′) − m(r)c(r − r′)]δρ(r′,t)

− δρ(r,t)∇∫

r′c(r − r′)δρ(r′,t). (8)

Note that the effect of confinement is contained only inthe background density field m(r); U itself does not appearexplicitly. The hydrodynamic equations (2) and (3) with (8)then readily yield the dynamical equation [35](

∂2t + k∂t + kBT k2

S(0)k

)δρk = −ik · (Fk + fk)

≡ −ik · (Fmρ

k + Fρρ

k + fLk

)(9)

for the spatial Fourier transform δρk(t) of the density field,with force densities

Fmρ

k (t) = ikBT

∫q

[kcq + (k − q)/ρ0S

(0)q

]δmk−qδρq(t) (10)

arising from interaction with the static inhomogeneousbackground, to first order in δm(r), which suffices for theone-loop treatment we will present [36], and

Fρρ

k (t) = i

2kBT

∫q[qcq + (k − q)ck−q]δρq(t)δρk−q(t) (11)

061501-3

Page 4: Mode-coupling glass transition in a fluid confined by a periodic potential

NANDI, BHATTACHARYYA, AND RAMASWAMY PHYSICAL REVIEW E 84, 061501 (2011)

from the pairwise interaction of fluid density fluctuations. InEq. (9) fL

k (t) is the longitudinal part of the Fourier transformof the bare noise f in Eq. (3), and S

(0)q in Eqs. (9) and (10) is the

static structure factor of the bulk fluid without confinement.In Eqs. (10) and (11) and hereafter

∫q denotes

∫ddq/(2π )d .

The autocorrelation of fLk (t) is linked to the bare longitudinal

damping k ≡ DLk2, and the Kubo formula [37] tells us thatthe excess damping due to interactions, expressed in the timedomain, is given by the memory function

Mk(t) = 1

kBT V〈Fk(0) · F−k(t)〉, (12)

where V → ∞ is the system volume, and Fk(t) is asin Eqs. (9)–(11).

We define translation-invariant correlation functionsSk(t) ≡ ∫

ddr exp(−ik · r)[〈δρ(r0)(0)δρ(r0 + r)(t)〉]r0 where[ ]r0 denotes an average over r0 over a period of the back-ground. Within a Gaussian decoupling approximation (9)–(12)lead to [38]

φk(t) + DLk2φk(t) + kBT k2

S(0)k

φk(t)

+∫ t

0Mk(t − τ )φk(τ ) dτ = 0 (13)

for φk(t) ≡ Sk(t)/Sk(0), with

Mk(t) = C1

2k2

∫q[k · qcq + k · (k − q)ck−q]2Sk−q(t)Sq(t)

+ C1

k2

∫q[k · qcq + k · (k − q)/ρ0]2Sb

k−qSq(t),

(14)

where C1 = kBTρ0 and Sbk ≡ ∫

ddr exp(−ik · r)[δm(r0)δm(r0 + r)]r0 is the structure factor of the backgrounddensity field. Equations (13) and (14) are the equations ofmode-coupling theory (MCT) for a confined fluid, as aninitial value problem for Sk(t) given the static structure factorSk ≡ Sk(t = 0) of the confined fluid. These equations areidentical to those obtained by Krakoviack [14] through theprojection operator formalism for the case of a fluid confinedby a porous medium. Our approach offers a simple andtransparent derivation of these results.

Irrespective of the details of the confining medium, (14)leads us to our first result, an important general featureof confined MCT. Of the two contributions to Mk(t) onthe right-hand side of Eq. (14), the second, coming fromthe interaction of dynamic density fluctuations with thebackground density field, is nonvanishing for wave vectork → 0. The interaction of the fluid with inhomogeneitiesof the confining medium damps the flow even at zero wavenumber, and MCT provides a convenient way to calculate thiseffective “Darcy” damping. It is this contribution that leadsto an effective no-slip condition on confining walls endowedwith structure, periodic or otherwise. As a consequence, forlong time scales and small k the third and fourth terms onthe left-hand side of Eq. (13) dominate, reducing it, as isto be expected [39,40], to a diffusion equation for φk(t)with collective diffusivity kBT /Sk=0M00, where M00 is theFourier transform of the memory function at zero wave

number and frequency. The crossover to Brownian ratherthan Newtonian dynamics assumed for convenience in Refs.[14,41,42] emerges at small wave number in a calculable formas a result of interaction with the confining medium.

III. MODE COUPLING IN ONE DIMENSION IN APERIODIC POTENTIAL

The theory presented in the previous section is quitegeneral and can, in principle, be applied in any dimen-sion. The mode-coupling calculation for a system withspatially periodic confinement in two or three dimensionsmakes enormous demands on computational time, because ofthe loss of isotropy. We therefore chose to display the principleof the calculation by working in one dimension. We areaware that a simple fluid of, say, hard rods in one spacedimension has no crystalline phase and therefore cannotbe supercooled or overcompressed, rendering it a poor can-didate for a glass transition. As MCT is relatively insensitiveto spatial dimensionality, we expect nonetheless that the trendsin our one-dimensional MCT treatment will be found in calcu-lations in dimensions �2. We will return to the question of thebehavior of one-dimensional systems at the end of the paper.

In order to implement our calculation we need an expressionfor Sk for the confined fluid. Rather than taking it fromexperiment or liquid theory, we express it in terms of thestructure factor S

(0)k of the bulk fluid that we treat as given. We

expand (6) around the static inhomogeneous density m(r); Sk

0 5 10 15 200

0.5

1

1.5

2

2.5

k

Sk(0

)

1 1.2 1.4 1.6 1.80

2

4

6

8

l

M ×

104

ρ0=0.70

ρ0=0.60

ρ0=0.70

0 2 4 60.7

0.74

0.78

0.8

M × 104

ρ 0

l=1.30l=1.10l=0.80l=0.55

0 2 4 60.1

0.3

0.5

0.7

0.8

M × 104

ρ 0

l=1.30l=1.10l=0.80l=0.55

l=1.9

l=0.55

(a) (b)

(d)

LIQUID

LIQUID

GLASS

GLASS

GLASS

LIQUID

l=1.3

(c)

FIG. 3. (Color online) (a) Static structure factor S(0)k for an

unconfined one-dimensional fluid at a density ρ0 = 0.70. The variousl values marked on the k axis in the figure correspond to the peakposition of δmk [Fig. 1(c)]. (b) The phase diagram for M vs � fortwo different densities. The threshold of M is quite sensitive to � forρ0 = 0.70, whereas it is less sensitive at ρ0 = 0.60. (c) Phase diagramfor ρ0 vs M for four values of �. (d) As in (c), for a larger range ofρ0. In both (c) and (d) filled and open symbols respectively indicatecontinuous and discontinuous transitions. For clarity, the extensionof the discontinuous transition line beyond the point where it crossesthe continuous transition is not shown here [see Fig. 5(a)].

061501-4

Page 5: Mode-coupling glass transition in a fluid confined by a periodic potential

MODE-COUPLING GLASS TRANSITION IN A FLUID . . . PHYSICAL REVIEW E 84, 061501 (2011)

−4 −2 0 2 4 60

0.2

0.4

0.6

0.8

1

log10

(time)

φ k=5.

0(t)

M=7.0×10−4

M=7.2×10−4

M=7.6×10−4

M=7.8×10−4

M=8.2×10−4

M=8.5×10−4

FIG. 4. (Color online) At density ρ0 = 0.60 and � = 0.80, thenonergodicity parameter fk = limt→∞ φk(t) shows a continuousonset to the MCT glass transition with increasing strength of theexternal potential. The plot shows the relaxation of φk(t) for k = 5.0.

is then determined by the coefficient of the term quadratic inδρ. This crude approximation is adequate for the purpose ofillustration and can be improved upon if necessary. To linearorder in δmk it is straightforward to see that

Sk = S(0)k + 1

ρ0S

(0)k δmkS

(0)k . (15)

We must solve the closed set of equations (13)–(15)numerically to obtain the behavior of our confined fluid. Asremarked above, we do this for space dimension d = 1. Thecalculation can be viewed as a schematic MCT for the confinedfluid problem in which the structure of the confining walls istaken into account in a simple manner.

We first solved the MCT equations (13) and (14) withoutconfinement, in one dimension, using the direct correlationfunction for a one-dimensional hard-rod fluid [43] as inputand found an MCT glass transition at ρ0 = 0.7726 [Fig. 2(a)].The nature of the MCT calculation in d = 1 deserves someexplanation. Through the quadratic nonlinearity, one mode atwave vector k couples to two others at k1, k2 with k1 + k2 = k.For d � 2 all three of these modes can be taken to lie on thefirst shell of maxima |k| = k0 of the structure factor S(k). Howthen do we understand the MCT glass transition qualitativelyfor d = 1, where the “shell” is two points? The answer: Fora one-dimensional fluid the higher-order peaks of S(k) carrysubstantial weight; the dominant triple of modes coupled byMCT must consist of two with |k| = k0 and one with |k| = 2k0.Having established an MCT glass transition in d = 1, we nowproceed to examine its modification by confinement.

Consider one-dimensional confinement by a periodic poten-tial U (x) = U sin 2πx

�, in units of kBT , where the parameters

U and � control the strength and periodicity, respectively, ofthe potential. Given u(x) we can construct the backgrounddensity field δm(x) = m(x) − ρ0 as in Eq. (7). Instead ofspecifying u(x), we will therefore characterize our confiningmedium by specifying δm(x), which we take for simplicityto be sinusoidal, δm(x) � M sin 2πx/�. We set the densityρ0 = 0.70, for which the system in the absence of a confiningpotential is in the fluid state, and take � = 1.3, which meansthe period of the external potential is 2π over the wave numberat which the structure factor of the bulk one-dimensional fluid

has its primary peak. As M is increased past a threshold ofabout 12.2 × 10−4 we find, solving (13) and (14) with (15),that the mean relaxation time obtained from φk(t) diverges,and the nonergodicity parameter fk ≡ limt→∞ φk(t) jumps toa nonzero value [Fig. 2(b)]. This is the one-dimensional MCTglass transition induced by confinement, i.e., by a periodicpotential. Keeping ρ0 fixed, we ask how changing � affectsthe threshold M for the transition. In Fig. 3(b) we scanaround the first peak of the structure factor and find that athigh density, the threshold value of M is the lowest, i.e.,the strongest enhancement of the transition is achieved, ata value of � = 1.3 corresponding to the peak. Values of �

on either side of 1.3 require a larger threshold value of M .However, at low density, the threshold of M is very insensitiveto �. This can be understood as follows. At higher densities,within MCT, the particles form a cage and � = 1.3, beingcompatible with the preferred interparticle distance, facilitatesthis caging. For other values of � the two length scales aredifferent and, hence, δmk is less effective. At low density,there is no caging, so varying � has little effect. It is importantto note that the transition at ρ0 = 0.60 is always continuous,whereas at ρ0 = 0.70 it can be continuous or discontinuousdepending on the value of � [Fig. 3(d)].

We list several features of our phase diagram that areconsistent with the findings of [14] for fluids in disorderedporous media. For moderately high-density ρ0, the effect ofthe external potential can be seen as enhancing the correlationsalready present in the fluid, i.e., promoting the effect of thequadratic term in the MCT equation (14). At substantiallylower densities, where correlations in the bulk fluid are weak,it is still possible within MCT to get a transition to a nonergodicstate by increasing M , but now the dominant role is presumablyplayed by the term linear in Sq(t) in Eq. (14). Correspondingly,for densities ρ0 lower than a value ρc that depends on �, thetransition turns continuous (Fig. 4). Quantitatively, however,the medium has a weaker effect at high densities than inRef. [14]. More puzzling, it is substantially more effectiveat low densities than in Ref. [14]. Reentrance is anotherfeature observed in common with Ref. [14]: At densitiessomewhat below ρc the threshold value of M decreases withdecreasing density. It is surprising that we observe it inour simple one-dimensional periodic system, and we suggesta simple interpretation. At low densities, the glassy state,if it exists, presumably comes from a feedback-enhancedslowing down of single-particle crossings of barriers posedby the external potential. As density is increased, interparticlerepulsion becomes more important, and particles typicallyoccupy regions not near the minima of the external potential.Thus, repulsion lowers the effective kinetic barrier to motion.Possibly our explanation is related to that proposed in Ref. [14],but this is unclear.

On the continuous MCT-glass transition line, there arepossible difficulties with infrared divergences in the mode-coupling integral (14) because we are working in one di-mension. This issue arises in a pronounced manner for asingle particle in a disordered medium [44], where there isno transition in d = 1 because the particle is always localized.There are, however, important differences between our systemand that of [44]. Our model is probably free from the k → 0problem because the background medium in our case is

061501-5

Page 6: Mode-coupling glass transition in a fluid confined by a periodic potential

NANDI, BHATTACHARYYA, AND RAMASWAMY PHYSICAL REVIEW E 84, 061501 (2011)

0 2 4 6 8 10

0.6

0.65

0.7

0.75

0.8

M×104

ρ 0

−5 0 5 100

0.2

0.4

0.6

0.8

1

log10

(time)

φ k=5.

0(t)

M=6.60×10−4

M=6.90×10−4

M=7.20×10−4

M=7.40×10−4

M=7.46×10−4

M=7.48×10−4

M=7.49×10−4

GLASS

GLASS

(b)(a)

D

C

A

LIQUIDB

FIG. 5. (Color online) (a) For 0.65 < ρ0 < 0.72, � = 0.8, a continuous transition from liquid to glass (segment BD) is followed by adiscontinuous glass-glass transition (segment BC). Note, however, that we do not resolve the phase boundary close to point B. (b) If the densityρ0 is close to but lower than ρc, the threshold density where the transition switches from discontinuous to continuous, the density correlationfunction shows a three-step relaxation for strong enough background density. The figure shows the relaxation of φk=5(t) at a density ρ0 = 0.70as a function of log(time) as we increase the strength of background density with � = 0.80.

periodic, not disordered, and lacks weight at k = 0. We havenevertheless checked that varying the lowest wave numberover a limited range (from 0.1 to 0.2) does not alter, to 0.1%accuracy, the MCT transition density as estimated by the onsetof a time-persistent density correlator.

For ρ0 moderately large, but still below ρc, fk undergoestwo onsets as a function of M . Across a first threshold valueof M , fk rises continuously from zero, presumably drivenby the linear term in the MCT equation. Upon increasingM past a second threshold, a discontinuous jump in fk isseen, which amounts to a glass-glass transition [Fig. 5(a)].A hint of such behavior has been reported in Ref. [14]; theeffect appears far stronger in our case. The relaxation of thenormalized intermediate scattering function in this secondglassy state [Fig. 5(b)] shows two plateaux followed by athird nonzero asymptotic value. Such dynamical behavior hasbeen reported for the self-intermediate scattering function inmolecular dynamics studies of Lennard-Jones fluids underplanar confinement to a thickness of about three moleculesby structured walls [24], where it was called three-steprelaxation. More recently similar dynamics has been reportedfor the coherent part of the intermediate scattering functionfor hard spheres in random media [25]. The possibility of suchmultistep relaxation was predicted in the context of variouspossible MCT integrals [45,46], and obtained by Krakoviack[15] for tagged-particle motion, from MCT in a disorderedmedium. Such a complex relaxation scenario has also beenreported in a variety of other contexts [47–49].

IV. SUMMARY AND DISCUSSION

We have succeeded in providing an economical descriptionof the slowing down of a fluid under confinement. Wederived the mode-coupling equations through the dynamicdensity-wave approach [50] of fluctuating hydrodynamics, andimplemented confinement in the form of an externally imposedperiodic potential. In the spirit of a one-loop approach, we

retained interactions between the fluid and the backgrounddensity only to bilinear order, but this is not a serious difficulty;we could work with the background molecular field instead.In order to make our numerical calculations manageable wechose to work with a one-dimensional model and to deal withproperties averaged over one period of the potential. Once werelax the period-averaging constraint, we can predict the de-gree of slowing down as a function of location in the potential.This is the nearest analog, in our one-dimensional model, tocalculating properties as a function of distance from confiningwalls. Remarkably, our model calculation reproduces all thefeatures observed in more detailed treatments, including acrossover to a continuous transition, the phenomenon ofreentrance, at lower densities, and a glass-glass transition asa function of confinement strength at intermediate densities.Unlike in Ref. [14], quenched disorder plays no role, and thedensity dynamics is diffusive at small wave number, in contrastto Ref. [19]. In future work we will consider in detail theproblem of confinement in a planar geometry with structuredwalls, thereby improving on the treatments presented here.

Apart from the glass-glass transition, which appears as aprominent feature in our treatment unlike in Ref. [14] whereit is barely detectable, our most novel predictions are (1) thateven a one-dimensional fluid, at least within an MCT approach,can undergo a glass transition; (2) that an imposed periodicpotential causes this transition to occur at lower densitiesand higher temperatures; (3) that in a calculable range ofdensities, the relaxation of the density correlation functionshould take place through a three-step process, an effect forwhich there is evidence in molecular dynamics studies oftwo types of confined fluid systems [24,25], and as shownin MCT with confinement in a disordered medium [15] for thetagged-density correlation function.

The initial experimental observations of the effect ofconfinement were done in shear flow measurements. It istherefore imperative to extend our approach to include shear.We have done this in a simplified implementation of the ideas

061501-6

Page 7: Mode-coupling glass transition in a fluid confined by a periodic potential

MODE-COUPLING GLASS TRANSITION IN A FLUID . . . PHYSICAL REVIEW E 84, 061501 (2011)

of Ref. [51]. Our preliminary results [52] on an isotropizedversion of the calculation find that the shear-thinning movesto very low imposed flow-rates in the presence of theconfining potential. Further results require imposing shear withplanar confinement, preferably with structured walls. We areconfident that our results will change only quantitatively whenthese elements of greater realism are introduced.

Last, should one expect a glass transition, or at least anMCT-style enhancement of viscosity, in experiments or sim-ulations on one-dimensional fluids, confined or otherwise? Itmight seem unlikely, given that a one-dimensional system hasno crystalline phase past which one can supercool it to enter aregion of metastability and glassiness. However, supercoolingis not necessary for viscosity increase of the MCT variety [53],

and numerical calorimetry [54] finds precursors of glassybehavior even for one-dimensional Lennard-Jones mixtures.It might therefore be worth testing our ideas in simulationson one-dimensional fluids, perhaps multicomponent or withcomplicated interactions, with and without confinement. In anycase, we look forward to tests of our predictions in experimentsand simulations, whether in one or higher dimensions.

ACKNOWLEDGMENTS

We thank G. Ayappa, M. Cates, M. Das, C. Dasgupta,M. Fuchs, W. Kob, V. Krakoviack, and K. Miyazaki fordiscussions and the DST, India.

[1] A. L. Demirel and S. Granick, Phys. Rev. Lett. 77, 2261(1996).

[2] S. Granick, Science 253, 1374 (1991).[3] J. Klein and E. Kumacheva, Science 269, 816 (1995).[4] J. N. Israelachvili, P. M. McGuiggan, and A. M. Homola, Science

240, 189 (1988).[5] J. Gao, W. D. Luedtke, and U. Landman, Phys. Rev. Lett. 79,

705 (1997).[6] K. G. Ayappa and R. K. Mishra, J. Phys. Chem. B 111, 14299

(2007).[7] P. Scheidler, W. Kob, and K. Binder, Europhys. Lett. 59, 701

(2002).[8] P. Scheidler, W. Kob, and K. Binder, J. Phys. Chem. B 108, 6673

(2004).[9] C. Ghatak and K. G. Ayappa, Phys. Rev. E 64, 051507 (2001).

[10] P. A. Thompson and M. O. Robbins, Science 250, 792 (1990).[11] P. A. Thompson, M. O. Robbins, and G. S. Grest, Isr. J. Chem.

35, 93 (1995).[12] S. T. Cui, P. T. Cummings, and H. D. Cochran, J. Chem. Phys.

114, 7189 (2001).[13] S. T. Cui, C. McCabe, P. T. Cummings, and H. D. Cochran,

J. Chem. Phys. 118, 8941 (2003).[14] V. Krakoviack, Phys. Rev. E 75, 031503 (2007).[15] V. Krakoviack, Phys. Rev. E 79, 061501 (2009).[16] W. Gotze, Complex Dynamics of Glass-Forming Liquids:

A Mode-Coupling Theory (Oxford University Press, Oxford,2009).

[17] D. R. Reichman and P. Charbonneau, J. Stat. Mech. (2005)P05013.

[18] S. P. Das, Rev. Mod. Phys. 76, 785 (2004).[19] S. Lang, V. Botan, M. Oettel, D. Hajnal, T. Franosch, and

R. Schilling, Phys. Rev. Lett. 105, 125701 (2010).[20] S. Karmakar, C. Dasgupta, and S. Sastry, Proc. Natl. Acad. Sci.

USA 106, 3675 (2009).[21] G. Biroli, J.-P. Bouchaud, K. Miyazaki, and D. R. Reichman,

Phys. Rev. Lett. 97, 195701 (2006).[22] B. N. J. Persson and E. Tosatti, Phys. Rev. B 50, 5590 (1994).[23] S. K. Nandi, Phys. Rev. B 74, 167401 (2006).[24] S. H. Krishnan, Ph.D. thesis, Indian Institute of Science (2005).[25] K. Kim, K. Miyazaki, and S. Saito, Eur. Phys. J. Special Topics

189, 135 (2010).

[26] T. V. Ramakrishnan and M. Yussouff, Phys. Rev. B 19, 2775(1979).

[27] K. Miyazaki and D. R. Reichman, J. Phys. A: Math. Gen. 38,L343 (2005).

[28] B. Bagchi, Physica A 145, 273 (1987).[29] D. S. Dean, J. Phys. A: Math. Gen. 29, L613 (1996).[30] A. Basu and S. Ramaswamy, J. Stat. Mech. (2007) P11003.[31] T. R. Kirkpatrick and D. Thirumalai, J. Phys. A 22, L149

(1989).[32] M. Fuchs (personal communication).[33] A comparison to the work of Scheidler et al. [7,8] would require

averaging over realizations of wall configurations, which wehave not done. Even in Refs. [7,8], in a given realization of thewalls, the structure of the fluid is affected by confinement.

[34] We set the masses of the constituent particles to unity andtherefore do not distinguish between the mass density andnumber density.

[35] We neglect the advective nonlinearity ρv · ∇v in Eq. (3). Fora discussion of its effects in an MCT context see M. E. Catesand S. Ramaswamy, Phys. Rev. Lett. 96, 135701 (2006), andE. Zaccarelli et al., J. Phys. Condens. Matter 14, 2413 (2002).

[36] We could instead directly work in terms of the potential U , inwhich case the first-order approximation will not be necessary.Our choice is in order to make contact with Ref. [14].

[37] R. Kubo, J. Phys. Soc. Jpn. 12, 570 (1957).[38] K. Kawasaki, J. Stat. Phys. 110, 1249 (2003).[39] S. Ramaswamy and G. F. Mazenko, Phys. Rev. A 26, 1735

(1982).[40] M. H. J. Hagen, I. Pagonabarraga, C. P. Lowe, and D. Frenkel,

Phys. Rev. Lett. 78, 3785 (1997).[41] T. Franosch, M. Fuchs, W. Gotze, M. R. Mayr, and A. P. Singh,

Phys. Rev. E 55, 7153 (1997).[42] See Eqs. (22) and (23) of Ref. [14].[43] M. S. Wertheim, J. Math. Phys. 5, 643 (1964).[44] S. Schnyder, F. Hofling, T. Franosch, and T. Voigtmann, J. Phys.

Condens. Matter 23, 234121 (2011).[45] W. Gotze and R. Haussmann, Z. Phys. B 72, 403 (1988).[46] W. G. M. Fuchs, I. Hofacker, and A. Latz, J. Phys. Condens.

Matter 3, 5047 (1991).[47] A. Crisanti, L. Leuzzi, and M. Paoluzzi, Eur. Phys. J. E 34, 98

(2011).

061501-7

Page 8: Mode-coupling glass transition in a fluid confined by a periodic potential

NANDI, BHATTACHARYYA, AND RAMASWAMY PHYSICAL REVIEW E 84, 061501 (2011)

[48] A. Crisanti and L. Leuzzi, Phys. Rev. B 76, 184417 (2007).[49] P. Chaudhuri, L. Berthier, P. I. Hurtado, and W. Kob, Phys. Rev.

E 81, 040502(R) (2010).[50] The limitations of a continuum approach to these phenomena

that take place at the scale of a few molecular layers areultimately no worse than those of any continuum theory of theliquid-solid or liquid-glass transition.

[51] M. Fuchs and M. E. Cates, J. Rheol. 53, 957(2009).

[52] S. K. Nandi (unpublished, 2010).[53] P. Taborek, R. N. Kleiman, and D. J. Bishop, Phys. Rev. B 34,

1835 (1986).[54] R. Bruning, D. St-Onge, S. Patterson, and W. Kob, J. Phys.

Condens. Matter 21, 035117 (2009).

061501-8


Recommended