15
Approach to the glass transition studied by higher order correlation functions N. Laˇ cevi´ c and S.C. Glotzer Departments of Chemical Engineering and Materials Science and Engineering, University of Michigan, Ann Arbor, MI 48109, USA Abstract. We present a theoretical framework based on a higher order density correlation function, analogous to that used to investigate spin glasses, to describe dynamical heterogeneities in simulated glass-forming liquids. These higher order correlation functions are a four-point, time-dependent density correlation function g 4 (r, t) and corresponding “structure factor” S 4 (q,t) which measure the spatial correlations between the local liquid density at two points in space, each at two different times. g 4 (r, t) and S 4 (q,t) were extensively studied via MD simulations of a binary Lennard- Jones mixture approaching the mode coupling temperature from above in Refs. [1–5]. Here, we examine the contribution to g 4 (r, t), S 4 (q,t), and corresponding dynamical correlation length, as well as the corresponding order parameter Q(t) and generalized susceptibility χ 4 (t) from localized particles. We show that the dynamical correlation length ξ SS 4 (t) of localized particles has a maximum as a function of time t, and the value of the maximum of ξ SS 4 (t) increases steadily in the temperature range approaching the mode coupling temperature from above.

Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

Approach to the glass transition studied by higher

order correlation functions

N. Lacevic and S.C. Glotzer

Departments of Chemical Engineering and Materials Science and Engineering,University of Michigan, Ann Arbor, MI 48109, USA

Abstract.We present a theoretical framework based on a higher order density correlation

function, analogous to that used to investigate spin glasses, to describe dynamicalheterogeneities in simulated glass-forming liquids. These higher order correlationfunctions are a four-point, time-dependent density correlation function g4(r, t) andcorresponding “structure factor” S4(q, t) which measure the spatial correlationsbetween the local liquid density at two points in space, each at two different times.g4(r, t) and S4(q, t) were extensively studied via MD simulations of a binary Lennard-Jones mixture approaching the mode coupling temperature from above in Refs. [1–5].Here, we examine the contribution to g4(r, t), S4(q, t), and corresponding dynamicalcorrelation length, as well as the corresponding order parameter Q(t) and generalizedsusceptibility χ4(t) from localized particles. We show that the dynamical correlationlength ξSS

4 (t) of localized particles has a maximum as a function of time t, and the valueof the maximum of ξSS

4 (t) increases steadily in the temperature range approaching themode coupling temperature from above.

Page 2: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

2

1. Introduction

One of the most challenging problems in condensed matter physics is understanding the

dynamics and thermodynamics of glass formation [6]. On cooling from high temperature,

liquids may crystallize at Tm, or if a liquid is cooled so that crystallization is avoided,

it may become supercooled. As a supercooled liquid is further cooled, the particles

move more and more slowly, and their motion is slowed down so drastically that, at

some low T , particles will not be able to rearrange. This rearrangement is necessary

for a liquid to find its equilibrium, and at this temperature particles in a liquid appear

to be “frozen” or “jammed”, at least for the time scale of an experiment. At Tg the

supercooled liquid is no longer in equilibrium. This is called a glass transition, and

the temperature at which this occurs is called the glass transition temperature Tg. An

extensive treatment of the dynamics and thermodynamics of supercooled liquids can be

found in the textbooks by Debenedetti [7], Harrison [8] and Knight [9]. There are many

conference proceedings [10–15] on experimental developments, theoretical advances and

the role of computer simulation in the search for a universal theory of supercooled liquids

and the glass transition.

We are still left with unanswered questions about the dramatic slowing down in

the molecular motion as temperature is lowered to Tg, which would not be surprising

alone, but this process is accompanied with no significant change in the long-range

structure. How is it possible that dynamics changes so dramatically while the structure,

as measured by traditional static two-point correlation functions, remains almost

unchanged, and what type of transition is the glass transition are some of the crucial

questions in the field of glass transition research. To be able to answer those questions,

we must fully understand the dynamics of particles on the microscopic level, i.e. search

for the patterns in particle motion and relate them to a mechanism for slowing down.

The prominent features of supercooled liquids are the non-Arrhenius temperature

dependence of the viscosity and structural relaxation time [16], non-exponential

character of the structural relaxation time [17], absence of changes in static structural

quantities [18–21] and decoupling of the transport coefficients [22–25]. In particular,

the non-exponential character of the relaxation of density correlation functions and

decoupling of the transport coefficients can be rationalized with the existence of spatially

heterogeneous dynamics (SHD) or “dynamical heterogeneity” now well established in

experiment and computer simulation [26–30]. Therefore, understanding of the origin

of dynamical heterogeneity is directly related to understanding the origin of the glass

transition. We refer to a system as dynamically heterogeneous if it is possible to select

a dynamically distinguishable subset of particles by experiment or computer simulation

[31]. From the theoretical point of view, the question “what is dynamical heterogeneity”

still has different answers depending on the theoretical framework used to describe it.

Recently, a new theoretical approach to the problem of cooperativity demonstrated how

spatially heterogeneous dynamics (SHD) can arise in simple systems with cooperative

dynamics. This theory predict a growing correlation length on decreasing T [32]. Here

Page 3: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

3

we use a four-point, time-dependent density correlation function formalism to select a

dynamically distinguishable subset of particles. In particular, we investigate properties

of those particles deemed to be localized.

2. Method and model

The simulation method we use to generate data for our analyses is molecular dynamics

(MD). This is a widely used method in the investigation of supercooled liquids and

glasses that provides static and dynamic properties for a collection of particles.

The code we use in our simulations is LAMMPS [33], a parallel MD code based on

spatial decomposition parallel technique.

We study a 50/50 binary mixture of particle types “A” and “B” that interact via

the Lennard-Jones potential

Vαβ(r) = 4εαβ

[(σαβ

r

)12

−(

σαβ

r

)6]. (1)

This system has been previously studied by Wahnstrom [34] and Schrøder [35]. Following

these authors, we use length parameters σAA = 1, σBB = 5/6, and σAB = (σAA+σBB)/2,

and energy parameters εAA = εBB = εAB = 1. The masses of the particles are chosen

to be mA = 2 and mB = 1. We shift the potential and truncate it so it vanishes at

r = 2.5σAB.

We simulate a system of N = 8000 particles using periodic boundary conditions

in a cubic box of length L = 18.334 in units of σAA, which yields a density of

ρ = N/L3 = 1.296 for all state points. We report time in units of τ = (mBσ2AA/48εAA)

12 ,

length in units of σAA, and temperature, T , in units of εAA/kB, where kB is Boltzmann’s

constant. We simulate eight state points at temperatures ranging from T = 2.0 to

T = 0.59, following a path similar to that followed in Refs. [2, 35–37]. The simulations

are performed in the NV E ensemble. We estimate the mode coupling temperature

TMCT = 0.57 ± 0.01, (the glass transition temperature Tg is typically in the range

0.6TMCT < Tg < 0.9TMCT [38]) and the Kauzmann temperature T0, which can be

considered a lower bound for the glass transition temperature Tg, T0 = 0.48±0.02. How

we estimate these temperatures and other simulation details can be found in Refs. [4,5].

3. Background

In this section, we briefly review the theoretical framework of the four-point,

spatiotemporal density correlation function and corresponding structure factor. Detailed

derivation of these quantities can be found in Refs. [1–5].

We consider a liquid of N particles occupying a volume V with density ρ(r, t) =∑δ(r− ri(t)), and investigate a quantity

Q(t) =∫

dr1r2ρ(r1, 0)ρ(r2, t)w(|r1 − r2|) =N∑

i=1

N∑

j=1

w(|ri(0)− rj(t)|), (2)

Page 4: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

4

which measures the number of particles that, in time t, either remain within a distance

a of their original position, or are replaced by another particle (“overlapping particles”).

The reason for introducing an “overlap” function w is to eliminate weakly correlated

vibrational motion of the particles (for more details see e.g. Ref. [4]). The fluctuation

in Q(t) may be defined as

χ4(t) =βV

N2[〈Q(t)2〉 − 〈Q(t)〉2]. (3)

Expressing χ4(t) in terms of the four-point correlation function G4(r1, r2, r3, r4, t), we

obtain

χ4(t) =βV

N2

∫dr1dr2dr3dr4 G4(r1, r2, r3, r4, t), (4)

where

G4(r1, r2, r3, r4, t) = 〈ρ(r1, 0)ρ(r2, t)w(|r1 − r2|)ρ(r3, 0)ρ(r4, t)w(|r3 − r4|)〉− 〈ρ(r1, 0)ρ(r2, t)w(|r1 − r2|)〉× 〈ρ(r3, 0)ρ(r4, t)w(|r3 − r4|)〉. (5)

Note that in the case of both the mean-field, p-spin model and a liquid in the hypernetted

chain approximation [39–42], the time dependence of χ4(t) was calculated numerically

from an analytic expression in Ref. [1]. Those calculations provide the first analytical

prediction of the growth of a generalized dynamical susceptibility and, by inference, a

corresponding dynamical correlation length ξ4(t) in a model glass-forming system.

We wish to radially average the four-point correlation function in Eq. (5) to obtain

a function g4(r, t) that depends only on the magnitude r of the distance between two

particles at time t = 0. We start from the requirement that

χ4(t) = β∫

drg4(r, t), (6)

and obtain

g4(r, t) =1

⟨ ∑

ijkl

δ(r− rk(0) + ri(0))w(|ri(0)− rj(t)|)w(|rk(0)− rl(t)|)⟩

−⟨Q(t)

N

⟩2. (7)

Assuming an isotropic, homogeneous system, g4(r, t) is a function of r = |r| (i.e.

g4(r, t)). Details of above derivation can be found in Refs. [4, 5]. g4(r, t) describes

spatial correlations between overlapping particles separated by a distance r at the initial

time (using information at time t to label the overlapping particles). The first term in

g4(r, t) is a pair correlation function restricted to the subset of overlapping particles. The

second term represents the probability of any two randomly chosen particles overlapping

at times 0 and t. The structure factor S4(q, t) that corresponds to g4(r, t) is its Fourier

transform

S4(q, t) =∫

g4(r, t)exp[−iq · r]dr. (8)

Page 5: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

5

Eq. (8) implies that

limq→0

S4(q, t) =χ4(t)

β. (9)

Eq. (8) is analogous to the static structure factor S(q), but “scatters” off of overlapping

particles using information on overlapping particles at time t to label particles at time

0.

3.1. Self and distinct contributions to Q(t), χ4(t), g4(r, t), and S4(q, t)

The contribution of a given particle i to Q(t) is a result of three possible events: (i)

particle i remains within a distance a of its original position; (ii) particle i moves and

is replaced (within a distance a) by another particle; or (iii) particle i moves a distance

greater than a and is not replaced by another particle. Case (iii) does not count as

an overlap, and thus does not contribute to Q(t). Cases (i) and (ii) count as overlaps

and contribute to the value of Q(t). Case (ii) and (iii) belong to the set of delocalized

particles. However, the case (i) and case (ii) clearly represent two very different physical

situations. To elucidate the various contributions to the four-point correlation function,

we separate Q into self and distinct components,

Q(t) = QS(t) + QD(t) =N∑

i=1

w(|ri(t)− ri(0)|)

+N∑

i=1

N∑

i6=j

w(|ri(0)− rj(t)|). (10)

The self part, QS(t), measures the number of particles that move less than a distance

a in a time interval t; we call these “localized” particles. The distinct part, QD(t)

measures the number of particles replaced within a radius a by another particle in time

t; we call these “replaced” particles.

Following the scheme of decomposing Q(t), χ4(t) can be decomposed into self

χSS(t), distinct χDD(t), and cross χSD(t) terms: χ4(t) = χSS(t)+χDD(t)+χSD(t), where

χSS(t) ∝ 〈Q2S(t)〉−〈QS(t)〉2, χDD(t) ∝ 〈Q2

D(t)〉−〈QD(t)〉2, and χSD(t) ∝ 〈QS(t)QD(t)〉−〈QS(t)〉〈QD(t)〉. Thus χSS(t) is the susceptibility arising from fluctuations in the number

of localized particles, χDD(t) is the susceptibility arising from fluctuations in the number

of particles that are replaced by a neighboring particle, and χSD(t) represents cross

fluctuations between the number of localized and replaced particles.

We also consider “delocalized” particles, that is, particles that in a time t are more

than a distance a from their original location. As was pointed out in Ref. [2], substituting

1− w for w in Eq. (2) gives the delocalized order parameter QDL(t) = N −QS(t), and

as a result, χDL(t) ≡ χSS(t).

Likewise, we can find terms in g4(r, t) and S4(q, t) that correspond to localized,

replaced and delocalized particles, e.g. g4(r, t) and S4(q, t) of localized particles

correspond to gSS4 (r, t) and SSS

4 (q, t) and for i = j and l = k in Eq. (7) and

Eq. (8), respectively. In the next section we present numerical results for SSS4 (q, t)

and corresponding correlation length ξSS4 (t).

Page 6: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

6

4. Results

Figure 1 shows the time and temperature dependence of Q(t) and χ4(t) and their terms

described in Subsection 3.1. Figure 1 shows that for all sufficiently low T , 〈Q(t)/N〉and 〈QS(t)/N〉, are characterized by a two-step relaxation, commonly observed in the

intermediate scattering function [26], as a result of the transient caging of particles.

〈QDL(t)/N〉 has the opposite time dependence from 〈QS(t)/N〉 due to the fact that

it measures the number of particles that moved a distance greater than a. The same

applies to 〈QD(t)/N〉 since those particles constitute the subset of delocalized particles.

At short times, particles oscillate in a region smaller than the overlap radius a, and so

〈Q(t)/N〉 = 1 and 〈QS(t)/N〉 = 1. We observe a short, initial relaxation of Q(t) and

QS(t), and a longer, secondary relaxation. χ4(t), χSS(t), and χDL(t) are zero at short

time, attain a maximum at some intermediate time tmax4 , and decay at long time to zero

in the thermodynamic limit.

χ4(t) and its terms measure the correlated motion between pairs of particles,

calculated equivalently from fluctuations in the number of localized, replaced and

delocalized particles or from the corresponding four-point correlation functions. The

behavior of χ4(t) demonstrates that correlations are time dependent, with a maximum

at a time tmax4 . Similar behavior was reported for the same and other model liquids in

Refs. [2, 43, 44] for a generalized susceptibility related to a displacement-displacement

correlation function χU(t), which measures the correlations between displacements of

particles as a function of time. In these works, SHD was observed to be most pronounced

in the α-relaxation regime. We find that the correlations measured by χ4(t) are also

most pronounced in the α-relaxation regime (see e.g. Ref. [4]).

Figure 2 shows the time dependence of χSS(t) at T = 0.62. There are nine points

marked as open circles that correspond to the times at which we show localized and

delocalized particles in Figure 3.

The four-point structure factor of localized particles SSS4 (q, t) calculated from

Eq. (8) is plotted vs q in Figure 4, at T = 0.60. We also find that at very early

times (when 〈Q(t)N〉 = 1) SSS

4 (q, t) = S(q). We find that while S(q) shows no change at

small q (see e.g the static structure factor in Ref. [3]), SSS4 (q, t) develops a peak at small

q which grows (Figure 4(a)) and decays in time (Figure 4(b)), indicating the presence

of long-range correlations in the locations of overlapping particles.

Inspired by the Ornstein-Zernike theory, OZT [45], which describes, e.g., density

fluctuations near a liquid-gas transition, we use the following function,

SSS4 (q, t) =

SSS4 (0, t)

(1 + (qξSS4 (t))2)

, (11)

where SSS4 (0, t) and ξSS

4 (t) are fitting parameters. The fitting was performed using an

interior-reflective Newton method in Matlab, and setting the termination tolerance of

the function value to 0.1. We find a good fit to the data in the q range from q = 0.34 to

q = 1.9, for each T and time. The observed narrowing of the peak directly reveals the

growing range of gSS4 (r, t) with decreasing T .

Page 7: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

7

The time and temperature dependence of ξSS4 (t) obtained from this fit is plotted for

several state points in Figure 5. We see that the qualitative behavior of ξSS4 (t) is similar

to that of χSS4 (t): ξSS

4 (t) has a maximum in time that coincides with the maximum in

χSS4 (t), and as T decreases, the amplitude and time of this maximum increase. The

highest values of ξSS4 (t) for T = 0.60 exceed half the simulation box size. The fit at

these points depends strongly on the number of points used, initial parameter guesses,

and other details and can yield large values (e.g. > 40) depending on these details.

Since these values greatly exceed the range over which we can meaningfully interpret

the resulting correlation length, we make no attempt to rigorously define the upper error

bounds at these points, but the data is well bounded from below. The fits at all other

points and temperatures are well constrained. The length scale ξSS4 (t) characterizes the

typical distance over which localized particles are spatially correlated.

5. Discussion

In this paper, we have focused on a four-point, time-dependent density correlation

function gSS4 (r, t) and corresponding time dependent structure factor SSS

4 (q, t), and

demonstrated that those functions are sensitive to dynamical heterogeneity in a model

glass-forming liquid. As derived in previous works [1–3,46,47], this correlation function

is related to an order parameter Q(t) corresponding to the number of ”overlapping”

particles in a time window t, where the term ”overlap” is used to denote a particle that

was either localized or replaced in a time t.

We calculated the correlation length of localized particles ξSS4 (t), characterizing the

range of gSS4 (r, t), and showed that it depends on time, and attains its maximum value

in the α-relaxation regime. We also showed that this maximum grows to exceed half of

the simulation box size, close to TMCT. This length scale characterizes the typical size

of dynamically homogeneous domains.

The characteristic length scale calculated here is related to length scales calculated

from the displacement-displacement correlation function [44], cluster size [48], and other

measures of correlated particle motion and dynamical heterogeneity [49–51]. ξSS4 (t) is

essentially the same as that obtained by considering the delocalized particles (the set of

particles that in any time window t move beyond a distance a) due to the mathematical

identity between χ4 for localized and delocalized particles. This suggests a picture of

fluctuating domains of temporarily localized and delocalized particles, perhaps similar

to that proposed by Stillinger and Hodgedon [24].

Finally, we note that all quantities presented here can be measured in dense colloidal

suspensions using confocal microscopy studies [52,53].

6. References

[1] S. Franz, C. Donati, G. Parisi, and S. C. Glotzer, Philos. Mag. B 79 1827 (1999); C. Donati, S.Franz, S. C. Glotzer, and G. Parisi, Journal of Non-Crystalline Solids 307, 215 (2002).

[2] S. C. Glotzer, V. N. Novikov, and T. B. Schrøder, Journal of Chemical Physics 112, 509 (2000).

Page 8: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

8

[3] N. Lacevic, F. W. Starr, T. B. Schrøder, et al., Physical Review E 66, 030101 (2002).[4] N. Lacevic, F. W. Starr, T. B. Schrøder and S. C. Glotzer submitted to Journal of Chemical Physics.[5] N. Lacevic, “Dynamical heterogeneity in simulated glass-forming liquids studied via a four-point

spatiotemporal density correlation function”, dissertation, The Johns Hopkins University, (2003).[6] H. Weintraub, M. Ashburner, P. N. Goodfellow, et al., Science 267, 1609 (1995).[7] P. G. Debenenedetti, Metastable Liquids: Concepts and principles, Princeton University Press,

Princeton, N. J. (1996).[8] G. Harrison, Dynamics Properties of Supercooled Liquids, Academic Press, London (1976).[9] C. Knight, The freezing of supercooled liquids, Princeton N.J.: Published for the Commission on

College Physics by Van Nostrand (1967).[10] J. T. Fourkas, editor, Supercooled liquids: advances and novel applications, ACS symposium series,

0097-6156; 676, American Chemical Society, Washington DC (1997).[11] S. C. Glotzer, editor, Glasses and the glass transition challenges in materials theory and simulation,

Computational Materials Science, 4 (4) 1995.[12] S. Franz, S. C. Glotzer, and S. Sastry, editors, Special issue containing articles from the ICTP-NIS

Conference on “Unifying concepts in glass physics”, Journal of Physics-Condensed Matter 12,(2000).

[13] M. Giordano, D. Leporini and M. P. Tosi, editors, Non equilibrium phenomena in supercooledfluids, glasses and amorphous materials, World Science, Singapore; River Edge, NJ (1996).

[14] K. J. Strandburg, editor; foreword by D. R. Nelson, Bond-orientational order in condensed mattersystems, Springer-Verlag, New York (1992).

[15] A. J. Liu and S. R. Nagel, editors, Jamming and rheology: constrained dynamics on microscopicand macroscopic scales, Taylor and Francis, London; New York (2001).

[16] D. J. Ferry, Viscoelastic properties of polymers, John Wiley, New York, (1980).[17] P. K. Dixon, L. Wu, S. R. Nagel, et al., Physical Review Letters 65, 1108 (1990).[18] A. Tolle, H. Schober, J. Wuttke, et al., Physical Review E 56, 809 (1997).[19] B. Frick, D. Richter, and C. Ritter, Europhysics Letters 9, 557 (1989).[20] E. Kartini, M. F. Collins, B. Collier, et al., Physical Review B 54, 6292 (1996).[21] R. L. Leheny, N. Menon, S. R. Nagel, et al., Journal of Chemical Physics 105, 7783 (1996).[22] D. D. Deppe, R. D. Miller, and J. M. Torkelson, Journal of Polymer Science Part B-Polymer

Physics 34, 2987 (1996).[23] D. B. Hall, A. Dhinojwala, and J. M. Torkelson, Physical Review Letters 79, 103 (1997).[24] J. A. Hodgdon and F. H. Stillinger, Physical Review E 48, 207 (1993).[25] F. H. Stillinger and J. A. Hodgdon, Physical Review E 50, 2064 (1994).[26] H. Sillescu, Journal of Non-Crystalline Solids 243, 81 (1999).[27] S. C. Glotzer, Journal of Non-Crystalline Solids 274, 342 (2000).[28] M. D. Ediger, Annual Review of Physical Chemistry 51, 99 (2000).[29] R. Bohmer, Current Opinion in Solid State and Materials Science 3, 378 (1998).[30] R. Richert, Journal of Physics-Condensed Matter 14, R703 (2002).[31] R. Bohmer, R. V. Chamberlin, G. Diezemann, et al., Journal of Non-Crystalline Solids 235, 1

(1998).[32] J. P. Garrahan and D. Chandler, Physical Review Letters 89, 03570 (2002).[33] Steve Plimpton, Sandia National Labs, www.cs.sandia.gov/∼sjplimp[34] G. Wahnstrom, Physical Review A 44, 3752 (1991).[35] T. B. Schrøder, Hopping in Disordered Media: A Model Glass Former and A Hopping Model,

cond-mat/0005127.[36] T. B. Schrøder, S. Sastry, J. C. Dyre, et al., Journal of Chemical Physics 112, 9834 (2000).[37] T. B. Schrøder and J. C. Dyre, Journal of Non-Crystalline Solids 235, 331 (1998).[38] V. N. Novikov and A. P. Sokolov, Physical Review E 67, 031507 (2003).[39] S. Franz and G. Parisi, J. Phys. Cond. Matt. 12, 6335 (2000).[40] T. R. Kirkpatrick and P. G. Wolynes, Phys. Rev. A 35, 3072 (1987); 36 852 (1987); T. R.

Page 9: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

9

Kirkpatrick and D. Thirumalai, Phys. Rev. B 36, 5388 (1987); Transp. Theory and Stat. Phys.24, 927 (1995).

[41] A. Crisanti, H. Horner and H. J. Sommers, Z. Phys. B 92, 257 (1993).[42] A review of the p-spin model can be found in A. Barrat, cond-mat/9701031.[43] C. Bennemann, C. Donati, J. Baschnagel, and S. C. Glotzer, Nature 399, 246 (1999).[44] S. C. Glotzer, C. Donati and P. H. Poole, Springer Proceedings in Physics 84, Computer Simulation

Studies in Condensed-Matter Physics XI, Eds. D.P. Landau and H.-B. Schuttler, Springer-Verlag,Berlin Heidelberg, 212 (1999).

[45] H. E. Stanley, Introduction to Phase Transitions and Critical Phenomena, New York, OxfordUniversity Press, (1971).

[46] C. Dasgupta, Europhysics Letters 15, 467 (1991).[47] S. Franz and G. Parisi, cond-matt/9804084.[48] Y. Gebremichael, T. B. Schrøder, F. W. Starr, et al., Physical Review E 64, 051503 (2001).[49] B. Doliwa and A. Heuer, Physical Review E 61, 6898 (2000).[50] R. Yamamoto and A. Onuki, Physical Review E 58, 3515 (1998).[51] A. I. Melcuk, R. A. Ramos, H. Gould, et al., Physical Review Letters 75, 2522 (1995).[52] E. R. Weeks, J. C. Crocker, A. C. Levitt, et al., Science 287, 627 (2000).[53] W. K. Kegel and A. van Blaaderen, Science 287, 290 (2000).

Page 10: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

10

Figure captions

Figure 1. Temperature and time dependence of 〈Q(t)/N〉, 〈QS(t)/N〉, 〈QD(t)/N〉,〈QDL(t)/N〉, χ4(t), χSS(t), χDD(t) and χDL(t) ≡ χSS(t).

Figure 2. Time dependence of χSS4 (t) at T = 0.62. Times that are encircled and

labeled with appropriate fraction of tmax4 correspond to the snapshots in Figure 3.

Figure 3. Localized (white) and delocalized (gray) particles at times 0.007tmax4 ,

0.04tmax4 , 0.2tmax

4 , 0.5tmax4 , tmax

4 , 1.6tmax4 , 3.3tmax

4 , 6.3tmax4 , 10.2tmax

4 .

Figure 4. Time dependence of SSS4 (q, t) at T = 0.62. SSS

4 (q, t) is shown at timesidentical to those shown for g4(r, t) in Figure 2. Note that the height of the firstdiffraction peak in SSS

4 (q, t) decreases monotonically as a function of time. Thisis because it depends on the number of localized particles, which also decreasesmonotonically in time (see Figure 1).

Figure 5. Time and temperature dependence of ξSS4 (t) obtained from the fits to

Eq. (11). The data shown are smoothed over successive groups of five points.

Page 11: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

11

0.20.40.60.8

1<

Q(t

)/N

>

5

10

15

20

χ 4(t)

0.20.40.60.8

<Q

S(t)/

N>

5

10

15

χ SS(t

)0.05

0.1

0.15

<Q

D(t

)/N

>

0.10.20.30.40.5

χ DD(t

)

10-2

10-1

100

101

102

103

104

t

0.20.40.60.8

<Q

DL(t

)/N

>

10-2

10-1

100

101

102

103

104

t

5

10

15

χ DL(t

)

T=0.60

T=0.94

T=0.60

T=0.60

T=0.60

T=0.94

T=0.94

T=0.94

T=0.94

T=0.94

T=0.94

T=0.94

T=0.60

T=0.60

T=0.60

T=0.60

Page 12: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

12

10-2

10-1

100

101

102

103

104

t

0

2

4

6

8

10

12χ 4S

S(t

)

t4

max

6.3t4

max

1.6t4

max

10.2t4

max0.04t

4

max

0.02t4

max

0.007t4

max

0.5t4

max

3.3t4

max

Page 13: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

13

Page 14: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

14

2468

1012

S4SS

(q,t)

0.007t4

max

0.04t4

max

0.2t4

max

0.5t4

max

t4

max

0 2 4 6 8q

2468

10

S4SS

(q,t)

t4

max

1.6t4

max

3.3t4

max

6.3t4

max

10.2t4

max

correlations are growing in time

correlations are decaying in time

(a)

(b)

Page 15: Approach to the glass transition studied by higher order ...glotzerlab.engin.umich.edu/home/publications-pdfs/2003/2003_glotz… · Approach to the glass transition studied by higher

15

100

101

102

103

104

t

0

2

4

6

8

10ξ 4S

S(t

)

T=0.60T=0.62T=0.64T=0.69T=0.94