121
Molecular Simulation Study of Homogeneous Crystal Nucleation in n-alkane Melts by Peng Yi B.S., Physics, Tsinghua University (1999) M.S., Physics, Tsinghua University (2002) Submitted to the Department of Physics in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY September 2011 © 2011 Massachusetts Institute of Technology. All rights reserved. Signature of A uthor ........................................... .......... Deptfnent of Physics August 31, 2011 C ertified by .................................................... ......... Gregory C. Rutledge Lammot du Pont Professor of Chemical Engineering Thesis Supervisor C ertified by ....................................... ............... Mehran Kardar Francis Friedman Prof sor of Physics Tjesis Supervisor A ccepted by ........................................................... ................... l Acceptd bya Rajagopal Professor of Physics Associate Department Head for Education

Molecular Simulation Study of Homogeneous Crystal Nucleation

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Molecular Simulation Study of Homogeneous Crystal Nucleation

Molecular Simulation Study of Homogeneous Crystal Nucleationin n-alkane Melts

by

Peng Yi

B.S., Physics, Tsinghua University (1999)M.S., Physics, Tsinghua University (2002)

Submitted to the Department of Physicsin partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHYat the

MASSACHUSETTS INSTITUTE OF TECHNOLOGYSeptember 2011

© 2011 Massachusetts Institute of Technology. All rights reserved.

Signature of A uthor ........................................... ..........Deptfnent of Physics

August 31, 2011

C ertified by .................................................... .........Gregory C. Rutledge

Lammot du Pont Professor of Chemical EngineeringThesis Supervisor

C ertified by ....................................... ...............Mehran Kardar

Francis Friedman Prof sor of PhysicsTjesis Supervisor

A ccepted by ........................................................... ...................lAcceptd bya Rajagopal

Professor of PhysicsAssociate Department Head for Education

Page 2: Molecular Simulation Study of Homogeneous Crystal Nucleation
Page 3: Molecular Simulation Study of Homogeneous Crystal Nucleation

Molecular Simulation Study of Homogeneous Crystal Nucleation

in n-alkane Melts

by

Peng Yi

Submitted to the Department of Physics on August 31, 2011

in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

Abstract

This work used molecular dynamics (MD) and Monte Carlo (MC) method to study thehomogeneous crystal nucleation in the melts of n-alkanes, the simplest class of chain molecules.Three n-alkanes with progressive chain length were studied, n-octane (C8), n-eicosane (C20),and C150, using a united atom force field, which is able to reproduce physical quantities relatedto the solid-liquid phase transition in n-alkanes.

Using a 3D Ising model, we proved that the size of the largest nucleus in the system, nmax, is thecontrolling reaction coordinate during the nucleation process. We have made direct observationof the homogeneous crystal nucleation using MD simulation at as small as 15% under-cooling.We calculated the nucleation rate and identified the critical nucleus through a mean-first-passagetime (MFPT) analysis. At about 20% under-cooling, the critical nucleus size n* is around 100united atoms, and is slightly decreasing as the chain length increases. Abnormal temperaturedependence of n* against classical nucleation theory was found in C150 system. This behaviorcould possibly be explained by the high viscosity of the melt formed by long chain molecules.

The crystal nucleus has a cylindrical shape. We have observed the change of the structure of thecrystal nucleus as the chain length increases. For C8, the chains attach to and detach from thecrystal nucleus as a whole, and the chains end at the end surface of the cylindrical nucleus. ForC20, the partial participation of chains in the crystal nucleus became apparent, where the criticalnucleus consists of a bundle of crystal segments with the tails on the same chains extending intothe amorphous melt. For C150, chain folding was observed during the nucleation stage.

3

Page 4: Molecular Simulation Study of Homogeneous Crystal Nucleation

A cylindrical nucleus model was adopted to characterize the crystal nucleus. The nucleus freeenergy AG(n) was sampled using MC, and was used to calculate the solid-liquid interfacial freeenergies based on classical nucleation theory. The end surface free energy ae is about 4 mJ/m 2

and the side surface free energyoa is about 10 mJ/m 2 . Their values are insensitive to the chainlength.

Thesis Supervisor: Gregory C. Rutledge

Title: Lammot du Pont Professor of Chemical Engineering

Thesis Supervisor: Mehran Kardar

Title: Francis Friedman Professor of Physics

Thesis committee member: Thomas Greytak

Title: Lester Wolfe Professor of Physics, Emeritus

Thesis committee member: J. David Litster

Title: Professor of Physics

4

Page 5: Molecular Simulation Study of Homogeneous Crystal Nucleation

Acknowledgements

I would like to first thank my thesis supervisor Professor Rutledge for his guidance of my thesiswork. His enthusiasm for scientific research and teaching has greatly motivated me. Heprovided me with great freedom in research, and he always helped and encouraged me towardevery milestone. He is a good tutor and has a genuine concern for the students of their academiccareers.

I also want to thank my co-supervisor Professor Kardar, for his knowledge in physics and manyvery helpful discussions on my research. He is always ready to offer help whenever I have anyquestion.

I also thank the current and former Rutledge group members, Pedja, Junmo, Ateeque, Sezen,Fred, Ahmed, Vikram, Sanghun, Pieter, Numan and MinJae. We are good companions on ourresearch and we also formed great friendship. I want to specifically thank Matthew for editingand proofreading some parts of this thesis.

I also want to thank Professor Greytak and Professor Kleppner and their former group members,Bonna, Cort, Julia, Kendra, Lia, and Tomo. When I first came to this country, I received warmwelcome from them. They have helped me settle down and start my MIT life. I also thank myacademic advisor Professor John Joannopoulos for his encouragement.

The Church in Cambridge is like my home. The spiritual and practical care I received from themis priceless.

I want to thank my parents. Without their unconditional love and sacrifice, I could have neverbeen here today. I also want to thank my wife Esther. She is always loving and supporting me.My parents and my wife have embraced me and carried me through many difficult times.

Last but not the least; I want to acknowledge the financial support from NSF (National ScienceFoundation) through CAEFF (Center for Advanced Engineering Fibers and Films) andExxonMobil for this project.

5

Page 6: Molecular Simulation Study of Homogeneous Crystal Nucleation

6

Page 7: Molecular Simulation Study of Homogeneous Crystal Nucleation

Table of Contents

Chapter 1 Introduction........................................................................................................ 9

Chapter 2 Fundam entals of hom ogeneous crystal nucleation ............................................. 15

2.1 Nucleation Theories ........................................................................................................... 16

2.1.1 Classical Nucleation Theory (CNT)........................................................................ 16

2.1.2 Density Functional Theory (DFT)........................................................................... 19

2.2 Experim ental m ethods ................................................................................................... 21

2.3 Sim ulation studies.............................................................................................................. 23

2.3.1 Dynam ical approach............................................................................................... 24

2.3.2 Therm o-dynam ical approach.................................................................................. 28

Chapter 3 Homogeneous crystal nucleation in n-octane (C8) melts .................................... 39

3.1 Cylindrical nucleus m odel ............................................................................................ 39

3.2 Sim ulation m ethods ........................................................................................................ 41

3.2.1 System ........................................................................................................................ 41

3.2.2 M D sim ulation............................................................................................................ 43

3.2.3 M onte Carlo sim ulation........................................................................................... 45

3.3 Results and discussion .................................................................................................... 48

3.3.1 Determ ining the crystal structure ............................................................................ 48

3.3.2 Determining the equilibrium melting point and the heat of fusion ............. 51

3.3.3 M olecular dynam ics sim ulation of the nucleation process...................................... 55

3.3.4 Monte Carlo sampling of the nucleation free energy barrier.................................. 61

3.4 Conclusion ......................................................................................................................... 69

7

Page 8: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 4 Bundle-like crystal nucleation in n-eicosane (C20) melts ................................... 71

4.1 M ethods.............................................................................................................................. 71

4.1.1 Crystal nucleus definition........................................................................................ 71

4.1.2 M olecular dynam ics sim ulation.............................................................................. 72

4.1.3 M onte Carlo sim ulation........................................................................................... 73

4.2 Results and discussion ................................................................................................... 74

4.2.1 M elting point and heat of fusion............................................................................. 74

4.2.2 M D sim ulation of the nucleation process............................................................... 76

4.2.3 Monte Carlo sampling of the nucleation free energy barrier.................................. 84

4.3 Conclusion ......................................................................................................................... 93

Chapter 5 Homogeneous crystal nucleation in entangled C 150 melts ................................. 95

5.1 M ethod ............................................................................................................................... 96

5.1.1 Force field................................................................................................................... 96

5.1.2 M olecular dynam ics sim ulation............................................................................. 96

5.1.3 Initial configurations............................................................................................... 97

5.1.4 System size determ ination.......................................................................................... 97

5.1.5 Equilibration ............................................................................................................... 99

5.2 Results and discussion ..................................................................................................... 100

5.2.1 M D sim ulation of the nucleation process................................................................. 100

Chapter 6 Conclusion and recommendations for future work............................................... 111

Chapter 7 Bibliography ......................................................................................................... 115

8

Page 9: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 1 Introduction

Polymers are long chain molecules. In the melt, the chain molecules form random coils; and

under freezing, the random coils of chain molecules might retain disorder or become partly

ordered (classified as semi-crystalline) [1]. The fraction of ordered region in semi-crystalline

polymers could range from 10% to 80%. The disordered region remains uncrystallizable due to

the topological frustration (entanglements) between the chains. Semi-crystallinity has been

proven by numerous experimental studies, including: reduced and broadly distributed melting

temperature, intermediate density between crystal phase and amorphous phase, and the crystal

segments exhibiting much smaller dimensions than the extended chain length as determined by

X-ray scattering. Semicrystaline polymers possess both the rigidity of the crystal phase and the

flexibility of the amorphous phase; therefore they have found broad applications. Nowadays

semicrystalline polymers constitute about 70% of all polymer products in the world. Examples

of semi-crystalline polymers are polyethylene (PE), polyethylene terephthalate (PET),

polytetrafluoroethylene (PTFE), and isotactic polypropylene (iPP).

How the semicrystalline polymer chains arrange themselves near or at the surface of crystallites

has been a long standing question to the polymer science community[1]. The first model

proposed was the fringed-micelle model, in which polymer chains pass through multiple

9

Page 10: Molecular Simulation Study of Homogeneous Crystal Nucleation

crystallites that are are stringed and spread over the volume, interspaced by non-crystalline

region. The fringed-micelle model faced strong criticism upon the discovery of thin crystal

lamellae formed from dilute solution. The lamellae are formed by aligned segments of chain

molecules. These parallel segments intersect the lamellar plane at an angle between 450 and 900.

The thickness of lamellae is usually in the range of 5-50nm, depending on the degree of under-

cooling, branching, etc.

The thickness of the lamellae is just a fraction of the extended length of the chains that

crystallize. Therefore in the dilute solution, the lamellae are formed by chain folding as

otherwise the chains would have nowhere else to place themselves. It is now generally believed

that the lamellae formed from dilute solution have their chains fold with "adjacent reentry", i.e.,

a single chain returns to the lamella from where it left.

The lamella structure in the melt phase is more complicated and there is considerable

controversy. First of all, there is no requirement for the chains to stay in one lamella as in the

dilute solution; however, thermodynamically, since there is a density difference between the

crystal phase and the amorphous phase, if the chains do not fold back, the splaying out of chain

ends would cause increasing strain to the crystal surface as the lateral dimension of the crystal

increases. Chain folding is certainly one possible way to relieve the strain. In the mean time, as

the "spherulite" crystal structure was found from polymer crystallization in the melt, the fringe-

micelle model came back into the picture as the chain-folded lamellae are tied together by

molecules running from one lamella to another to form so-called "spherulite" crystal. There are

both chain folds and ties (chains joining two lamellae), and chain folding is less regular, as

described by a "random switchboard" model.

10

Page 11: Molecular Simulation Study of Homogeneous Crystal Nucleation

These microscopic details are important factors in determining the morphology and property of

polymer crystals, e.g., the ratio between folds and ties affects the mechanical stability of

spherulites. Even though these questions were raised more than 40 years ago, they still remain

very much unanswered (mainly due to the difficulty in experimental observations) and remain an

active area of research [3, 5].

Polyethylene (PE) is the most widely used semicrystalline polymer, and it is the polymer that we

see most in daily life. About 200 billion pounds of different polyethylene are made globally

each year. The main applications are as plastics and films, e.g., grocery bags, shampoo bottles,

children's toys, and even bulletproof vests. For such a versatile material, it has a very simple

structure, the simplest of all commercial polymers. A polyethylene molecule is prepared from

ethylene (CH 2-CH 2) monomers:

nCH2 =CH 2 -+ ......- CH 2-CH 2-CH 2-CH 2-CH 2-CH 2-CH 2-CH 2......

We write PE as (-CH2-CH 2-)n, where n is called the degree of polymerization (DP). The DP for

commercial polyethylene is in the range of 1041-05, so the molecular weights are in the range of

105-106 (g/mol). Polyethylene is normally produced with a molecular weight distribution

(polydispersity), which arises because of the statistical addition of monomers to growing

polymer chains.

Compared to the large molecular weight for PE, alkanes (also called paraffins or saturated

hydrocarbons) are the simplest organic chemical compounds, and the general formula is CnH 2n+2.

For a given value of n there exist many structural isomers - various arrangements of n carbon

atoms. For n=1, n=2, and n=3, these arrangements are unique. For n=4 there exists two isomers,

11

Page 12: Molecular Simulation Study of Homogeneous Crystal Nucleation

and for n=5 three isomers. For n >5, the number of isomeric alkanes rapidly increases. Isomers

can have different chemical and physical properties. The simplest isomer of an alkane is the one

in which the carbon atoms are arranged linearly in a single chain with no branches. This isomer

is sometimes called the n-isomer (n for "normal" here, e.g., n-butane is the normal isomer of

C4H1 o).

N-alkanes are the building blocks of various important molecules such as alcohols, lipids, and

polyethylene, the archetype of a large class of polymers. The availability of mono-disperse n-

alkanes with molecular weights from a few tens up to several thousands of methylene groups

makes them good candidates for experiments to elucidate the crystalline structure and

crystallization kinetics of chain molecules.[2-4] Nucleation experiments of n-alkanes from the

melt date back to 1960s;[5-9] however, the resolution of the experiments does not allow one to

describe the nucleation process at the molecular level, as the embryonic nucleus often contains

only a few hundred methylene groups. For this reason the microscopic mechanism of chain

molecule crystallization remains a subject of debate. [10-13]

Therefore we choose n-alkane as the study system, expecting the results of our study to shed

light in the understanding of polymer crystallization in general. We are focusing on the

homogeneous crystal nucleation in quiescent melt for the following reasons:

1. Crystallization can be decomposed into the processes of nucleation and growth. While

our understanding of the growth of chain molecule crystals, especially polymers, has

enjoyed considerable progress over the past several decades [14-16], nucleation of the

incipient crystal phase is far less developed.

12

Page 13: Molecular Simulation Study of Homogeneous Crystal Nucleation

2. Nucleation is the time limiting step in crystallization and the initial structure formed

during nucleation has significant impact to the subsequent crystal growth.

3. Although nucleation is a rare event and involves a long time scale, it occurs in a small

length scale, which is suitable for computer simulation.

4. Homogeneous nucleation is the most basic (not necessarily the most common) nucleation

process. Although industrial processing of polymer crystallization often is accompanied

by external field such as flow [18], the principles are the same as in the quiescent case

[19].

Molecular simulation has become a powerful tool for the study of crystallization of chain

molecules, and several recent reviews of computer simulations of crystallization of long chain

molecules are available;[17-20] however, there were only a few studies on the initial nucleation

event that leads to crystal phase formation from quiescent homogeneous melts of the model

compound n-alkanes. Esselink et al [21], Takeuchi [22] and Fujiwara and Sato [23] have all

used united-atom force fields in molecular dynamics simulation to observe the formation of a

crystal phase from the melt for short n-alkanes. Meyer and Muller-Plathe [24] used a coarse-

grained bead-spring model to study the chain folding crystallization of poly(vinyl alcohol).

However, the induction period and the corresponding nucleation event were never clearly

identified in any of these studies, probably due to the artificial rigidity of the force fields

employed[25]. As a consequence, none of the works cited above have calculated the nucleation

rate or examined closely the nucleation event itself, and little is known about either the structure

or free energy of the critical nucleus, two of the most important properties required for

describing the crystallization of chain molecules.

13

Page 14: Molecular Simulation Study of Homogeneous Crystal Nucleation

The goal of this thesis is to use molecular simulation method to investigate the homogeneous

crystal nucleation of chain molecules and study the temperature dependence and chain length

dependence. Particularly we are looking for the answers to the following questions:

1. What is the size and shape of the critical nucleus? What is the solid-liquid interfacial free

energy?

2. When does chain folding set in during the crystallization process?

3. What is the morphology of the crystal-liquid interface?

4. What are the chain length dependence and the temperature dependence of the nucleation

process?

For this purpose, we organize this thesis as follows: Chapter 2 reviews the general methods used

for homogeneous crystal nucleation study, especially the simulation methods developed in the

past 20 years. ' Chapter 3-5 present our application of these methods to n-alkanes of three

progressive lengths, n-octane (C8) 2, n-eicosane (C20) 3 and C150, respectively. Chapter 6

provides a summary of this thesis and gives recommendations to the future work.

The content of this chapter will appear in a review article submitted to Annual Review of Chemical andBiomolecular Engineering2 The content of this chapter was published in Journalof Chemical Physics, 131, 134902 (2009)3 The content of this chapter was published in Journal of Chemical Physics, 135, 024903 (2011)

14

Page 15: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 2 Fundamentals of homogeneous crystal nucleation

Homogeneous nucleation is a process that governs a broad spectrum of physical-chemical

phenomena. The theoretical formulation of nucleation process has been covered by numerous

textbooks and reviews, e.g., Zettlemoyer [26], Skripov[27], Oxtoby [28], Debenedetti [29] and

Kashchiev [30]. However, due to the difficulty in the experimental observation, the microscopic

mechanism of homogeneous nucleation, especially crystal nucleation from the liquid/melts,

remains poorly understood. Computer simulation became a useful tool in the study of

homogeneous nucleation in the past 20 years, during which new concepts and methods were

developed, often by trial and error. This chapter therefore will review the development of

simulation methods of homogeneous nucleation. To give a complete picture, a brief recap of the

basic theories and common experimental methods will also be presented.

Although the focus here is the homogeneous crystal nucleation, there are also many interesting

topics closely related, including heterogeneous crystal nucleation, crystal nucleation induced by

external forces, nucleation under confinement, cross-over of homogeneous nucleation to spinodal

decomposition, homogeneous nucleation in glasses, etc. These topics are also in the research

frontier of the science community.

15

Page 16: Molecular Simulation Study of Homogeneous Crystal Nucleation

2.1 Nucleation Theories

2.1.1 Classical Nucleation Theory (CNT)

Classical nucleation theory (CNT) [31] has been widely applied to study homogeneous

nucleation. It was first developed by Gibbs [32], Volmer and Weber [33], Becker and Ddring

[34], Zeldovich [35] and others based on the condensation of a vapor to a liquid, and this

treatment can be extended to the crystal nucleation from melts and solutions. Based on CNT, a

crystal nucleus consisting of the thermodynamically most stable phase is separated from the

surrounding liquid by a sharp, infinitely thin interface. For temperatures below the melting point,

the competition between the free energy gain of the interior of the nucleus and the free energy

cost of the interface creates a free energy barrier. For a spherical nucleus of radius r, the free

energy of formation AG can be written as

AG= 4rr 2- 47rr' AG, (2.1)3

where o- is the crystal-liquid interfacial free energy per unit area and AG, is the Gibbs free energy

difference per unit volume between the liquid and crystal phases at the under-cooling

temperature. The top of the free energy barrier corresponds to the critical nucleus with radius r*,

and the critical free energy AG* is the free energy of formation for the critical nucleus. AG* and

r* are obtained by maximizing AG with respect to r in Eq.(2.1):

AG* = AG(r*) = 16;7 a 2 (2.2)3 (AG,)

16

Page 17: Molecular Simulation Study of Homogeneous Crystal Nucleation

and

r 2 (2.3)AG,

Correspondingly, the size of the critical nucleus for the spherical nucleus model is

n 32 (2.4)3 3 AG,

According to some conversions, the term "nucleus" is only used to refer to a nucleus of size

equal to or greater than n*; and "embryo" is used to refer to a nucleus of size smaller than n*.

Therefore the term "nucleation" refers to the formation of one nucleus that serves as a stable

center for further crystal growth. However, we do not make such distinction here. Any nucleus

in our conversion can have size as small as 1.

The rate of nucleation, i.e., the number of critical nuclei formed per unit time per unit volume,

can be expressed in the form of the Arrhenius equation:

I = Ie-AG*/kT, (2.5)

where Io is a kinetic prefactor, and kB is the Boltzmann constant. Io is given as

I4 ~ Nv, (2.6)

where N, is the molecule number density in the melt state, and v is the frequency of molecular

transport at the nucleus surface. Furthermore, v can be approximated using the Stokes-Einstein

relation

17

Page 18: Molecular Simulation Study of Homogeneous Crystal Nucleation

SkB (2.7)

where ao is the molecular diameter and r7 is the viscosity. [5]

For a small degree of under-cooling, AG, can be approximated by

AG, ~ pFAHf AT / T (2.8)

where AHf is the heat of fusion per molecule at the equilibrium melting temperature Tm, AT

(equal to Tm - ) is the under-cooling, and p, is the molecule number density of the crystal phase.

For deeper under-cooling, more precise approximation is needed,

AG,~ pAHfATT / T,,2 . (2.9)

Using the first order approximation, the temperature dependence of the critical nucleus size and

the critical free energy is given by

r* oc (A T)-',(2. 10)

and

AG* oc (AT)- 2 . (2.11)

When the under-cooling AT increases, AG* decreases but the viscosity r7 increases. The

combination of these two factors results in a maximum of the nucleation rate I at a temperature

Tma somewhere between the melting point Tm and the glassy transition temperature Tg. Tmax in

general depends on the material.

18

Page 19: Molecular Simulation Study of Homogeneous Crystal Nucleation

Assuming that the molecular diffusion to the surface of nucleus is also an activated process, we

can further express Eq.(2.5) into

I = Ae-E /kBT e-AG*IkBT - A-(Ed+AG*)IkBT (2.12)

where A is a temperature independent factor and Ed is the diffusion free energy barrier.

2.1.2 Density Functional Theory (DFT)

One of the main criticisms to CNT is the capillary approximation, that is, small portions of the

new phase are treated as if they represent macroscopic regions of space, and that material at the

center of the nucleus behaves like the new phase in bulk; and that the surface free energy of a

small cluster is the same as that of an infinite planar surface. These assumptions can only be

deemphasized when the nuclei are big enough, making CNT best for nucleation near

coexistence. Far away from the coexistence, especially close to spinodal decomposition, CNT

could become very inaccurate. As a matter of fact, the nucleation rate obtained in experiments

and predicted by CNT, using independent measured thermodynamics quantities such as

interfacial free energies and heat of fusion, often differs by orders of magnitude. Under such

circumstance, more sophisticated theory is needed.

Density functional theory is a quantum mechanical modeling method used in physics and

chemistry to investigate the electronic structure of many-body system. It has also been proved a

powerful approach to study nonclassical nucleation of the gas-to-liquid transition by Cahn and

Hilliard [36], Abraham [37], and Oxtoby [38]. Their approach expresses the free energy as a

19

Page 20: Molecular Simulation Study of Homogeneous Crystal Nucleation

functional of radial density profile p(r). The density varies from the center of the nucleus

outward, and the density at the center of the nucleus does not have to be the same as that of the

bulk new phase, nor does it have to behave like a planar interface.

Oxtoby [39] has shown that DFT yields more accurate results than CNT. For the gas-liquid

nucleation, DFT has clarified issues for weakly polar liquids and has given rise to explanations

of the behavior of non polar fluids. In binary condensation, Oxtoby and Kashchiev [40] have

proven the nucleation theorem, a relationship between the effect of pressure, or chemical

potential, on the free energy of the critical nucleus and its size and composition.

Harrowell and Oxtoby [41] extended DFT method to the solid-liquid transition. Since solid-

liquid transition involves not only the density change, but also an order change. The free energy

is a functional of not only the density profile, but also the Fourier components of the lattice

structure [42]. They have observed that the properties of a critical nucleus can differ

significantly from those of the stable bulk phase that eventually forms. They have applied this

theory in metal alloy crystallization. Another application is to protein crystallization from

aqueous solution in which protein concentration and crystal structure evolve together but not at

the same rate. The density functional approach applied to these situations should yield more

information on these systems and resolve many other problems inherent in the classical

approaches. However, due to the simplicity of classical nucleation theory, the discussion that

follows will be limited within the scope of classical nucleation theory.

20

Page 21: Molecular Simulation Study of Homogeneous Crystal Nucleation

2.2 Experimental methods

Homogeneous crystal nucleation occurs in the interior of an under-cooled liquid, making it

difficult for any experimental equipment to detect. Moreover, impurities induce the

crystallization and quickly drive the whole system to crystallize. The droplet technique was

proposed by Vonnegut [43] to address this problem, and was used by Turnbull with much

success [44]. The sample liquid is dispersed into a large number of tiny, normally micron in size,

droplets, exceeding the number of impurities in the liquid. A significant number of droplets are

therefore impurity-free and could be used for homogenous nucleation. The volume of each

droplet is so small that the one nucleation event automatically precludes other nucleation events

in the same droplet. Once a nucleation event happens in a droplet, since the crystal growth is

much faster than nucleation rate, this droplet almost immediately crystallizes completely. The

crystallization process can be monitored by using X-ray scattering, dilatometry, differential

scanning calorimetry (DSC), or visual method. Thus we are able to estimate the homogeneous

nucleation rate L

Crystal nucleation can still occur on the surface even the droplets are impurity free. Special

procedures thus must be taken to ensure that the nucleation happens in the bulk. One simple test

is to check whether the nucleation rate is proportional to the volume of the droplets or to the

surface.[45]

Unlike the gas-liquid transition, where the interfacial free energy is equal to the interfacial

tension and it is easy to measure. The solid-liquid interfacial free energies are very difficult to

21

Page 22: Molecular Simulation Study of Homogeneous Crystal Nucleation

measure, especially away from coexistence. The available data are rare. [46] Homogeneous

nucleation experiment allows us to measure the solid-liquid interfacial free energy. There are

several different methods, depending on the under-cooling condition, isothermal or with a finite

cooling rate.

For the isothermal nucleation experiments, combining Eq.(2.8),(2.2) and (2.5) we have

1 6,c T2 o3In I= n Io 1(2.13)

3kB p,2AH2 AT2 (1

Therefore if the nucleation rate I is measured as a function of temperature, the interfacial free

energy o can be obtained from the slope of a plot of InI against 1/AT2T and the kinetic prefactor

1o is the intercept. This method has been adopted by Turnbull and his coworkers to measure the

interfacial free energy o-for some organic and inorganic materials.

For nucleation experiments with finite cooling rate, an alternative approach is available. [47] The

temperature at which any given fraction of the droplets are solidified can be related to the

cooling rates ri, r2 by the equation

r 2AG*AT(

r2n kBTAv(2

where ATav is the average undercooling at the chosen fraction crystallized and ATd is the

difference in this temperature at the two rates.

22

Page 23: Molecular Simulation Study of Homogeneous Crystal Nucleation

2.3 Simulation studies

The development of computer technology has made numerical simulation a very useful tool to

study condensed systems, allowing scientists to work with nanoscale time and space resolution.

In practice, Ising model [48], Lennard-Jones model [49], soft-sphere model [50] and hard-sphere

model [51] are among the most studied because of their simplicity and representation for a large

group of real systems. Simulation studies in 2D systems, e.g., 2D Ising model and hard-disk

system were also available, although we will focus on 3D systems. There are two main

approaches to the nucleation problem, kinetic and thermodynamics. Molecular dynamics

method belongs to the former; Monte Carlo method the latter, often being used to sample the free

energy. Monte Carlo method was also used to "flip" the spins in an Ising model to study the

"kinetics" of a lattice system.

Ising model has been used in physics community for many years to study phase transition. The

discrete positions of spins and the simple form of interaction make computation much less costly

compared to off lattice models. In addition, analytical results are often available. Binder [52]

has carried out extensive theoretical and simulation study on nucleation using the Ising model.

Even in the past two decades, it is still a very useful model to examine the nucleation process.

Ising model is particularly suitable for Monte Carlo not only to sample the free energy landscape,

but also to study the real dynamics, e.g. the Kinetic Monte Carlo (KMC) method [53-55] was

developed to estimate the real transition time if the transition rate for all possible directions are

known and can be tabulated. Nevertheless, in the study of solid-liquid transition, realistic off-

lattice models still draw more attention.

23

Page 24: Molecular Simulation Study of Homogeneous Crystal Nucleation

2.3.1 Dynamical approach

Nucleation is intrinsically a non-equilibrium, dynamic process, and the most convincing

simulation study is molecular dynamics simulation. The first molecular dynamics study of

crystal nucleation (in a Lennard Jones system) was reported by Mandell et al. [49]. They used a

small system of only 108 particles with periodic boundary condition, which raised the question

of finite size effect to nucleation. This finite size effect was examined later by Honeycutt and

Andersen [56], and a further simulation used 15,000 and then 106 Lennard Jones particles was

reported by Swope and Andersen [57], where a Voronoi analysis was adopted to define the

crystal region and thus crystal nucleus. Compare to gas-liquid transition where only

densification is involved, the translational ordering for solid-liquid transition requires a more

sophisticated definition of crystal phase. There have been different ways in practice to define a

crystal nucleus [58-60], and the effect of different choices of definition needs to be considered in

each individual numerical study.

With the identification of crystal nuclei, Swope and Andersen [57] were able to measure the

steady state nucleus size distribution Pst(n). They fitted Pst(n) to a polynomial, and by finding

the maximum of Pst(n) they estimated the critical nucleus size n*. It is not a rigorous approach, a

more systematic procedure to identify the critical nucleus size and induction time through MD

simulation was later introduced as a mean-first-passage time (MFPT) method[61-63].

Swope and Andersen [57] claimed to have observed nucleation. However not only did they use

a doubtful ensemble, i.e., canonical ensemble that in principle prohibits a phase transition; they

also did not observe a clear induction period. The induction period is the signature of a

24

Page 25: Molecular Simulation Study of Homogeneous Crystal Nucleation

nucleation event, or a rare event in general. An under-cooled liquid is in a metastable

equilibrium state and the crystal phase is in the stable equilibrium state. These two states are

separated by a free energy barrierAG(x), where x is the reaction coordinate, and the top of this

barrier is called the transition state, corresponding to the presence of one critical nucleus. The

waiting time for a system to produce a fluctuation big enough to overcome that free energy

barrier is the induction time. The timescale accessible to computer simulations is normally

between ns to ps, far shorter than the induction time in real experiments. Furthermore, the

induction time scales inversely proportional to the system volume, making the direct observation

of nucleation in MD simulation very difficult. Accelerated molecular dynamics methods, e.g.,

metadynamics [64], were proposed to help the system find and overcome the free energy barrier.

They are summarized in ref.[64].

The induction time r measures the ability of a system to stay in metastable equilibrium state. It

depends on the observables used to determine whether the system still remains in the metastable

state. During the induction period, all, not just some, system variables should be fluctuating at a

level corresponding to the metastable equilibrium. However, lacking an induction period does

not always mean the absence of nucleation because r* is a system size dependent quantity.

When the system is big enough, the chance of finding one critical nucleus could becomes so high

that the induction time is too short to catch by an observer.

Depending on the definition, r" might or might not include a transient timer'for the system to

adopt the under-cooling before the metastable equilibrium is reached. This r'could be evaluated

and then subtracted from *, making use of the fact that nucleation is a Poisson process, so that

25

Page 26: Molecular Simulation Study of Homogeneous Crystal Nucleation

the probability distribution P( r*) should be independent of where to start timing z-*, once the

transient period r' is past. [27]

The MFPT method was proposed to determine the induction time and to extract other useful

thermodynamics information from the MD simulation results.[61-63] This method is presented

in different ways, but they are interrelated.[65] The approach by Wedekind et al.[63] makes

particularly clear the link between the classical theoretical treatment and the quantities available

by MD simulation, and was introduced here. According to this method, the mean first passage

time of a chosen reaction coordinate x, nmax in our case, takes the form

r(nm.) = 0.5r* [1+ erf(Z-(nm. - n*))], (2.15)

where r* is the average induction time, Z is the Zeldovich factor and

1 d 2 AG(nm)Zk = k B" " . (2 .1 6 )2gckBT dnnma

The critical value n*, corresponds to the transition state. Therefore MFPT method allows us to

estimate nm*., Z, and r* from MD simulations. The MFPT method was furthermore extended by

Wedekind et al. [66, 67] to reconstruct the free energy curve of the system AG(nma). Although

in ref. [66] the authors made a mistake by using nucleus size n rather than nma as the reaction

coordinate, it was then corrected later. [67]

The induction time r* is related to the nucleation rate I as

= 1 (2.17)IV

26

Page 27: Molecular Simulation Study of Homogeneous Crystal Nucleation

where V is the volume of the system. Through this relation, the nucleation rate can be calculated

when r* and V are both known. Eq.(2.17), however, is only limited to the so-called mononuclear

mechanism of nucleation[30], which is applicable to systems undergoing phase transition

through the appearance of only one critical nucleus. As the opposite, the polynuclear mechanism

is for the systems undergoing phase transition through the appearance of statistically multiple

nuclei. An unified formula for the induction time, considering both mechanisms, is (Eq.(29.12)

in ref. [30])

* 1 / IV + [(1 + vd)ad / (C gGvdi)1/(1+vd). (2.18)

where d is the dimensionality of the system, cg is a shape factor in the order of 1, G, is the crystal

growth rate, v is the growth component in the order of 1 and ad is the detectable fraction of

crystallized volume. When the system is large, Eq.(2.18) reduces to polynuclear case; under the

opposite condition, Eq.(2.18) reduces to mononuclear case. According to Eq.(2.18), in

polynuclear case, the role of I is weaker and the induction time does not depend on the volume.

This could explain why we observe nucleation in a small simulation box in nanoseconds, while

in real experiments the induction time is in seconds or longer. One directly application is to

calculate the nucleation rate I from a small simulation in a straightforward way as Eq.(2.17), and

then use its value to predict r* in a much larger system based on Eq.(2.18).

The MFPT method still relies on unbiased, direct observation of nucleation in MD simulation.

Although it does have the advantage that the trajectory is not affected by any pre-chosen reaction

coordinate, its applicability to systems close to coexistence, i.e., with very high nucleation free

27

Page 28: Molecular Simulation Study of Homogeneous Crystal Nucleation

energy barrier, is limited. The thermo-dynamical approach was developed mostly to study the

nucleation process with a very high free energy barrier.

2.3.2 Thermo-dynamical approach

Another approach to study the nucleation problem is to sample the free energy curve. This

approach is particularly useful when the free energy barrier is very high, that the brute-force

molecular dynamics simulation needs a very long time to simulate one nucleation event. In

sampling the free energy, biasing techniques are very helpful and they are reviewed in ref. [64].

Among them the most well-known is the umbrella sampling method.[68, 69] The basic thought

is to bias the Boltzmann factor near the region toward the transition state. When the reaction

coordinate(s) could be explicitly expressed as a function of coordinates, e.g., the distance

between a protein molecule and the substrate in an adsorption problem, the umbrella sampling

can be carried out in molecular dynamics. Otherwise, Monte Carlo is more suitable for umbrella

sampling method.

The first practice was to expand the free energy as a function of a number of order parameters as

formulated in Landau's mean field theory. There is little difficulty in finding the appropriate

order parameter in Ising model [52], while for off-lattice model, it is less straightforward.

Structural order parameter was used by Alexander and McTague [70] for a Landau expansion to

compare the relative stability of FCC and BCC crystal structures. Steinhardt [71] has used

spherical harmonics to develop the bond orientation order parameters Q4, Q6, etc, for similar

analysis. This Q6 order parameter measures the average order of the system. It is very unique in

that its value is almost the same for FCC, HCP and BCC lattices, which enables one to

28

Page 29: Molecular Simulation Study of Homogeneous Crystal Nucleation

distinguish liquid from crystal regardless of the crystal structure, and to examine the famous

Ostwald's rule of metastable state transition.[72, 73] Q6 parameter was later used by Frenkel and

his coworkers to study the homogeneous nucleation in a soft spheres system [50] (Fig. 2.1) and a

Lenneard Jones liquid system [74].

C)

0.0

Aq

DBcC

BCC

02 0.4

FIG 2.1 Gibbs free energy of a under-cooled soft sphere system as a function of Q6.

(Duijneveldt and Frenkel, 1992)[50].

It was soon discovered that there are two problems of using a global order parameter like Q6 to

monitor the nucleation process, (1) Q6 value for the transition state varies with system size [75],

and (2) the "entropic breakdown" [76], which means, with the same Q6, the system will prefer a

collection of small nuclei rather than a big nucleus. The intrinsic reason for these two problems

29

50 1-

..100F

Page 30: Molecular Simulation Study of Homogeneous Crystal Nucleation

is that nucleation is a local event but Q6 is a global parameter. Unless the system is very small, a

global parameter is not a good candidate for nucleation study. Despite the effort of trying to

improve the global order parameters [77, 78], a local reaction coordinate replaced the role of

global order parameter to describe the nucleation process [51, 76, 79].

Reaction coordinate is a term used in transition state theory (TST) to measure the process of a

chemical reaction, where the reactants are normally simple atoms and the distance (coordinate)

between them naturally serves as an indicator of reaction process. As a comparison, an order

paramter in statistical mechanics is more often used in an average sense to describe the relative

stability of a bulk phase, Therefore in our discussion of nucleation process, we will use reaction

coordinates to describe the state of the system.

AG(n)/AG(nmax)

,, n*~~-nmax*

(a) AG(nmax*)

AG*=AG(n*)._ .

(b) (c) n/nmax

FIG 2.2 The free energy of a system according to classical nucleation theory. (a) A

system containing only one nucleus of size n. (b) A system containing multiple nuclei,

the largest of which has size nmax. (c) AG(n) and AG(nmax) as a function of n and nmax,

respectively.

30

Page 31: Molecular Simulation Study of Homogeneous Crystal Nucleation

The free energy AG(n) introduced in classical nucleation theory is the free energy change due to

the change of the size of the only one nucleus in the system. However, when a real system is

under- cooled, there are always multiple nuclei presented. It is only natural to believe that once

the largest nucleus reaches the critical size n*, the whole system will be quickly driven to

crystallization. Therefore, the most reasonable choice of reaction coordinate for this system is

the size of the largest one, nmz, and the free energy difference of the system is expressed as

AG(nma). The transition state, denoted by nmax*, therefore corresponds to the presence of a

critical nucleus n*, i.e., nmax~n*, as illustrated in Fig. 2.2 (c).

Assuming the mononuclear mechanism, when the largest nucleus is much bigger than the rest of

the nuclei, AG(nma) is expected to follow AG(n), i.e., the free energy of the system is dominated

by the largest nucleus with the other nuclei play as a unchanging background. Assuming that the

upper limit of this equilibrium background is n', if nmax is greater than n', then AG(nma) is well

approximated by AG(n); if nma n', then the nucleus size distribution will be changed and that

AG'(nma) will deviated from AG(n). The exchange entropy of multiple nuclei results in a finite

most probable nma value corresponding to the metastable equilibrium, as shown in Fig 2.2 (c).

It is the free energy AG(nmj) that directly determines how fast a system experiences nucleation,

although AG(nmax) is closely related to AG(n).

Some research groups [51, 76, 79] applied the concept of nucleus size in the consideration of a

local reaction coordinate, but did not offer clear explanation. In order to understand the role of

nucleus size as the reaction coordinate, we performed nucleation simulations using 3D Ising

31

Page 32: Molecular Simulation Study of Homogeneous Crystal Nucleation

model. The cubic lattice has length L, and the whole system contains L xL xL spins. Periodic

boundary condition was applied to all three dimensions. The Hamiltonian of the system is

expressed by

H J B-= --_ I sisi Is, (2.19)kBT kBT,,,' kBT

where the first sum runs over all nearest-neighbor pairs and the second sum over all sites. J is

the interaction coefficient and B is the magnetic field. The critical temperature for 3D Ising

model is kBTc~4.51J. We chose the system temperature to be T=0.6Te, which means that

J/kBT0.3694. The initial configuration was all "down" spins. At time 0, an "up" magnetic field

B was switched on, which served as "under-cooling". The system will remain in the metastable

"melt" state for a period of time, before a critical nucleus of opposite spins drives the spin system

to a complete reversal. The nucleus is defined as that two nearest neighbor "up" spins belong to

the same nucleus of "solid" phase. A metropolis algorithm was used to update the configuration

of the system by "flipping" one spin at a time.

Using umbrella sampling Monte Carlo method, we sampled the free energy of the system as a

function of nmax, AG(nmax). After correcting the bias introduced by umbrella sampling method,

AG(nm.ax) was straightforwardly calculated from the relative probability of a system to have

certain value of nma, P(nmax). Since at any given moment, the system always contains multiple

nuclei of various sizes, we also sampled the nucleus size distribution P(n) and calculated the free

energy AG(n) defined by classical nucleation theory (Eq.(2.1)) [74, 79],

AG(n) / kBT = -In P(n) + constant. (2.20)

32

Page 33: Molecular Simulation Study of Homogeneous Crystal Nucleation

It must be stated that the nucleus size distribution P(n) used to calculate AG(n) by Eq.(2.20) must

be the one under metastable equilibrium, Peq(n). Sometimes it is only possible to measure the

steady state nucleus size distribution Pst(n), then it has to first be converted to Peq(n) by (Eq.(13.2)

in ref.[30])

(n)= P,(n)[1- erf (ZvE(n - n*))] (2.21)[I ( [1-f(Z;(1-n))]

It is clear from Eq.Error! Reference source not found. that Peq(n*)=2Pst(n*). Since AG(n)/kBT

is the minus logarithm of Peq(n), the final estimate of AG(n)/kBT will not be affected by much, if

AG(n) itself is much greater than kBT.

We have confirmed our simulation by reproducing the results by Pan and Chandler [79]. The

free energy AG(nmax) and AG(n) were plotted in Fig. 2.3. For a fixed volume system, we should

use the Helmholtz free energy AF, but we still used AG for the consistency of notation in the

discussion. According to Fig. 2.3, AG(n) has the shape as predicted by CNT, however, AG(nmax)

has a local minimum at a small but finite nmax. Fig. 2.3 resembles Fig. 2.2 (c) well.

According to Fig. 2.3, the free energy barrier of AG(nmax) and AG(n) overlaps, i.e. n ~ n*,

which confirmed that the presence of a critical nucleus n* is the transition state of the system.

This fact provides a way to approximate n* by n* , if the latter is easier to calculate or measure,

for example, by the aforementioned MFPT method.

33

Page 34: Molecular Simulation Study of Homogeneous Crystal Nucleation

As shown by Fig.2.3, AG(nma) is system size dependent, while AG(n) is not. When the system

size is very big compared to the crystal nuclei, the system free energy barrier AG(nm,*), defined

in Fig. 2.2 (c), could be very small. As a matter of fact,

AG(n*) = AG(n*)- kBTInV, (2.22)

where V=L 3 is the volume of the system. The result of a small AG(nma*) is a very fast phase

transition, which is to be expected because the chance to find a critical nucleus is bigger in a

bigger system. This phenomenon, i.e., faster phase transition in bigger system, has been

mistakenly interpreted as spinodal decomposition assisted nucleation [78]. But as we see from

Fig.2.3, the nucleus free energy AG* is independent of system size. As long as AG* >> kBT is

satisfied, it is still nucleation.

If the controlling reaction coordinate is nma, then the nucleation rate should be the rate to

overcome a free energy barrier AG(nmax*) for this particular system of size V,

kk =-4-exp(-AG(n~m)/kBT) (2.23)

V

where ko is a pre-factor independent of, or weakly dependent on, temperature T. Inserting

Eq.(2.22) into Eq.(2.23) we obtained

k = ko exp(-AG(n*) / kBT) = k0 exp(-AG* / kBT), (2.24)

and the nucleation rate equation Eq.(2.5). Therefore we have proved that nma is the controlling

reaction coordinate of the nucleation process. The difference between AG(nma) and AG(n) was

also discussed somewhere [67, 80, 81]. After all, when considering a local reaction coordinate,

34

Page 35: Molecular Simulation Study of Homogeneous Crystal Nucleation

we realized that the largest nucleus size, nmax, is the most relevant quantity to describe the

nucleation process in a system.

Eq.(2.22) was also reported by Wedekind et al. [67]. They further claimed that this relation

should hold true for all n, not just n*, namely,

AG(n. ) = AG(n)-k 8 T In V. (2.25)

which is certainly not correct as they failed to notice that the metastable equilibrium state has a

most probable but not-zero nmax.

B=0.55J, T=0.6TC

20

lZ o

m 15

Ema*IS 10 -(AG(n max L=32

- - -AG(n ), L=25

00

I o 5 -0 AG(n ),L=20 -O max

< 0 ---- AG(n ),x L=16

0 AG(n)00-0 .o, i . . . I . I .- I .

0 20 40 60 80 100 120 140 160

n /nmax

FIG 2.3 The comparison between AG(nmax) (lines, system size increases from bottom to

top) and AG(n) (open circles), from a 3D Ising model simulation.

35

Page 36: Molecular Simulation Study of Homogeneous Crystal Nucleation

Although the thermodynamics approach predicts the same nmax* for different system sizes. The

nmx* estimated by MFPT method increases as system size increases, as shown in Fig 2.4 and

Table 2.1. This difference lies at the approximation used by MFPT method, as well as the

problematic nature of using a thermo-dynamical approach to study the intrinsically dynamics

process. Further investigation of the relation between the kinetics and the thermodynamics of

the nucleation process is needed. Under such circumstance, whether nmax* is still a good

approximation of n* is unclear.

FIG 2.4 System size dependence of the

a function of nmax for the 3D Ising model

a

4x10

3x103

2x10

Ix10

300 350 400

mean-first-passage time (Monte Carlo Steps) as

with T=0.6Tc and B=0.55J.

36

0

0 50 100 150 200 250n

max

Page 37: Molecular Simulation Study of Homogeneous Crystal Nucleation

TABLE 2.1 The system size dependence of nma* estimated using MFPT method Eq.(2.15).

37

Ln L max

16 85

20 89

25 96

32 119

Page 38: Molecular Simulation Study of Homogeneous Crystal Nucleation

38

Page 39: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 3 Homogeneous crystal nucleation in n-octane (C8) melts

3.1 Cylindrical nucleus model

For crystal nucleation of chain molecules, a cylindrical nucleus model is more suitable compared

to the spherical model as described by Eq.(2.1).[5, 7, 47] A cylindrical model distinguishes two

types of surface: the chain-end surface and the chain-side surface or lateral surface. Similar to

Eq. (2.1), the free energy of formation of a cylindrical nucleus with radius r and thickness 1 can

be written as

AG(r,l) = 27rr2 J, + 2;rrla, - ifr2 lAG, (3.1)

where oe and as are the crystal-liquid interfacial free energies per unit area for the end surface

and the side surface, respectively. If both the radius and the thickness of the cylinder vary, the

critical free energy AG* is

2

AG* =8zc Qae (3.2)(AG,) 2 '

and the critical nucleus has thickness l* and radius r* given by

1* = 4e / AG,, (3.3)

39

Page 40: Molecular Simulation Study of Homogeneous Crystal Nucleation

and

r* = 2o-, / AG , (3.4)

respectively. Thus the critical nucleus size n* for the cylindrical nucleus model of variable

thickness is

n =16)r -U "Pn (3.5)(AG,) 3

A simple relation exists between the critical nucleus size n' and the critical free energy AG*:

n= 2p AG* (3.6)AG,

which also holds true from spherical nucleus model.

For short chain molecules such as n-octane, the critical cylinder thickness 1* can be greater than

the extended length of the respective chain molecule length lo. In this case, a more appropriate

model for the cylindrical nucleus may be the one in which the thickness of the nucleus is fixed

and equal to lo. In this model, the free energy of formation is expressed as

AG(r) = 21rr 2, + 2rrloacr-, cr 2lAG,. (3.7)

By finding the maximum of AG with respect to r in Eq.(3.7), the critical free energy for the

cylindrical nucleus model of fixed thickness is found to be

AG* = 2' (3.8)(l0 AG, -2a,)

40

Page 41: Molecular Simulation Study of Homogeneous Crystal Nucleation

and the critical nucleus size is

n* = 0 " ' . (3.9)(l 0 AG, - 2o)

Conversely, if the values of AG* and n' are known, then the interfacial free energies can be

calculated from

p.AG-*2as= , (3.10)

and

e = J AG -P"G*J (3.11)2 n

3.2 Simulation methods

3.2.1 System

We simulated a system containing either 480 or 960 n-octane chains using the isothermal-

isobaric (NPT) ensemble with the pressure P set at 1 atm. Simulations in which the crystal phase

spanned one or more dimensions of the simulation box were conducted with fully variable side

lengths and angles of the simulation box; for all other simulations, the angles were fixed at 900

and only the side lengths were allowed to vary independently. Periodic boundary conditions

were employed in all three directions.

41

Page 42: Molecular Simulation Study of Homogeneous Crystal Nucleation

In polymer crystallization, kinetics plays an as important role as the thermodynamics. Therefore

a realistic chain model is necessary. We used a united-atom (UA) force field proposed originally

by Paul, Yoon and Smith[82] and modified subsequently by Waheed et al.,[83, 84] which we

designate PYS. This force field was parameterized using experimental data and quantum

calculations on short n-alkanes and has been shown to describe the static and dynamics

properties of polyethylene melts accurately.[82, 85-87] This model is able to represent the

average local mobility with sufficient accuracy to yield reasonably good results for local and

global dynamics (within 20%-30% of experimental values).

In this force field, polyethylene and alkane chains are composed of spherical beads, or "united

atoms," each representing a CH2 group (or CH3 for the terminal beads). CH 2 and CH 3 beads

differ only in mass. Each bead interacts through bonded and nonbonded potentials. The bond

stretching potential between two adjacent beads is

Ebond =kb(b -bo) 2 , (3.12)

where b is the length of bond, kb is the bond stretching constant, which is equal to 1.46 x 105

kJ/mol/nm 2, and bo is 0.153 nm. The bond angle bending potential among three adjacent beads

is

Eange =k,9 ( -0 0)2 (3.13)

where 0 is the complement of the bond angle, ke is the angle bending constant, which is 251.04

kJ/mol/rad2, and 66 is 1.187 rad. The bond torsion potential among four adjacent beads is

Jt 1Eto = -[k(1- cos p)+ k2(1- cos 2p)+k 3(1- cos 3p)], (3.14)

242

Page 43: Molecular Simulation Study of Homogeneous Crystal Nucleation

where p is the torsion angle, the torsion constant ki is 6.78 kJ/mol, k2 is -3.60 kJ/mol, and k3 is

13.56 kJ/mol. The nonbonded interactions are described by a Lennard-Jones 12-6 potential for

all intermolecular interactions between beads on different chains and for intramolecular

interactions between beads on the same chain that are separated by four or more bonds,

Ej = 4c - .(3.15)r ) r)

where eis 0.39 kJ/mol, and a-is 0.401 nm. The Lennard-Jones potential was truncated at 2.50,

and tail corrections were added for potential energy and pressure that assume the radial

distribution function g(r) =1 beyond this cutoff.

3.2.2 MD simulation

Molecular dynamics (MD) simulations are suitable for obtaining a direct, unbiased, kinetic

description of the nucleation process for small molecules at moderate to large supercooling.

However, for large molecules or small supercooling, if the critical free energy of nucleation is

too high, the spontaneous crossing of the free energy barrier becomes very unlikely, and

nucleation cannot be observed in the timescale accessible by brute-force MD simulations.

We carried out MD simulations using open source code for the DLPOLY package [88] and the

LAMMPS (Large-scale Atomic/Molecular Massively Parallel Simulator) package.[89, 90] The

DLPOLY package was used for simulations to determine the crystal structure, melting point

and heat of fusion of n-octane, where independent variation of the box angles was required; for

all other MD simulations, LAMMPS was used. In the MD simulations, the initial velocities of

43

Page 44: Molecular Simulation Study of Homogeneous Crystal Nucleation

all beads were generated from the Maxwell-Boltzmann distribution according to the desired

temperature, and the equations of motion were integrated using the velocity Verlet method with

an integration time step At equal to 2fs. Waheed et al. [84] and Lavine et al. [91] have previously

evaluated the effect of the integration time step during the MD simulation of n-alkanes with this

PYS force field. They found that changing At from 1 fs to 5 fs increased the bond energy, the

angle-bending energy, and the relaxation time for the chain-orientation autocorrelation function

by approximately 10%. The consequence of these errors was mitigated through the use of a

thermostat, such that no detectable difference in crystallization kinetics was observed for

simulations with At between 1 and 5 fs.

For a reaction coordinate to characterize crystallization, we monitored the size of the largest

crystal nucleus in the system, nma, during the simulation. This choice was made because the

dynamics of the nucleation process is dominated by the biggest nucleus in the system. [51] In

calculating the nucleus sizes n, we adopted the definition of nucleus used by Esselink: [21] if two

chain molecules have the same orientation and are neighbors, then they belong to the same

nucleus. They are considered to have the same orientation if the angle between their main axes is

less than or equal to 10 degrees, and they are neighbors if their centers of mass are less than or

equal to 1.5 a apart. The main axis of a molecule is the principal axis with the smallest moment

of inertia. By this definition, one molecule is either part of the nucleus or not; it cannot be

"partially crystalline." Because the PYS force field predicts a persistence length of about 0.8

nm,[85] which is comparable to the extended length of one n-octane chain, 0.82 nm, assuming

that every chain of n-octane joins the crystal as a whole, rather in segments, is reasonable.

44

Page 45: Molecular Simulation Study of Homogeneous Crystal Nucleation

In addition to nma, two other variables that measure the order of the system were monitored.

The first variable is the global orientation order parameter, P 2, which is defined as

P2 =. , O2 U (3.16)2 i,j~iI

where Oy is the angle between the vector from the (i - 1 )th bead to the (i + 1 )th bead, and the

vector from the (j - 1 )th bead to the (j + 1 )th bead. The average is taken over all pairs of chords

and over all molecules. The second variable is the fraction of torsions in the trans state in the

system, Ptrans. A trans state for a torsion angle is defined as a state in which the torsion angle is

between -60* and +600. P 2 and Ptrans are standard measures of the global order in a chain

molecule system. In comparison, nmax is a measure of the local order.

3.2.3 Monte Carlo simulation

Nucleation at small supercooling is not accessible by brute-force MD simulation because of the

high free-energy barrier. This difficulty can be alleviated by the use of biased MD simulations,

as demonstrated previously for a Lennard-Jones system.[75] The application of biased MD

simulations, however, is limited because the biasing parameter needs to be an explicit function of

particle coordinates for the biasing force to be calculated. By contrast, biased MC simulation

does not have such a limitation; therefore biasing techniques are easier to implement in MC

simulations than in MD. Furthermore, MC simulation allows the use of unphysical moves, e.g.,

45

Page 46: Molecular Simulation Study of Homogeneous Crystal Nucleation

end-bridging moves,[92] to sample phase space more efficiently. This sampling efficiency is

essential to equilibrate systems of complex molecules like polymer melts.

In our Monte Carlo simulations, each Monte Carlo cycle consisted of Nbeads trial moves, where

Nbeads was the total number of beads in the system. The trial moves were randomly chosen from

three types: (1) local displacement of one bead, (2) reptation of one end bead, and (3)

configuration-biased re-growth[93]. These three moves were chosen with relative probabilities

of 50:45:5. In addition, each Monte Carlo cycle also contained five volume change moves. The

relative probabilities of all four types of Monte Carlo moves were chosen to obtain rapid

equilibration, as measured by the mean-squared displacement (MSD) of beads and decay of the

chain end-to-end vector autocorrelation. The acceptance ratios for the displacement and volume

change moves were both 50%, controlled through the choice of the maximum bead displacement

and maximum volume change. For configuration-biased re-growth, the number of beads

displaced was chosen uniformly and randomly between 1 and 8. No attempt was made to

optimize the acceptance ratios of the configuration-biased re-growth or the reptation moves.

Their acceptance ratios under these simulation conditions were 16% and 0.3%, respectively.

The umbrella sampling technique[68] was used to sample the free energy of formation of crystal

nuclei during MC simulation. In umbrella sampling, a biasing potential energy is added to

improve the sampling of configurations with small Boltzmann factors; the bias is subsequently

removed during analysis of the results.

We chose a fixed Ebias (x) = 0.5kx(x - Xtaget )2 in our free energy sampling, where x is the chosen

reaction coordinate. The center and width of the sampling window depended on xtarget and kx ,

respectively. We divided the whole sampling range [Xinit, xrma] into a series of overlapping

46

Page 47: Molecular Simulation Study of Homogeneous Crystal Nucleation

windows. The initial configuration for each window was extracted from a MD trajectory that

exhibited nucleation, such that the initial configuration had an x value in the corresponding

window.

To implement umbrella sampling, we first carried out a sequence of m Monte Carlo moves

without the biasing potential; then we calculated the value of x and accepted or rejected the

whole sequence based on the change in the biasing potential, exp(-AEias / kBT), where AEbia is

the difference in the biasing potential before and after the sequence:

A.Ebias = Ebias (after) - Eblas (before) (3.17)

We used a sequence consisting of one MC cycle. In principle the biasing potential could be

applied after every Monte Carlo move, but, in this case, the calculation of the reaction coordinate

x was computationally expensive, and the value of x was strongly correlated from one Monte

Carlo move to the next, so that the statistics could not be improved much using a shorter

sequence. Sampling was performed once per sequence.

The reaction coordinate x was chosen to be the size of the largest crystal nucleus in the system,

nmax. We divided the reaction coordinate range 0 nma 40 into approximately eight

overlapping sampling windows and used kx = 0.05 kBT for all windows. In each window, an

initial configuration was relaxed for 216 = 65,536 MC cycles, and then statistics were taken over

a run of 218 = 262,144 MC cycles.

47

Page 48: Molecular Simulation Study of Homogeneous Crystal Nucleation

3.3 Results and discussion

3.3.1 Determining the crystal structure

In order to test the performance of the PYS force field in the crystal phase, we prepared a system

of 480 chains with the experimentally determined crystal structure[94, 95] and equilibrated it at

200K using MD simulation. Table 3.1 compares the simulated crystal structure with the

experimental one.

Being a UA force field, the PYS force field does not treat hydrogen atoms explicitly, thus

resulting in a tilted hexagonal structure (7=120*). In addition, the ordered phase generated by

this force field at 200K is, in fact, a "rotator" phase [96, 97], as shown in Fig. 3.1 and confirmed

by the distribution of chain orientation shown in Fig. 3.2, rather than a perfect representation of

the n-octane crystal. The chain orientation in Fig. 3.2 is defined as the azimuthal angle of a

vector in the x-y plane. This vector points from the average of the projection of all odd beads on

one chain on the x-y plane to that of all even beads on the same chain. Translational registry of

the chain centers of mass is maintained in all three directions, which precludes this phase being a

liquid crystal phase. The differences between our simulated crystal structure and the

experimental one can be remedied by employing an all-atom (AA) force field[98] but not by an

anisotropic united-atom model (AUA)[99, 100]. Nevertheless, nucleation of the ordered, rotator

phase is considered to be a sufficiently close approximation to that of the crystalline phase in n-

alkanes for purposes of this study.

48

Page 49: Molecular Simulation Study of Homogeneous Crystal Nucleation

TABLE 3.1 Crystal structure of n-octane at 200K.

: a, b, c, a, P, y are lattice constants of a triclinic crystal.

49

Experiments MD simulation

Crystal structure* Triclinic Triclinic

a (nm)* 0.422 ± 0.002 0.47 ± 0.003

b (nm) 0.479 ± 0.002 0.47 ± 0.003

c (nm) 1.102 ± 0.002 1.226± 0.005

a (degree) 94.7 ±0.3 81.7± 1.0

P (degree) 84.3 ±0.3 101.1 ±1.0

y (degree) 105.8 ±0.3 120.0± 1.0

density (g/cm 3 ) 0.858 0.826 ± 0.003

Page 50: Molecular Simulation Study of Homogeneous Crystal Nucleation

111111111 111 11 il

(a) (b)

101

164 4'4 1#- 14.4

FIG 3.1 A snapshot of 480 n-octane chains simulated at 200K. (a) viewed along the a-

axis of the triclinic unit cell; (b) viewed along the b-axis of the triclinic unit cell; (c)

viewed along chain direction. This figure was rendered by VMD [101].

50

Page 51: Molecular Simulation Study of Homogeneous Crystal Nucleation

0.012

0.010

0.0080

CU)

LL 0.006

0.004

0.002-180 -135 -90 -45 0 45 90 135 180

Chain orientation angle (degree)

FIG 3.2 The chain orientation distribution for a system of 480 n-octane chains at 200K in

the rotator phase. The chain orientation is defined as the azimuthal angle of a vector in

the x-y plane. This vector points from the average of the projection of all odd beads on

one chain on the x-y plane to that of all even beads on the same chain.

3.3.2 Determining the equilibrium melting point and the heat of fusion

The equilibrium melting temperature Tm is a crucial reference point for subsequent analysis of

crystallization behavior. Ko et al. [102] used a force field similar to PYS and indirectly estimated

the Tm of C400 to be within 10K of the experimental value, 410K[103]. Waheed et al.[84] also

used the PYS force field and reported the melting of n-eicosane around 345K for a system with

51

Page 52: Molecular Simulation Study of Homogeneous Crystal Nucleation

periodic boundaries, compared to the experimental value of 31OK[9]. The discrepancy in

Waheed's study was explained to be due in part to the superheating required to nucleate the melt

phase within an essentially infinite crystal (because of the periodic boundary conditions in the

simulation); in real systems, melting typically proceeds from the surface inward.

To avoid this problem, we determined Tm using a simulation with a crystal-melt interface, as

proposed by Bai et al. [104] For this purpose, we first created a system comprising a perfect

crystal of 480 chains and then increased the Lennard-Jones parameter V for half of the system by

a factor of two, effectively raising the melting point for this half of the system to some Tm' > Tm.

Then, by trial-and-error, we chose a temperature T1 between Tm' and Tm such that the unmodified

half of the system melts, thus generating a system with a flat crystal-melt interface parallel to the

(100) crystal facet. For this purpose, T was set to 300K. Then all beads were restored to the

original c value, and the system was quenched to a lower temperature T2 between 200K and

220K. (The experimental value of Tm for n-octane is 216.4K.) We then monitored the

displacement of the crystal-melt interface at T2 (Fig. 3.3).

The crystal fraction in the simulation was quantified using the global orientation order parameter

P2. An increase of the crystal domain at the expense of the melt domain is signaled by an

increase of P2 , indicating that the temperature of that simulation is lower than Tm. If P2 decreases,

then the simulation temperature is higher than T.. Thus, Tm is identified with the value of T2 at

which P2 remains essentially unchanged with time. At each T2, four MD simulations were

performed using randomized initial velocities, with similar results. Fig. 3.4 shows representative

trajectories at five different T2s. Through this procedure, we determined that Tm = 212 ± 1K for

n-octane using the PYS force field, which agrees well with the experimental value 216.4K.

52

Page 53: Molecular Simulation Study of Homogeneous Crystal Nucleation

(a)

(b)

(c)

FIG 3.3 Determination of the equilibrium melting temperature with a crystal-melt

interface. 480 n-octane chains are rendered as line models. Boundaries are periodic in

all three directions. Two periodic images in the horizontal direction are shown to

emphasize more clearly the interfaces between the amorphous melt and ordered

crystalline regions. (a) at T = 210K, the interface moves toward the melt region; (b) at

T=212K, the interface stays stationary for at least 3 ns; (c) at T = 214K, the interface

moves toward the crystalline region, and melting of the crystal can be seen. This figure

was rendered by VMD.

The heat of fusion per molecule at the melting temperature is an important quantity for phase

transition studies. It is calculated from

AHf = AE + PAV, (3.18)

53

Page 54: Molecular Simulation Study of Homogeneous Crystal Nucleation

where the changes of energy and volume are due to the phase transformation at constant pressure

and constant temperature. At constant temperature, the kinetic energy does not change, so we

only measured the potential energy and average density at the simulated equilibrium melting

temperature Tm = 212K for crystal and melt states, respectively. These values and the average

density of both crystal and melt states at several temperatures are presented in Table 3.2. The

calculation yields AHf = 12.7 ± 0.2 kJ/mol. As a comparison, the value cited by Oliver et al.[7]

is AHf = 20.68 kJ/mol. We attribute this difference to the fact that our simulated crystal phase is

a "rotator" phase, rather than a perfect crystal phase.

0.8 - - -....-o-208K-o- 210K-e- 212K

0.6 A 214K-v- 216K

0.4

0.2

0.00.0 0.5 1.0 1.5 2.0 2.5 3.0

Time (ns)

FIG 3.4 Global orientation order parameter P2 as a function of time at five different

temperatures, for a system of 480 n-octane chains with a crystal-melt interface. Within

the precision achievable in 3 ns, the equilibrium melting temperature Tm is 212 ± 2K.

54

Page 55: Molecular Simulation Study of Homogeneous Crystal Nucleation

TABLE 3.2 Potential energy per chain and average density of n-octane systems at pressure P = 1 atm and

several different temperatures.

T potential energy per chain average density

(K) (kJ/mol) (g/cm 3)

crystal state melt state crystal state melt state

212 -37.60 ± 0.34 -24.86 ± 0.34 0.818 ± 0.003 0.745 ± 0.003

190 -40.68 ± 0.28 -28.17 ± 0.30 0.831 ± 0.003 0.760 + 0.003

180 -41.92± 0.26 -29.83 + 0.29 0.836 ± 0.003 0.767 ± 0.003

170 -43.09± 0.24 -31.59 + 0.32 0.840 ± 0.003 0.775 ± 0.003

3.3.3 Molecular dynamics simulation of the nucleation process

Having determined the equilibrium melting temperature Tm, we performed MD simulations to

study crystal nucleation in n-octane melts. First the system was equilibrated in the melt state at

250K for 1 ns; then it was quenched to 170K, which is about 20% supercooling. In a typical

trajectory after quenching (Fig. 3.5), three time periods can be observed: (i) an initial period

from t = 0 to 2 ns, during which the potential energy decreases rapidly to re-establish

equipartition of energy after the quench; (ii) an induction period from t = 2 to 31 ns, during

55

Page 56: Molecular Simulation Study of Homogeneous Crystal Nucleation

which the system is metastable and there is no evidence of a nucleation event; and (iii) a period

of crystal growth after a nucleation event occurs (t > 31 ns). Around t = 28 ns, a nucleus with n

~ 27 apparently forms but is short-lived. Around t = 31 ns, another nucleus forms that grows

rapidly to a size of n - 25 and then serves as the object from which the rest of the system

crystallizes. This observation is consistent with the picture of classical nucleation theory. As

demonstrated below, the top of the free energy barrier is relatively flat, and there is a finite

probability that any particular nucleus of size comparable to the critical value, n*, will either re-

cross the barrier and melt or else proceed to form a stable crystal phase. The onset of nucleation

is clearly represented by a sudden increase of nmax, whereas overall density, potential energy, P 2,

and Ptrans are all less sensitive to the nucleation event and show delayed response to nucleation.

56

Page 57: Molecular Simulation Study of Homogeneous Crystal Nucleation

0

CD

0EL

CD

0)

C0

.4-0-

MF

0-o

0.

0 5 10 15 20 25 30 35 40

Time (ns)

0.26 <0

0.25 ~CD3j

0.24

0-23

0.30a

0.2 7

30.1 CD

0.0

0

CD

0

FIG 3.5. Evolution of characteristic variables for a system of 960 n-octane chains during a

typical MD simulation, after quenching from 250K to 170K at t = 0. (a) potential energy

and volume per chain; (b) size of the largest nucleus, nmax, and the global orientation

order parameter, P2; (c) fraction of trans states, Ptrans*

In order to evaluate how the nucleus definition influences this result, we re-analyzed the same

MD trajectory using two additional sets of cutoffs in Esselink's nucleus definition, i.e., (9e = 15*,

re = 1.8a) and (Oe = 5*, re = 1.3a) (Fig. 3.6). The nmax curve with cutoffs (01 = 5*, re = 1.3a)

57

Page 58: Molecular Simulation Study of Homogeneous Crystal Nucleation

displays considerable fluctuation after the nucleation event occurs, at around 32 ns; this

fluctuation suggests that the size of a nucleus is overly sensitive to small motions at its surface,

and that the nucleus definition is too restrictive. On the other hand, the nmax curve with cutoffs

(Oe = 150, re = 1.8a) displays "spikes" after the nucleation occurs, which we trace to the

"merger" and subsequent "splitting" of two different nuclei; thus, this nucleus definition is too

lenient. Therefore we confirmed (Oe = 10*, re = 1.5a) to be a good empirical choice for the

nucleus definition for purpose of our study.

500

M f)-- = 15*, r = 1.8a, Y-axis offset = 100E C c

400 -- o- = 100, r = 1.5a, Y-axis offset = 50

-9 = 50, r = 1.3a, Y-axis offset = 0C~ 30C

U> 3 300 -

)L-M~ 200

U)

o 100 -4)N

UI)0 - '

0 5 '10 15 20 25 30 35 40

Time (ns)

FIG 3.6 The maximum nucleus size nmax as a function of time, using three different sets

of cutoff angle and cutoff radius in Esselink's definition of nucleus. Vertical offsets were

made for purposes of clarity.

58

Page 59: Molecular Simulation Study of Homogeneous Crystal Nucleation

In all, we performed 48 independent MD simulations for a system of 480 chains and 24

independent MD simulations for a system of 960 chains, each quenched from 250K to 170K at

time t = 0. The systems with 480 chains were simulated for 60 ns; 30 of the 48 simulations

exhibited nucleation, as typified by Fig. 3.5. The systems with 960 chains were simulated for 30

ns; 20 of the 24 simulations exhibited nucleation.

0 10 20 /7*

n30 40 50

max

FIG 3.7 The mean first passage time (MFPT) of maximum nucleus size nmax from 24 MD

simulations of a system of 960 n-octane chains quenched from 250K to 170K at t = 0.

The open circles are simulation data, and the solid line is the formula of Eq.(2.15), with

n*, r* and b parameterized to fit the simulation data.

Wedekind et al.'s MFPT method[63] was used to estimate the critical nucleus size n* and the

average induction time . The results are n*= 19 and v*= 24± 12 ns for a system of 480 chains,

and n' = 21 and / = 16± 10 ns for a system of 960 chains. Fig. 3.7 illustrates how to estimate n*

59

24

20

16

12

8

U,)

E

I-a-LL

4

0

Page 60: Molecular Simulation Study of Homogeneous Crystal Nucleation

and z * through a fit of Eq.(2.15) to the MFPT of nma, Z(nma). Fig. 3.7 shows a large statistical

error in z(nmax) for nmx> n*; this error is due to the limited number of MD trajectories used for

averaging. Moreover, some MD simulations failed to exhibit nucleation throughout their whole

simulation time, so r is underestimated, more so for the system of 480 chains than for the

system of 960 chains because the former has a higher percentage of simulations that failed to

exhibit nucleation. Therefore, the induction time T* estimated for a system of 960 chains is more

reliable.

The induction time r* is equal to the inverse of the product of the nucleation rate I and the

volume of the system V:

r* (3.19)IV

Plugging r = 16 ± 10 ns into Eq.(3.19), we obtain a nucleation rate I equal to (2.7 ± 0.6) x 1026

cm-3sec-1 for a system of 960 chains at 170K, where the system volume is calculated from the

melt state density in Table II. We further calculate the critical free energy AG* using Eq.(2.5),

where the kinetic prefactor Io was calculated by Uhlmann et al. to be 3.71 x 1032 cm-3sec'.[5]

The calculated critical free energy AG* is 14.1 ± 0.4 kBT for a system of 960 chains at 170K.

However, this value of AG* should be treated with caution because it is hard to determine Io

precisely. Uhlmann et al. calculated Io using Eq.(2.6)-(2.7), an approximation that is good for

molecules of isotropic shape, e.g., fused salts, metals, and simple organic liquids. However, the

n-octane molecule is anisotropic, and only those molecules that reach the nucleus surface with

60

Page 61: Molecular Simulation Study of Homogeneous Crystal Nucleation

certain orientations contribute to effective attachment. Therefore Io is overestimated by Eq. (2.6)

-(2.7), and thus AG* is also overestimated. It is desirable, then, to employ another simulation

method from which one can calculate the critical free energy AG* directly, without having to

invoke additional assumptions or approximations.

3.3.4 Monte Carlo sampling of the nucleation free energy barrier

The free energy of formation, or reversible work, of a n-sized nucleus, AG(n), can be calculated

from the equilibrium nucleus size distribution, similar to Eq.(2.20), as follows: [51]

AG(n) /kB = -In N(n) / N, + constant (3.20)

where N(n) is the number of n-sized nuclei observed during the simulation, and N, is the total

number of sites in the system on which nuclei can form, which is equal to the total number of

chains in the system. Because a 1-sized nucleus correspond to a molecule in the melt state by

our nucleus definition, the constant in Eq. (3.20) can be determined by equating AG(n=1) to zero.

For sufficiently small values of n, multiple nuclei of various n are observed to form

spontaneously within a system of finite size. For larger values of n, especially those approaching

the critical size n*, umbrella sampling is required to ensure adequate statistics for estimating the

relative probabilities of these nuclei. In general, only one nucleus of size nm is observed in

each of the biased simulations, and nnm, is well separated from those small values of n where

multiple nuclei spontaneously occur. Thus, within a given simulation in which umbrella

61

Page 62: Molecular Simulation Study of Homogeneous Crystal Nucleation

sampling is used, we can approximate the differences in free energy for nuclei of different sizes

using Eq.(3.20), replacing n by nmax and N, by the number of systems sampled. Simulations with

different biases are chosen to ensure overlapping ranges of nmax, so that all of the free energy

curves can be shifted subsequently to form a single, continuous, universal curve of AG(n) vs n.

In this way, we constructed the whole free energy curve without loss of accuracy. We have

checked this approach by reproducing an Ising model nucleation simulation by Chandler et

al.[79]

Fig. 3.8 shows AG(n) versus n at 170K. In order to evaluate the finite size effect, we present the

results for systems of 480 chains and 960 chains. The nucleation free energy curve does not

change with system size, indicating that a system of 480 n-octane chains is sufficiently large to

be free from finite size effects under these simulation conditions.

62

Page 63: Molecular Simulation Study of Homogeneous Crystal Nucleation

-- 10(D

£ 80CU,E 6

0 4

2

) 01 0(DL_ I I I 4

N 480 n-octane chainso 960 n-octane chains

I I * I I I I I i

UL 0 5 10 15 20 25 30 35 40 45

Nucleus size n

FIG 3.8 The free energy of formation for a crystal nucleus in a melt of n-octane chains at

170K. System with 480 chains (stars); system with 960 chains (open circles). In both

cases the critical nucleus size is 18 chains and the free energy barrier height is 9.3 ± 1.0

kBT.

From the critical free energy AG* obtained by Monte Carlo and the nucleation rate I reported

above from MD simulations, both at 170K, we determined the kinetic prefactor lo using Eq.(2.5),

to be 1o = (2.95 ± 3.61) x 1030 cm-3sec~1. This is about two orders of magnitude lower than the

value estimated by Uhlmann et al. using Eq. (2.6)-(2.7). Since Io is only weakly temperature-

dependent, we can use this Io to estimate the nucleation rate at other temperatures, if the critical

free energy AG* for those temperatures is known. This is particularly valuable for temperatures

close to the melting point, where brute-force MD simulations are too inefficient to study the

nucleation event directly.

63

mom

Page 64: Molecular Simulation Study of Homogeneous Crystal Nucleation

Similar free energy sampling has been performed at 180K and 190K. The critical nucleus size n'

and the critical free energy AG* for an n-octane melt at 170K, 180K, and 190K, obtained from

the maxima in the curves for AG(n)/kBT vs n, are summarized in Table 3.3. At 170K, the

critical nucleus size obtained by MC is in reasonable agreement with that indicated by the MD

results. As the supercooling AT decreases toward zero, both the critical size n'* and the critical

free energy AG* increase, which is consistent with classical nucleation theory. We did not

observe a free energy barrier at 190K, probably because AT was so small. In this case, the

critical nucleus may have been too big for our finite-size system.

TABLE 3.3

containing

The crystal nucleation free energy as a function of nucleus size n for an n-octane melt

960 chains.

*: We did not observe a critical nucleus in simulations at 190K.

64

T AT/Tm n AG*

(K) (kBI)

170 19.8% 18 ±3 9.3 ±1.0

180 15.1% 23 ±3 12.5 ±1.0

190 10.4%

Page 65: Molecular Simulation Study of Homogeneous Crystal Nucleation

The crystal-liquid interfacial free energy a-of n-octane has been calculated by Uhlmann et al. [5]

and by Oliver et al [7] from their experimental measurements of nucleation rate at A T/Tm values

of 0.111 and 0.138, respectively. By assuming a spherical nucleus model, they estimated a to be

between 10.3 and 13.4 mJ/m 2 . We also used the spherical nucleus model to calculate a from the

AG* determined by Monte Carlo simulation, using Eq.(2.2), and obtained a value for a of 12.0

mJ/m 2 at 170K. The values of Tm and AHf in Eq. (2.2) both come from our MD simulations,

reported above. The a determined in this way from the simulation data is consistent with the

experimental studies. The free energy curves of the spherical nucleus model are shown in Fig.

3.9, and illustrate the sensitivity of the critical free energy to the value of surface energy.

65

Page 66: Molecular Simulation Study of Homogeneous Crystal Nucleation

-- Spherical model (Oliver)'A 16 ---- Spherical model (Uhlmann)-

- This work.)~ --- - Spherical model

< - - Cylindrical model

CU

O -8

4 -0 \

)

LL 0 5 10 15 20 25 30 35 40 45 50

Nucleus size n

FIG 3.9 The free energy of formation for a crystal nucleus in a melt of 960 n-octane

chains at 170K. Simulation data (filled circles); spherical nucleus model using surface

free energy a of Uhlmann et al. (dot line); spherical nucleus model using surface free

energy a of Oliver et al. (dash line); spherical nucleus model fit using surface free energy

o= 12.0 mJ/m 2 , chosen to match the critical nucleus free energy from simulation (dash

dot line); cylindrical nucleus model using surface free energies a-=5.4mJ/m 2 for the end

surface and o;=6.8mJ/m 2 for the lateral surface, chosen to fit both the critical nucleus

free energy and the critical nucleus size of the simulation data (solid line).

Although the spherical nucleus model can be used to calculate an interfacial free energy o- from

the critical free energy AG*, the shape of the free energy curve deviates from the simulation data

significantly (Fig. 3.9). Visual inspection of nuclei indicates that the spherical model is a poor

description of the actual nuclei for n-octane. Fig. 3.10 is a snapshot of a crystal nucleus

66

Page 67: Molecular Simulation Study of Homogeneous Crystal Nucleation

containing 18 n-octane chains at 170K; its shape is more cylindrical than spherical. Furthermore,

the critical nucleus size n* calculated from AG* based on Eq.(3.6) is significantly smaller than the

n* measured in simulation, proving that the spherical nucleus model is not an appropriate model

for the n-octane crystal nuclei. Because the variable-thickness cylinder model also has to

satisfy Eq.(3.6), we conclude that neither the spherical nucleus model nor the variable-thickness

cylindrical nucleus model describes the crystal nucleus of n-octane chains well.

(a) (b)

FIG 3.10 A snapshot of a crystal nucleus thatconsists of 18 n-octane chains from a Monte

Carlo simulation of a melt of 960 n-octane chains at 170K. (a): side view; (b): top view.

This figure was rendered by VMD.

Therefore we turn to the fixed-thickness cylinder model. The length of an extended n-octane

molecule 10 is 0.82 nm. Applying Eq.(3.10)-(3.1 1), we obtain the interfacial free energy of the

end surface o- = 5.4 mJ/m2 and of the side surface o-s = 6.8 mJ/m 2 at 170K. Using these free

energy values, the fixed-thickness cylinder model fits our simulation data well over the whole

range of n (Fig. 3.9). Similarly, the interfacial free energies at 180K and 190K are calculated

(Table 3.4). Although we did not observe any critical nuclei at 190K, we can still parameterize

67

Page 68: Molecular Simulation Study of Homogeneous Crystal Nucleation

the fixed-thickness cylinder model, Eq.(3.7), to match the pre-critical part of the simulated free

energy curve and to obtain the interfacial free energies. The interfacial free energies at 190K

predict a critical nucleus size n' of about 113, which is too big to observe reliably in our system

of only 960 chains. The fixed-thickness cylindrical nucleus model captures quantitatively the

formation of a crystal nucleus in the n-octane system for all three supercooling temperatures (Fig.

3.11).

24 - -a-190K22 -<-180K20 - -o- 170K18 -

0 16 -

Cfj14

E 120

o 8CY, 6

Q) 2

LL -2 ' '-5 0 5 10 15 20 25 30 35 40 45 50 55 60

Nucleus size n

FIG 3.11. The free energy of formation for a crystal nucleus in a melt of 960 n-octane

chains at 170K, 180K, and 190K, respectively. Simulation data (open symbols); fixed-

thickness cylindrical nucleus model using Eq.(3.7) (solid curves). The values of the

interfacial free energy cr and o, used in generating the modeled curves are presented in

Table 3.4.

68

Page 69: Molecular Simulation Study of Homogeneous Crystal Nucleation

Table 3.4 Crystal-liquid interfacial free energy of n-octane molecules.

As AT decreases toward zero, the interfacial free energy does not change much for the side

surface; however, it decreases significantly for the end surface. This difference is attributed to

the longitudinal chain motion, which increases dramatically with increasing temperature; the

transverse chain motion is not sensitive to temperature. This temperature dependence of chain

mobility is consistent with previous observations by Ryckaert et al. [96, 97]

3.4 Conclusion

In this chapter, we have studied homogeneous nucleation of the crystal phase from an n-octane

melt, using both MD and MC methods. For the first time, the critical nucleus of a chain

69

T Ge us

(K) (mJ/m2) (mJ/m 2)

170 5.4 ±0.5 6.8 ± 0.7

180 3.3 ±0.3 8.4 ± 0.8

190 2.9 ±0.3 8.0 ±0.8

Page 70: Molecular Simulation Study of Homogeneous Crystal Nucleation

molecule system was identified, both kinetically using an unbiased MD method and analysis of

mean first passage times, and thermodynamically using the MC method with umbrella sampling.

The results of both methods are in reasonable agreement with each other and with the available

experimental data. Within the framework of the classical nucleation theory, a cylindrical nucleus

model provides a reliable description of the dependence of free energy on nucleus size, AG(n). It

allows us to estimate the critical nucleus size even at small degrees of supercooling, from

simulation results for pre-critical nuclei, and to calculate the solid-liquid interfacial free energies

for both the end-surface and side-surface of the nucleus. The decoupling of these two surfaces is

important to understand the crossover from extended chain nucleation to folded chain nucleation

in chain molecule systems.

Homogeneous nucleation of the crystal phase from a quiescent melt of mono-disperse short

chains represents an idealized case. However, we believe that the concepts and methods

employed in this study can be extended to study more complicated nucleation behavior typical of

longer chains, e.g., integer-folded nucleation and flow-induced nucleation, because the properties

of the critical nucleus should also be the controlling factors therein.

70

Page 71: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 4 Bundle-like crystal nucleation in n-eicosane (C20) melts

Previously we have studied the crystal nucleation from n-octane (C8) melt. At the crystallization

temperature, each C8 molecule behaves like a rod, and the crystal nucleus can be well described

by a fixed-thickness cylindrical nucleus model. In this chapter, we extended our study to a

longer molecule, n-eicosane (C20), where the flexibility of chain will allow the molecules to

participate the crystal nucleus only partially. The force field used here is the same as chapter 3,

and there are 336 n-eicosane chains in the simulation box.

4.1 Methods

4.1.1 Crystal nucleus definition

In our previous study of nucleation in n-octane, we adopted the algorithm of Esselink et. al. [21]

to define a crystal nucleus; in that algorithm, two chains belong to the same nucleus if the angle

between their principle axes is less than 10 degrees and their centers of mass are less than 1.5a

apart (where a is the length parameter in the Lennard-Jones potential). This algorithm is

acceptable for chains that are short compared to the persistence length of the chain, but yields

spurious results when applied to longer chains. For this reason, we introduce here a new

71

Page 72: Molecular Simulation Study of Homogeneous Crystal Nucleation

algorithm that is applicable to flexible chains, based on a local p2 order parameter defined for

each UA, or "bead". The local p2 order parameter of the i-th bead is defined as

P2(0) = 2ji1(4-1)

where Oy is the angle between the vector from the (i - 1)th bead to the (i + 1)th bead, and the

vector from the (i - 1)th bead to the (i + 1)th bead. The average in Eq.(4. 1) is taken over all of the

j neighboring beads of bead i that lie within a cutoff distance ry<rp2. By contrast, the global

orientation order parameter P2 is defined as

P2 K3cos' /ij, (4.2)2 i#j

where the average is taken over all pairs of (ij) in the system regardless of the distance between

i andj. A threshold P2,th can be chosen so that beads with p2 greater than P2,th are assigned to the

crystal phase. Two crystal beads i andj are assigned to the same crystal nucleus if ry is less than

a threshold value rth. Thus, suitable values for three parameters, rp2, rth and p2,th, are needed.

Determination of numerical values for these parameters is presented in the results section. The

size of a crystal nucleus is the number of beads that constitute this crystal nucleus.

4.1.2 Molecular dynamics simulation

MD simulations were performed using open source code for the LAMMPS package[89]

exclusively. In the MD simulations, the initial velocities of all beads were generated from the

Maxwell-Boltzmann distribution according to the desired temperature, and the equations of

72

Page 73: Molecular Simulation Study of Homogeneous Crystal Nucleation

motion were integrated using the velocity Verlet method with an integration time step At equal to

5fs. Waheed et al. [84] and Lavine et al.[91] have previously evaluated the effect of using larger

integration time steps during the MD simulation of n-alkanes with this PYS force field. They

found that changing At from 1 to 5 fs increased the bond energy, the angle-bending energy, and

the relaxation time for the chain-orientation autocorrelation function by approximately 10%.

The consequence of these errors is mitigated through the use of a thermostat, such that no

detectable difference in crystallization kinetics was observed for simulations with At between 1

and 5 fs. Temperature and pressure were maintained at their set point values using the

Nose/Hoover thermostat and barostat; the damping frequencies were (oT =1/(1OAt) for the

thermostat and (op =1/(1OOOAt) for the barostat. Correspondingly the thermostat "mass" is

Q=NfkB T/T 2 and the barostat "mass" is W=(N+d)kBT/op2, where Nf is the number of degrees of

freedom and d =3 is the dimensionality of the system.[105]

4.1.3 Monte Carlo simulation

The types of MC moves are the same as in Chapter 3: (1) local displacement of one bead, (2)

reptation of one end bead, (3) configuration-biased re-growth and (4) volume change. In each

Monte Carlo cycle, Nbead moves were chosen from the first three types with relative probabilities

of 40:40:20, where Nbead was the total number of beads in the system. Each Monte Carlo cycle

also contained five volume change moves. The relative probabilities of all four types of moves

were optimized to obtain rapid equilibration, as measured by the mean-squared displacement

(MSD) of beads and decay of the chain end-to-end vector autocorrelation. The acceptance ratios

for the bead displacement and volume change were both 50%, controlled by adjusting the

73

Page 74: Molecular Simulation Study of Homogeneous Crystal Nucleation

maximum bead displacement and maximum volume change. For configuration-biased re-growth,

the number of re-grown beads was chosen between 1 and 20, uniformly and randomly. No

attempt was made to optimize the acceptance ratios of the configuration-biased re-growth or the

reptation moves.

The umbrella sampling method was also the same as Chapter 3. The reaction coordinate nmax

was used as the biasing parameter. The reaction coordinate range was divided into a series of

overlapping sampling windows at intervals of around 25 in xtaget, and k, = 0.0025 kBT was used

for all windows. The initial configuration for each window was extracted from a MD trajectory

that exhibited nucleation, such that the initial configuration had an nma value within the

corresponding window. In each window, an initial configuration was relaxed for 16384 MC

cycles, and then statistics were taken over a run 65536 MC cycles.

4.2 Results and discussion

4.2.1 Melting point and heat of fusion

We first determine the melting point and heat of fusion predicted for n-eicosane by the PYS

force field. The equilibrium melting temperature Tm in particular is a crucial reference point for

accurate evaluation of the degree of under-cooling, AT. For this purpose, we use the method of

Bai et. al., [104] in which a simulation containing coexisting crystal and melt phases separated

by a flat interface is constructed, and the growth of one phase at the expense of the other is

monitored over time for a range of different temperatures. The equilibrium melting temperature

is then identified as the temperature at which neither phase grows, and the interfaces remain

74

Page 75: Molecular Simulation Study of Homogeneous Crystal Nucleation

essentially unmoved. We previously employed this method in our study of n-octane, and were

able to identify Tm to an accuracy of a few degrees. Our procedure for constructing the

coexistence of phases and subsequently monitoring the growth or depletion of the crystal phase

is described in detail in that work.[106] Using this method, the PYS force field predicts Tm for n-

eicosane to be 310+2K, which is in remarkably good agreement with the experimental value

309.75K. [107]

As was the case for n-octane, the crystal phase for n-eicosane predicted by the PYS force field is

an RI rotator phase, in which molecules are aligned and in register, but there is no orientational

correlation in the plane transverse to the chain axes. In real n-eicosane, the RI rotator phase is

only metastable with respect to the triclinic crystal phase.[8] This discrepancy between the

stable ordered phase in the real material and that predicted by the PYS force field is a known

shortcoming of UA force fields. Although our simulations do not reproduce the ultimate stability

of the triclinic crystal phase, nevertheless, we believe the rotator phase predicted in our

simulations is relevant in the nucleation of n-eicosane, where it has been proposed that the

rotator phase arises as an intermediate phase. [8, 108]

Similar to Chapter 3, in order to calculate the heat of fusion, MD simulations were performed for

the crystal phase and the melt phase separately at several temperatures and the values of the

potential energy and average density are presented in Table 4.1. The PYS force field predicts a

value of AH = 49.16 ± 2.59 kJ/mol for n-eicosane. Experimental measurements for n-eicosane

indicate that the heat of fusion between the melt and the triclinic crystal phase is 69.87

kJ/mol,[107] while the heat of fusion between the melt and the RI rotator phase is 46.17

75

Page 76: Molecular Simulation Study of Homogeneous Crystal Nucleation

kJ/mol.[109] The simulated value is in reasonable agreement with the experimental value for the

rotator phase.

TABLE 4.1 Potential energy per chain and average density of n-eicosane systems at pressure P = 1 atm

and several different temperatures.

T (K) Potential energy per chain (kJ/mol) Average density (g/cm3 )

crystal state melt state crystal state melt state

310 -72.55± 1.37 -23.43± 1.22 0.910 ±0.003 0.803 ±0.003

280 -84.64 ± 1.06 -36.09 ± 1.00 0.928 ± 0.002 0.819 ± 0.003

265 -89.88 ± 1.19 -42.44 ± 0.86 0.934± 0.002 0.828 ± 0.003

250 -94.79± 1.11 -49.31 ±0.90 0.939 ±0.002 0.836 ±0.003

4.2.2 MD simulation of the nucleation process

Before analyzing the simulation for nucleation, we need to determine appropriate parameters to

define the size of the crystal nuclei, that is, to determine the three parameters, rp2, rth and P2,th,

introduced previously in the Method section. For this purpose, we first measured P(p2(i)), the

probability distribution of the local order parameter p 2(i), in the crystal-melt coexistence

simulation used to determine Tm. As expected, the distribution is bimodal (Fig.4. 1). The two

peaks in the distribution correspond to a disordered melt phase (lower P2) and an ordered crystal

phase (higher p2). The P2 value that corresponds to the minimum of the distribution, P2,th, serves

naturally as the dividing point to distinguish beads in the melt phase from those in the crystal

phase for numerical analysis. Both the probability distribution and P2,th depend on the cutoff76

Page 77: Molecular Simulation Study of Homogeneous Crystal Nucleation

radius rp2 with which we calculate all p2(i)'s. Fig. 4.1 shows the probability distributions

obtained using several different values of rp2 for calculating p2(i). We found that for rp2 in the

range of 1.5 to 3.8 a-, the smallest values of P(p2,th) are obtained, with p2, in the vicinity of 0.4.

e0

0

0

-Q

0

0.60

0.45

0.30

0.15

0.00 I_-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0

p 2(/)

FIG 4.1 Probability distributions of local order parameter P2(i), calculated using five

different values for the cutoff radius rp2.

In order to aggregate crystal beads into nuclei, any two crystal beads that are within a distance

rth=1.3 a are considered to be in the same nucleus; 1.3 a is slightly larger than the first nearest-

neighbor distance for beads on different chains in the crystal phase. For most of this work we

used the parameter set (rp2=2 .5a-, rh=1.3 , P2,th=0.4 ). Below we show how this choice of

parameters can be refined based on knowledge of the first passage time behavior.

77

Page 78: Molecular Simulation Study of Homogeneous Crystal Nucleation

Having determined the equilibrium melting temperature Tm and all of the parameters required to

determine nm,, we performed MD simulations to study crystal nucleation in n-eicosane melts.

First the system was equilibrated in the melt at 400K; then at time t--0, it was quenched to 250K,

which corresponds to about 19% supercooling. In a typical trajectory after quenching (Fig. 4.2),

three time periods can be observed: (i) an initial period from t=0 to -3 ns, during which the

volume and potential energy re-equilibrate after the quench; (ii) an induction period from t=3 to

40 ns, during which the system is metastable and there is no evidence of a nucleation event; and

(iii) a period of crystal growth after a nucleation event occurs (t>40 ns). Around t--20 ns, a

nucleus with n- 10 apparently forms but is short-lived. Around t-40 ns, another nucleus forms

that grows rapidly to a size of n-200 and then serves as the object from which the rest of the

system crystallizes. The onset of nucleation is represented by a rapid increase in nma. This

observation is consistent with the picture of CNT. As demonstrated below, the top of the free

energy barrier is relatively flat, and there is a finite probability that any particular nucleus of size

comparable to the critical value, n*, will either re-cross the barrier and then melt, or else proceed

to grow and form a stable crystal phase.

78

Page 79: Molecular Simulation Study of Homogeneous Crystal Nucleation

20.0

0.0 ~(a) Potential energy per chain 0.62E~ o Volume per chain0

-20.0 0.60

-40.0 -05

(b) n 0-08600

-0C-P

- 2 (D0

0.064000

-00 0

10 20 30 40Time (ns)

FIG 4.2 Evolution of characteristic variables for a system of 336 n-eicosane chains during

a typical MD simulation, after quenching from 400 to 250K at t=0. (a) Potential energy

and volume per chain. (b) Size of the largest nucleus, nlma, and the global orientation

order parameter P2.

We performed 56 independent MD simulations for the system, each quenched from 400K to

250K at time t=0. The mean-first-passage-time (MFPT) method of Wedekind et al.[63] was used

to estimate the critical nucleus size n' and the average induction time r* However, instead of

averaging the first-passage-time Z(nlm.) before the fitting of Eq.(2. 15), we first fitted Eq.(2. 15) to

Z(nlmax) for each of the 56 trajectories. This permits us to collect some statistics on the

79

Page 80: Molecular Simulation Study of Homogeneous Crystal Nucleation

distributions of r*, n' and Z One typical first-passage-time curve and the fitting result are shown

in Fig. 4.3, where the fitting parameters r*=155 ns, n*=126 and Z=O.012 were used.

In Fig.4.3, the first-passage time i(nma) shows a sigmoidal shape that reaches a plateau at

nfma~'200. The presence of this plateau indicates that the rate limiting step in crystallization is

the fluctuation to realize a nucleus of critical size, as the time required for the subsequent growth

of the nucleus is negligible (nm. grows from 200 to 400 in less than 5ns) compared to the

induction time of 155ns.

Fig. 4.4 shows histograms for r*, n' and Z from the 56 simulations, all of which were run until

crystallization had occurred. Making use of the fact that nucleation is a process involving

random events, the probability distribution of r* is expected to follow an exponential distribution,

f(r*) (4.3)

It must be borne in mind, however, that the observed induction time includes not only the true

induction time, but also a waiting time, r', during which the system re-equilibrates after the

quench and the nucleation process is not stationary.[27, 110] As a first approximation, we

assume -r' 3x rr, where zr is the time constant of the end-to-end vector autorcorrelation function

in the melt. We determined zr = 3.25±0.02 ns for C20 at 250K. The two systems that nucleated

within the first 10 ns were therefore discarded, and r'= 10 ns was subtracted from the observed

induction times for the purposes of fitting Eq.(4.3) to the distribution shown in Fig. 4.4a. Using

GnuPlot for nonlinear least squares fitting, the parameter of Eq. (4.3) was determined to be

80

Page 81: Molecular Simulation Study of Homogeneous Crystal Nucleation

<- >=80.6±8.8 ns. As a further check, we also determined <r*> using nonlinear least squares

fitting for other values of r'such that 3x -r, <r'< 70 ns and obtained the same estimate for <r*>

within statistical error.

160 -

V140 -

M 120 -

100

E 80 -

) 60 -

40 -

20 - MD sirrulation-- Fitting

0 50 100 150 200 250 300 350 400

Largest nucleus size nmx

FIG 4.3 The first-passage time of the largest nucleus size, nmax, from one MD simulation

of a system of 336 n-eicosane chains quenched from 400 to 250K at t=0. The open

circles are simulation data, and the solid line is the formula of Eq.(2.15), with n*, r, and

Z parameterized to fit the simulation data.

We assume that n* and Z are Gaussian distributed, as shown in Fig. 4.4b and 4.4c. Their

statistical averages are estimated to be <n*>=1 15±4 and <Z>=0.0082±0.0001. Both n* and Z

exhibit significant tails on the high sides of the distributions; other functional forms for these fits

can be imagined, but these are not pursued further here. The two systems that nucleate during

the waiting time r'= 10 ns were excluded in calculating <n*> and <Z>.

81

Page 82: Molecular Simulation Study of Homogeneous Crystal Nucleation

The average nucleation rate I is readily obtained from the average induction time < r> and the

system volume V according to Eq.(3.19). Using values of <7'>=80.6±8.8 ns and V =

(1.882±0.006)x10-19 cm3 from Table I, we obtain I= (6.59±0.72 )x10 25 cm-3 s at 250K.

7'=10 ns

0.14

0.12Simulation data

0.10 - Exponential fitting

0.08

0.06

0.04

0.02

0.000 50 100 150 200 250 300 350 400

Induction time (ns)

Simulation dataGaussian fitting

0.00 0.05 0.10 .35 0.40

Zeldovich factor Z

a0C0

.0

U,V

(0n00~

450

0.30(b)

0.25 = Simulation data- Gaussian fitting

0.20

0.15

0.10 -

0.05

0.00.0 50 100 150 200 250 300

Critical nucleus size n*

FIG 4.4 The probability distribution of z , n' and Z, parameters obtained by fitting the

first-passage time Eq.(2.15) to each of the 56 independent MD simulations: (a) r* and

the exponential fitting; (b) n' and the Gaussian fitting and (c) Z and the Gaussian fitting.

A waiting time z-'= 10 ns was subtracted from ; for the fitting and two systems that

nucleate within r'were excluded from all fittings.

82

0

-0

(0-o.

N

C0

(0CL

0.35

0.30

0.25

0.20

0.15

0.10

0.05

0.00

Page 83: Molecular Simulation Study of Homogeneous Crystal Nucleation

Next, we tested the sensitivity of the results of the first-passage time analysis to our choice of

parameters used to identify crystal beads and cluster them into nuclei. Fig. 5 shows the first-

passage time of the same MD trajectory shown in Fig. 3, but using three different sets of cutoff

radii (rp2, rth, P2,th). As expected, the critical nucleus size is sensitive to the choice of cutoff radii

used to identify crystal nuclei, but the induction time is for all practical purposes unaffected.

Thus the nucleation free energy barrier is insensitive to the details of the nucleus definition used

in the simulation. On the premise that the most relevant transition path through the critical state

is the trajectory of steepest descent on the free energy surface, our preferred reaction coordinate

is that which yields the transition path with the largest curvature around the free energy barrier,

that is, largest Z defined by Eq.(2.16). According to this criterion, the parameter set (rp2=3.00,

rth 1.3a, P2,th=0. 4 ) could be an improvement over the default values used in this work, but the

two sets are sufficiently similar that we do not expect this refinement to alter in any significant

way the conclusions of this work.

83

Page 84: Molecular Simulation Study of Homogeneous Crystal Nucleation

160

140

120

T' 100

E 80-0

) 60 -_-0) r =2.0a, r*=1.5a, p2*=0.5

CU40 po r 2 =2.5, r*=1.3a, p 2*= 0.420 a r 2=3.0T, r*=1.3, op 2*=0.4

00 50 100 150 200 250 300 350 400

Largest nucleus size nmax

FIG 4.5 The first-passage time of the largest nucleus size, nma, from one MD simulation

of a system of 336 n-eicosane chains quenched from 400 to 250K at t=0, using three

different parameter sets to define the crystal nucleus.

4.2.3 Monte Carlo sampling of the nucleation free energy barrier

We performed MC sampling of the nucleation free energy barrier at three different temperatures:

250K, 265K and 280K. At each temperature the nucleus size distribution P(n) was sampled to

calculate the nucleation free energy AG(n) based on Eq.(13). The free energy curves are shown

in Fig. 4.6, and the critical size n' and critical free energy AG* are summarized in Table 4.2.

84

Page 85: Molecular Simulation Study of Homogeneous Crystal Nucleation

Using GnuPlot and a nonlinear least squares algorithm, a third-order polynomial

AG(n)=ao+ain+a2n 2+a 3n3 was fitted to the free energy curve around the top of the barrier at

250K, and the fitting parameters were determined to be ao=2.67±0.16, a=O. 110±0.006, a2=(-

7.07±0.64)x 10 4 and a3=(1.27±0.1 9)x 10-6, all in units of kBT. From this, the critical nucleus size

n' and the Zeldovich factor Z are readily calculated to be n*=1 10±3 and Z=0.010±0.002, at 250K.

These values are in good agreement with those calculated from MD simulations, n*=1 15±4 and

Z=0.0082±0.0001.

-*1

0

E0

0

16

14

12

10

8

6

4

2

00 50 100 150 200 250 300 350 400

Nucleus size n

FIG 4.6 The free energy of formation for a crystal nucleus in a melt of 336 n-eicosane

chains at 250, 265, and 280K, respectively.

85

Page 86: Molecular Simulation Study of Homogeneous Crystal Nucleation

TABLE 4.2 The crystal nucleation free energy and critical nucleus size at several temperatures for an n-

eicosane melt containing 336 chains.

T (K) AT/Tm n AG*(kBT)

250 19.4% 110 7.3 ±1.0

265 14.5% 140 9.5 ± 1.0

280 9.7% 300 13.2 ± 1.0

Given the nucleation rate I obtained by MD simulation and the critical free energy AG* obtained

by MC simulation with umbrella sampling, we used Eq.(2.5) to calculate the nucleation prefactor

to be Io ~ 1029 cm-3 sec at 250K, or 19% of supercooling. The only reported Io from

experiment for C20 is about 1018 cm-3 sec-1, obtained at 4% of supercooling by Kraack et. al. [9].

However, experimentally determined values of Io appear to be significantly dependent upon

cooling rate, with higher cooling rate yielding lower Io values.[9] For example, Turnbull and

Cormia [6] obtained Io - 1031 cm-3sec-1 for C17 and C18, respectively, under isothermal

conditions, while Kraack et al. [9] reported Io - 1024 and 1025 cm-3sec-1, respectively, at a cooling

rate of 0.2K/min, and Oliver and Calvert. [7] reported Io - 1022 and 1019 cm-3 sec-1 at a cooling

rate of 2.5K/min. The isothermal measurements of Turnbull should yield the most reliable value

of Io. Our results for C20 agree reasonably well with those of Turnbull and Cormia [6], based on

their C17 and C18 results.

86

Page 87: Molecular Simulation Study of Homogeneous Crystal Nucleation

Previously we obtained Io(C8) ~ 1030 cm~3 sec~1 for n-octane, at 20% of supercooling. Io(C20) is

about one order of magnitude smaller than Io(C8) at the same degree of supercooling, which

suggests that the diffusion of the beads at the surface of the crystal nucleus is drastically slowed

down as the chain length increases.

SirS

(a) (b) (c)

FIG 4.7 A snapshot of a crystal nucleus that consists of 130 united atoms from a MC

simulation of a melt of 336 n-eicosane chains at 250K: (a) end view, showing only those

united atoms that are part of the nucleus; (b) side view, showing only those united

atoms that are part of the nucleus; (c) side view, showing all united atoms belonging to

chains that participate in the nucleus. United atoms in the crystal phase are black, while

those in the melt phase are red.

Snapshots of a representative nucleus near the critical nucleation condition at 250K are shown in

Fig. 4.7. Using the site-wise definition of crystal beads for flexible chains described in the

Methods section, we can distinguish sections of chains that are crystalline and noncrystalline.

87

Page 88: Molecular Simulation Study of Homogeneous Crystal Nucleation

We observe that an n-eicosane chain does not join a crystal nucleus as a whole, but rather may

do so only partially, with shorter crystal stems commonly observed at the periphery of the

nucleus. In order to extract thermodynamic quantities like surface energy, we use the cylindrical

model to analyze the shape and size of the crystal nuclei. The volume of each nucleus is

estimated from the total size (number of beads) and the bulk density of the crystal phase at the

relevant temperature. The thickness / of each nucleus is calculated as the average length of the

portion of each chain that participates in the crystal nucleus. Finally, the radius r is obtained by

using the formula for the volume of a cylinder.

From the nucleus shape parameters I and r determined in this manner, a least-squares method

was used to fit the cylindrical nucleus model (Eq.(3.1)) to the free energy curve AG(n)

determined by MC simulation. Two different schemes were employed for this purpose. In the

first scheme, we fixed the value of AGv according to Eq.(2.8) and used only o and o as fitting

parameters. As shown in Fig. 4.8, the fitted curves match the simulation results well at 280K

and 265K, but not so well at 250K. This suggests that the validity of Eq.(2.8) begins to break

down for supercooling around 19% or greater. For this reason, we performed a second fit in

which AG, was also treated as a fitting parameter. With three fitting parameters the cylinder

model matches the free energy curve very well at all three temperatures. The results of the fitting

parameters or, o and AG, are reported in Table 4.3.

88

Page 89: Molecular Simulation Study of Homogeneous Crystal Nucleation

10 2.4

9(-) 2.2

8 2.0

1.87

1.66 1.4

4 - 1.0

1- 0.8

" Ir %:0.62

- 0.4

1 -T=250K 0.2

0 0.00 25 50 75 100 125 150 175 200 225

Nucleus size n

16 -

14

12

10

8

6

4

2

0-0 50 100 150 200 250

Nucleus size n

, 2.4

+ 1.8

1.6

1.4

1.2

0.6

0.4

280K 0.2

300 3500

z

0X,

C)

D

3

0

10

4)>)

4)

12 2.4

11 (b) -2.2

10 2.0

9 T1 1 1.8

8 1.6

3 -0.6

2 00.4

0 0.0

0 - 1 ' - ' - ' - ' - ' - 0.00 25 50 75 100 125 150 175 200 225 250 275

Nucleus size n

z

0

(D

CL

3'

FIG 4.8 The free energy of formation, the thickness and the radius as a function of

nucleus size, n, for a crystal nuclei observed in a melt of 336 n-eicosane chains by Monte

Carlo simulation at three different temperatures, (a) 250K, (b) 265K and (c) 280K. (+)=

nucleus free energy from simulation; (A)=thickness I of nucleus; (0)=radius r of nucleus.

Two fitting schemes to the cylinder model Eq. (1) were used: dashed curve = two-

parameter (a,, a,) fitting and solid curve = three-parameter (cr., o, AG,) fitting.

From the results reported in Table 4.3, we obtain a value of about 4 mJ/m 2 for the end surface

free energy a in n-eicosane. It is relatively independent of temperature. By comparison, o for

89

zC

Co

0(D

0.

C-

Page 90: Molecular Simulation Study of Homogeneous Crystal Nucleation

n-octane was found to decrease with increasing temperature, as the longitudinal chain motion

roughens the end surface much more for short chains than for the longer ones studied here.[106]

For the side surface free energy as, our best estimate is around 10 mJ/m 2 and is relatively

insensitive to temperature. This conclusion is in fair agreement with our previous results for n-

octane, where a relatively temperature-independent value of s=6.8 to 8.4 mJ/m 2 was

obtained.[106] We have furthermore confirmed that the interfacial energies obtained for C8

using the site-wise definition of crystal beads are essentially the same as those reported

previously using Esselink's algorithm. Turnbull et al. [111] calculated the interfacial free energy

for C20 to be around 14 mJ/m 2 , assuming a spherical nucleus model. Hoffmann and Miller [14]

report values of q=1 1.8 mJ/m 2 , based on studies of growth rates of polyethylene lamellae. It is

worth pointing out that interfacial free energy may depend on the size of the nucleus [112],

decreasing with increasing nucleus size. Therefore we infer that the interfacial free energy

obtained by Hoffman and Miller from studies of lamellar growth is an upper bound on the value

relevant to critical nucleation.

Several groups [5, 7, 9] have calculated an interfacial free energy value, (as2 e)1 /3 , for n-alkanes

from nucleation experiments. n-Octane (C8) was found to have a higher interfacial free energy

value than n-eicosane (C20). By contrast, our simulations show the opposite. We attribute this

discrepancy to a shortcoming of the C8 simulations, namely, that the PYS force field predicts a

rotator phase for C8, whereas no such rotator phase is observed for C8 experimentally. Most

likely, the presence of the rotator phase in C8 in simulations leads to an underestimation of the

interfacial free energy in that case. As mentioned previously, C20 does exhibit a rotator phase

experimentally, and this rotator phase is believed to play an important role as an intermediate

90

Page 91: Molecular Simulation Study of Homogeneous Crystal Nucleation

phase during crystallization. Thus, the results presented here are thought to be more reliable than

those for C8.

TABLE 4.3 Crystal-liquid interfacial free energy of n-eicosane molecules.

aAGv was calculated using Eq. (2) and was fixed during the fitting procedure.

In contrast to the interfacial free energies, the bulk free energy difference AGv is sensitive to

temperature, but not as strongly as suggested by Eq.(2.8). The weaker temperature dependence

91

Two-parameter cylindrical model fit Three-parameter cylindrical model fitAT/Tm

T (K) oe as AGv a oS AG,(%) ae (mJ/m 2)

(mJ/m 2) (mJ/m 2) (101 mJ/m3) (mJ/m 2) (1 01 mJ/m3)

5.08 15.50 8.94250 19.4 3.08 3.86 ±0.15 1.71±0.03

±0.49 ±0.69 ±0.26

5.60 12.56 11.42265 14.5 2.43 4.30 ±0.11 1.98±0.02

±0.16 ±0.23 ±0.14

2.64 13.54 8.78280 9.7 1.62 3.17 ±0.14 1.08±0.02

±0.22 ±0.31 ±0.31

Page 92: Molecular Simulation Study of Homogeneous Crystal Nucleation

of AG, is consistent with decreases in the differences in enthalpy and entropy between the

supercooled melt and the crystal, as discussed by Hoffman[ 113].

In closing this discussion, we observe that in any nucleation study, there is inevitably some

arbitrariness in the definition of a crystal nucleus. This is because the interface is not infinitely

sharp, and the interior of the nucleus is not necessarily the same as in the bulk crystal. Some

quantities will be affected by this definition, for example, AGv which measures the chemical

potential difference between the beads in the crystal nucleus and those in the melt. However, the

critical free energy AG* is entirely thermodynamic in origin and does not depend on the

particular definition of a crystal nucleus; this is confirmed indirectly by the results of Fig. 4.5,

i.e., the induction time r* is determined primarily by AG*. Therefore according to Eq.(3.2), the

quantity aas /(AGy) 2 should also be more or less independent of the definition of a crystal

nucleus, and n * is then inversely proportional to AGy, according to Eq.(3.6). Thus, it is only the

mapping of the free energy surface onto a particular order parameter, in our case the nucleus size

n, that depends on how the crystal nucleus is defined; by adopting more restrictive definitions of

what constitutes a crystal bead, we obtain smaller values for the critical nucleus size n*, but at the

same time larger values of AG. Since o 2 o /(AGy) 2 is invariant under different definitions, we

can conclude that as o also increases with a more restrictive definition of what constitutes a

crystal bead. We infer that the quantities a2 o and AGv become unambiguous only in the

thermodynamic limit of large crystal size, where variations in the definition of crystal beads are

unlikely to have a noticeable effect on the determination of the crystal size itself. For crystals of

finite (molecular) dimensions, we conclude only that a o is proportional to (AG)2 , regardless

of nucleus definition.

92

Page 93: Molecular Simulation Study of Homogeneous Crystal Nucleation

4.3 Conclusion

We have studied the nucleation of crystals of n-eicosane from the melt by molecular simulation.

In order to keep track of the attachment and detachment of individual UA beads to the crystal

surface, we developed a crystal nucleus definition based on local ordering around each bead.

The size and shape of the crystal nuclei are well-described by the cylindrical nucleus model. We

obtained the crystal-melt interfacial free energies, o and oe, for the side surface and end surface,

respectively, by fitting this model to simulation results for the free energy of formation of a

crystal nucleus. Very limited temperature dependence was observed for the interfacial free

energies, which suggests that a significant contribution to the critical free energy comes from the

change in enthalpy between the crystal and the melt. For the temperature range we investigated,

or is about 2.5 times greater than a; as a consequence, lateral growth occurs more readily than

thickening growth, resulting in bundle-like nuclei.

The end surface free energy e for the bundle-like nucleus in C20 is found here to be around 4

mJ/m2 . This is much smaller than the value of 280 mJ/m 2 previously estimated by Zachmann

[114] and used to argue against the existence of bundle-like nuclei relative to chain-folded nuclei.

We believe the value determined here is more credible, and supports the contention, also

suggested by Kraack et.al., [115], that bundle-like nuclei are more prevalent than previously

thought, even for long chain polyethylene.

Finally, we have identified an ensemble of critical states by calculating the induction time, zr, the

critical size, n*, and the Zeldovich factor Z for each of 56 nucleation trajectories. The induction

time r", and therefore the critical free energy AG*, is an intrinsic thermodynamic property of the93

Page 94: Molecular Simulation Study of Homogeneous Crystal Nucleation

nucleation process that is insensitive to the details of nucleus definition. Since Z is related to the

second derivative of free energy with respect to nucleus size n, the narrow distribution of Z (Fig.

4.4c) suggests that our proposed nucleus definition serves well to distinguish crystal sites from

melt sites.

94

Page 95: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 5 Homogeneous crystal nucleation in entangled C150 melts

One of the most remarkable features of flexible polymers is the chain-folded lamellae structure

in crystallization. [116] There are still many open questions about this phenomenon [117].

Existing crystal growth rate theories such as Lauritzen-Hoffman (L-H) theory [14] and Sadler's

theory [118], with defects as reviewed by [83], focus on explaining the temperature dependence

of the thickness of the lamellae and developing the crystal growth rate equation, yet how the

chain folding starts was not addressed. In the mean time, it is still very difficult for experiments

to provide the required resolution to understand the microscopic mechanism of chain folding.

This is the point where molecular simulation may give complementary information.

The simulation of chain folding on a predefined crystal front was performed by Monte Carlo on a

lattice [119-122] as well as by molecular dynamics [123]. Simulation of chain folding without

existing crystal front was reported by Liu and Muthukumar [124], Mayer and Muller-Plathe[24],

and Yamamoto[125]. Liu and Muthukumar [124] observed PE chain folding in the vacuum

using Langevin dynamics simulation. Yamamoto [125] studied the homogeneous nucleation

from Cioo melts using a UA model where the trans-gauge barrier was approximated by a bending

potential. This is a model that is similar to what was adopted by Mayer et.al. [24]. The rigidity

of their models is believed to accelerate the crystallization [25]. As a comparison, Yamamoto's

95

Page 96: Molecular Simulation Study of Homogeneous Crystal Nucleation

model [125] produced a radius of gyration Rg2 about 50% higher than that of a flexible chain

model [86].

As a continuation of the previous studies on C8 and C20, we will investigate the homogeneous

crystal nucleation from an under-cooled melt for a long n-alkane and study how the chain folding

affects the crystal nucleation process. As reported by Ungar et.al [3], the chain folding of n-

alkanes sets in for chains longer than C150. We therefore chose to study C150, which we

believe to be long enough to represent polyethylene chains while at the same time still simple

enough for computation simulation.

5.1 Method

5.1.1 Force field

The force field is the same as Chapter 3 and 4. Based on our previous experience with C8 and

C20, we believed that this force field reproduces the experimental melting point, which is about

400K for C150.[126]

5.1.2 Molecular dynamics simulation

MD simulations were performed using the same method as Chapter 4. The configurations were

dumped every 5x10 4 time steps, or 250 ps. For the equilibration runs, a cubic box is applied; for

crystallization runs, three dimensions were allowed to change independently. Periodic boundary

condition in all three directions is applied at all time.

96

Page 97: Molecular Simulation Study of Homogeneous Crystal Nucleation

5.1.3 Initial configurations

Previously the initial configurations for C8 and C20 melt were obtained by melting a crystal.

However, C150 chains are very long chain and relax much slower than C8 or C20. In order to

eliminate any possible thermal memory of crystal phase, the amorphous melt was created at a

low density 0.3g/cm 3 using the algorithm proposed by Theodorou and Suter, when they studied

the glassy state of polymer.[127] The initial configurations then were equilibrated using MD

simulations at 550K for subsequent simulations.

5.1.4 System size determination

Finite size could result in pseudo-folding due to the limited number of chains available in the

system. In an extreme case, any crystal nucleus formed in a simulation system containing only

one chain is guaranteed to contain chain folding. We evaluate this effect by measuring the

system size dependence of the average number of neighboring chains for each united atom. The

neighborhood is determined by a distance cutoff Rb, which corresponds to the length scale of

critical nuclei. Based on our previous studies on n-octane and n-eicosane, the critical nucleus

contains around 200 united atoms, corresponding to a length scale of about 1.0 nm, or 2.5c,

where a=0.401 nm is the Lennard-Jones interaction parameter in our forcefield. Yu et.al. [128]

have also used similar measurement to identify the degree of entanglement by calculating the

ratio between the number of neighboring beads from other chain and total number of neighboring

beads within a neighborhood defined by 0.5nm, noting that they used the Dreiding II force field,

which is a rigid chain model.97

Page 98: Molecular Simulation Study of Homogeneous Crystal Nucleation

We perform simulations with 10, 20, 30, 40, 60, and 80 chains at three temperatures 280K, 400K

and 550K. The system was equilibrated for 10 ns and statistics were collected for 20 ns. We

found that the radius of gyration and the number of neighboring chains approach asymptote

when the system size is greater than 60 chains (Fig. 5.1). Therefore we choose our system size

to be 60 chains. We have measured Rg2 to be about 410 A2 at 550K and about 440 A2 at 280K.

We have also measured the squared end-to-end distance Ro2 to be about 2720 A2 at 550K and

about the same at 280K. This result is consistent with Harmandaris et al. [86] When counting

the number of neighboring chains, the periodic image chain should be counted as a different

chain. However, since the radius of gyration is about 5.5 a and the box dimension is about 15 ,

the chance of seeing one chain and its image at the same time within the range of 2.5 a is small,

therefore we did not take special care of this effect.

18 -(a) -(b) 280KLO3 +o 400K

C 16 - -31 1 550K

30

14 - 29A 286

= 12 - -o26 --

0 21 10 - 280K 5

+ 400KE -a-550K 23M 8 22

Z D 21 I0 10 20 30 40 50 60 70 80 90 0 10 20 30 40 50 60 70 80 90

Number of chains in the system Number of chains in the system

FIG 5.1 (a) The number of neighboring chains with in the neighborhood of 2.50; and (b)

the system size dependence of radius of gyration (in units of a).

98

Page 99: Molecular Simulation Study of Homogeneous Crystal Nucleation

5.1.5 Equilibration

The amorphous melt generated previously was equilibrated 550K. The autocorrelation function

of the end-to-end vector was used to measure the memory of the initial configuration. Fitting the

autocorrelation function to an exponential decay, the time constant for 550K is about 5 ns (Fig.

5.2 and Table 5.1). After the initial 10 ns, configurations were extracted every 10 ns as the

independent input for the nucleation simulation.

L..0

C(D

C(D

0C0

:P

0~0100-1

0 20 40 60 80 100 120 140 160

Time (ns)

FIG 5.2 Autocorrelation functions of end-to-end vector in C150 melt at different

temperatures.

99

Page 100: Molecular Simulation Study of Homogeneous Crystal Nucleation

TABLE 5.1 Autocorrelation time of end-to-end vector in C150 melt.

T (K) Autocorrelation time of end-to-end vector TR (ns)

280 435.50

300 234.45

320 176.58

340 83.77

360 52.20

380 33.77

550 4.63

5.2 Results and discussion

5.2.1 MD simulation of the nucleation process

As we discussed in Chapter 2, the nucleation rate reaches maximum at a temperature between the

melting point and the glassy transition temperature Tg. The melting point Tm for C150 is about

400K. The glass transition temperature Tg can be determined by monitoring the specific volume

as a function of temperature. The value of Tg for polyethylene (PE)spreads a broad range, based

on a number of factors, the specific PE component, the thermal history, the cooling rate, and the

semicrystalline nature of PE, etc. The experiments reported Tg around 240K[129]. Computer

100

Page 101: Molecular Simulation Study of Homogeneous Crystal Nucleation

simulations have also been carried out to estimate Tg for n-alkanes of a few hundreds to one

thousand of methylene groups. Han et al.[130] reported Tg between 200K and 250K, Takeuchi

and Roe [131] reported 200K, Capaldi et al.[132] reported 280K, which was likely an

overestimation due to a very high cooling rate, as illustrated by Figure 1 in Han et al. [130]

Therefore a reasonable estimate of Tg for C150 is between 200K and 250K. We chose to

perform the nucleation study at several temperatures between 280K and 340K, corresponding to

a degree of undercooling from 30% to 15%. The crystal nucleus was defined the same way as

we did previously in n-eicosane [133]. Fig.5.3 shows a typical nucleation trajectory. The

reaction coordinate nma shows two short-lived fluctuations at around 50 ns and 150 ns,

respectively, before a successful nucleation event occurs at around 220 ns. The other parameters,

P2 , potential energy and volume all reflect the nucleation event but are less sensitive to nma. The

elongation of the box was due to the later growth of the largest nucleus and the elongation

direction is the same as the alignment of the chains in that nucleus. Fig. 5.4 is a snapshot

illustrating the crystal nuclei, a chain that bridges two nuclei and a chain that contains a fold.

101

Page 102: Molecular Simulation Study of Homogeneous Crystal Nucleation

Tc = 300K Tc = 300K900 0.020 400.0 4.5

800 . Potential energy per chain - 4.4

700 0.015 200.0 - Volume per chain 43 <

600 - 0.010 0.0 4.3

500 42

0.005 -600.0(D 400 - 1 41

0.000 -800.03

200 -- 4000000 20000.0

10 -3.8 ..

0 -0.010 -800.0 3.80 50 100 150 200 250 0 50 100 150 200 250a.

Time (ns) Time (ns)

FIG 5.3 Evolution of characteristic variables for a system of 60 C150 chains during a

typical MD simulation, after quenching from 550K to 300K at t=0. (a) Size of the largest

nucleus, nnm,,, and the global orientation order parameter P2. (b) Potential energy and

volume per chain.

102

Page 103: Molecular Simulation Study of Homogeneous Crystal Nucleation

(b)

(c)

0ji0

(d)

(e)

FIG 5.4 A snapshot of the crystal nuclei in the C150 melt at 280K. (a) All the chains

participating in the largest nucleus, which is of size 91. (b) The largest nucleus (blue),

the 2nd largest nucleus (green), one chain to fold into the largest nucleus (cyan), and one

chain to bridge two nuclei (red). (c) Side view of the largest nucleus. (d) Top view of the

largest nucleus. (e) One chain participating the largest nucleus, crystal segment(blue),

amorphous segment (yellow).

103

(a)

Page 104: Molecular Simulation Study of Homogeneous Crystal Nucleation

(a)

250

150

LL100

50

00 100 200 300

nmax

(b)

250

200

00

50

00 100 200 300

nmax

400 500 600

400 500 600

FIG 5.5 (a) Symbols: Mean-first-passage time (MFPT) for four different temperatures;

solid lines: the linear growth rate fitting. (b) Dash lines: MFPT fitting using Wedekind's

method for nucleation; solid lines: the overall fitting including both growth and

nucleation.

104

Page 105: Molecular Simulation Study of Homogeneous Crystal Nucleation

The MFPT curve as a function of nma was shown in Fig. 5.5. When the under-cooling, AT,

increases, the nucleation rate also increases due to the lower free energy barrier, and the growth

rate decreases due to the slower diffusivity. Therefore at low temperatures, the growth rate and

nucleation rate become comparable and both contribute to the MFPT curve.

We used the MFPT data to measure the growth rate assuming planar growth on a base of 4

parallel segments each of 9 beads long (9 beads is the average length of crystal segment and the

lateral dimension of nuclei of size 400 is about 3 chain spacing in crystal phase.) The growth

rate measurements are summarized in Table 5.2. The growth rate was estimated to be in the

order of 0.1 nm/ns. As a comparison, Waheed et al.[83], using the same force field, measured a

growth rate around 0.0001 nm/ns for C100 at 350K, which is about two orders of magnitude

lower that our results. The reason for this is that our measured measurement was for small nuclei,

where faceting has not happened and the growth was faster than the planar growth.

TABLE 5.2 Linear growth rate estimated using MFPT curve.

T (K) AT/Tm Slope (ns/bead) Bead/ns Growth rate Intercept (ns)

(nm/ns)

280 30% 0.145 0.001 6.9 0.081 160.30 ±0.47

300 25% 0.094 ± 0.002 10.6 0.124 203.55 ± 0.82

320 20% 0.071 0.001 14.1 0.165 214.76 ±0.29

340 15% 0.0176 ± 0.0002 56.8 0.663 233.51 ± 0.09

105

Page 106: Molecular Simulation Study of Homogeneous Crystal Nucleation

TABLE 5.3. Temperature dependence of the induction time r* and the critical nucleus size n*

(approximated by nmax) obtained using MFPT method.

Tc (K) AT/Tm No. of samples Z r* (ns) n

280 30% 32 0.0115 0.0001 160.55 ±0.16 96.5 0.3

300 25% 24 0.0128 ± 0.0002 202.99 ± 0.38 96.9 ± 0.6

320 20% 24 0.0186 ±0.0002 213.25 0.22 71.5 ±0.3

340 15% 16 0.0306 ± 0.0003 232.60 ± 0.21 48.5 ± 0.2

After correcting the contribution of the linear growth, the MFPT curve was fit to extract the

induction time r* and n' (Table 5.3). Contrary to the classical nucleation theory, the critical

nucleus size n* decreases with decreasing under-cooling. We have also measured the steady

state nucleus size distribution Pst(n). Combining Pst(n) and MFPT results, we used Eq. (2.21) to

constructed the free energy AG(n), shown in Fig.5.6.

106

Page 107: Molecular Simulation Study of Homogeneous Crystal Nucleation

8

|280K|

4

(k

-2

0 50 100 150 200 250 300

Nuclei size n

FIG 5.6 AG(n) reconstructed by using Eq.(2.21) for four different temperatures.

One possible explanation for this abnormal temperature dependence of n* is that, similar to

previous Ising model simulation (Table 2.1), the nma* estimated from MFPT method is not

consistent with the free energy sampling. Another possible explanation is by the diffusion free

energy barrier Ed introduced by Eq.(2.12). Every attachment of particle to the surface of crystal

nucleus involves a barrier crossing, as pointed out by Turnbull and Fisher[134]. Therefore, the

steady state nucleus size distribution Pst(n) must also reflect the influence of Ed. Ed might not be

important for simple molecules. However, polymer melt is highly viscous and the contribution

of the diffusion barrier is not negligible.

AG(n) reconstructed from Pst(n) includes an accumulation of Ed(1,2), Ed(2,3), ... , Ed(n-l,n),

where Ed(i,i+l) is the diffusion barrier for adding one particle to a nucleus of size i. Since the

exact form of Ed(i,i+l) is unknown, we assume that the cumulative effect of Ed's for a nucleus of

107

Page 108: Molecular Simulation Study of Homogeneous Crystal Nucleation

size n is represented by AGA(n), which has a simple form k(T)n2 /3. Therefore, we obtained the

true thermodynamics free energy curve AG 0(n) by subtracting k(T)n2 /3 from AG(n), which is

calculated from Pst(n). By tuning amplitude of k(T), the CNT picture, that is, n' and AG* both

increase as AT decreases, was recovered, as shown in Fig. 5.7. The temperature dependence of

k(T) must be related to the viscosity of the melt and the relaxation time of the chain molecules in

the system.

6

4

2

0

-20 50 100 150

Nucleus size n

FIG 5.7 The thermodynamics free energy barrier AG4(n), after subtracting the diffusion

barrier AGd(n) from AG(n), AG0(n)= AG(n)-AGd(n). The pre-factor of AGd(n) is chosen to

be k(T)=0.33, 0.27, 0.18, and 0 for 280K, 300K, 320K and 340K, respectively.

108

Page 109: Molecular Simulation Study of Homogeneous Crystal Nucleation

We have also examined the solid-liquid interface using a segment analysis. There are four types

of segments: fold (loop), tie (bridge), tail and xseg. A fold (loop) is a segment in the amorphous

region with two ends attached to the same crystal nucleus. A tie (bridge) is a segment in the

amorphous region with two ends attached to two different crystal nuclei. A tail is a segment in

the amorphous region that has only one end attached to a crystal nucleus. An xseg is a segment

in the crystal phase with two ends on the solid-liquid interface. Given any crystal nucleus, there

is a simple relation between the numbers of these four types of segments:

2 nfold+ n,ie+ naii = 2 nxseg (5.1)

Shown in Fig.5.8 is a snapshot of a simulation box at the later stage of crystallization. We

performed the segment analysis to the bigger nucleus and found that it has 60 folds, 31 ties, 30

tails and 105 xsegs. Therefore the ratio between folds and ties on the solid-liquid interface of

this particular nucleus is about 2:1.

109

Page 110: Molecular Simulation Study of Homogeneous Crystal Nucleation

FIG 5.8 A snapshot of a simulation box with one period image on each side. (Blue) the

bigger crystal nucleus, (Red) the smaller crystal nucleus, (Cyan) chain segments in the

amorphous region.

110

Page 111: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 6 Conclusion and recommendations for future work

Using molecular simulation method, the homogeneous crystal nucleation in the melt

for three n-alkanes, n-octane (C8), n-eicosane (C20) and C150. We have found that

united atom force field was capable of reproducing the physical quantities

crystallization, including the equilibrium melting temperature, heat of fusion and

interfacial free energies.

was studied

the realistic

related to

solid-liquid

Both the dynamics method (MFPT) and the thermodynamic method (free energy analysis) were

applied to interpret the simulation data. In order to properly apply these methods, we have

examined the classical nucleation theory using a 3D Ising model simulation. By studying the

nucleation free energy we proved that the size of the largest nucleus in the system, nmax, is the

controlling reaction coordinate in the nucleation process.

We have found that the critical nucleus has a cylindrical shape and contains around 100 united

atoms. The crystal-liquid interfacial free energies for C8 and C20 were estimated to be o ~ 4

mJ/m2 for the end surface and o - 10 mJ/m 2 for the side surface. Since the size and shape of the

critical nucleus in C150 system are similar to C8 and C20, the interfacial free energies should

also be close. Our measured o is much smaller than the estimations for bundle-like interface

(-280 mJ/m2) and chain-folding interface (-170 mJ/m2)[114]; however, considering that the111

Page 112: Molecular Simulation Study of Homogeneous Crystal Nucleation

interfacial free energy is dependent on the nucleus size, we expect oe to grow as the lateral

dimension of the crystal increases.

There were very few experimental measurements of interfacial free energy for PE. Cormia et al.

[47] studied the crystal nucleation of PE, and they found the critical nucleus of rod shape with

length of 19.2 nm and radius of 1.1 nm at 56K under-cooling. Such a big n* corresponds to very

high interfacial free energies (q, = 9.6 mJ/m2, o =168 mJ/m2). It is very hard to imagine such a

high oa for a crystal nucleus that is only 1.1 nm wide in cross section. Their oC value is close to

the values quoted by Kraack et al. [115] The oe values quoted there were not from homogeneous

nucleation experiment, but from measuring the melting point of existing crystal lamellae. Again,

the dimension matters, and the comparison of ae from different measurements require careful

examination of the details of the measurements.

We have observed chain folding during the crystal nucleation in the C150 melt. The

configuration of the folds supports the random switchboard reentry model over the adjacent

reentry model. The fact that chain folding starts at a fairly early stage suggests that chain folding

is probably more a result of the configuration of chain segments in the under-cooled melt rather

than being driven by the extra strain due to the accumulation of bundle-like free tails, the

argument that was used against the fringe micelle model, and we found both chain folds and

bundle-like tails and ties on the solid-liquid interface. It will be very interesting to see if the

configuration of one chain changes significantly before and after it enters a crystallite.

The chain length dependence of nucleation process was studied. For all three chain lengths, at

around 20% of under-cooling, AT/Tm, the critical nucleus contains around 70-160 united atoms,

112

Page 113: Molecular Simulation Study of Homogeneous Crystal Nucleation

and less than 1 nm in diameter. The small n* suggests that homogeneous nucleation is a strictly

local event. Within the range of chain length in our study, we have found that at the same degree

of under-cooling, n* slightly decreases as the chain length increase. Since the heat of fusion per

unit volume is relatively constant for different chain length, the interfacial free energies are

therefore smaller for C150 than for C20. It is particularly meaningful to compare C20 and C150.

C20 has bundle-like nucleus, i.e., chains contribute to the crystal do not fold; C150 has nucleus

with both chain fold and bundle-like free tails. It was suggested [114] that chain fold is more

energetically preferable than dangling free ends in the amorphous region, i.e., the bundle-like

structure, and our observation is consistent with this argument.

We have also studied the temperature dependence of the nucleation process. We have found that

in C8 and C20, the critical nucleus size n* and critical free energy AG* both decrease as the

under-cooling increases, as predicted by classical nucleation theory. However, in C150 melt, the

critical nucleus size n* increases as the under-cooling increases. This abnormal behavior is one

of the remaining questions from this study. It could potentially be resolved in two directions: (1)

the deviation of the dynamics approach from the thermodynamics approach; and (2) the high

viscosity of the melts for long n-alkanes, due to a diffusion free energy barrier.

Another remaining question is the trend of slight decrease of n* as the chain length increases.

We have argued above that it might be associated with the onset of chain folding. Therefore one

practical proposal is to perform a simulation on a system containing only one very long chain,

e.g., a C9000 chain. In this way the total number of united atom is the same as the C150 system

we studied, at the same time there could be no bundle-like free tails.

113

Page 114: Molecular Simulation Study of Homogeneous Crystal Nucleation

One more interesting project for the future would be to study the interface between the crystal

and amorphous regions, particularly the surface that crystal segments intersect. The constituent

of the interface, i.e., folds, ties, tails can significantly change the physical properties of the

semicrystalline product. The temperature and the concentration of chain molecules will

determine the relative probability of seeing these different types of segments. We could measure

the interfacial free energies directly [60, 135, 136] since so far there is no such measurement

available from experiments.

In our study we applied a step quench, corresponding to a delta function cooling rate. However,

the cooling rate is almost always finite in real experiment. It has been pointed out in Chapter 4

that the cooling rate can significantly change the kinetic pre-factor, I0, of the nucleation rate. It is

worth to study how the cooling rate affects the nucleation process.

Polymer crystallization is a complicated process and there are still many aspects of this field

need to be understood, as pointed out in the beginning of Chapter 2. We believe that our work

has laid a foundation toward a complete understanding of the polymer crystallization on a

microscopic level.

114

Page 115: Molecular Simulation Study of Homogeneous Crystal Nucleation

Chapter 7 Bibliography

1. Li, C.Y. and S.Z.D. Cheng, Semicrystalline Polymers, in Encyclopedia of Polymer Science andTechnology. 2002, John Wiley & Sons, Inc.

2. Teckoe, J. and D.C. Bassett, The crystallization kinetics of monodisperse C98H198 from the melt.Polymer, 2000. 41(5): p. 1953-1957.

3. Ungar, G., et al., The Crystallization of Ultralong Normal Paraffins: The Onset of Chain Folding.Science, 1985. 229(4711): p. 386-389.

4. Hoffman, J.D., Growth rate of extended-chain crystals: the lateral surface free energy of pure n-C94H190 and a fraction .apprx.C207H416. Macromolecules, 1985. 18(4): p. 772-786.

5. Uhlmann, D.R., et al., Crystal nucleation in normal alkane liquids. The Journal of ChemicalPhysics, 1975. 62(12): p. 4896-4903.

6. Turnbull, D. and R.L. Cormia, Kinetics of Crystal Nucleation in Some Normal Alkane Liquids. TheJournal of Chemical Physics, 1961. 34(3): p. 820-831.

7. Oliver, M.J. and P.D. Calvert, Homogeneous nucleation of n-alkanes measured by differentialscanning calorimetry. Journal of Crystal Growth, 1975. 30(3): p. 343-351.

8. Sirota, E.B. and A.B. Herhold, Transient Phase-induced Nucleation. Science, 1999. 283(5401): p.529-532.

9. Kraack, H., E.B. Sirota, and M. Deutsch, Measurements of homogeneous nucleation in normal-alkanes. The Journal of Chemical Physics, 2000. 112(15): p. 6873-6885.

10. Muthukumar, M. and P. Welch, Modeling polymer crystallization from solutions. Polymer, 2000.41(25): p. 8833-8837.

11. Strobl, G., From the melt via mesomorphic and granular crystalline layers to lamellar crystallites:A major route followed in polymer crystallization? The European Physical Journal E: Soft Matterand Biological Physics, 2000. 3(2): p. 165-183.

12. Olmsted, P.D., et al., Spinodal-Assisted Crystallization in Polymer Melts. Physical Review Letters,1998. 81(2): p. 373.

13. Allegra, G. and S.V. Meille, The bundle theory for polymer crystallization. Physical ChemistryChemical Physics, 1999. 1(22): p. 5179-5188.

14. Hoffman, J.D. and R.L. Miller, Kinetic of crystallization from the melt and chain folding inpolyethylene fractions revisited: theory and experiment. Polymer, 1997. 38(13): p. 3151-3212.

15. Muthukumar, M., Molecular Modelling of Nucleation in Polymers. Philosophical Transactions:Mathematical, Physical and Engineering Sciences, 2003. 361(1804): p. 539-556.

16. Strobl, G., Crystallization and melting of bulk polymers: New observations, conclusions and a

thermodynamic scheme. Progress in Polymer Science, 2006. 31(4): p. 398-442.17. Hu, W., Intramolecular Crystal Nucleation, in Progress in Understanding of Polymer

Crystallization, G. Reiter and G. Strobl, Editors. 2007, Springer Berlin / Heidelberg. p. 47-63.18. Waheed, N., M. Ko, and G. Rutledge, Atomistic Simulation of Polymer Melt Crystallization by

Molecular Dynamics, in Progress in Understanding of Polymer Crystallization, G. Reiter and G.Strobl, Editors. 2007, Springer Berlin / Heidelberg. p. 457-480.

115

Page 116: Molecular Simulation Study of Homogeneous Crystal Nucleation

19. Muthukumar, M., Shifting Paradigms in Polymer Crystallization, in Progress in Understanding ofPolymer Crystallization, G. Reiter and G. Strobl, Editors. 2007, Springer Berlin / Heidelberg. p. 1-18.

20. Yamamoto, T., Molecular Dynamics Modeling of the Crystal-Melt Interfaces and the Growth ofChain Folded Lamellae, in Interphases and Mesophases in Polymer Crystallization Ill, G. Allegra,Editor. 2005, Springer Berlin / Heidelberg. p. 37-85.

21. Esselink, K., P.A.J. Hilbers, and B.W.H. van Beest, Molecular dynamics study of nucleation andmelting of n-alkanes. The Journal of Chemical Physics, 1994. 101(10): p. 9033-9041.

22. Takeuchi, H., Structure formation during the crystallization induction period of a short chain-molecule system: A molecular dynamics study. The Journal of Chemical Physics, 1998. 109(13): p.5614-5621.

23. Fujiwara, S. and T. Sato, Molecular Dynamics Simulation of Structural Formation of ShortPolymer Chains. Physical Review Letters, 1998. 80(5): p. 991.

24. Meyer, H. and F. Muller-Plathe, Formation of Chain-Folded Structures in Supercooled PolymerMelts Examined by MD Simulations. Macromolecules, 2002. 35(4): p. 1241-1252.

25. Miura, T., et al., Effect of rigidity on the crystallization processes of short polymer melts. PhysicalReview E, 2001. 63(6): p. 061807.

26. Nucleation, ed. A.C. Zettlemoyer. 1969, New York: Dekker.27. Skripov, V.P., Metastable Liquids. 1974, New York: Wiley.28. Oxtoby, D.W., Homogeneous nucleation: theory and experiment. Journal of Physics: Condensed

Matter, 1992. 4(38): p. 7627-7650.29. Debenedetti, P.G., Metastable liquids : concepts and principles 1996, Princeton, N.J. : Princeton

University Press.30. Kashchiev, D., Nucleation: Basic Theory with Applications. 2000, Oxford: Butterworth-

Heinemann.31. Kelton, K.F., Crystal Nucleation in Liquids and Glasses. Solid State Physics, ed. H. Ehrenreich and

D. Turnbull. Vol. 45. 1991, Boston: Academic.32. Gibbs, J.W., The Collected works of J. Willard Gibbs. Vol. 1. 1928, New York: Longmans-Green.33. Volmer, V. and A. Weber, Nucleus formation in supersaturated systems. Z. Phys. Chem., 1926.

119: p. 277.34. Becker, R. and W. Doring, Ann. Phys., 1935. 24: p. 719.35. Zeldovich, J.B., Acta Physicochim. USSR, 1943. 18: p. 1.36. Cahn, J.W. and J.E. Hilliard, Free Energy of a Nonuniform System. Ill. Nucleation in a Two-

Component Incompressible Fluid. Journal of Chemical Physics, 1959. 31(3): p. 688.37. Abraham, F.F., On the thermodynamics, structure and phase stability of the nonuniform fluid

state. Physics Reports, 1979. 53(2): p. 93-156.38. Oxtoby, D.W. and R. Evans, Nonclassical nucleation theory for the gas-liquid transition. J. Chem.

Phys., 1988. 89(12): p. 7521.39. Oxtoby, D.W., Nucleation of First-Order Phase Transitions. Accounts of Chemical Research, 1998.

31(2): p. 91-97.40. Oxtoby, D.W. and D. Kashchiev, A General Relation between the Nucleation Work and the Size of

the Nucleus in Multicomponent Nucleation. Journal of Chemical Physics, 1994. 100(10): p. 7665-7671.

41. Harrowell, P. and D.W. Oxtoby, A molecular theory of crystal nucleation from the melt. TheJournal of Chemical Physics, 1984. 80(4): p. 1639-1646.

42. Shen, Y.C. and D.W. Oxtoby, Nucleation of Lennard - Jones fluids: A density functional approachJ. Chem. Phys., 1996. 105(15): p. 6517.

116

Page 117: Molecular Simulation Study of Homogeneous Crystal Nucleation

43. Vonnegut, B., Variation with temperature of the nucleation rate of supercooled liquid tin andwater drops. Journal of Colloid Science, 1948. 3(6): p. 563-569.

44. Turnbull, D., Formation of Crystal Nuclei in Liquid Metals. Journal of APplied Physics, 1950.21(10): p. 1022-1028.

45. Carvalho, J. and K. Dalnoki-Veress, Surface nucleation in the crystallisation of polyethylenedroplets. The European Physical Journal E: Soft Matter and Biological Physics, 2011. 34(1): p. 1-6.

46. Jones, D.R.H., The free energies of solid-liquid interfaces. Journal of Materials Science, 1974. 9(1):p. 1-17.

47. Cormia, R.L., F.P. Price, and D. Turnbull, Kinetics of Crystal Nucleation in Polyethylene. TheJournal of Chemical Physics, 1962. 37(6): p. 1333-1340.

48. Stauffer, D., A. Coniglio, and D.W. Heermann, Monte Carlo Experiment for Nucleation Rate in theThree-Dimensional Ising Model. Physical Review Letters, 1982. 49(18): p. 1299.

49. Mandell, M., J. McTague, and A. Rahman, Crystal nucleation in a three?dimensionalLennard?Jones system: A molecular dynamics study. The Journal of Chemical Physics, 1976. 64(9):p. 3699.

50. van Duijneveldt, J.S. and D. Frenkel, Computer simulation study of free energy barriers in crystalnucleation. The Journal of Chemical Physics, 1992. 96(6): p. 4655-4668.

51. Auer, S. and D. Frenkel, Numerical prediction of absolute crystallization rates in hard-spherecolloids. The Journal of Chemical Physics, 2004. 120(6): p. 3015-3029.

52. Binder, K. and H. MUller-Krumbhaar, Investigation of metastable states and nucleation in thekinetic Ising model. Physical Review B, 1974. 9(5): p. 2328.

53. Bortz, A.B., M.H. Kalos, and J.L. Lebowitz, A new algorithm for Monte Carlo simulation of Isingspin systems. Journal of Computational Physics, 1975. 17(1): p. 10-18.

54. Fichthorn, K.A. and W.H. Weinberg, Theoretical foundations of dynamical Monte Carlosimulations The Journal of Chemical Physics, 1991. 95(2): p. 1090.

55. Gillespie, D.T., A general method for numerically simulating the stochastic time evolution ofcoupled chemical reactions. Journal of Computational Physics, 1976. 22(4): p. 403-434.

56. Honeycutt, J.D. and H.C. Andersen, Small system size artifacts in the molecular dynamicssimulation of homogeneous crystal nucleation in supercooled atomic liquids. The Journal ofPhysical Chemistry, 1986. 90(8): p. 1585-1589.

57. Swope, W.C. and H.C. Andersen, 106-particle molecular-dynamics study of homogeneousnucleation of crystals in a supercooled atomic liquid. Physical Review B, 1990. 41(10): p. 7042.

58. Angioletti-Uberti, S., et al., Solid-liquid interface free energy through metadynamics simulations.Physical Review B, 2010. 81(12): p. 125416.

59. Morris, J.R. and X. Song, The melting lines of model systems calculated from coexistencesimulations. The Journal of Chemical Physics, 2002. 116(21): p. 9352-9358.

60. Hoyt, J.J., M. Asta, and A. Karma, Method for Computing the Anisotropy of the Solid-LiquidInterfacial Free Energy. Physical Review Letters, 2001. 86(24): p. 5530.

61. ter Horst, J.H. and D. Kashchiev, Determination of the nucleus size from the growth probability ofclusters. The Journal of Chemical Physics, 2003. 119(4): p. 2241-2246.

62. Bartell, L.S. and D.T. Wu, A new procedure for analyzing the nucleation kinetics of freezing incomputer simulation. The Journal of Chemical Physics, 2006. 125(19): p. 194503-4.

63. Wedekind, J., R. Strey, and D. Reguera, New method to analyze simulations of activatedprocesses. The Journal of Chemical Physics, 2007. 126(13): p. 134103-7.

64. Alessandro, L. and L.G. Francesco, Metadynamics: a method to simulate rare events andreconstruct the free energy in biophysics, chemistry and material science. Reports on Progress inPhysics, 2008. 71(12): p. 126601.

117

Page 118: Molecular Simulation Study of Homogeneous Crystal Nucleation

65. Kashchiev, D., Interrelation between cluster formation time, cluster growth probability, andnucleation rate. The Journal of Chemical Physics, 2007. 127(6): p. 064505-8.

66. Wedekind, J. and D. Reguera, Kinetic Reconstruction of the Free-Energy Landscape. The Journalof Physical Chemistry B, 2008. 112(35): p. 11060-11063.

67. Wedekind, J., et al., Crossover from nucleation to spinodal decomposition in a condensing vapor.The Journal of Chemical Physics, 2009. 131(11): p. 114506-11.

68. Torrie, G.M. and J.P. Valleau, Monte Carlofree energy estimates using non-Boltzmann sampling:Application to the sub-critical Lennard-Jones fluid. Chemical Physics Letters, 1974. 28(4): p. 578-581.

69. Torrie, G.M. and J.P. Valleau, Nonphysical sampling distributions in Monte Carlo free-energyestimation: Umbrella sampling. Journal of Computational Physics, 1977. 23(2): p. 187-199.

70. Alexander, S. and J. McTague, Should All Crystals Be bcc? Landau Theory of Solidification andCrystal Nucleation. Physical Review Letters, 1978. 41(10): p. 702.

71. Steinhardt, P.J., D.R. Nelson, and M. Ronchetti, Bond-orientational order in liquids and glasses.Physical Review B, 1983. 28(2): p. 784.

72. ten Wolde, P.R., M.J. Ruiz-Montero, and D. Frenkel, Numerical Evidence for bcc Ordering at theSurface of a Criticalfcc Nucleus. Physical Review Letters, 1995. 75(14): p. 2714.

73. Ten Wolde, P.R. and D. Frenkel, Homogeneous nucleation and the Ostwald step rule. PhysicalChemistry Chemical Physics, 1999. 1: p. 2191-2196.

74. ten Wolde, P.R., M.J. Ruiz-Montero, and D. Frenkel, Numerical calculation of the rate of crystalnucleation in a Lennard-Jones system at moderate undercooling. The Journal of Chemical Physics,1996. 104(24): p. 9932-9947.

75. Leyssale, J.M., J. Delhommelle, and C. Millot, A molecular dynamics study of homogeneouscrystal nucleation in liquid nitrogen. Chemical Physics Letters, 2003. 375(5-6): p. 612-618.

76. ten Wolde, P.R., M.J. Ruiz-Montero, and D. Frenkel, Simulation of homogeneous crystalnucleation close to coexistence. Faraday Discussions, 1996. 104: p. 93-110.

77. Leyssale, J.-M., J. Delhommelle, and C. Millot, Atomistic simulation of the homogeneousnucleation and of the growth of N[sub 2] crystallites. The Journal of Chemical Physics, 2005.122(10): p. 104510-10.

78. Trudu, F., D. Donadio, and M. Parrinello, Freezing of a Lennard-Jones Fluid: From Nucleation toSpinodal Regime. Physical Review Letters, 2006. 97(10): p. 105701-4.

79. Pan, A.C. and D. Chandler, Dynamics of Nucleation in the Ising Model. J. Phys. Chem. B, 2004.108(51): p. 19681-19686.

80. Bhimalapuram, P., S. Chakrabarty, and B. Bagchi, Elucidating the Mechanism of Nucleation nearthe Gas-Liquid Spinodal. Physical Review Letters, 2007. 98(20): p. 206104.

81. Lundrigan, S.E.M. and |. Saika-Voivod, Test of classical nucleation theory and mean first-passagetime formalism on crystallization in the Lennard-Jones liquid. The Journal of Chemical Physics,2009. 131(10): p. 104503-7.

82. Paul, W., D.Y. Yoon, and G.D. Smith, An optimized united atom model for simulations ofpolymethylene melts. The Journal of Chemical Physics, 1995. 103(4): p. 1702-1709.

83. Waheed, N., M.J. Ko, and G.C. Rutledge, Molecular simulation of crystal growth in long alkanes.Polymer, 2005. 46(20): p. 8689-8702.

84. Waheed, N., M.S. Lavine, and G.C. Rutledge, Molecular simulation of crystal growth in n-eicosane. The Journal of Chemical Physics, 2002. 116(5): p. 2301-2309.

85. Paul, W., G.D. Smith, and D.Y. Yoon, Static and Dynamic Properties of a n-C100H202 Melt fromMolecular Dynamics Simulations. Macromolecules, 1997. 30(25): p. 7772-7780.

118

Page 119: Molecular Simulation Study of Homogeneous Crystal Nucleation

86. Harmandaris, V.A., V.G. Mavrantzas, and D.N. Theodorou, Atomistic Molecular DynamicsSimulation of Polydisperse Linear Polyethylene Melts. Macromolecules, 1998. 31(22): p. 7934-7943.

87. Mavrantzas, V.G. and D.N. Theodorou, Atomistic Simulation of Polymer Melt Elasticity:Calculation of the Free Energy of an Oriented Polymer Melt. Macromolecules, 1998. 31(18): p.6310-6332.

88. Smith, W. and T.R. Forester, DLPOLY_2.0: A general-purpose parallel molecular dynamicssimulation package. Journal of Molecular Graphics, 1996. 14(3): p. 136-141.

89. Plimpton, S., Fast Parallel Algorithms for Short-Range Molecular Dynamics. Journal ofComputational Physics, 1995. 117(1): p. 1-19.

90. LAMMPS (large-scale atomic/molecular massively parallel simulator). [cited 2009 April 22];Available from: http://lammps.sandia.gov/.

91. Lavine, M.S., N. Waheed, and G.C. Rutledge, Molecular dynamics simulation of orientation andcrystallization of polyethylene during uniaxial extension. Polymer, 2003. 44(5): p. 1771-1779.

92. Pant, P.V.K. and D.N. Theodorou, Variable Connectivity Method for the Atomistic Monte CarloSimulation of Polydisperse Polymer Melts. Macromolecules, 1995. 28(21): p. 7224-7234.

93. Siepmann, J.|., A method for the direct calculation of chemical potentials for dense chain systems.Molecular Physics, 1990. 70(6): p. 1145-1158.

94. Norman, N. and H. Mathisen, CRYSTAL STRUCTURE OF LOWER n-PARAFFINS .1. n-Octane. ActaChemica Scandinavica, 1961. 15(8): p. 1747-1754.

95. Mathisen, H., N. Norman, and B.F. Pedersen, CRYSTAL STRUCTURE OF LOWER PARAFFINS .4.Refinement of the Crystal Structures of Pentane and Octane. Acta Chemica Scandinavica, 1967.21(1): p. 127-135.

96. Ryckaert, J.P., M.L. Klein, and |.R. Mcdonald, Disorder at the Bilayer Interface in thePseudohexagonal Rotator Phase of Solid N-Alkanes. Physical Review Letters, 1987. 58(7): p. 698-701.

97. Ryckaert, J.P., |.R. Mcdonald, and M.L. Klein, Disorder in the Pseudohexagonal Rotator Phase ofNormal-Alkanes - Molecular-Dynamics Calculations for Tricosane. Molecular Physics, 1989. 67(5):p. 957-979.

98. Poison, J.M. and D. Frenkel, Numerical prediction of the melting curve of n-octane. The Journalof Chemical Physics, 1999. 111(4): p. 1501-1510.

99. Toxvaerd, S., Molecular dynamics calculation of the equation of state of alkanes. The Journal ofChemical Physics, 1990. 93(6): p. 4290-4295.

100. Toxvaerd, S., Equation of state of alkanes /I. The Journal of Chemical Physics, 1997. 107(13): p.5197-5204.

101. Humphrey, W., A. Dalke, and K. Schulten, VMD: Visual molecular dynamics. Journal of MolecularGraphics, 1996. 14(1): p. 33-38.

102. Ko, M.J., et al., Characterization of polyethylene crystallization from an oriented melt bymolecular dynamics simulation. The Journal of Chemical Physics, 2004. 121(6): p. 2823-2832.

103. Stack, G.M., L. Mandelkern, and |.G. Voigt-Martin, Crystallization, melting, and morphology of

low molecular weight polyethylene fractions. Macromolecules, 1984. 17(3): p. 321-331.104. Bai, X.-M. and M. Li, Calculation of solid-liquid interfacial free energy: A classical nucleation

theory based approach. The Journal of Chemical Physics, 2006. 124(12): p. 124707-12.105. Martyna, G.J., et al., Explicit reversible integrators for extended systems dynamics. Molecular

Physics: An International Journal at the Interface Between Chemistry and Physics, 1996. 87(5): p.1117 - 1157.

119

Page 120: Molecular Simulation Study of Homogeneous Crystal Nucleation

106. Yi, P. and G.C. Rutledge, Molecular simulation of crystal nucleation in n-octane melts. TheJournal of Chemical Physics, 2009. 131(13): p. 134902.

107. Schaerer, A.A., et al., Properties of Pure Normal Alkanes in the C17 to C36 Range. Journal of theAmerican Chemical Society, 1955. 77(7): p. 2017-2019.

108. van Miltenburg, J.C., H.A.J. Oonk, and V. Metivaud, Heat Capacities and Derived ThermodynamicFunctions of n-Nonadecane and n-Eicosane between 10 K and 390 K. Journal of Chemical &Engineering Data, 1999. 44(4): p. 715-720.

109. Claudy, P. and J.M. Letoffe, Transition de Phase dans les n-Alcanes Pairs CnH2n+2 n=16-28. . 22-emes Journees de Calorimetrie et d'Analyse Thermique (through NIST), 1991. 22: p. 281-290.

110. Baidakov, V.G., et al., Crystal nucleation rate isotherms in Lennard-Jones liquids. The Journal ofChemical Physics, 2010. 132(23): p. 234505-9.

111. David Turnbull, F.S., Crystal nucleation and the crystalmelt interfacial tension in linearhydrocarbons. Journal of Polymer Science: Polymer Symposia, 1978. 63(1): p. 237-243.

112. Tolman, R.C., The Effect of Droplet Size on Surface Tension. The Journal of Chemical Physics,1949. 17(3): p. 333-337.

113. Hoffman, J.D., Thermodynamic Driving Force in Nucleation and Growth Processes. The Journal ofChemical Physics, 1958. 29(5): p. 1192-1193.

114. Zachmann, H.G., The Crystalline State of Macromolecular Substances. 1974, H~thig & WepfVerlag. p. 244-252.

115. Kraack, H., M. Deutsch, and E.B. Sirota, n-Alkane Homogeneous Nucleation: Crossover toPolymer Behavior. Macromolecules, 2000. 33(16): p. 6174-6184.

116. Keller, A., Philosophical Magazine, 1957. 2: p. 1171.117. Progress in Understanding of Polymer Crystallization. Lecture Notes in Physics, ed. G. Reiter and

G.R. Strobl. 2007: Springer Berlin / Heidelberg.118. Sadler, D.M., New explanation for chain folding in polymers. Nature, 1987. 326(6109): p. 174-

177.119. Sadler, D.M. and G.H. Gilmer, Rate-Theory Model of Polymer Crystallization. Physical Review

Letters, 1986. 56(25): p. 2708.120. Doye, J.P.K. and D. Frenkel, Mechanism of Thickness Determination in Polymer Crystals. Physical

Review Letters, 1998. 81(10): p. 2160.121. Doye, J.P.K., Computer simulations of the mechanism of thickness selection in polymer crystals.

Polymer, 2000. 41(25): p. 8857-8867.122. Anderson, K.L. and G. Goldbeck-Wood, Simulation of thickening growth in polymer

crystallisation. Polymer, 2000. 41(25): p. 8849-8855.123. Yamamoto, T., Molecular dynamics simulation of polymer crystallization through chain folding

The Journal of Chemical Physics, 1997. 107(7): p. 2653.124. Liu, C. and M. Muthukumar, Langevin dynamics simulations of early-stage polymer nucleation

and crystallization. The Journal of Chemical Physics, 1998. 109(6): p. 2536-2542.125. Yamamoto, T., Molecular dynamics simulations of polymer crystallization in highly supercooled

melt: Primary nucleation and cold crystallization. The Journal of Chemical Physics, 2010. 133(3):p. 034904-10.

126. Mandelkern, L., et al., Melting temperature of the n-alkanes and the linear polyethylenes.Macromolecules, 1990. 23(15): p. 3696-3700.

127. Theodorou, D.N. and U.W. Suter, Detailed molecular structure of a vinyl polymer glass.Macromolecules, 1985. 18(7): p. 1467-1478.

120

Page 121: Molecular Simulation Study of Homogeneous Crystal Nucleation

128. Yu, X., B. Kong, and X. Yang, Molecular Dynamics Study on the Crystallization of a Cluster ofPolymer Chains Depending on the Initial Entanglement Structure. Macromolecules, 2008. 41(18):p. 6733-6740.

129. Gaur, U. and B. Wunderlich, The Glass Transition Temperature of Polyethylene. Macromolecules,11980. 13(2): p. 445-446.

130. Han, J., R.H. Gee, and R.H. Boyd, Glass Transition Temperatures of Polymers from MolecularDynamics Simulations. Macromolecules, 1994. 27(26): p. 7781-7784.

131. Takeuchi, H. and R.-J. Roe, Molecular dynamics simulation of local chain motion in bulkamorphous polymers. /1. Dynamics at glass transition Journal of Chemical Physics, 1991. 94(11):p. 7458-7465.

132. Capaldi, F.M., M.C. Boyce, and G.C. Rutledge, Molecular response of a glassy polymer to activedeformation. Polymer, 2004. 45(4): p. 1391-1399.

133. Yi, P. and G.C. Rutledge, Molecular simulation of bundle-like crystal nucleation from n-eicosanemelts. The Journal of Chemical Physics, 2011. 135(2): p. 024903-11.

134. Turnbull, D. and J.C. Fisher, Rate of Nucleation in Condensed Systems. The Journal of ChemicalPhysics, 1949. 17(1): p. 71-73.

135. Morris, J.R. and X. Song, The anisotropicfree energy of the Lennard-Jones crystal-melt interface.The Journal of Chemical Physics, 2003. 119(7): p. 3920-3925.

136. HLtter, M., P. Veld, and G. Rutledge, Monte Carlo Simulations of Semicrystalline Polyethylene:Interlamellar Domain and Crystal-Melt Interface, in Progress in Understanding of PolymerCrystallization, G. Reiter and G. Strobl, Editors. 2007, Springer Berlin / Heidelberg. p. 261-284.

121