SlideShare a Scribd company logo
Melting Phenomena in Aluminium Nanowires
Thomas Simon Aldershof
Department of Physics
The University of Auckland
Supervisor: Professor Shaun Hendy
Dissertation submitted in partial fulļ¬lment of the requirements for the degree of BSc(Hons) in
Physics, The University of Auckland, 2016.
Abstract
The melting transition of a cylindrical, metal nanowire can progress through a number of diļ¬€er-
ent pathways. The three pathways identiļ¬ed in this study are surface melting, where a molten
layer envelops a solid core, linear melting, where the wire melts in the longitudinal direction
and fact induced melting, where the wire melts from the surface maintaining at least one solid-
vapour interface. The diļ¬€erent pathways are compared both through a simple model based on
free energy considerations, developed here for each pathway, and through molecular dynamics
simulations to identify the main parameters inļ¬‚uencing what pathway is taken
From the free energy models, the wire radius and the diļ¬€erence in the energy density of
the diļ¬€erent phase interfaces, āˆ†Ī³ are found as the deciding parameters. A discussion on the
diļ¬€erence in nucleation cost and importance of non-melting crystal facets is also included.
In the simulations, Face Centric Cubic, aluminium nanowires undergo the diļ¬€erent melting
pathways using molecular dynamics. From these the radius is conļ¬rmed to be an important
parameter deciding between the diļ¬€erent pathways. The interface energy of the crystal facets on
the surface of the wire are also identiļ¬ed as important parameters. The latter was accompanied
by shrinkage of the non-melting surfaces through the melting of the wetting surfaces, which the
free energy model does not take account of. The observed melting pathway was also sensitive
to the strength of the coupling of the heat bath, which decided whether the system acts more
canonically or micro-canonically.
The conclusion of this work is that linear melting occurs for wires with a diameter of at most
12 lattice spacings. Surface melting and facet induced melting occurred for thicker wires treated
more canonically or micro-canonically respectively. An accurate theory predicting the onset of
the melting transition needs to take the faceting of the wire into account as the non-melting
surfaces play an important role.
1
2
Acknowledgements
First and foremost I would like to thank my supervisor Professor Shaun Hendy for suggesting
this subject and allowing me to explore the interesting world of post-graduate research. I would
like to thank him for ensuring the regular meetings were not easily bypassed, even if my results
were meagre. He is also responsible for introducing me to condensed matter through the third
year course he taught together with Associate Professor Malcolm Grimson and Dr. Geoļ¬€ Will-
mott, who also have my gratitude.
I would further like to thank Dr. Geoļ¬€ Willmott, as well as Professor Geoļ¬€ Austin for
introducing me to academic research through the summer research scholarships they oļ¬€ered me
in successive summers.
Next I need to thank Kannan Ridings, who always seemed to be available to answer my
many questions, whether they were technical or theoretical. I am positive he spend more time
on my work than his own on some days and I would like to thank him for this dedication. The
code he provided me at the start of this project allowed me to hit the ground running and
without him this work would be far less complete.
Finally I would like to thank my lecturers, family, friends and peers as they made this year
the most enjoyable, albeit the most demanding year of my university career yet. Special mention
among these is my partner, Rebecca Bub, who always made an eļ¬€ort to help and understand
me in this project and gave her support when I needed it most, whether she was near or far.
3
4
Contents
Abstract 1
1 Introduction 7
2 Theory 9
2.1 Nucleation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Non-melting surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Wire geometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.1 Free energy evaluation of surface melting . . . . . . . . . . . . . . . . . . 11
2.3.2 Free energy evaluation of linear melting . . . . . . . . . . . . . . . . . . . 12
2.3.3 Free energy evaluation of facet induced melting . . . . . . . . . . . . . . . 13
2.4 Preferred melting pathway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3 Methodologies 17
3.1 Molecular dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.1 Initial state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1.2 Embedded Atom Method Potentials . . . . . . . . . . . . . . . . . . . . . 18
3.2 Numerical considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.1 Simulation dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.2 Time integration and the Langevin thermostat . . . . . . . . . . . . . . . 20
3.3 Post processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.3.1 Steinhardt paramaters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4 Analysis and results 23
4.1 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.1.1 Initialistion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
4.1.2 Simulation parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
4.2 Observed melting pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2.1 Observation of surface melting . . . . . . . . . . . . . . . . . . . . . . . . 25
4.2.2 Observation of linear melting . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.2.3 Observation of facet induced melting . . . . . . . . . . . . . . . . . . . . . 29
4.3 Comparison to theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
5 Discussion 33
6 Conclusions 35
5
6 CONTENTS
Chapter 1
Introduction
In recent years nanowires have been of increasing interest in both academic and industrial ļ¬elds
such as opto-electronics, nano computing and integrated circuits. The one-dimensional nature
and the high surface to volume ratio of these wires cause them to diļ¬€er from both bulk materials
and nanoclusters in a number of diļ¬€erent ways, prompting some to call metallic nanowires mag-
ical.1 Due to the potential of nanowires in nanoscale electronics, research has focused on their
electric properties. However, the potential range of applications is huge and extends to extreme
thermal conditions. Thus an extensive investigation into the thermal behaviour is warranted.
This document will focus on the melting phenomenon of a metal nano-wire. In particular,
the diļ¬€erent pathways through which a metal nanowire can melt and the parameters inļ¬‚uencing
this process. In previous work metal nanowires have been found to display several novel melting
phenomena and processes.2 This work seeks to determine the processes through which a metal
nanowire can melt and to investigate the parameters that inļ¬‚uence the wireā€™s preference for one
pathway over an other.
Figure 1.1: Diļ¬€erent pathways of melting investigated in this study. Dark colouring indicates a
particle classiļ¬ed as solid state while light colouring indicates a liquid. Solid lines connect the
geometries of the main pathways and dotted lines show possible intermediate states. The initial
state is taken to be the solid wire, which can melt through, top to bottom, surface, linear or
facet induced melting to the fully liquid sate. The wire may go through a state with premelted
facets ļ¬rst (left) or break up into spheres (right) before fully melting.
7
8 CHAPTER 1. INTRODUCTION
The pathways discussed in this document are given by ļ¬gure 1.1, deļ¬ning the diļ¬€erent melt-
ing geometries encountered. Notice that although the facet pre-melting geometry on the top-left
and the wire break-up on the top-right are encountered in this study, the focus is on the three
melting pathways shown in the centre. Surface melting, where the wire is completely encap-
sulated with a liquid layer. In the linear melting pathway the wire melts in the longitudinal
direction and never becomes fully encapsulated by the melt. Finally, in facet induced melting
the wire melts from the surface but maintains at least one crystal facet as a solid-vapour inter-
face. The ļ¬nal state of the system is always the fully melted case, although it may go through
a geometry where the solid wire has broken up into multiple spheres within the molten wire.
Chapter 2 summarises some of the theory vital to this study. First it revises classical nu-
cleation theory. Then it evaluates how the interface free energy densities inļ¬‚uence melting on
diļ¬€erent surfaces. With these two elements we then perform free energy evaluations for the
diļ¬€erent pathways. At the time of writing we are not aware of a similar evaluation for linear
or facet induced melting in the literature. We then combine these calculations in an attempt
to identify the preferred melting pathway and the parameters inļ¬‚uencing this preference. This
again has not been found in literature.
Chapter 3 covers the numerical methods used in this study. The chapter begins with a
discussion of molecular dynamics. This includes a justiļ¬cation of the potential used and lays
out the initial state of the system. It also lists the choices we made to obtain a smooth and
physical simulation that is computationally manageable. Finally it reviews the method used to
distinguish solid and liquid atoms.
Chapter 4 provides an exert of the simulations run showing a representative simulation for
each of the pathways. It also shows the initial state of the system and tests some of the physical
aspects of the simulations. Chapters 5 and 6 provide a summary of our result and point the
way for future research.
Chapter 2
Theory
In this chapter we evaluate some of the theory associated with the melting phenomena and
we attempt to predict the occurrence of the diļ¬€erent pathways. We will ļ¬rst review classical
nucleation theory and introduce some relations used in the later sections and then discuss non-
melting facets of a crystal lattice. After this is set-up we engage in a free energy evaluation of
the diļ¬€erent melting pathways. This method has been tested for surface melting but is new for
linear and facet induced melting. Finally we compare the relations obtained for the diļ¬€erent
pathways to identify parameters that may aļ¬€ect the preference of one melting pathway over the
other.
2.1 Nucleation Theory
An interesting and much-studied aspect of thermodynamics are phase transitions, which is de-
ļ¬ned as a qualitative change in the properties of an object or material. The most commonly
known examples are melting and vaporisation, the ļ¬rst of which is under discussion in this
document. The onset of this melting transition, as other types of transitions, is described by
classical nucleation theory (CNT).
According to to this theory the melting transition is always initiated at a localised nucleation
site. At this point a stable cluster of liquid needs to be formed from which the melting is then
promoted. One can make the distinction between two types, homogeneous and heterogeneous
nucleation. The former occurs when the nucleation site is within the bulk of the material and
away from any impurity. The latter occurs when the nucleation is encouraged by an impurity of
boundary of the material. In the nanowire, nucleation will always occur on or near the surface
and will thus undergo heterogeneous nucleation.
Both types of nucleation and the following melting transition describe the formation of a
liquid cluster of a certain critical size,l from which the melting transition progresses sponta-
neously. For the nucleation to occur the energetic cost of forming the extra phase interface is
less than the energy gained from the formation of the volume of liquid. Classically the growth
of these nucleation clusters is described by atomic vibrations as ļ¬rst proposed by Lindemann
in 1910. These atomic vibrations increase in amplitude with increasing temperature. However,
this description breaks down for a nanosized object. This is due to the low volume to surface
ratioand thus the relatively high number of less restricted atoms on the surface of the wire.3
9
10 CHAPTER 2. THEORY
The lower energy state results in a depression of the melting temperature of a nano-sized
object. This phenomenon has been extensively veriļ¬ed in the literature for many diļ¬€erent
systems and is still an active area of research.2,4 When comparing the liquid and solid state
of a cylindrical nanowire which, for simplicity, maintains its shape and density one can ļ¬nd
section 2.1 for the melting temperature.5
Tm = Tc 1 āˆ’
2(Ī³sv āˆ’ Ī³lv)
ĻRLm
(2.1)
Where the temperature dependence was introduced using the relation in equation (2.2). Note
also that the melting temperature is lower than the bulk melting temperature whenever Ī³sv < Ī³lv
.
fs āˆ’ fl = āˆ’ĻLm(1 āˆ’
T
Tc
) (2.2)
2.2 Non-melting surfaces.
At a nano-scale surface the orientation of the lattice with respect to this surface becomes im-
portant. The direction the surface is cut from the bulk lattice aļ¬€ects a number of properties,
such as the smoothness, packing ratio and surface energy density of the ļ¬rst layer of atoms.
The last of these is of the most interest in this study. In this document, we will specify this
orientation by Miller indices.
The surface energy density is speciļ¬c to the type of interface the surface is part of. Numer-
ous studies have tried to determine the surface energies of pure metals with varying success.
Experimentally this can often only be found for the solid-vapour energy, Ī³sv, although there is a
large range of reported values. Additionally as the lattice is only present in the solid state, the
liquid-vapour interface, Ī³lv, often has an accepted value independent of the lattice orientation.
However, the solid-liquid interface energy density, Ī³lv, is often diļ¬ƒcult to determine as it is
not often a stable state. Even if this is an equilibrium state, atoms will continuously melt and
crystallise and therefore will often not have a well-deļ¬ned facet. We can deļ¬ne the free energy
density diļ¬€erence.6ā€“10
āˆ†Ī³ = Ī³sv āˆ’ Ī³sl āˆ’ Ī³lv (2.3)
This quantity gives the free energy cost of replacing a solid-vapour interface with a solid-
liquid and a liquid-vapour interface per unit area. In other words, the cost to have a layer of
melt form on the solid surface. Like the interfacial energy densities, this quantity depends on
the surface orientation and had been found for some crystal facets of some metals. However,
when āˆ†Ī³ > 0, a liquid layer will form on this facet and measurements encounter the same
problems as with Ī³sl. In this case āˆ†Ī³ often can not be determined beyond the negative sign.6
The case where āˆ†Ī³ > 0 for only some facets of the surface is worth investigating by itself.
As some facets may have this condition satisļ¬ed and thus the form a liquid layer (called wetting
surfaces), while others still have positive āˆ†Ī³ and thus maintain a solid-vapour interface. The
latter quickly gained the name of non-melting surfaces and have been found to cause some in-
teresting phenomena.11ā€“13 In this study, they are important in the occurrence of facet induced
melting, as outlined in section 2.3.3. They also cause the facet melted geometry in ļ¬gure 1.1.
In Aluminium the (110) facet is a wetting surface, while the (100) and (111) facets are
non-melting surfaces. We thus expect the facet melted geometry to have (110) facets melted
with (100) or (111) intact. Due to the geometry of a nanowire, we cannot have all three facets
as part of the surface.
2.3. WIRE GEOMETRIES 11
2.3 Wire geometries
Figure 2.1 shows some of the possible geometries a nanowire may take during the melting tran-
sition. We can see that for each of the displayed geometries, the melting transitions can be
described by only one changing variable, r, l or h respectively. The following is a canonical free
energy calculation based on existing models for surface melting.5,11 They are adjusted for the
new pathways as we believe these calculations have not been done before ļ¬r linear and facet
induced melting. In section 2.4 we will compare the results of these calculations to identify the
parameters inļ¬‚uencing the preferred pathway.
(a) Surface melting (b) Linear melting (c) Facet induced melting
Figure 2.1: Abstracted wire geometries that can occur during the melting process with some
parameters shown. The solid state is dark, while the liquid state is lighter gray.
2.3.1 Free energy evaluation of surface melting
When an object undergoes surface melting a thin layer of liquid forms on the solid-vapour inter-
face before reaching the bulk melting temperature. This process was ļ¬rst proposed by Micheal
Faraday to describe the fusion of ice particles. The transition has later been observed in electron
scattering experiments that also saw the ļ¬rst nonmelting surfaces discussed before.7,14
When surface melting occurs, the solid-vapour interface is replaced by a solid-liquid and a
liquid-vapour interface, separated by a thin layer of liquid. As this transition will only spon-
taneously occur when it lowers the free energy we need Ī“sv + Fs > Ī“sl + Ī“lv + Fl + Fs. The
thickness of the layer, R āˆ’ r, changes continuously from R āˆ’ r = 0 from the unmelted case.
Thus Fl is initially negligible with no signiļ¬cant change in Fs for very small R āˆ’ r. This gives
rise to the condition also found in section 2.2, āˆ†Ī³ > 0.15
We assume the solid core retains the same shape as the original nanowire with radius r < R.
The liquid layer thus is a cylindrical shell with inner radius r and outer radius R. We can thus
write down the free energy as below, again taking account of the crystal facets where necessary.
Fsurf (r) = Ļ€r2
Lfs + Ļ€(R2
āˆ’ r2
)Lfl + 2Ļ€RLĪ³lv + 2Ļ€rL Ī³sl + āˆ†Ī³e
āˆ’Rāˆ’r
Ī¾ (2.4)
12 CHAPTER 2. THEORY
We assume the nanowire to be at equilibrium at all times. This allows us to ļ¬nd the
transition temperature by ļ¬nding the extrema. Thus we need dF
dr = 0.
dFsurf
dr
= 2Ļ€L r (fs āˆ’ fl) + Ī³ls + āˆ†Ī³ 1 +
r
Ī¾
e
āˆ’Rāˆ’r
Ī¾ = 0 (2.5)
Where the last term represents the energy cost of nucleation, with Ī¾, a critical coherence
length of the liquid layer speciļ¬c to the metal. As before we use equation (2.2) on the unmelted,
initial state, i.e. r = R. Thus we obtain the onset temperature of surface melting, equation (2.6),
in terms of the bulk melting temperature, Tc.
Tsurf = Tc 1 āˆ’
Ī³sv āˆ’ Ī³lv
ĻLmR
1 +
R
Rc
(2.6)
Where we deļ¬ned the critical radius Rc = Ī³svāˆ’Ī³lv
āˆ†Ī³ Ī¾ and used equation (2.3) for āˆ†Ī³. It is easy
to see that as long as Ī³sv > Ī³lv and R
Rc
> āˆ’1 the onset temperature of surface melting is below
the bulk melting temperature. To ensure this is an energetically stable state we evaluate the
second derivative of the free energy at r = R, T = Tsurf .
d2Fsurf
dr2
= 2Ļ€L
āˆ†Ī³
Ī¾
R2
+ 2Ī¾ R āˆ’
1
2
Rc (2.7)
This is greater than zero for āˆ†Ī³ > 0 and R > Rc. This ensures that we have been working in
a free energy minimum and surface melting can be an energetically favourable state.
2.3.2 Free energy evaluation of linear melting
Linear melting occurs when a thin cross sectional disk of width l melts before reaching the
bulk melting temperature. As the melting progresses l increases from 0 to L, the length of
the wire. The liquid-vapour interface and the melted volume grow linearly with l as the solid-
vapour interface and solid volume decrease similarly in size. The liquid-solid interface remains
constant and is given by the sum of the two cross sectional areas at either side of the melted disk.
Returning to nucleation theory it is easy to see that this method has a signiļ¬cantly higher
energy barrier than surface melting. Firstly the new interface is mostly away from the sur-
face of the nanowire. This makes the initial nucleation more energy expensive. Secondly this
type of melting does not immediately replace the solid-vapour interface, as we did above and
therefore maintains the, presumably, more expensive interface for longer. We will however still
employ a similar free energy analysis of this pathway. Through similar geometric arguments
as we employed in section 2.3.1 the free energy of the linear melting geometry can be found.
Equation (2.8) gives this as a function of the melting parameter, l.
Flin(l) = Ļ€R2
(L āˆ’ l)fs + Ļ€R2
lfl + 2Ļ€R(L āˆ’ l)Ī³sv + 2Ļ€RlĪ³lv + 2Ļ€R2
Ī³sl + āˆ†Ī³e
āˆ’ l
Ī¾ (2.8)
Again the nanowire is assumed to be in an equilibrium state. Finding the minima of the free
energy will thus enable us to ļ¬nd the transition temperature again. Note that this does assume
the nanowire has not undergone surface melting, just as in the previous section we implicitly
assumed the wire had not undergone linear melting.
dFlin
dl
= Ļ€R (fl āˆ’ fs) R + 2 (Ī³lv āˆ’ Ī³sv) āˆ’ 2āˆ†Ī³
R
Ī¾
e
āˆ’ l
Ī¾ = 0 (2.9)
2.3. WIRE GEOMETRIES 13
As in the previous section, the last term represents the energy barrier due to the nucleation
process. We again set the wire to the initial, unmelted state, l = 0, to obtain the transition
temperature.
Tlin = Tc 1 āˆ’ 2
Ī³sv āˆ’ Ī³lv
ĻLmR
1 +
R
Rc
(2.10)
Where Rc = Ī³svāˆ’Ī³lv
āˆ†Ī³ Ī¾, as before.
Again this transition temperature is below the bulk temperature for Ī³sv > Ī³lv and R
Rc
> āˆ’1.
For this to be an energetically accessible state the above parameters need to correspond to a
minimum in the free energy. That is, d2F
dl2 > 0 at l = 0, T = Tlin.
d2Flin
dl2
= 2Ļ€āˆ†Ī³
R2
Ī¾2
(2.11)
This is greater than zero for āˆ†Ī³ > 0 and R > Rc. Thus linear melting is an energetically
accessible state.
2.3.3 Free energy evaluation of facet induced melting
Facet induced melting (FI) is an eļ¬€ect of the interplay between wetting and non-melting surfaces.
While the non-melting surfaces of the nanowire remain unmelted, the wetting facets melt. This
allows wetting facet to grow in size. At some point it may be favourable to maintain only
one facet of solid-vapour interface while relinquishing the others in favour of maximizing the
solid-liquid interface of the wetting facets. This geometry is shown in 2.1c. Once again the free
energy can be written down using geometrical arguments.
Fpw(h) = 2RL Ī±Rfs +
Ļ€
2
āˆ’ Ī± Rfl + Ī±Ī³sv + (Ļ€ āˆ’ Ī±) Ī³lv + Ī± Ī³sl + āˆ†Ī³e
āˆ’2Rāˆ’h
Ī¾ (2.12)
Where, cos(Ī±) = Rāˆ’h
R gives half the angle the solid-vapour interface spans.
As in the previous sections the ļ¬rst derivative of the free energy is set to zero to ļ¬nd the minima
of this system. Note that dĪ±
dh = (h(2R āˆ’ h))āˆ’0.5 .
dFpw
dh
=
2RL
h(2R āˆ’ h)
R(fs āˆ’ fl) + 2Ī³sl + āˆ†Ī³ 1 +
2Ī±
Ī¾
h(2R āˆ’ h) + 1 e
āˆ’2Rāˆ’h
Ī¾ = 0
(2.13)
We use equation (2.2) to eliminate the volumetric terms and set the system to the onset of
melting, h = R, to obtain the temperature according to this free energy evaluation.
Tpw = Tc 1 āˆ’ 2
Ī³sv āˆ’ Ī³ls
ĻLmR
1 +
Ļ€
2
R
Rc
(2.14)
With Rc as in section 2.3.1.
Note that this is less than the bulk melting temperature, Tc, for Ī³sv > Ī³lv and Ļ€R
2Rc
> āˆ’1 Again
we ensure that the parameters above describe a minimum in the free energy by evaluating the
second derivative with respect to the melting parameter at h = R.
d2Fpw
dh2
= 4
L
Ī¾
2 + Ļ€
R
Ī¾
āˆ†Ī³ (2.15)
This has the sign of āˆ†Ī³ and thus we have a minimum for āˆ†Ī³ > 0. Thus from these free
energy calculations we ļ¬nd that the FI geometry can be an energetically favourable state.
14 CHAPTER 2. THEORY
2.4 Preferred melting pathway
In this section we aim to determine the preferred pathway of melting based on the free energy
calculations in the previous section. To do this we will compare the melting temperatures
found in equations (2.6), (2.10) and (2.14). We again ignore the dependence of the surface
energy density on the crystal facet. Otherwise one would use the found values of Ī³sv and āˆ†Ī³
for each facet and the facet-independent value for Ī³lv to ļ¬nd the diļ¬€erent surface energies. First,
we compare surface and linear melting by deļ¬ning āˆ†Tsl, the diļ¬€erence in the onset temperatures
of the two pathways.
āˆ†Tsl = Tsurf āˆ’ Tlin =
Ī³sv āˆ’ Ī³lv
ĻLmR
1 +
R
Rc
Tc (2.16)
Thus the onset of surface melting precedes the onset of linear melting when equation (2.16)is
negative. As Ī³sv > Ī³lv and R > 0 we deduce that surface melting is preferred when 1 + R
Rc
< 0.
Inversely the linear melting pathway is taken when 1+ R
Rc
> 0. This condition for surface melting
is the inverse of that between surface and the bulk melting temperature, 2.3.1. That is, linear
melting always precedes surface melting according to this canonical calculation. However, due to
the expensive nucleation stage, we can still observe surface melting. This would occur when the
energy cost of nucleation outweighs the energy gain from melting. We can summarise the found
conditions below, while deļ¬ning the transition radius Rsl, where a lower onset temperature
indicates a preference for this pathway, based on these free energy evaluations.
Rsl = āˆ’Rc =
Ī³lv āˆ’ Ī³sv
āˆ†Ī³
Ī¾
R < Rsl ā‡’ Tlin < Tsurf < Tc
R > Rsl ā‡’ Tc < Tsurf < Tlin
(2.17)
These conditions may be satisļ¬ed for some facets of the nanowire while not on the other.
This suggests that surface melting will have its initial nucleation site on a diļ¬€erent facet than
linear melting.
We can similarly compare surface melting and FI melting. Deļ¬ning āˆ†Tsf , the diļ¬€erence in
the onset temperatures of surface and FI melting.
āˆ†Tsf = Tsurf āˆ’ Tpw =
Ī³sv āˆ’ Ī³lv
ĻLmR
1 + (Ļ€ āˆ’ 1)
R
Rc
Tc (2.18)
The onset of surface melting precedes the onset of FI melting when āˆ†Tsf < 0. Thus when
1 + (Ļ€ āˆ’ 1) R
Rc
< 0 , the wire melts through the surface melting pathway. This now culminates
into the following conditions, deļ¬ning the transition radius Rsf .
Rsf = āˆ’
Rc
Ļ€ āˆ’ 1
= āˆ’
Ī³sv āˆ’ Ī³lv
(Ļ€ āˆ’ 1)āˆ†Ī³
Ī¾
R < Rsf ā‡’ Tlin < Tsurf
R > Rsf ā‡’ Tsurf < Tlin
(2.19)
2.4. PREFERRED MELTING PATHWAY 15
Note that these conditions do not include the bulk melting temperature. As Rsf is smaller
than Rsl and is smaller than the bulk melting temperature, it is possible to have the transition
between the two pathways before reaching the bulk melting temperature. However to actually
predict at what value this occurs we need values for the free energy densities of the interfaces.
These quantities have had a large range of reported values and at present time it is not possible
to predict the melting pathway other than for a speciļ¬c simulation where these can be extracted
from the potential.
Finally we can compare the linear and FI melting pathways. Deļ¬ning āˆ†Tlf , the diļ¬€erence
in onset temperature between these two pathways.
āˆ†Tlf = Tlin āˆ’ Tpw = (Ļ€ āˆ’ 2)
Ī³sv āˆ’ Ī³lv
ĻLmRc
Tc = (Ļ€ āˆ’ 2)
āˆ†Ī³
ĻLmĪ¾
Tc (2.20)
Notice that unlike equations (2.16) and (2.18) this is constant with respect to R. Note that
the sign of āˆ†Tlf is the sign of āˆ†Ī³. Thus linear melting is preferred for negative āˆ†Ī³ and inversely,
FI melting is preferred when āˆ†Ī³ > 0. We are unable to deļ¬ne a Rlf as we did for the previous
two comparisons, however we can still write down the following set of conditions. Where again,
a lower onset temperature indicates a preference for the associated melting pathway.
āˆ†Ī³ < 0 ā‡’ Tlin < Tpw
āˆ†Ī³ > 0 ā‡’ Tpw < Tlin
(2.21)
Combining the three conditions in equations (2.17), (2.19) and (2.21) we can summarise the
results of the last two sections in table 2.1. With Rsl and Rsf as before.
Preferred pathway Radius āˆ†Ī³ Caveat
Surface melting > Rsl > 0 Low nucleation cost
Linear melting < Rsl > 0 High nucleation cost
facet induced melting < Rsf > 0 Preceded by faceting
Bulk melting > Rc < 0
Table 2.1: Conditions for the diļ¬€erent melting pathways as obtained from the free energy
evaluations in sections 2.3 and 2.4
16 CHAPTER 2. THEORY
Chapter 3
Methodologies
This section describes the numerical methods used. First, we take a look at molecular dynamics
and the choice of some parameters within this method. We evaluate the elements of a single
simulation to provide reproducibility. This includes a discussion of the computational consider-
ations and simulation parameters. We ļ¬nish this chapter discussing the Steinhardt parameters
and their role in the solid-liquid classiļ¬cation.
3.1 Molecular dynamics
The model we employed in chapter 2 assumes an approximately cylindrical shape. When the
radius of the wire is suļ¬ƒciently small, faceting of the nanowire surface renders this assumption
invalid. This puts a lower limit on R, the initial radius of the solid wire. To have a one di-
mensional nano-scale object we need the radius to be less than 100 nm with the length greater
than 100nm. Time integration with a timescale of the order of nanoseconds is required to
have suļ¬ƒcient resolution of the melting process. These constraints ļ¬t best within a molecular
dynamics framework. Alternatives would be Monte Carlo or electronic structure methods for
longer timescales or smaller systems respectively.
The ļ¬rst study that can be classiļ¬ed as Molecular Dynamics (MD) was published posthu-
mously by Enrico Fermi.16 The name ā€molecular dynamicsā€ was coined by Alder and Wain-
wright 4 years later.17 It is now recognised as an eļ¬€ective method to model many-body systems,
which can be larger and larger as computing power increases, where Fermi started with a mere
64 particles. MD evaluates Newtonā€™s laws on one particle due to a potential created by the
particles in the close environment of this particle. It ignores electronic structure and other
quantum eļ¬€ects, which can be justiļ¬ed by using the Born-Oppenheimer approximation and
assuming that the electrons are in equilibrium with the nucleus.18
3.1.1 Initial state
As outlined in section 3.2 the number of atoms needs to be approximately constant. To obtain
the initial state, these atoms were arranged in an FCC lattice from which we cut a cylinder
along the (100) direction to obtain a nanowire. An example of this initial wire is shown in
ļ¬gure 3.1.
This wire is placed in a unit cell with periodic boundary conditions in the z-direction such
that we obtain an inļ¬nitely long nanowire. We thus eliminate any eļ¬€ects due to the ļ¬nite length
of the wire and look solely at the nano-dimensions of our wire. For similar reasons there are
solid boundaries in the x and y directions. This ensures there are no interactions in the radial
17
18 CHAPTER 3. METHODOLOGIES
Figure 3.1: Showing the intial state of a nanowire as cut from a perfect FCC lattice. The left
hand ļ¬gure shows the the (111) plane. The right hand ļ¬gure faces the (001) plane with the
(100) and (110) planes marked.
direction outside the nanowire.
3.1.2 Embedded Atom Method Potentials
As mentioned before classic MD does not take quantum eļ¬€ects in account. In these methods
intermolecular forces are modelled by hard spheres or more commonly a Lennard-Jones poten-
tial.19 This is radially symmetric, pairwise potential. The equations of motion can now be
obtained from the potential through classical mechanics. Which gives the force between any
pair of atoms as described in equation (3.1). Where the we substituted the Lennard-Jones
potential, Ī¦LJ .
F(r) = āˆ’ Ī¦LJ = 2
rm
r
6
āˆ’
rm
r
12
(3.1)
Although the LJ potential is an improvement over the hard sphere model it is still merely a
pair potential model, based only on the inter-atomic separation. The Embedded-Atom Method
(EAM) potentials were developed by Daw and Baskes to model mesoscopic phenomenon such
as micro-fractures in metals. The pair potential had insuļ¬ƒcient detail, while methods explicitly
taking electrons into account could not handle the meso scale required.20,21
In EAM methods atoms are embedded into a ā€œJeliumā€, that is, they are embedded into
a potential deļ¬ned not only by the other atoms but also by the electron cloud tied to these
atoms. To avoid solving the SchrĀØodinger equation they assumed that each atom experiences
a locally uniform electron density, deļ¬ned by the atoms already in the direct environment of
the introduced atom. This electron density is determined empirically. That is, the model is
set up implicitly including this electron density after which it is ļ¬tted to experimentally known
parameters. This makes the EAM potentials semi-empirical. It also makes EAM potentials
signiļ¬cantly more accurate than pair potentials while maintaining relatively cheap numerics.
EAM potentials thus need to be tailored to the material under investigation. Furthermore, for
any material there is a number of diļ¬€erent potential each suited to a diļ¬€erent purpose.22
3.2. NUMERICAL CONSIDERATIONS 19
In this study we investigate phenomena that happen close to the melting point and we thus
hope to model the melting process accurately. For this we need a potential that closely ļ¬ts the
melting temperature and the latent heat per atom. As these seem to be key parameters from
our study in chapter 2. The potential we chose also accurately models the lattice contraction
or relaxation at the surface. This lattice property can be monitored experimentally through
ion scattering and has been shown to be closely related to the surface energies of the diļ¬€erent
crystal facets, another important parameter from our theoretical studies.8,23
All simulations were run on the PAN cluster, maintained by NESI. This allowed us to run
the simulations on parallel cores to maximise the available computing power. The LAMMPS
software, used to run the simulations, quite naturally adopts to parallel computing by dividing
the unit cell into smaller sections. Each of these sections then is assigned to a core. These cores
then only need to communicate about the dynamics close to the box boundaries. LAMMPS
is an open-source software package developed by Sandia National Laboratories and currently
maintained by this institution and Temple University. It is speciļ¬cally designed for molecular
dynamics and is compatible with the above potential.24
3.2 Numerical considerations
In any numerical simulations one has to make a choice between computational eļ¬ƒciency and
physical completeness. In a simulation one can control almost any aspect of the dynamics,
changing the outcome by changing a single parameter. It is therefore vital that the parameter
set chosen agrees with the physics dominating the real world. One needs to compromise this
absolute accuracy to maintain a simulation with a feasible computing cost. This section dis-
cusses some of the choices this study has made to balance physical accuracy and computing cost.
3.2.1 Simulation dimensions
First among these choices was the size of the initial system already discussed in section 3.1.1.
As outlined in that section a unit cell of any simulation handled in this study contains approxi-
mately 25000 atoms. Note that the EAM potential formally scales as N2 as for any particle the
potential and thus the force on that particle is deļ¬ned by the positions of all other particles in
the simulation, relative to the particle under consideration. To reduce the computational cost
we only take into account particles that have a signiļ¬cant contribution to this potential, that is
particles in the near neighbourhood of the particle. This near neighbourhood was deļ¬ned to be
a ball with a radius of 2.5 times the lattice spacing, which is around 10ĖšAfor Aluminium. The
list of atoms present in this near neighbourhood is updated regularly.
This choice restricts the volume and causes the length and radius of the wire to be directly
related. As the radius is one of the parameters we allow to vary, we should ensure that a
changing length has a minimal impact on the simulated dynamics. Even though our wire is
theoretically inļ¬nite, through the periodic boundary condition at the z-boundary, all dynamics
are still contained to just one simulation cell. This is especially worrying for a wire of relatively
large radius, and thus short length as some phenomena may require group dynamics over a
large distance. One of these long range phenomena is the break-up of the molten wire into balls
of liquid to minimise the surface area. This requires the bulk movement of the molten metal
and the momentary increase in surface area as the wire breaks the cylindrical shape. These two
properties make this process a longer wavelength phenomenon and we can use this to test our
system in this respect.
20 CHAPTER 3. METHODOLOGIES
We refer to the wires treated in sections 4.2.1 and 4.2.2 from the results chapter ahead as
the thick and thin wire respectively. The caloric curve of the thin wire shows a rise and fall in
the energy per atom after it has completely molten, corresponding to the break-up of the liquid
wire as discussed above. The caloric curve corresponding to the thicker wire does not show
these dynamics which was some cause of concern. To put these concerns to rest we allowed
the thick wire to continue to heat for half again the simulation time for the thick wire. This
simulation showed a similar dynamic did occur at a higher temperature. The lack hereof in
the regular simulation time could just be due to the fact the thicker wire needs to overcome a
greater energy barrier to break up into the spherical shape. Some conļ¬dence that the length
scale is suļ¬ƒcient to capture any melting dynamics can be taken from ļ¬gure 4.4 where we see
the melting take place at two diļ¬€erent sites. This suggest that the wire in that simulation is
signiļ¬cantly longer than the interaction wavelength of any of the melting dynamics.
3.2.2 Time integration and the Langevin thermostat
The next choices concern the time integration of the simulation. Molecular dynamics studies
modelling metals often use a time step of the order of femtoseconds. Any larger than this may
cause discontinuous dynamics where smaller timesteps provide no signiļ¬cant change. In this
particular study we use a timestep of 2 femtoseconds for a million timesteps, giving a total
simulation time of 2 nanoseconds. This simulation length is necessary to capture the complete
melting transition.
To heat the system we use a Langevin thermostat. Physically, the Langevin thermostat can
be thought of as a heat bath of very light solvent molecules in which the system is immersed.
The temperature of this heat bath is controlled exactly. The system attempts to equilibrate
with the heat bath through collisions with the solvent particles. This avoids the system being
treated as a vacuum and allows the temperature of the system to be controlled in a natural
manner through the heat bath. The dynamics for the ith solvent element can be captured by
the following stochastic equation.
miĀØxi(t) = Fi ({x(t)}) āˆ’
mi
Ī»i
Ė™xi(t) + Ri(t) (3.2)
Where mi and xi are the mass and positions of the system particles. Fi is the force de-
ļ¬ned by the EAM potential, which thus depends on the positions of the nearby particles. R
is a stochastic term with a Gaussian distribution of mean zero. mi
Ī»i
Ė™x forms the coupling with
the heat bath. The strength of this coupling with respect to the momentum of the particle is
determined by Ī»i, the Langevin parameter, which has units of time and can be seen as the time
necessary for the system to equilibrate with the heat bath. In general this Langevin parameter
can depend on the particle, however in this work it will be constant for all atoms in any given
simulation. Note that the Langevin parameter is inversely proportional to the coupling strength
between the system and the heat bath. Thus a low Langevin parameter will cause the temper-
ature to be strictly controlled by the temperature of the solvent, thus mimicking a canonical
system. Inversely a high Langevin parameters will mimic a micro-canonical system.25,26
3.3. POST PROCESSING 21
3.3 Post processing
During an MD simulation the state of the system is written to ļ¬les at a ļ¬xed time interval.
These ļ¬les contain the atomic positions, the kinetic and potential energies and the temperature
from the system. The latter two are determined from the velocities of the particles where the
potential is determined by the positions of the atoms within the EAM potential. The energies
and temperature together allow us to plot caloric curves to monitor how the transition diļ¬€ers
from bulk melting.
The atomic positions allow us to make qualitative judgements about the state of the system
and they are instrumental in distinguishing between the diļ¬€erent melting pathways. As the
transition progresses a group of atoms changes from being in a strictly ordered lattice to a
disordered, ļ¬‚uid state. This transition is less obvious when looking at a single atom and the
atomic position of one atom is not suļ¬ƒcient to determine the phase this atom is in. However,
we would like to determine whether any atom in our system is part of the solid or liquid phase.
We thus need to evaluate the surroundings of the atom and determine whether it is an ordered
lattice or a disordered ļ¬‚uid.
3.3.1 Steinhardt paramaters
From the many methods in use to do this solid-liquid classiļ¬cation, this study uses the Steinhardt
parameters. These are named after Paul Steinhardt, who published a paper discussing bond
order parameters.27 These are quantities evaluating the symmetry of the ā€bondsā€ around a
particular atom based on spherical harmonics. Here bonds are characterised not only by the
proximity to the atom under consideration but also the direction with respect to other bonds
from this atom. As spherical harmonics are distinguished by their degree, l, there are diļ¬€erent
degrees of the Steinhardt parameters, denoted ql. In this study only the order parameter of the
sixth degree is of interest.
Figure 3.2: Snapshots showing the progression with a consistent solid-liquid distribution in
cross-section throughout the length of the wire with atoms classiļ¬ed as either solid (copper)
or liquid (blue). Each snapshot is accompanied by the distribution of the ĀÆq6 Steinhardt order
parameter for that frame showing the classiļ¬cation as either solid (green) or liquid (blue).
Together they form an empirical conļ¬rmation of the validity of the Steinhardt parameters as
a method for solid-liquid classiļ¬cation. The last snapshot has all atoms classiļ¬ed as liquid
removed, to highlight a shortcoming of this method.
22 CHAPTER 3. METHODOLOGIES
The Steinhardt parameters can thus be used to distinguish between diļ¬€erent lattice types as
well as between the solid and liquid state. Locally this is captured by a complex valued vector
for each atom describing the length and direction of each bond to this atom. This vector can be
converted to real numbers by taking the Fourier invariant of the ļ¬eld of neighbouring complex
vectors. Finally these are averaged to smoothen the distribution of thermal ļ¬‚uctuations. These
average values are denoted as ĀÆql, where we use the ĀÆq6 order parameter in this study.27,28
Thus in short the Steinhardt parameters are real values assigned to each atom quantifying
the relation this atom has to its neighbourhood. The values of these ĀÆq6 corresponding to dif-
ferent crystal structures and thermodynamic phases can be found in literature. The latter of
these makes these parameters useful in following a phase transition. It is worth noting that we
need to deļ¬ne a radius of the near neighbourhood of an atom, for this method of liquid-solid
classiļ¬cation to function. This distance is usually chosen to be the ļ¬rst minimum of the ra-
dial distribution function, which can be found either from the atomic positions or the literature.
To test the solid-liquid classiļ¬cation weā€™ve picked a simulation that had a fairly consistent
cross-sectional picture. In other words, for any particular frame, the image seen for the cross-
sectional snapshot does not change signiļ¬cantly as the viewpoint moves along the length of the
wire. For such a wire the proportion of liquid to solid atoms in the cross-sectional snapshot
should be approximately the same as that proportion for the complete wire. Representative
snapshots of these cross sections are shown in ļ¬gure 3.2 along with the distribution of the ĀÆq6
parameter for the wire at the time of that snapshot.
We can now visually conļ¬rm that the atoms classiļ¬ed as liquid, actually seem to be in a
disordered state. Similarly, the atoms classiļ¬ed as solid are ordered into a crystal lattice. Some
small calculations also give that the distribution of the ĀÆq6 parameter gives the same liquid frac-
tion as counting the atoms in a representative number of layers in the cross-sectional snapshot.
However the last snapshot shows a seemingly fully melted wire with some atoms classiļ¬ed as
solid still present along the surface of the wire. These atoms appear to be misclassiļ¬ed due to
their position at the physical boundary of the system. Aside from this caveat, we believe the
Steinhardt order parameter to be a valid method of solid-liquid classiļ¬cation.
Chapter 4
Analysis and results
In molecular dynamics the position of all atoms can be tracked throughout the simulation.
This allows us to take snapshots at any given time to evaluate the state of the system. Exam-
ples of these snapshots can be found throughout this chapter. This becomes especially useful
after solid-liquid classiļ¬cation as discussed in section 3.3. In these simulation it was chosen
to take snapshots every 4000 fs and thus produce 500 snapshots for a 2ns simulation. With
these snapshots we also evaluate the temperature, from the average speed of the particles, as
well as the kinetic and potential energies of the system and per atom. Where the energies are
found from average speeds and the positions of the atom within the EAM potential respectively.
We can use all this information to evaluate our simulations in a more qualitative manner.
In this chapter there are two main parameters that are varied. The ļ¬rst is the size of nanowire,
deļ¬ned by its radius. The total length is trivially held constant through the periodic boundary
condition in the z-direction and is thus not a relevant parameter. Note that the length of the
wire within a unit cell may change between simulations, as described in section 3.2.1.
4.1 Preliminaries
Before running the simulations of wires along the diļ¬€erent melting pathways we need to en-
sure that our simulations model a physical and stable system. We ļ¬rst ensure that the initial
system is a stable and feasible conļ¬guration We also check the simulation parameters such as
the Langevin parameter and the heating rate are within a reasonable range and provide the
required resolution to monitor the changing system. Finally we explore some of the output
parameters and visualisation tools available. This include verifying the Steinhard parameter
accurately classiļ¬es solid and liquid atoms.
4.1.1 Initialistion
As discussed in chapter 3 we ļ¬rst need to equilibrate the wire and allow the lattice to relax.
To allow this we performed an NVT minimisation before starting each simulation. We notice
a slight, but important change in the wire geometry after this relaxation period. Figure 4.1
shows a (001) cross-sectional view of the initial wire before and after relaxation at 600 Kelvin,
the initial temperature of our simulations. To ensure that the right hand side was a stable state
a simulation was run unheated for the length of the later simulations, 2 ns. This preserved the
shape of the wire suļ¬ƒciently and no un-physical behaviour was noticed.
23
24 CHAPTER 4. ANALYSIS AND RESULTS
We can see that before the relaxation period the wire is cut from a perfect lattice, as none
but the ļ¬rst layer of atoms are visible. Although the cross section is not perfectly circular, the
layering of the atoms is such that it models this shape as closely as possible. This perfection
disappears after relaxation. First we note the slight thermal expansion. We also see the atoms
that were hidden behind the ļ¬rst layer due to the stochastic elements of the potential and the
non-uniform initial velocity distribution. However, the wire still has a distinct crystal structure,
which is re-established by the fact that almost all atoms are still classiļ¬ed as solid. The few
atoms classiļ¬ed as liquid can be seen as outliers that will likely soon crystallise again.
A more subtle eļ¬€ect of the relaxation is the increased faceting of the cross section. We
see that in particular the diagonal facets have ļ¬‚attened as to maximize the (110) facets. This
results in the cross section becoming almost octagonal with the (110) and (100) facets being
the only two substantial facets on the surface of the wire. This is likely to have an eļ¬€ect as
chapter 2 found the surface energies of these facets to be critical. We still expect the theory to
hold reasonably well as we knew the perfectly cylindrical shape was an approximation in that
chapter.
4.1.2 Simulation parameters
As outlined in section 3.2 the total heating done and the total duration of the simulations is
held constant. It seems to follow that the heating rate remains constant across simulations.
This only holds true for the heat bath connected to our simulation. The Langevin parameter,
Ī», inļ¬‚uences the rate at which this heat enters the nanowire system. As this parameter is the
means to control what ensemble the simulation represents we cannot avoid this discrepancy.
Instead we will investigate the inļ¬‚uence of the Langevin parameter on the heating rate.
Figure 4.1: Initial wire as created before the start of the simulation (left) and after an initial
relaxation period (right) with the crystal facets indicated. Atoms classiļ¬ed as solid are indicated
copper while those classiļ¬ed as liquid are blue.
4.2. OBSERVED MELTING PATHWAYS 25
4.2 Observed melting pathways
In this section the diļ¬€erent pathways outlined in the theory of chapter 2 are found and followed
in simulation. We managed to ļ¬nd each of the pathways discussed in the theory and will thus
discuss them again and compare the theory to the simulations.
4.2.1 Observation of surface melting
As described in section 2.3.1, surface melting is the forming of a liquid layer on the nanowire
surface. This liquid layer encapsulates the remaining solid core fully for there to be complete
surface melting. Figure 4.2 shows the progression of the transition through snapshots. Fig-
ure 4.3 shows the caloric curve and liquid fractions of two simulations showing surface melting
at diļ¬€erent Langevin parameters. The left graph is of the same simulation as the snapshots.
Note that the actual melting transition is relatively short. It is preceded and followed by a
signiļ¬cant heating period characterised by a linear trend in the caloric curve. This shows the
wire is a physical and stable system.
Figure 4.2: Snapshots for the progression of surface melting left to right, top to bottom. This
simulation has a relatively thick wire and low Langevin parameter Ī» = 10ps. Atoms identiļ¬ed
as solid are copper coloured, blue indicates atoms in a liquid state. Each snapshots has 4
projections of the wire, two side on views, facing the (110) plane(left) and two cross sections,
facing the (001) plane(right). The top row shows the full wire where the bottom row shows
only the atoms classiļ¬ed as solid.
26 CHAPTER 4. ANALYSIS AND RESULTS
We can now compare the snapshots with the caloric curve through the liquid fraction. Each
snapshot has 4 projections of the wire as outlined in the caption. First note that snapshot A
is not the initial state of the wire, which looks more like ļ¬gure 3.1. Snapshot A is just before
the onset of surface melting. We see that the (110) facets have already melted and now form a
solid-liquid interface. These facets have increased in size by melting the sides of the (100) facets,
which are also part of the (110) facet. This period of facet melting can be seen as a relatively
slow, but accelerating rise of the liquid-solid fraction to about 10 percent. At the same time
we see the caloric curve deviating from the linear heating characteristic for a fully solid system.
This is due to the transition of these atoms from solid and liquid and the associated latent heat.
Snapshot B shows the wire just after the onset of surface melting. We see the wire fully
encapsulated in a liquid layer with solid core tending to maximise the (110)-facets. This is
reinforced in snapshot C which shows the same wire at a later stage of surface melting. Keep in
mind that all atoms are displayed in these snapshots, gradually fading as they are further away.
Thus although in snapshot C there seems to be a signiļ¬cant (100) facet on the right hand side of
the wire, especially obvious in the solid only cross section. However, it is quite thin and we see
that there is a second facet further towards the centre where we see a far greater density of the
displayed atoms, showing that this later facet extends along a larger portion of the wire. The
melting of the solid core also seems to be fairly symmetric, allowing the solid core to remain
in the centre. This is backed up by the linear trend in the liquid fraction between 10 and 90
per. We see this corresponds to the ļ¬‚at section in the caloric curve, as one would expect from
a canonical system.
The ļ¬nal snapshot, D, shows the wire after the solid core has broken up and there is only
a small cluster remaining. This cluster would melt away quickly as it contains less than 10
percent of the original atoms. We see that the liquid fraction ļ¬‚attens again as the number
of solid atoms remaining becomes of the order of misclassiļ¬ed atoms near the surface. After
the liquid fraction reaches this maximum value the caloric curve again assumes a linear trend,
showing the heating of a pure liquid.
The second graph in ļ¬gure 4.3 shows the caloric curve and liquid fraction for the same wire
in a simulation with a higher Langevin parameter. These follow a very similar trend as do the
snapshots, hence why these are not shown. However, the caloric curve starts to show a deļ¬nite
dip in temperature, after reaching a maximum. This shows the changing ensemble as we move
away from the canonical treatment. When the Langevin parameter further we start to see facet
induced melting. This is discussed further in section 4.2.3.
Figure 4.3: Caloric curves (red) with liquid fraction (blue) for a simulation of surface melting.
These simulations have diļ¬€erent Langevin paramaters, Ī» = 10ps and Ī» = 40ps, respectively.
4.2. OBSERVED MELTING PATHWAYS 27
4.2.2 Observation of linear melting
Similar to the previous section ļ¬gure 4.4 shows the progression of linear melting in snapshots
while ļ¬gure 4.5 shows the caloric curve and liquid fraction of two diļ¬€erent simulations showing
linear melting. The snapshots are for a simulation with Langevin parameter Ī» = 200ps. Thus
the right caloric curve corresponds to the same simulations as the snapshots. This particular
value was chosen because the snapshots obtained from this simulation best showed the range
of phenomena encountered. Linear melting was observed for the full range of the Langevin
parameter. In other words for both canonical and micro-canonical systems. Note that this is
not true for surface melting or for facet induced melting in thicker wires, as we will encounter in
the next section. As before the actual melting transition is a relatively small part of the caloric
curve, to ensure the system is physical. At the end of the caloric curve we see the forming of a
liquid ball and the associated drop in energy, as discussed in section 3.2.
First we examine the snapshots. The ļ¬rst thing to note that the wire needed to be signif-
Figure 4.4: Snapshots showing the progression of linear melting left to right, top to bottom.
This simulation has a relatively thin wire and an intermediate Langevin parameter Ī» = 100ps.
Atoms identiļ¬ed as solid are copper coloured, blue indicates atoms in a liquid state. Each
snapshots has 4 projections of the wire, two side on views, facing the (110) plane(left) and two
cross sections, facing the (001) plane(right). The top row shows the full wire where the bottom
row shows only the atoms classiļ¬ed as solid.
28 CHAPTER 4. ANALYSIS AND RESULTS
icantly thinner to clearly and reliably show linear melting. This agrees with what we saw in
chapter 2 where the distinguishing feature between linear and surface melting seemed to be the
radius, Rsl. However a caveat was placed by this ļ¬nding as linear melting has a signiļ¬cantly
higher nucleation cost and thus may not be preferred when close to this critical radius. This is
reļ¬‚ected in snapshot A, which again is not the initial state, where a signiļ¬cant portion of the
atoms have already melted but the wire still has not cut through. The energy cost in the latent
heat of these atoms is related to the nucleation cost of this pathway.
In most snapshots we see some faceting is present in the cross-section of the full wire. How-
ever, it is signiļ¬cantly less strong and from the solid only pictures we see that the faceting
only occurs along sections of the wire as these cross-sections remains fairly round for the full
duration of simulation. Snapshot B shows the wire just after it has been cut through and we
see that it seems to be cut in two places, one just out of view at the left. The lower section
in between these cuts is also completely detached from the lower surface. I.e. it is undergoing
facet-induced melting. This is reinforced by snapshot C, which also shows this part of the wire
has melted signiļ¬cantly faster, which is partly due to the smaller size in snapshot B and partly
due to the greater melting surface for facet induced melting.
The fact that we see these two diļ¬€erent pathways supports the hypothesis that the nucle-
ation cost of linear melting is an important factor. Since nucleation is ultimately a random
process, it is to be expected that we see seemingly ambiguous results as ļ¬gure 4.4. When we
compare the snapshots with the right-hand liquid fraction we see this reļ¬‚ected as a kink be-
tween two fairly linear trends. This kink is the moment the facet induced melting has ļ¬nished
and we are left with only linear melting. The initial nucleation is reļ¬‚ected in the initial curved
section, which now does not represent faceting. We see a similar trend in the caloric curve.
initially there is a linear trend as we are heating a pure solid. When the liquid fraction starts to
increase, the caloric curve deviates from this linear trend. In the period with the two melting
types simultaneous we see a sharp drop in te caloric curve as energy is released as latent heat
faster than the system is allowed to heat up under this high Langevin parameter. After the
facet induced melting is completed the temperature slowly increases again as the latent heat of
the linear melting alone is not enough to maintain the negative gradient.
The linear melting pathway was observed only for relatively thin wires, with a diameter of
at most 12 lattice spacings. However, when linear melting seems largely independent of the
Langevin parameter. When linear melting was observed for a simulation with low Ī» it would
generally also been seen for a simulation with Ī» large.
Figure 4.5: Caloric curves (red) with liquid fraction (blue) for a simulation of linear melting.
These simulations have diļ¬€erent Langevin paramaters, Ī» = 10ps and Ī» = 200ps, respectively.
4.2. OBSERVED MELTING PATHWAYS 29
4.2.3 Observation of facet induced melting
As in the previous two sections the snapshots shown in ļ¬gure 4.6 and the caloric curve and liquid
fraction found in ļ¬gure 4.7 follow a nanowire undergoing facet induced melting. The snapshots
are for a simulation with a high Langevin parameter, Ī» = 200ps. Thus the corresponding caloric
curve is on the right. Unlike what was observed for the linear melting pathway, the Langevin
parameter needs to be suļ¬ƒciently high to observe facet induced melting. In other words, we
observe it only for more micro-canonical systems.
In section 2.3.3 we deļ¬ned facet induced melting (FI) as a melting transition that forms a
liquid layer on most of the surface while maintaining a solid-vapour interface on at least one
facet for the duration of the transition. This is indeed what we see in the snapshots below.
In snapshot A there signiļ¬cant faceting occurs along the full length of the wire. However, the
faceting is not symmetric as it was for surface melting. Only the (110) facets on the right
hand side have melted along the full side of the wire. This asymmetry was present in all FI-
simulations, even though it seemed random which side of the wire would melt.
Note that the side-on view has been rotated from the cross section such that the right hand
Figure 4.6: Snapshots for the progression of facet induced melting left to right, top to bottom.
This simulation has a relatively thick wire and high Langevin parameter Ī» = 200ps. Atoms
identiļ¬ed as solid are copper coloured, blue indicates atoms in a liquid state. Each snapshots
has 4 projections of the wire, two side on views, facing the (110) plane(left) and two cross
sections, facing the (001) plane(right). The top row shows the full wire where the bottom row
shows only the atoms classiļ¬ed as solid.
30 CHAPTER 4. ANALYSIS AND RESULTS
side, the side showing the initial melting, faces towards the viewer. Snapshot B shows the same
asymmetry. The wire has now completely disconnected from both the (110) and (100) facets
on the right hand side of the cross-sections. Examining the side-on view, we see that the solid
phase still extends the full width of the wire. There is still a signiļ¬cant fraction of these (110)
facets that maintains a solid-vapour interface. This part resembles the geometry used in the
chapter 2, but the rest of the wire does not. Showing an almost spherical shape, ļ¬‚attened by
last remaining (110) solid-vapour interface. Recall that in the chapter 2 we did not properly
take account of the wetting and NM facets, which is likely to be the source of this discrepancy
between the theory and simulation.
The liquid fraction follows a fairly un-eventful trend looking a lot like that for surface melt-
ing. We initially see the initial melting of the (110) facets in the almost quadratic increase in
liquid fraction. As the wire starts the main melting process after forming the liquid layer on the
(100) facet the liquid fraction follows a linear trend, which continues until a near 100 percent
liquid fraction is reached. The caloric curve however is slightly diļ¬€erent showing a deļ¬nite dip in
temperature as the melting progresses. This is again explained by the high Langevin parameter.
This makes the transition occur in a more micro-canonical ensemble. Thus the kinetic energy
of the atoms, inseparable from the temperature, is used directly to supply the heat required for
the melting transition rather than the heat added to the system through the heat bath.
Snapshot C shows the wire after a period of FI melting. The solid remainder is now dis-
connected from all of the wire surface except for the (100) facet on the far side of the side-on
view. Based on cross section, it seems there is still strong faceting present, despite the constant
melting at the solid-liquid interface. Snapshot C also shows the wire as it is about to break up.
Notice that the break-up occurs for a signiļ¬cantly thicker wire than seen for surface melting
earlier. After the break-up the solid remainder is melted while still maintaining a solid-liquid
frame all the way to snapshot D, which was one of the last frames before the completely molten
state.
The FI pathway was observed for wires of the same radius as those for surface melting, but
at a signiļ¬cantly higher Langevin parameter, generally with Ī» > 100ps. When the radius of the
wire was made smaller this pathway is no longer preferred over linear melting and we no longer
see the facet induced melting pathway.
Figure 4.7: Caloric curves (red) with liquid fraction (blue) for a simulation of facet induced
melting. These simulations have diļ¬€erent Langevin paramaters, Ī» = 120ps and Ī» = 200ps,
respectively.
4.3. COMPARISON TO THEORY 31
4.3 Comparison to theory
The simulations discussed in this chapter have managed to reproduce each of the pathways dis-
cussed in chapter 2. We identiļ¬ed some of the parameters governing which pathway is observed.
The parameters found in this chapter are summarised in table 4.1. This section will evaluate
the agreement between the two approaches and discuss the diļ¬€erences.
Preferred pathway Radius Langevin parameter, Ī» Caveat
Surface melting > 8a < 40 Shrinkage of Non-melting surfaces
Linear melting < 6a > 0, < 200 No pre-melting facets
facet induced melting > 8a > 80 Preceded by assymetric faceting
Bulk melting - - Not invesigated in this chapter
Table 4.1: Conditions for the diļ¬€erent melting pathways as observed in molecular dynamics
simulations. The lattice parameter, a is 4.050ĖšAfor FCC aluminium.
Note that the conditions on the radius and Langevin parameter are not closely ļ¬tting. In
between these boundary values we would often observe a strong competition between two path-
ways. Some examples of which are included in ļ¬gure 4.8. The competition between linear and
facet induced melting was already shown in ļ¬gure 4.4.
The deciding parameter between linear melting and both surface and facet induced melting
seems to be the radius, as we found only for the former of these comparisons in the free energy
evaluation. From table 4.1, one is tempted to estimate Rsl to be between 6 and 8 lattice spacings.
However, we have not been able to identify the nucleation cost of the diļ¬€erent pathways, and
closer investigation may reveal a higher critical radius. We have not been able to identify the
surface energy as an explicit parameter but have seen the important role of the crystal facets
and the Langevin parameter. These ļ¬ndings suggest that for facet induced melting the simple
model employed in chapter 2 is not suļ¬ƒcient.
Figure 4.8: Two examples of competing pathways, with atoms classiļ¬ed as solid (liquid) shown
in copper (blue). Snapshot A show part of the wire surrounded by melt while the other half
of the wire attempts to maintain a solid-vapour interface, showing the competition between
surface and facet induced melting. Snapshot B shows competition between surface and linear
melting as the solid phase becomes fully encapsulated in the melt after a cross-sectional cut.
32 CHAPTER 4. ANALYSIS AND RESULTS
Chapter 5
Discussion
In this study we investigated three diļ¬€erent pathways through which a nanowire can melt.
Chapter 2 compared the pathways by means of a free energy evaluation, calculating the free
energy cost and predicting a melting temperature for each wire. This approach made a number
of approximations which overlooked possibly important details. Foremost among these was the
shape of the nanowire. As we assume a cylindrical shape the faceting seen in the simulations
is not taken account of. This is vital in predicting the melting temperature as this depends
heavily on the phase interface energy densities. These energy densities depend on the crystal
facet at which the interface is located.
A more complete investigation may be made when an octagonal cylinder is assumed. The
sides of the octogonal can then be taken to be the diļ¬€erent crystal facets and one may acquire an
expression for the temperature in terms of the speciļ¬c interface energies of the diļ¬€erent facets.
Allowing experimental or simulation parameters to be substituted to predict the melting tem-
perature and pathway of a given wire. The cross section may also be allowed to change into
an irregular octagonal by letting the wetting surface to melt the sides of non melting surfaces.
This would make for signiļ¬cantly more complicated expressions but would model a number of
the interesting phenomena encountered in simulation more closely.
Another source of discrepancy between theory and simulation originated from ignoring the
nucleation cost of the diļ¬€erent pathways. Based on the free energy evaluation in chapter 2 the
linear pathway is always preferred below the bulk melting temperature. However, in chapter 4
we certainly saw both pathways. This is likely due to the linear melting transition needing to
overcome a greater energy barrier in the nucleation stage during which the simulation is still
heated. Thus if this barrier cannot be overcome before reaching a temperature at which surface
melting is possible we may see either pathway.
Future work may simulate wires with a diļ¬€erently orientated lattice. Of special interest
would be a wire with only (110) and (111) facets which should highlight any diļ¬€erence between
the two non-melting facets in Aluminium, after the (100) facet was examined here. Another
interesting case would be a geometry with only the non-melting surfaces (111) and (100) at the
surface, although this is not possible in a cylindrical nanowire.
It should also be noted that this study observed multiple wires in which more than one
pathway was followed. This co-occurrence of two pathways may be of interest to another study
as it changed the stability of the solid signiļ¬cantly, often melting signiļ¬cantly quicker in the
boundary areas. An example of co-occurrence of linear and facet induced melting is given in
section 4.2.2.
33
34 CHAPTER 5. DISCUSSION
Finally future studies could investigate similar behaviour for nanowires consisting of diļ¬€er-
ent materials. The authors believe that any FCC metal with two non-melting surfaces should
behave qualitatively similar to the aluminium studied here. However, a metal with a diļ¬€erent
crystal lattice or an FCC lattice with one or zero non-melting surfaces may behave quite diļ¬€er-
ently.
Chapter 6
Conclusions
Simulations using molecular dynamics were carried out to investigate melting phenomena in
metal nanowires. The main focus was on the diļ¬€erent pathways through which a nanowire can
melt. In this study surface melting was observed for wires with relatively large radius in canoni-
cal simulations. Facet induced melting was observed for similar wires treated micro-canonically.
Linear melting was observed for both canonical and micro-canonical systems for wires with
a diameter smaller than 12 lattice spacings, where the system ensemble was decided by the
Langevin parameter, Ī».
Another melting phenomena investigated, albeit less in-depth, was the faceting in melting
due to the presence of both wetting and non-melting facets at the nanowire surface. The pres-
ence of this state before the system seems to be associated with the melting pathway taken.
This state was not signiļ¬cantly present before linear melting, while the facets melt symmetri-
cally for surface melting and showed asymmetry for the facet induced melting.
An attempt was made to explain and predict the pathways using free energy evaluations for
a simple model of a nanowire. This managed to predict the dependence of the melting pathway
on the radius of the wire and the importance of the interface energy diļ¬€erence, āˆ†Ī³. However it
was discovered that an accurate theory describing these phenomena needs to take the faceting
dynamics and the cost of nucleation into account.
This work thus identiļ¬ed three diļ¬€erent pathways for the melting transition in a nanowire.
It also identiļ¬ed some of the main parameters inļ¬‚uencing the pathway taken by any particular
wire. This work is of importance as electronic components reach the nanoscale and are pushed
to their thermodynamic limits. Future research directions are outlined in chapter 5.
35
36 CHAPTER 6. CONCLUSIONS
Bibliography
1 Y. Kondo and K. Takayanagi, ā€œSynthesis and characterization of helical multi-shell gold
nanowires,ā€ Science, vol. 289, no. 5479, pp. 606ā€“608, 2000.
2 A. Xhang, G. Zheng, and C. Lieber, Nanowires, Building Blocks for Nanoscience and Nan-
otechnology, vol. 1 of NanoScience and Technology. Springer, 2016.
3 F. A. Lindemann, ā€œUeber die berechnung molekularer eigenfrequenzen,ā€ Phys. Z, vol. 11,
pp. 609ā€“612, 1910.
4 P.-L. Sun, S.-P. Wu, and T.-S. Chin, ā€œMelting point depression of tin nanoparticles embedded
in a stable alpha-alumina matrix fabricated by ball milling,ā€ Materials Letters, vol. 144,
pp. 142 ā€“ 145, 2015.
5 K. M. Ridings, ā€œThe breakup and melting of metal nanowires using molecular dynamics,ā€
University of Auckland, 2015.
6 A. M. Molenbroek and J. W. M. Frenken, ā€œAnharmonicity but absence of surface melting on
al(001),ā€ Phys. Rev. B, vol. 50, pp. 11132ā€“11141, Oct 1994.
7 j. W. M. Frenken and J. F. van der Veen, ā€œObservation of surface melting,ā€ Physical review
letters, vol. 54, pp. 827ā€“830, 1985.
8 A. M. Rodrguez, G. Bozzolo, and J. Ferrante, ā€œMultilayer relaxation and surface energies of
fcc and bcc metals using equivalent crystal theory,ā€ Surface Science, vol. 289, no. 1, pp. 100
ā€“ 126, 1993.
9 L. Vitos, A. Ruban, H. Skriver, and J. Kollr, ā€œThe surface energy of metals,ā€ Surface Science,
vol. 411, no. 12, pp. 186 ā€“ 202, 1998.
10 Y. Mishin, D. Farkas, M. J. Mehl, and D. A. Papaconstantopoulos, ā€œInteratomic potentials
for monoatomic metals from experimental data and ab initio calculations,ā€ Phys. Rev. B,
vol. 59, pp. 3393ā€“3407, Feb 1999.
11 D. Schebarchov and S. C. Hendy, ā€œSuperheating and solid-liquid phase coexistence in nanopar-
ticles with nonmelting surfaces,ā€ Phys. Rev. Lett., vol. 96, p. 256101, Jun 2006.
12 D. Schebarchov and S. C. Hendy, ā€œSuperheating and solid-liquid phase coexistence in nanopar-
ticles with nonmelting surfaces,ā€ Phys. Rev. Lett., vol. 96, p. 256101, Jun 2006.
13 J. A. Alonso, Structure and Properties of Atomic Nanoclusters. Imperial College Press, 2005.
14 M. Faraday, ā€œOn certain conditions of freezing water,ā€ Journal of the Franklin Institute,
vol. 5, pp. 283ā€“284, 1850.
15 L. Hui, F. Pederiva, B. L. Wang, and B. L. Wang, ā€œHow does the nickel nanowire melt?,ā€
Appl. Phys. Lett., vol. 86, 2005.
37
38 BIBLIOGRAPHY
16 E. Fermi, J. Pasta, S. Ulam, and M. Tsingou, ā€œStudies of nonlinear problems,ā€ Los Alamos,
1955.
17 B. Alder and T. Wainwright, ā€œStudies in molecular dynamics. i. general method,ā€ The Journal
of Chemical Physics, pp. 459 ā€“ 466, 1959.
18 D. Frenkel and B. Smit, Understanding Molecular Simulations: from algorithms to applica-
tions. Computational sciences series, Academic Press, Hartcourt, 2002.
19 J. E. Jones, ā€œOn the determination of molecular ļ¬elds. ii. from the equation of state of a
gas,ā€ Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering
Sciences, vol. 106, no. 738, pp. 463ā€“477, 1924.
20 M. S. Daw and M. I. Baskes, ā€œSemiempirical, quantum mechanical calculation of hydrogen
embrittlement in metals,ā€ Phys. Rev. Lett., vol. 50, pp. 1285ā€“1288, Apr 1983.
21 M. S. Daw and M. I. Baskes, ā€œEmbedded-atom method: Derivation and application to im-
purities, surfaces, and other defects in metals,ā€ Phys. Rev. B, vol. 29, pp. 6443ā€“6453, Jun
1984.
22 M. S. Daw, S. M. Foiles, and M. I. Baskes, ā€œThe embedded-atom method: a review of theory
and applications,ā€ Materials Science Reports, pp. 251ā€“310, 1993.
23 Y. Mishin, D. Farkas, M. J. Mehl, and D. A. Papaconstantopoulos, ā€œInteratomic potentials
for monoatomic metals from experimental data and ab initio calculations,ā€ Phys. Rev. B,
vol. 59, pp. 3393ā€“3407, Feb 1999.
24 S. Plimpton, ā€œFast parallel algorithms for short-range molecular dynamics,ā€ Journal of Com-
putational Physics, vol. 117, no. 1, pp. 1 ā€“ 19, 1995.
25 T. Schneider and E. Stoll, ā€œMolecular-dynamics study of a three-dimensional one-component
model for distortive phase transitions,ā€ Phys. Rev. B, vol. 17, pp. 1302ā€“1322, Feb 1978.
26 I. Snook, The Langevin and Generalised Langevin Approach to the Dynamics of Atomic,
Polymeric and Colloidal Systems. Amsterdam: Elsevier, 2007.
27 P. J. Steinhardt, D. R. Nelson, and M. Ronchetti, ā€œBond-orientational order in liquids and
glasses,ā€ Phys. Rev. B, vol. 28, pp. 784ā€“805, Jul 1983.
28 A. S. Keys, C. R. Iacovella, and S. C. Glotzer, ā€œCharacterizing structure through shape
matching and applications to self-assembly,ā€ Annual Review of Condensed Matter Physics,
pp. 263ā€“285, March 2011.

More Related Content

Viewers also liked

Introduction to mla citation format
Introduction to mla citation formatIntroduction to mla citation format
Introduction to mla citation format
Essay Corner
Ā 
Introduction to apa citation style
Introduction to apa citation styleIntroduction to apa citation style
Introduction to apa citation style
Essay Corner
Ā 
AHRC_Presentation
AHRC_PresentationAHRC_Presentation
AHRC_PresentationThom Aldershof
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬
Codingclub
Ā 
Infineon Technologies Company Presentation
Infineon Technologies Company PresentationInfineon Technologies Company Presentation
Infineon Technologies Company PresentationRobert Popescu, MBA
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬
Codingclub
Ā 
IKEA(MARKETING STRATEGY)
IKEA(MARKETING STRATEGY)IKEA(MARKETING STRATEGY)
IKEA(MARKETING STRATEGY)
Sunidhi Garg
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬
Codingclub
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜
Codingclub
Ā 
Ye Lin Oo
Ye Lin OoYe Lin Oo
Ye Lin OoYe Lin Oo
Ā 
[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼
[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼
[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼
Codingclub
Ā 

Viewers also liked (11)

Introduction to mla citation format
Introduction to mla citation formatIntroduction to mla citation format
Introduction to mla citation format
Ā 
Introduction to apa citation style
Introduction to apa citation styleIntroduction to apa citation style
Introduction to apa citation style
Ā 
AHRC_Presentation
AHRC_PresentationAHRC_Presentation
AHRC_Presentation
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 3ģ£¼ģ°Ø]źø°ģˆ˜ģ •ė ¬/ź³„ģˆ˜ģ •ė ¬/ė²„ķ‚·ģ •ė ¬
Ā 
Infineon Technologies Company Presentation
Infineon Technologies Company PresentationInfineon Technologies Company Presentation
Infineon Technologies Company Presentation
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 2ģ£¼ģ°Ø]ė³‘ķ•©ģ •ė ¬/ķ€µģ •ė ¬/ķž™ģ •ė ¬
Ā 
IKEA(MARKETING STRATEGY)
IKEA(MARKETING STRATEGY)IKEA(MARKETING STRATEGY)
IKEA(MARKETING STRATEGY)
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 1ģ£¼ģ°Ø]ģ‚½ģž…ģ •ė ¬/ė²„ėø”ģ •ė ¬/ģ„ ķƒģ •ė ¬
Ā 
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜
[ģ•Œź³ ė¦¬ģ¦˜ ģŠ¤ķ„°ė”” 4ģ£¼ģ°Ø]LCS ģ•Œź³ ė¦¬ģ¦˜
Ā 
Ye Lin Oo
Ye Lin OoYe Lin Oo
Ye Lin Oo
Ā 
[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼
[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼
[ģ•„ė‘ģ“ė…ø ģ›Œķ¬ģƒµ 1ģ°Ø] ģ•„ė‘ģ“ė…ø ģ†Œź°œ / LED / ķ”¼ģ—ģ”° ė¶€ģ € / ė²„ķŠ¼
Ā 

Similar to Honours_disertation_Thom_Aldershof

Parameter estimation in a non stationary markov model
Parameter estimation in a non stationary markov modelParameter estimation in a non stationary markov model
Parameter estimation in a non stationary markov model
Longhow Lam
Ā 
Molecular Electronics
Molecular ElectronicsMolecular Electronics
Molecular Electronics
Valentina Ferro
Ā 
Study of Energy by charles-delacourt
Study of Energy by charles-delacourtStudy of Energy by charles-delacourt
Study of Energy by charles-delacourt
Next Gen Group of Company
Ā 
Graphene Quantum Dots
Graphene Quantum DotsGraphene Quantum Dots
Graphene Quantum Dots
Ricardo Camacho SƔnchez
Ā 
morten_bakkedal_dissertation_final
morten_bakkedal_dissertation_finalmorten_bakkedal_dissertation_final
morten_bakkedal_dissertation_finalMorten Bakkedal
Ā 
Protdock - Aatu Kaapro
Protdock - Aatu KaaproProtdock - Aatu Kaapro
Protdock - Aatu KaaproSwapnesh Singh
Ā 
S.Denega_thesis_2011
S.Denega_thesis_2011S.Denega_thesis_2011
S.Denega_thesis_2011Sergii Denega
Ā 
Photophysics of dendrimers colombi
Photophysics of dendrimers   colombiPhotophysics of dendrimers   colombi
Photophysics of dendrimers colombi
Giorgio Colombi
Ā 
thesis
thesisthesis
thesisYishi Lee
Ā 
Testo
TestoTesto
MasterThesis-YimingHu
MasterThesis-YimingHuMasterThesis-YimingHu
MasterThesis-YimingHuYiming Hu
Ā 
Dynamics of cold atoms in moving optical lattices (Version Dek8)
Dynamics of cold atoms in moving optical lattices (Version Dek8)Dynamics of cold atoms in moving optical lattices (Version Dek8)
Dynamics of cold atoms in moving optical lattices (Version Dek8)Nadal Sarkytbayev
Ā 
1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern Physics1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern Physics
Laurie Smith
Ā 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf
BEATRIZJAIMESGARCIA
Ā 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf
BEATRIZJAIMESGARCIA
Ā 

Similar to Honours_disertation_Thom_Aldershof (20)

Parameter estimation in a non stationary markov model
Parameter estimation in a non stationary markov modelParameter estimation in a non stationary markov model
Parameter estimation in a non stationary markov model
Ā 
Molecular Electronics
Molecular ElectronicsMolecular Electronics
Molecular Electronics
Ā 
Study of Energy by charles-delacourt
Study of Energy by charles-delacourtStudy of Energy by charles-delacourt
Study of Energy by charles-delacourt
Ā 
Graphene Quantum Dots
Graphene Quantum DotsGraphene Quantum Dots
Graphene Quantum Dots
Ā 
thesis
thesisthesis
thesis
Ā 
morten_bakkedal_dissertation_final
morten_bakkedal_dissertation_finalmorten_bakkedal_dissertation_final
morten_bakkedal_dissertation_final
Ā 
Protdock - Aatu Kaapro
Protdock - Aatu KaaproProtdock - Aatu Kaapro
Protdock - Aatu Kaapro
Ā 
S.Denega_thesis_2011
S.Denega_thesis_2011S.Denega_thesis_2011
S.Denega_thesis_2011
Ā 
Corina_Barbu-Dissertation
Corina_Barbu-DissertationCorina_Barbu-Dissertation
Corina_Barbu-Dissertation
Ā 
Thesis
ThesisThesis
Thesis
Ā 
Photophysics of dendrimers colombi
Photophysics of dendrimers   colombiPhotophysics of dendrimers   colombi
Photophysics of dendrimers colombi
Ā 
thesis
thesisthesis
thesis
Ā 
Testo
TestoTesto
Testo
Ā 
MasterThesis
MasterThesisMasterThesis
MasterThesis
Ā 
MasterThesis-YimingHu
MasterThesis-YimingHuMasterThesis-YimingHu
MasterThesis-YimingHu
Ā 
Dynamics of cold atoms in moving optical lattices (Version Dek8)
Dynamics of cold atoms in moving optical lattices (Version Dek8)Dynamics of cold atoms in moving optical lattices (Version Dek8)
Dynamics of cold atoms in moving optical lattices (Version Dek8)
Ā 
MyThesis
MyThesisMyThesis
MyThesis
Ā 
1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern Physics1000 Solved Problems In Modern Physics
1000 Solved Problems In Modern Physics
Ā 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf
Ā 
1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf1000-solved-problems-in-modern-physics.pdf
1000-solved-problems-in-modern-physics.pdf
Ā 

Honours_disertation_Thom_Aldershof

  • 1. Melting Phenomena in Aluminium Nanowires Thomas Simon Aldershof Department of Physics The University of Auckland Supervisor: Professor Shaun Hendy Dissertation submitted in partial fulļ¬lment of the requirements for the degree of BSc(Hons) in Physics, The University of Auckland, 2016.
  • 2.
  • 3. Abstract The melting transition of a cylindrical, metal nanowire can progress through a number of diļ¬€er- ent pathways. The three pathways identiļ¬ed in this study are surface melting, where a molten layer envelops a solid core, linear melting, where the wire melts in the longitudinal direction and fact induced melting, where the wire melts from the surface maintaining at least one solid- vapour interface. The diļ¬€erent pathways are compared both through a simple model based on free energy considerations, developed here for each pathway, and through molecular dynamics simulations to identify the main parameters inļ¬‚uencing what pathway is taken From the free energy models, the wire radius and the diļ¬€erence in the energy density of the diļ¬€erent phase interfaces, āˆ†Ī³ are found as the deciding parameters. A discussion on the diļ¬€erence in nucleation cost and importance of non-melting crystal facets is also included. In the simulations, Face Centric Cubic, aluminium nanowires undergo the diļ¬€erent melting pathways using molecular dynamics. From these the radius is conļ¬rmed to be an important parameter deciding between the diļ¬€erent pathways. The interface energy of the crystal facets on the surface of the wire are also identiļ¬ed as important parameters. The latter was accompanied by shrinkage of the non-melting surfaces through the melting of the wetting surfaces, which the free energy model does not take account of. The observed melting pathway was also sensitive to the strength of the coupling of the heat bath, which decided whether the system acts more canonically or micro-canonically. The conclusion of this work is that linear melting occurs for wires with a diameter of at most 12 lattice spacings. Surface melting and facet induced melting occurred for thicker wires treated more canonically or micro-canonically respectively. An accurate theory predicting the onset of the melting transition needs to take the faceting of the wire into account as the non-melting surfaces play an important role. 1
  • 4. 2
  • 5. Acknowledgements First and foremost I would like to thank my supervisor Professor Shaun Hendy for suggesting this subject and allowing me to explore the interesting world of post-graduate research. I would like to thank him for ensuring the regular meetings were not easily bypassed, even if my results were meagre. He is also responsible for introducing me to condensed matter through the third year course he taught together with Associate Professor Malcolm Grimson and Dr. Geoļ¬€ Will- mott, who also have my gratitude. I would further like to thank Dr. Geoļ¬€ Willmott, as well as Professor Geoļ¬€ Austin for introducing me to academic research through the summer research scholarships they oļ¬€ered me in successive summers. Next I need to thank Kannan Ridings, who always seemed to be available to answer my many questions, whether they were technical or theoretical. I am positive he spend more time on my work than his own on some days and I would like to thank him for this dedication. The code he provided me at the start of this project allowed me to hit the ground running and without him this work would be far less complete. Finally I would like to thank my lecturers, family, friends and peers as they made this year the most enjoyable, albeit the most demanding year of my university career yet. Special mention among these is my partner, Rebecca Bub, who always made an eļ¬€ort to help and understand me in this project and gave her support when I needed it most, whether she was near or far. 3
  • 6. 4
  • 7. Contents Abstract 1 1 Introduction 7 2 Theory 9 2.1 Nucleation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.2 Non-melting surfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2.3 Wire geometries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.3.1 Free energy evaluation of surface melting . . . . . . . . . . . . . . . . . . 11 2.3.2 Free energy evaluation of linear melting . . . . . . . . . . . . . . . . . . . 12 2.3.3 Free energy evaluation of facet induced melting . . . . . . . . . . . . . . . 13 2.4 Preferred melting pathway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3 Methodologies 17 3.1 Molecular dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 3.1.1 Initial state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 3.1.2 Embedded Atom Method Potentials . . . . . . . . . . . . . . . . . . . . . 18 3.2 Numerical considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 3.2.1 Simulation dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 3.2.2 Time integration and the Langevin thermostat . . . . . . . . . . . . . . . 20 3.3 Post processing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 3.3.1 Steinhardt paramaters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 4 Analysis and results 23 4.1 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 4.1.1 Initialistion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 4.1.2 Simulation parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 4.2 Observed melting pathways . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 4.2.1 Observation of surface melting . . . . . . . . . . . . . . . . . . . . . . . . 25 4.2.2 Observation of linear melting . . . . . . . . . . . . . . . . . . . . . . . . . 27 4.2.3 Observation of facet induced melting . . . . . . . . . . . . . . . . . . . . . 29 4.3 Comparison to theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 5 Discussion 33 6 Conclusions 35 5
  • 9. Chapter 1 Introduction In recent years nanowires have been of increasing interest in both academic and industrial ļ¬elds such as opto-electronics, nano computing and integrated circuits. The one-dimensional nature and the high surface to volume ratio of these wires cause them to diļ¬€er from both bulk materials and nanoclusters in a number of diļ¬€erent ways, prompting some to call metallic nanowires mag- ical.1 Due to the potential of nanowires in nanoscale electronics, research has focused on their electric properties. However, the potential range of applications is huge and extends to extreme thermal conditions. Thus an extensive investigation into the thermal behaviour is warranted. This document will focus on the melting phenomenon of a metal nano-wire. In particular, the diļ¬€erent pathways through which a metal nanowire can melt and the parameters inļ¬‚uencing this process. In previous work metal nanowires have been found to display several novel melting phenomena and processes.2 This work seeks to determine the processes through which a metal nanowire can melt and to investigate the parameters that inļ¬‚uence the wireā€™s preference for one pathway over an other. Figure 1.1: Diļ¬€erent pathways of melting investigated in this study. Dark colouring indicates a particle classiļ¬ed as solid state while light colouring indicates a liquid. Solid lines connect the geometries of the main pathways and dotted lines show possible intermediate states. The initial state is taken to be the solid wire, which can melt through, top to bottom, surface, linear or facet induced melting to the fully liquid sate. The wire may go through a state with premelted facets ļ¬rst (left) or break up into spheres (right) before fully melting. 7
  • 10. 8 CHAPTER 1. INTRODUCTION The pathways discussed in this document are given by ļ¬gure 1.1, deļ¬ning the diļ¬€erent melt- ing geometries encountered. Notice that although the facet pre-melting geometry on the top-left and the wire break-up on the top-right are encountered in this study, the focus is on the three melting pathways shown in the centre. Surface melting, where the wire is completely encap- sulated with a liquid layer. In the linear melting pathway the wire melts in the longitudinal direction and never becomes fully encapsulated by the melt. Finally, in facet induced melting the wire melts from the surface but maintains at least one crystal facet as a solid-vapour inter- face. The ļ¬nal state of the system is always the fully melted case, although it may go through a geometry where the solid wire has broken up into multiple spheres within the molten wire. Chapter 2 summarises some of the theory vital to this study. First it revises classical nu- cleation theory. Then it evaluates how the interface free energy densities inļ¬‚uence melting on diļ¬€erent surfaces. With these two elements we then perform free energy evaluations for the diļ¬€erent pathways. At the time of writing we are not aware of a similar evaluation for linear or facet induced melting in the literature. We then combine these calculations in an attempt to identify the preferred melting pathway and the parameters inļ¬‚uencing this preference. This again has not been found in literature. Chapter 3 covers the numerical methods used in this study. The chapter begins with a discussion of molecular dynamics. This includes a justiļ¬cation of the potential used and lays out the initial state of the system. It also lists the choices we made to obtain a smooth and physical simulation that is computationally manageable. Finally it reviews the method used to distinguish solid and liquid atoms. Chapter 4 provides an exert of the simulations run showing a representative simulation for each of the pathways. It also shows the initial state of the system and tests some of the physical aspects of the simulations. Chapters 5 and 6 provide a summary of our result and point the way for future research.
  • 11. Chapter 2 Theory In this chapter we evaluate some of the theory associated with the melting phenomena and we attempt to predict the occurrence of the diļ¬€erent pathways. We will ļ¬rst review classical nucleation theory and introduce some relations used in the later sections and then discuss non- melting facets of a crystal lattice. After this is set-up we engage in a free energy evaluation of the diļ¬€erent melting pathways. This method has been tested for surface melting but is new for linear and facet induced melting. Finally we compare the relations obtained for the diļ¬€erent pathways to identify parameters that may aļ¬€ect the preference of one melting pathway over the other. 2.1 Nucleation Theory An interesting and much-studied aspect of thermodynamics are phase transitions, which is de- ļ¬ned as a qualitative change in the properties of an object or material. The most commonly known examples are melting and vaporisation, the ļ¬rst of which is under discussion in this document. The onset of this melting transition, as other types of transitions, is described by classical nucleation theory (CNT). According to to this theory the melting transition is always initiated at a localised nucleation site. At this point a stable cluster of liquid needs to be formed from which the melting is then promoted. One can make the distinction between two types, homogeneous and heterogeneous nucleation. The former occurs when the nucleation site is within the bulk of the material and away from any impurity. The latter occurs when the nucleation is encouraged by an impurity of boundary of the material. In the nanowire, nucleation will always occur on or near the surface and will thus undergo heterogeneous nucleation. Both types of nucleation and the following melting transition describe the formation of a liquid cluster of a certain critical size,l from which the melting transition progresses sponta- neously. For the nucleation to occur the energetic cost of forming the extra phase interface is less than the energy gained from the formation of the volume of liquid. Classically the growth of these nucleation clusters is described by atomic vibrations as ļ¬rst proposed by Lindemann in 1910. These atomic vibrations increase in amplitude with increasing temperature. However, this description breaks down for a nanosized object. This is due to the low volume to surface ratioand thus the relatively high number of less restricted atoms on the surface of the wire.3 9
  • 12. 10 CHAPTER 2. THEORY The lower energy state results in a depression of the melting temperature of a nano-sized object. This phenomenon has been extensively veriļ¬ed in the literature for many diļ¬€erent systems and is still an active area of research.2,4 When comparing the liquid and solid state of a cylindrical nanowire which, for simplicity, maintains its shape and density one can ļ¬nd section 2.1 for the melting temperature.5 Tm = Tc 1 āˆ’ 2(Ī³sv āˆ’ Ī³lv) ĻRLm (2.1) Where the temperature dependence was introduced using the relation in equation (2.2). Note also that the melting temperature is lower than the bulk melting temperature whenever Ī³sv < Ī³lv . fs āˆ’ fl = āˆ’ĻLm(1 āˆ’ T Tc ) (2.2) 2.2 Non-melting surfaces. At a nano-scale surface the orientation of the lattice with respect to this surface becomes im- portant. The direction the surface is cut from the bulk lattice aļ¬€ects a number of properties, such as the smoothness, packing ratio and surface energy density of the ļ¬rst layer of atoms. The last of these is of the most interest in this study. In this document, we will specify this orientation by Miller indices. The surface energy density is speciļ¬c to the type of interface the surface is part of. Numer- ous studies have tried to determine the surface energies of pure metals with varying success. Experimentally this can often only be found for the solid-vapour energy, Ī³sv, although there is a large range of reported values. Additionally as the lattice is only present in the solid state, the liquid-vapour interface, Ī³lv, often has an accepted value independent of the lattice orientation. However, the solid-liquid interface energy density, Ī³lv, is often diļ¬ƒcult to determine as it is not often a stable state. Even if this is an equilibrium state, atoms will continuously melt and crystallise and therefore will often not have a well-deļ¬ned facet. We can deļ¬ne the free energy density diļ¬€erence.6ā€“10 āˆ†Ī³ = Ī³sv āˆ’ Ī³sl āˆ’ Ī³lv (2.3) This quantity gives the free energy cost of replacing a solid-vapour interface with a solid- liquid and a liquid-vapour interface per unit area. In other words, the cost to have a layer of melt form on the solid surface. Like the interfacial energy densities, this quantity depends on the surface orientation and had been found for some crystal facets of some metals. However, when āˆ†Ī³ > 0, a liquid layer will form on this facet and measurements encounter the same problems as with Ī³sl. In this case āˆ†Ī³ often can not be determined beyond the negative sign.6 The case where āˆ†Ī³ > 0 for only some facets of the surface is worth investigating by itself. As some facets may have this condition satisļ¬ed and thus the form a liquid layer (called wetting surfaces), while others still have positive āˆ†Ī³ and thus maintain a solid-vapour interface. The latter quickly gained the name of non-melting surfaces and have been found to cause some in- teresting phenomena.11ā€“13 In this study, they are important in the occurrence of facet induced melting, as outlined in section 2.3.3. They also cause the facet melted geometry in ļ¬gure 1.1. In Aluminium the (110) facet is a wetting surface, while the (100) and (111) facets are non-melting surfaces. We thus expect the facet melted geometry to have (110) facets melted with (100) or (111) intact. Due to the geometry of a nanowire, we cannot have all three facets as part of the surface.
  • 13. 2.3. WIRE GEOMETRIES 11 2.3 Wire geometries Figure 2.1 shows some of the possible geometries a nanowire may take during the melting tran- sition. We can see that for each of the displayed geometries, the melting transitions can be described by only one changing variable, r, l or h respectively. The following is a canonical free energy calculation based on existing models for surface melting.5,11 They are adjusted for the new pathways as we believe these calculations have not been done before ļ¬r linear and facet induced melting. In section 2.4 we will compare the results of these calculations to identify the parameters inļ¬‚uencing the preferred pathway. (a) Surface melting (b) Linear melting (c) Facet induced melting Figure 2.1: Abstracted wire geometries that can occur during the melting process with some parameters shown. The solid state is dark, while the liquid state is lighter gray. 2.3.1 Free energy evaluation of surface melting When an object undergoes surface melting a thin layer of liquid forms on the solid-vapour inter- face before reaching the bulk melting temperature. This process was ļ¬rst proposed by Micheal Faraday to describe the fusion of ice particles. The transition has later been observed in electron scattering experiments that also saw the ļ¬rst nonmelting surfaces discussed before.7,14 When surface melting occurs, the solid-vapour interface is replaced by a solid-liquid and a liquid-vapour interface, separated by a thin layer of liquid. As this transition will only spon- taneously occur when it lowers the free energy we need Ī“sv + Fs > Ī“sl + Ī“lv + Fl + Fs. The thickness of the layer, R āˆ’ r, changes continuously from R āˆ’ r = 0 from the unmelted case. Thus Fl is initially negligible with no signiļ¬cant change in Fs for very small R āˆ’ r. This gives rise to the condition also found in section 2.2, āˆ†Ī³ > 0.15 We assume the solid core retains the same shape as the original nanowire with radius r < R. The liquid layer thus is a cylindrical shell with inner radius r and outer radius R. We can thus write down the free energy as below, again taking account of the crystal facets where necessary. Fsurf (r) = Ļ€r2 Lfs + Ļ€(R2 āˆ’ r2 )Lfl + 2Ļ€RLĪ³lv + 2Ļ€rL Ī³sl + āˆ†Ī³e āˆ’Rāˆ’r Ī¾ (2.4)
  • 14. 12 CHAPTER 2. THEORY We assume the nanowire to be at equilibrium at all times. This allows us to ļ¬nd the transition temperature by ļ¬nding the extrema. Thus we need dF dr = 0. dFsurf dr = 2Ļ€L r (fs āˆ’ fl) + Ī³ls + āˆ†Ī³ 1 + r Ī¾ e āˆ’Rāˆ’r Ī¾ = 0 (2.5) Where the last term represents the energy cost of nucleation, with Ī¾, a critical coherence length of the liquid layer speciļ¬c to the metal. As before we use equation (2.2) on the unmelted, initial state, i.e. r = R. Thus we obtain the onset temperature of surface melting, equation (2.6), in terms of the bulk melting temperature, Tc. Tsurf = Tc 1 āˆ’ Ī³sv āˆ’ Ī³lv ĻLmR 1 + R Rc (2.6) Where we deļ¬ned the critical radius Rc = Ī³svāˆ’Ī³lv āˆ†Ī³ Ī¾ and used equation (2.3) for āˆ†Ī³. It is easy to see that as long as Ī³sv > Ī³lv and R Rc > āˆ’1 the onset temperature of surface melting is below the bulk melting temperature. To ensure this is an energetically stable state we evaluate the second derivative of the free energy at r = R, T = Tsurf . d2Fsurf dr2 = 2Ļ€L āˆ†Ī³ Ī¾ R2 + 2Ī¾ R āˆ’ 1 2 Rc (2.7) This is greater than zero for āˆ†Ī³ > 0 and R > Rc. This ensures that we have been working in a free energy minimum and surface melting can be an energetically favourable state. 2.3.2 Free energy evaluation of linear melting Linear melting occurs when a thin cross sectional disk of width l melts before reaching the bulk melting temperature. As the melting progresses l increases from 0 to L, the length of the wire. The liquid-vapour interface and the melted volume grow linearly with l as the solid- vapour interface and solid volume decrease similarly in size. The liquid-solid interface remains constant and is given by the sum of the two cross sectional areas at either side of the melted disk. Returning to nucleation theory it is easy to see that this method has a signiļ¬cantly higher energy barrier than surface melting. Firstly the new interface is mostly away from the sur- face of the nanowire. This makes the initial nucleation more energy expensive. Secondly this type of melting does not immediately replace the solid-vapour interface, as we did above and therefore maintains the, presumably, more expensive interface for longer. We will however still employ a similar free energy analysis of this pathway. Through similar geometric arguments as we employed in section 2.3.1 the free energy of the linear melting geometry can be found. Equation (2.8) gives this as a function of the melting parameter, l. Flin(l) = Ļ€R2 (L āˆ’ l)fs + Ļ€R2 lfl + 2Ļ€R(L āˆ’ l)Ī³sv + 2Ļ€RlĪ³lv + 2Ļ€R2 Ī³sl + āˆ†Ī³e āˆ’ l Ī¾ (2.8) Again the nanowire is assumed to be in an equilibrium state. Finding the minima of the free energy will thus enable us to ļ¬nd the transition temperature again. Note that this does assume the nanowire has not undergone surface melting, just as in the previous section we implicitly assumed the wire had not undergone linear melting. dFlin dl = Ļ€R (fl āˆ’ fs) R + 2 (Ī³lv āˆ’ Ī³sv) āˆ’ 2āˆ†Ī³ R Ī¾ e āˆ’ l Ī¾ = 0 (2.9)
  • 15. 2.3. WIRE GEOMETRIES 13 As in the previous section, the last term represents the energy barrier due to the nucleation process. We again set the wire to the initial, unmelted state, l = 0, to obtain the transition temperature. Tlin = Tc 1 āˆ’ 2 Ī³sv āˆ’ Ī³lv ĻLmR 1 + R Rc (2.10) Where Rc = Ī³svāˆ’Ī³lv āˆ†Ī³ Ī¾, as before. Again this transition temperature is below the bulk temperature for Ī³sv > Ī³lv and R Rc > āˆ’1. For this to be an energetically accessible state the above parameters need to correspond to a minimum in the free energy. That is, d2F dl2 > 0 at l = 0, T = Tlin. d2Flin dl2 = 2Ļ€āˆ†Ī³ R2 Ī¾2 (2.11) This is greater than zero for āˆ†Ī³ > 0 and R > Rc. Thus linear melting is an energetically accessible state. 2.3.3 Free energy evaluation of facet induced melting Facet induced melting (FI) is an eļ¬€ect of the interplay between wetting and non-melting surfaces. While the non-melting surfaces of the nanowire remain unmelted, the wetting facets melt. This allows wetting facet to grow in size. At some point it may be favourable to maintain only one facet of solid-vapour interface while relinquishing the others in favour of maximizing the solid-liquid interface of the wetting facets. This geometry is shown in 2.1c. Once again the free energy can be written down using geometrical arguments. Fpw(h) = 2RL Ī±Rfs + Ļ€ 2 āˆ’ Ī± Rfl + Ī±Ī³sv + (Ļ€ āˆ’ Ī±) Ī³lv + Ī± Ī³sl + āˆ†Ī³e āˆ’2Rāˆ’h Ī¾ (2.12) Where, cos(Ī±) = Rāˆ’h R gives half the angle the solid-vapour interface spans. As in the previous sections the ļ¬rst derivative of the free energy is set to zero to ļ¬nd the minima of this system. Note that dĪ± dh = (h(2R āˆ’ h))āˆ’0.5 . dFpw dh = 2RL h(2R āˆ’ h) R(fs āˆ’ fl) + 2Ī³sl + āˆ†Ī³ 1 + 2Ī± Ī¾ h(2R āˆ’ h) + 1 e āˆ’2Rāˆ’h Ī¾ = 0 (2.13) We use equation (2.2) to eliminate the volumetric terms and set the system to the onset of melting, h = R, to obtain the temperature according to this free energy evaluation. Tpw = Tc 1 āˆ’ 2 Ī³sv āˆ’ Ī³ls ĻLmR 1 + Ļ€ 2 R Rc (2.14) With Rc as in section 2.3.1. Note that this is less than the bulk melting temperature, Tc, for Ī³sv > Ī³lv and Ļ€R 2Rc > āˆ’1 Again we ensure that the parameters above describe a minimum in the free energy by evaluating the second derivative with respect to the melting parameter at h = R. d2Fpw dh2 = 4 L Ī¾ 2 + Ļ€ R Ī¾ āˆ†Ī³ (2.15) This has the sign of āˆ†Ī³ and thus we have a minimum for āˆ†Ī³ > 0. Thus from these free energy calculations we ļ¬nd that the FI geometry can be an energetically favourable state.
  • 16. 14 CHAPTER 2. THEORY 2.4 Preferred melting pathway In this section we aim to determine the preferred pathway of melting based on the free energy calculations in the previous section. To do this we will compare the melting temperatures found in equations (2.6), (2.10) and (2.14). We again ignore the dependence of the surface energy density on the crystal facet. Otherwise one would use the found values of Ī³sv and āˆ†Ī³ for each facet and the facet-independent value for Ī³lv to ļ¬nd the diļ¬€erent surface energies. First, we compare surface and linear melting by deļ¬ning āˆ†Tsl, the diļ¬€erence in the onset temperatures of the two pathways. āˆ†Tsl = Tsurf āˆ’ Tlin = Ī³sv āˆ’ Ī³lv ĻLmR 1 + R Rc Tc (2.16) Thus the onset of surface melting precedes the onset of linear melting when equation (2.16)is negative. As Ī³sv > Ī³lv and R > 0 we deduce that surface melting is preferred when 1 + R Rc < 0. Inversely the linear melting pathway is taken when 1+ R Rc > 0. This condition for surface melting is the inverse of that between surface and the bulk melting temperature, 2.3.1. That is, linear melting always precedes surface melting according to this canonical calculation. However, due to the expensive nucleation stage, we can still observe surface melting. This would occur when the energy cost of nucleation outweighs the energy gain from melting. We can summarise the found conditions below, while deļ¬ning the transition radius Rsl, where a lower onset temperature indicates a preference for this pathway, based on these free energy evaluations. Rsl = āˆ’Rc = Ī³lv āˆ’ Ī³sv āˆ†Ī³ Ī¾ R < Rsl ā‡’ Tlin < Tsurf < Tc R > Rsl ā‡’ Tc < Tsurf < Tlin (2.17) These conditions may be satisļ¬ed for some facets of the nanowire while not on the other. This suggests that surface melting will have its initial nucleation site on a diļ¬€erent facet than linear melting. We can similarly compare surface melting and FI melting. Deļ¬ning āˆ†Tsf , the diļ¬€erence in the onset temperatures of surface and FI melting. āˆ†Tsf = Tsurf āˆ’ Tpw = Ī³sv āˆ’ Ī³lv ĻLmR 1 + (Ļ€ āˆ’ 1) R Rc Tc (2.18) The onset of surface melting precedes the onset of FI melting when āˆ†Tsf < 0. Thus when 1 + (Ļ€ āˆ’ 1) R Rc < 0 , the wire melts through the surface melting pathway. This now culminates into the following conditions, deļ¬ning the transition radius Rsf . Rsf = āˆ’ Rc Ļ€ āˆ’ 1 = āˆ’ Ī³sv āˆ’ Ī³lv (Ļ€ āˆ’ 1)āˆ†Ī³ Ī¾ R < Rsf ā‡’ Tlin < Tsurf R > Rsf ā‡’ Tsurf < Tlin (2.19)
  • 17. 2.4. PREFERRED MELTING PATHWAY 15 Note that these conditions do not include the bulk melting temperature. As Rsf is smaller than Rsl and is smaller than the bulk melting temperature, it is possible to have the transition between the two pathways before reaching the bulk melting temperature. However to actually predict at what value this occurs we need values for the free energy densities of the interfaces. These quantities have had a large range of reported values and at present time it is not possible to predict the melting pathway other than for a speciļ¬c simulation where these can be extracted from the potential. Finally we can compare the linear and FI melting pathways. Deļ¬ning āˆ†Tlf , the diļ¬€erence in onset temperature between these two pathways. āˆ†Tlf = Tlin āˆ’ Tpw = (Ļ€ āˆ’ 2) Ī³sv āˆ’ Ī³lv ĻLmRc Tc = (Ļ€ āˆ’ 2) āˆ†Ī³ ĻLmĪ¾ Tc (2.20) Notice that unlike equations (2.16) and (2.18) this is constant with respect to R. Note that the sign of āˆ†Tlf is the sign of āˆ†Ī³. Thus linear melting is preferred for negative āˆ†Ī³ and inversely, FI melting is preferred when āˆ†Ī³ > 0. We are unable to deļ¬ne a Rlf as we did for the previous two comparisons, however we can still write down the following set of conditions. Where again, a lower onset temperature indicates a preference for the associated melting pathway. āˆ†Ī³ < 0 ā‡’ Tlin < Tpw āˆ†Ī³ > 0 ā‡’ Tpw < Tlin (2.21) Combining the three conditions in equations (2.17), (2.19) and (2.21) we can summarise the results of the last two sections in table 2.1. With Rsl and Rsf as before. Preferred pathway Radius āˆ†Ī³ Caveat Surface melting > Rsl > 0 Low nucleation cost Linear melting < Rsl > 0 High nucleation cost facet induced melting < Rsf > 0 Preceded by faceting Bulk melting > Rc < 0 Table 2.1: Conditions for the diļ¬€erent melting pathways as obtained from the free energy evaluations in sections 2.3 and 2.4
  • 18. 16 CHAPTER 2. THEORY
  • 19. Chapter 3 Methodologies This section describes the numerical methods used. First, we take a look at molecular dynamics and the choice of some parameters within this method. We evaluate the elements of a single simulation to provide reproducibility. This includes a discussion of the computational consider- ations and simulation parameters. We ļ¬nish this chapter discussing the Steinhardt parameters and their role in the solid-liquid classiļ¬cation. 3.1 Molecular dynamics The model we employed in chapter 2 assumes an approximately cylindrical shape. When the radius of the wire is suļ¬ƒciently small, faceting of the nanowire surface renders this assumption invalid. This puts a lower limit on R, the initial radius of the solid wire. To have a one di- mensional nano-scale object we need the radius to be less than 100 nm with the length greater than 100nm. Time integration with a timescale of the order of nanoseconds is required to have suļ¬ƒcient resolution of the melting process. These constraints ļ¬t best within a molecular dynamics framework. Alternatives would be Monte Carlo or electronic structure methods for longer timescales or smaller systems respectively. The ļ¬rst study that can be classiļ¬ed as Molecular Dynamics (MD) was published posthu- mously by Enrico Fermi.16 The name ā€molecular dynamicsā€ was coined by Alder and Wain- wright 4 years later.17 It is now recognised as an eļ¬€ective method to model many-body systems, which can be larger and larger as computing power increases, where Fermi started with a mere 64 particles. MD evaluates Newtonā€™s laws on one particle due to a potential created by the particles in the close environment of this particle. It ignores electronic structure and other quantum eļ¬€ects, which can be justiļ¬ed by using the Born-Oppenheimer approximation and assuming that the electrons are in equilibrium with the nucleus.18 3.1.1 Initial state As outlined in section 3.2 the number of atoms needs to be approximately constant. To obtain the initial state, these atoms were arranged in an FCC lattice from which we cut a cylinder along the (100) direction to obtain a nanowire. An example of this initial wire is shown in ļ¬gure 3.1. This wire is placed in a unit cell with periodic boundary conditions in the z-direction such that we obtain an inļ¬nitely long nanowire. We thus eliminate any eļ¬€ects due to the ļ¬nite length of the wire and look solely at the nano-dimensions of our wire. For similar reasons there are solid boundaries in the x and y directions. This ensures there are no interactions in the radial 17
  • 20. 18 CHAPTER 3. METHODOLOGIES Figure 3.1: Showing the intial state of a nanowire as cut from a perfect FCC lattice. The left hand ļ¬gure shows the the (111) plane. The right hand ļ¬gure faces the (001) plane with the (100) and (110) planes marked. direction outside the nanowire. 3.1.2 Embedded Atom Method Potentials As mentioned before classic MD does not take quantum eļ¬€ects in account. In these methods intermolecular forces are modelled by hard spheres or more commonly a Lennard-Jones poten- tial.19 This is radially symmetric, pairwise potential. The equations of motion can now be obtained from the potential through classical mechanics. Which gives the force between any pair of atoms as described in equation (3.1). Where the we substituted the Lennard-Jones potential, Ī¦LJ . F(r) = āˆ’ Ī¦LJ = 2 rm r 6 āˆ’ rm r 12 (3.1) Although the LJ potential is an improvement over the hard sphere model it is still merely a pair potential model, based only on the inter-atomic separation. The Embedded-Atom Method (EAM) potentials were developed by Daw and Baskes to model mesoscopic phenomenon such as micro-fractures in metals. The pair potential had insuļ¬ƒcient detail, while methods explicitly taking electrons into account could not handle the meso scale required.20,21 In EAM methods atoms are embedded into a ā€œJeliumā€, that is, they are embedded into a potential deļ¬ned not only by the other atoms but also by the electron cloud tied to these atoms. To avoid solving the SchrĀØodinger equation they assumed that each atom experiences a locally uniform electron density, deļ¬ned by the atoms already in the direct environment of the introduced atom. This electron density is determined empirically. That is, the model is set up implicitly including this electron density after which it is ļ¬tted to experimentally known parameters. This makes the EAM potentials semi-empirical. It also makes EAM potentials signiļ¬cantly more accurate than pair potentials while maintaining relatively cheap numerics. EAM potentials thus need to be tailored to the material under investigation. Furthermore, for any material there is a number of diļ¬€erent potential each suited to a diļ¬€erent purpose.22
  • 21. 3.2. NUMERICAL CONSIDERATIONS 19 In this study we investigate phenomena that happen close to the melting point and we thus hope to model the melting process accurately. For this we need a potential that closely ļ¬ts the melting temperature and the latent heat per atom. As these seem to be key parameters from our study in chapter 2. The potential we chose also accurately models the lattice contraction or relaxation at the surface. This lattice property can be monitored experimentally through ion scattering and has been shown to be closely related to the surface energies of the diļ¬€erent crystal facets, another important parameter from our theoretical studies.8,23 All simulations were run on the PAN cluster, maintained by NESI. This allowed us to run the simulations on parallel cores to maximise the available computing power. The LAMMPS software, used to run the simulations, quite naturally adopts to parallel computing by dividing the unit cell into smaller sections. Each of these sections then is assigned to a core. These cores then only need to communicate about the dynamics close to the box boundaries. LAMMPS is an open-source software package developed by Sandia National Laboratories and currently maintained by this institution and Temple University. It is speciļ¬cally designed for molecular dynamics and is compatible with the above potential.24 3.2 Numerical considerations In any numerical simulations one has to make a choice between computational eļ¬ƒciency and physical completeness. In a simulation one can control almost any aspect of the dynamics, changing the outcome by changing a single parameter. It is therefore vital that the parameter set chosen agrees with the physics dominating the real world. One needs to compromise this absolute accuracy to maintain a simulation with a feasible computing cost. This section dis- cusses some of the choices this study has made to balance physical accuracy and computing cost. 3.2.1 Simulation dimensions First among these choices was the size of the initial system already discussed in section 3.1.1. As outlined in that section a unit cell of any simulation handled in this study contains approxi- mately 25000 atoms. Note that the EAM potential formally scales as N2 as for any particle the potential and thus the force on that particle is deļ¬ned by the positions of all other particles in the simulation, relative to the particle under consideration. To reduce the computational cost we only take into account particles that have a signiļ¬cant contribution to this potential, that is particles in the near neighbourhood of the particle. This near neighbourhood was deļ¬ned to be a ball with a radius of 2.5 times the lattice spacing, which is around 10ĖšAfor Aluminium. The list of atoms present in this near neighbourhood is updated regularly. This choice restricts the volume and causes the length and radius of the wire to be directly related. As the radius is one of the parameters we allow to vary, we should ensure that a changing length has a minimal impact on the simulated dynamics. Even though our wire is theoretically inļ¬nite, through the periodic boundary condition at the z-boundary, all dynamics are still contained to just one simulation cell. This is especially worrying for a wire of relatively large radius, and thus short length as some phenomena may require group dynamics over a large distance. One of these long range phenomena is the break-up of the molten wire into balls of liquid to minimise the surface area. This requires the bulk movement of the molten metal and the momentary increase in surface area as the wire breaks the cylindrical shape. These two properties make this process a longer wavelength phenomenon and we can use this to test our system in this respect.
  • 22. 20 CHAPTER 3. METHODOLOGIES We refer to the wires treated in sections 4.2.1 and 4.2.2 from the results chapter ahead as the thick and thin wire respectively. The caloric curve of the thin wire shows a rise and fall in the energy per atom after it has completely molten, corresponding to the break-up of the liquid wire as discussed above. The caloric curve corresponding to the thicker wire does not show these dynamics which was some cause of concern. To put these concerns to rest we allowed the thick wire to continue to heat for half again the simulation time for the thick wire. This simulation showed a similar dynamic did occur at a higher temperature. The lack hereof in the regular simulation time could just be due to the fact the thicker wire needs to overcome a greater energy barrier to break up into the spherical shape. Some conļ¬dence that the length scale is suļ¬ƒcient to capture any melting dynamics can be taken from ļ¬gure 4.4 where we see the melting take place at two diļ¬€erent sites. This suggest that the wire in that simulation is signiļ¬cantly longer than the interaction wavelength of any of the melting dynamics. 3.2.2 Time integration and the Langevin thermostat The next choices concern the time integration of the simulation. Molecular dynamics studies modelling metals often use a time step of the order of femtoseconds. Any larger than this may cause discontinuous dynamics where smaller timesteps provide no signiļ¬cant change. In this particular study we use a timestep of 2 femtoseconds for a million timesteps, giving a total simulation time of 2 nanoseconds. This simulation length is necessary to capture the complete melting transition. To heat the system we use a Langevin thermostat. Physically, the Langevin thermostat can be thought of as a heat bath of very light solvent molecules in which the system is immersed. The temperature of this heat bath is controlled exactly. The system attempts to equilibrate with the heat bath through collisions with the solvent particles. This avoids the system being treated as a vacuum and allows the temperature of the system to be controlled in a natural manner through the heat bath. The dynamics for the ith solvent element can be captured by the following stochastic equation. miĀØxi(t) = Fi ({x(t)}) āˆ’ mi Ī»i Ė™xi(t) + Ri(t) (3.2) Where mi and xi are the mass and positions of the system particles. Fi is the force de- ļ¬ned by the EAM potential, which thus depends on the positions of the nearby particles. R is a stochastic term with a Gaussian distribution of mean zero. mi Ī»i Ė™x forms the coupling with the heat bath. The strength of this coupling with respect to the momentum of the particle is determined by Ī»i, the Langevin parameter, which has units of time and can be seen as the time necessary for the system to equilibrate with the heat bath. In general this Langevin parameter can depend on the particle, however in this work it will be constant for all atoms in any given simulation. Note that the Langevin parameter is inversely proportional to the coupling strength between the system and the heat bath. Thus a low Langevin parameter will cause the temper- ature to be strictly controlled by the temperature of the solvent, thus mimicking a canonical system. Inversely a high Langevin parameters will mimic a micro-canonical system.25,26
  • 23. 3.3. POST PROCESSING 21 3.3 Post processing During an MD simulation the state of the system is written to ļ¬les at a ļ¬xed time interval. These ļ¬les contain the atomic positions, the kinetic and potential energies and the temperature from the system. The latter two are determined from the velocities of the particles where the potential is determined by the positions of the atoms within the EAM potential. The energies and temperature together allow us to plot caloric curves to monitor how the transition diļ¬€ers from bulk melting. The atomic positions allow us to make qualitative judgements about the state of the system and they are instrumental in distinguishing between the diļ¬€erent melting pathways. As the transition progresses a group of atoms changes from being in a strictly ordered lattice to a disordered, ļ¬‚uid state. This transition is less obvious when looking at a single atom and the atomic position of one atom is not suļ¬ƒcient to determine the phase this atom is in. However, we would like to determine whether any atom in our system is part of the solid or liquid phase. We thus need to evaluate the surroundings of the atom and determine whether it is an ordered lattice or a disordered ļ¬‚uid. 3.3.1 Steinhardt paramaters From the many methods in use to do this solid-liquid classiļ¬cation, this study uses the Steinhardt parameters. These are named after Paul Steinhardt, who published a paper discussing bond order parameters.27 These are quantities evaluating the symmetry of the ā€bondsā€ around a particular atom based on spherical harmonics. Here bonds are characterised not only by the proximity to the atom under consideration but also the direction with respect to other bonds from this atom. As spherical harmonics are distinguished by their degree, l, there are diļ¬€erent degrees of the Steinhardt parameters, denoted ql. In this study only the order parameter of the sixth degree is of interest. Figure 3.2: Snapshots showing the progression with a consistent solid-liquid distribution in cross-section throughout the length of the wire with atoms classiļ¬ed as either solid (copper) or liquid (blue). Each snapshot is accompanied by the distribution of the ĀÆq6 Steinhardt order parameter for that frame showing the classiļ¬cation as either solid (green) or liquid (blue). Together they form an empirical conļ¬rmation of the validity of the Steinhardt parameters as a method for solid-liquid classiļ¬cation. The last snapshot has all atoms classiļ¬ed as liquid removed, to highlight a shortcoming of this method.
  • 24. 22 CHAPTER 3. METHODOLOGIES The Steinhardt parameters can thus be used to distinguish between diļ¬€erent lattice types as well as between the solid and liquid state. Locally this is captured by a complex valued vector for each atom describing the length and direction of each bond to this atom. This vector can be converted to real numbers by taking the Fourier invariant of the ļ¬eld of neighbouring complex vectors. Finally these are averaged to smoothen the distribution of thermal ļ¬‚uctuations. These average values are denoted as ĀÆql, where we use the ĀÆq6 order parameter in this study.27,28 Thus in short the Steinhardt parameters are real values assigned to each atom quantifying the relation this atom has to its neighbourhood. The values of these ĀÆq6 corresponding to dif- ferent crystal structures and thermodynamic phases can be found in literature. The latter of these makes these parameters useful in following a phase transition. It is worth noting that we need to deļ¬ne a radius of the near neighbourhood of an atom, for this method of liquid-solid classiļ¬cation to function. This distance is usually chosen to be the ļ¬rst minimum of the ra- dial distribution function, which can be found either from the atomic positions or the literature. To test the solid-liquid classiļ¬cation weā€™ve picked a simulation that had a fairly consistent cross-sectional picture. In other words, for any particular frame, the image seen for the cross- sectional snapshot does not change signiļ¬cantly as the viewpoint moves along the length of the wire. For such a wire the proportion of liquid to solid atoms in the cross-sectional snapshot should be approximately the same as that proportion for the complete wire. Representative snapshots of these cross sections are shown in ļ¬gure 3.2 along with the distribution of the ĀÆq6 parameter for the wire at the time of that snapshot. We can now visually conļ¬rm that the atoms classiļ¬ed as liquid, actually seem to be in a disordered state. Similarly, the atoms classiļ¬ed as solid are ordered into a crystal lattice. Some small calculations also give that the distribution of the ĀÆq6 parameter gives the same liquid frac- tion as counting the atoms in a representative number of layers in the cross-sectional snapshot. However the last snapshot shows a seemingly fully melted wire with some atoms classiļ¬ed as solid still present along the surface of the wire. These atoms appear to be misclassiļ¬ed due to their position at the physical boundary of the system. Aside from this caveat, we believe the Steinhardt order parameter to be a valid method of solid-liquid classiļ¬cation.
  • 25. Chapter 4 Analysis and results In molecular dynamics the position of all atoms can be tracked throughout the simulation. This allows us to take snapshots at any given time to evaluate the state of the system. Exam- ples of these snapshots can be found throughout this chapter. This becomes especially useful after solid-liquid classiļ¬cation as discussed in section 3.3. In these simulation it was chosen to take snapshots every 4000 fs and thus produce 500 snapshots for a 2ns simulation. With these snapshots we also evaluate the temperature, from the average speed of the particles, as well as the kinetic and potential energies of the system and per atom. Where the energies are found from average speeds and the positions of the atom within the EAM potential respectively. We can use all this information to evaluate our simulations in a more qualitative manner. In this chapter there are two main parameters that are varied. The ļ¬rst is the size of nanowire, deļ¬ned by its radius. The total length is trivially held constant through the periodic boundary condition in the z-direction and is thus not a relevant parameter. Note that the length of the wire within a unit cell may change between simulations, as described in section 3.2.1. 4.1 Preliminaries Before running the simulations of wires along the diļ¬€erent melting pathways we need to en- sure that our simulations model a physical and stable system. We ļ¬rst ensure that the initial system is a stable and feasible conļ¬guration We also check the simulation parameters such as the Langevin parameter and the heating rate are within a reasonable range and provide the required resolution to monitor the changing system. Finally we explore some of the output parameters and visualisation tools available. This include verifying the Steinhard parameter accurately classiļ¬es solid and liquid atoms. 4.1.1 Initialistion As discussed in chapter 3 we ļ¬rst need to equilibrate the wire and allow the lattice to relax. To allow this we performed an NVT minimisation before starting each simulation. We notice a slight, but important change in the wire geometry after this relaxation period. Figure 4.1 shows a (001) cross-sectional view of the initial wire before and after relaxation at 600 Kelvin, the initial temperature of our simulations. To ensure that the right hand side was a stable state a simulation was run unheated for the length of the later simulations, 2 ns. This preserved the shape of the wire suļ¬ƒciently and no un-physical behaviour was noticed. 23
  • 26. 24 CHAPTER 4. ANALYSIS AND RESULTS We can see that before the relaxation period the wire is cut from a perfect lattice, as none but the ļ¬rst layer of atoms are visible. Although the cross section is not perfectly circular, the layering of the atoms is such that it models this shape as closely as possible. This perfection disappears after relaxation. First we note the slight thermal expansion. We also see the atoms that were hidden behind the ļ¬rst layer due to the stochastic elements of the potential and the non-uniform initial velocity distribution. However, the wire still has a distinct crystal structure, which is re-established by the fact that almost all atoms are still classiļ¬ed as solid. The few atoms classiļ¬ed as liquid can be seen as outliers that will likely soon crystallise again. A more subtle eļ¬€ect of the relaxation is the increased faceting of the cross section. We see that in particular the diagonal facets have ļ¬‚attened as to maximize the (110) facets. This results in the cross section becoming almost octagonal with the (110) and (100) facets being the only two substantial facets on the surface of the wire. This is likely to have an eļ¬€ect as chapter 2 found the surface energies of these facets to be critical. We still expect the theory to hold reasonably well as we knew the perfectly cylindrical shape was an approximation in that chapter. 4.1.2 Simulation parameters As outlined in section 3.2 the total heating done and the total duration of the simulations is held constant. It seems to follow that the heating rate remains constant across simulations. This only holds true for the heat bath connected to our simulation. The Langevin parameter, Ī», inļ¬‚uences the rate at which this heat enters the nanowire system. As this parameter is the means to control what ensemble the simulation represents we cannot avoid this discrepancy. Instead we will investigate the inļ¬‚uence of the Langevin parameter on the heating rate. Figure 4.1: Initial wire as created before the start of the simulation (left) and after an initial relaxation period (right) with the crystal facets indicated. Atoms classiļ¬ed as solid are indicated copper while those classiļ¬ed as liquid are blue.
  • 27. 4.2. OBSERVED MELTING PATHWAYS 25 4.2 Observed melting pathways In this section the diļ¬€erent pathways outlined in the theory of chapter 2 are found and followed in simulation. We managed to ļ¬nd each of the pathways discussed in the theory and will thus discuss them again and compare the theory to the simulations. 4.2.1 Observation of surface melting As described in section 2.3.1, surface melting is the forming of a liquid layer on the nanowire surface. This liquid layer encapsulates the remaining solid core fully for there to be complete surface melting. Figure 4.2 shows the progression of the transition through snapshots. Fig- ure 4.3 shows the caloric curve and liquid fractions of two simulations showing surface melting at diļ¬€erent Langevin parameters. The left graph is of the same simulation as the snapshots. Note that the actual melting transition is relatively short. It is preceded and followed by a signiļ¬cant heating period characterised by a linear trend in the caloric curve. This shows the wire is a physical and stable system. Figure 4.2: Snapshots for the progression of surface melting left to right, top to bottom. This simulation has a relatively thick wire and low Langevin parameter Ī» = 10ps. Atoms identiļ¬ed as solid are copper coloured, blue indicates atoms in a liquid state. Each snapshots has 4 projections of the wire, two side on views, facing the (110) plane(left) and two cross sections, facing the (001) plane(right). The top row shows the full wire where the bottom row shows only the atoms classiļ¬ed as solid.
  • 28. 26 CHAPTER 4. ANALYSIS AND RESULTS We can now compare the snapshots with the caloric curve through the liquid fraction. Each snapshot has 4 projections of the wire as outlined in the caption. First note that snapshot A is not the initial state of the wire, which looks more like ļ¬gure 3.1. Snapshot A is just before the onset of surface melting. We see that the (110) facets have already melted and now form a solid-liquid interface. These facets have increased in size by melting the sides of the (100) facets, which are also part of the (110) facet. This period of facet melting can be seen as a relatively slow, but accelerating rise of the liquid-solid fraction to about 10 percent. At the same time we see the caloric curve deviating from the linear heating characteristic for a fully solid system. This is due to the transition of these atoms from solid and liquid and the associated latent heat. Snapshot B shows the wire just after the onset of surface melting. We see the wire fully encapsulated in a liquid layer with solid core tending to maximise the (110)-facets. This is reinforced in snapshot C which shows the same wire at a later stage of surface melting. Keep in mind that all atoms are displayed in these snapshots, gradually fading as they are further away. Thus although in snapshot C there seems to be a signiļ¬cant (100) facet on the right hand side of the wire, especially obvious in the solid only cross section. However, it is quite thin and we see that there is a second facet further towards the centre where we see a far greater density of the displayed atoms, showing that this later facet extends along a larger portion of the wire. The melting of the solid core also seems to be fairly symmetric, allowing the solid core to remain in the centre. This is backed up by the linear trend in the liquid fraction between 10 and 90 per. We see this corresponds to the ļ¬‚at section in the caloric curve, as one would expect from a canonical system. The ļ¬nal snapshot, D, shows the wire after the solid core has broken up and there is only a small cluster remaining. This cluster would melt away quickly as it contains less than 10 percent of the original atoms. We see that the liquid fraction ļ¬‚attens again as the number of solid atoms remaining becomes of the order of misclassiļ¬ed atoms near the surface. After the liquid fraction reaches this maximum value the caloric curve again assumes a linear trend, showing the heating of a pure liquid. The second graph in ļ¬gure 4.3 shows the caloric curve and liquid fraction for the same wire in a simulation with a higher Langevin parameter. These follow a very similar trend as do the snapshots, hence why these are not shown. However, the caloric curve starts to show a deļ¬nite dip in temperature, after reaching a maximum. This shows the changing ensemble as we move away from the canonical treatment. When the Langevin parameter further we start to see facet induced melting. This is discussed further in section 4.2.3. Figure 4.3: Caloric curves (red) with liquid fraction (blue) for a simulation of surface melting. These simulations have diļ¬€erent Langevin paramaters, Ī» = 10ps and Ī» = 40ps, respectively.
  • 29. 4.2. OBSERVED MELTING PATHWAYS 27 4.2.2 Observation of linear melting Similar to the previous section ļ¬gure 4.4 shows the progression of linear melting in snapshots while ļ¬gure 4.5 shows the caloric curve and liquid fraction of two diļ¬€erent simulations showing linear melting. The snapshots are for a simulation with Langevin parameter Ī» = 200ps. Thus the right caloric curve corresponds to the same simulations as the snapshots. This particular value was chosen because the snapshots obtained from this simulation best showed the range of phenomena encountered. Linear melting was observed for the full range of the Langevin parameter. In other words for both canonical and micro-canonical systems. Note that this is not true for surface melting or for facet induced melting in thicker wires, as we will encounter in the next section. As before the actual melting transition is a relatively small part of the caloric curve, to ensure the system is physical. At the end of the caloric curve we see the forming of a liquid ball and the associated drop in energy, as discussed in section 3.2. First we examine the snapshots. The ļ¬rst thing to note that the wire needed to be signif- Figure 4.4: Snapshots showing the progression of linear melting left to right, top to bottom. This simulation has a relatively thin wire and an intermediate Langevin parameter Ī» = 100ps. Atoms identiļ¬ed as solid are copper coloured, blue indicates atoms in a liquid state. Each snapshots has 4 projections of the wire, two side on views, facing the (110) plane(left) and two cross sections, facing the (001) plane(right). The top row shows the full wire where the bottom row shows only the atoms classiļ¬ed as solid.
  • 30. 28 CHAPTER 4. ANALYSIS AND RESULTS icantly thinner to clearly and reliably show linear melting. This agrees with what we saw in chapter 2 where the distinguishing feature between linear and surface melting seemed to be the radius, Rsl. However a caveat was placed by this ļ¬nding as linear melting has a signiļ¬cantly higher nucleation cost and thus may not be preferred when close to this critical radius. This is reļ¬‚ected in snapshot A, which again is not the initial state, where a signiļ¬cant portion of the atoms have already melted but the wire still has not cut through. The energy cost in the latent heat of these atoms is related to the nucleation cost of this pathway. In most snapshots we see some faceting is present in the cross-section of the full wire. How- ever, it is signiļ¬cantly less strong and from the solid only pictures we see that the faceting only occurs along sections of the wire as these cross-sections remains fairly round for the full duration of simulation. Snapshot B shows the wire just after it has been cut through and we see that it seems to be cut in two places, one just out of view at the left. The lower section in between these cuts is also completely detached from the lower surface. I.e. it is undergoing facet-induced melting. This is reinforced by snapshot C, which also shows this part of the wire has melted signiļ¬cantly faster, which is partly due to the smaller size in snapshot B and partly due to the greater melting surface for facet induced melting. The fact that we see these two diļ¬€erent pathways supports the hypothesis that the nucle- ation cost of linear melting is an important factor. Since nucleation is ultimately a random process, it is to be expected that we see seemingly ambiguous results as ļ¬gure 4.4. When we compare the snapshots with the right-hand liquid fraction we see this reļ¬‚ected as a kink be- tween two fairly linear trends. This kink is the moment the facet induced melting has ļ¬nished and we are left with only linear melting. The initial nucleation is reļ¬‚ected in the initial curved section, which now does not represent faceting. We see a similar trend in the caloric curve. initially there is a linear trend as we are heating a pure solid. When the liquid fraction starts to increase, the caloric curve deviates from this linear trend. In the period with the two melting types simultaneous we see a sharp drop in te caloric curve as energy is released as latent heat faster than the system is allowed to heat up under this high Langevin parameter. After the facet induced melting is completed the temperature slowly increases again as the latent heat of the linear melting alone is not enough to maintain the negative gradient. The linear melting pathway was observed only for relatively thin wires, with a diameter of at most 12 lattice spacings. However, when linear melting seems largely independent of the Langevin parameter. When linear melting was observed for a simulation with low Ī» it would generally also been seen for a simulation with Ī» large. Figure 4.5: Caloric curves (red) with liquid fraction (blue) for a simulation of linear melting. These simulations have diļ¬€erent Langevin paramaters, Ī» = 10ps and Ī» = 200ps, respectively.
  • 31. 4.2. OBSERVED MELTING PATHWAYS 29 4.2.3 Observation of facet induced melting As in the previous two sections the snapshots shown in ļ¬gure 4.6 and the caloric curve and liquid fraction found in ļ¬gure 4.7 follow a nanowire undergoing facet induced melting. The snapshots are for a simulation with a high Langevin parameter, Ī» = 200ps. Thus the corresponding caloric curve is on the right. Unlike what was observed for the linear melting pathway, the Langevin parameter needs to be suļ¬ƒciently high to observe facet induced melting. In other words, we observe it only for more micro-canonical systems. In section 2.3.3 we deļ¬ned facet induced melting (FI) as a melting transition that forms a liquid layer on most of the surface while maintaining a solid-vapour interface on at least one facet for the duration of the transition. This is indeed what we see in the snapshots below. In snapshot A there signiļ¬cant faceting occurs along the full length of the wire. However, the faceting is not symmetric as it was for surface melting. Only the (110) facets on the right hand side have melted along the full side of the wire. This asymmetry was present in all FI- simulations, even though it seemed random which side of the wire would melt. Note that the side-on view has been rotated from the cross section such that the right hand Figure 4.6: Snapshots for the progression of facet induced melting left to right, top to bottom. This simulation has a relatively thick wire and high Langevin parameter Ī» = 200ps. Atoms identiļ¬ed as solid are copper coloured, blue indicates atoms in a liquid state. Each snapshots has 4 projections of the wire, two side on views, facing the (110) plane(left) and two cross sections, facing the (001) plane(right). The top row shows the full wire where the bottom row shows only the atoms classiļ¬ed as solid.
  • 32. 30 CHAPTER 4. ANALYSIS AND RESULTS side, the side showing the initial melting, faces towards the viewer. Snapshot B shows the same asymmetry. The wire has now completely disconnected from both the (110) and (100) facets on the right hand side of the cross-sections. Examining the side-on view, we see that the solid phase still extends the full width of the wire. There is still a signiļ¬cant fraction of these (110) facets that maintains a solid-vapour interface. This part resembles the geometry used in the chapter 2, but the rest of the wire does not. Showing an almost spherical shape, ļ¬‚attened by last remaining (110) solid-vapour interface. Recall that in the chapter 2 we did not properly take account of the wetting and NM facets, which is likely to be the source of this discrepancy between the theory and simulation. The liquid fraction follows a fairly un-eventful trend looking a lot like that for surface melt- ing. We initially see the initial melting of the (110) facets in the almost quadratic increase in liquid fraction. As the wire starts the main melting process after forming the liquid layer on the (100) facet the liquid fraction follows a linear trend, which continues until a near 100 percent liquid fraction is reached. The caloric curve however is slightly diļ¬€erent showing a deļ¬nite dip in temperature as the melting progresses. This is again explained by the high Langevin parameter. This makes the transition occur in a more micro-canonical ensemble. Thus the kinetic energy of the atoms, inseparable from the temperature, is used directly to supply the heat required for the melting transition rather than the heat added to the system through the heat bath. Snapshot C shows the wire after a period of FI melting. The solid remainder is now dis- connected from all of the wire surface except for the (100) facet on the far side of the side-on view. Based on cross section, it seems there is still strong faceting present, despite the constant melting at the solid-liquid interface. Snapshot C also shows the wire as it is about to break up. Notice that the break-up occurs for a signiļ¬cantly thicker wire than seen for surface melting earlier. After the break-up the solid remainder is melted while still maintaining a solid-liquid frame all the way to snapshot D, which was one of the last frames before the completely molten state. The FI pathway was observed for wires of the same radius as those for surface melting, but at a signiļ¬cantly higher Langevin parameter, generally with Ī» > 100ps. When the radius of the wire was made smaller this pathway is no longer preferred over linear melting and we no longer see the facet induced melting pathway. Figure 4.7: Caloric curves (red) with liquid fraction (blue) for a simulation of facet induced melting. These simulations have diļ¬€erent Langevin paramaters, Ī» = 120ps and Ī» = 200ps, respectively.
  • 33. 4.3. COMPARISON TO THEORY 31 4.3 Comparison to theory The simulations discussed in this chapter have managed to reproduce each of the pathways dis- cussed in chapter 2. We identiļ¬ed some of the parameters governing which pathway is observed. The parameters found in this chapter are summarised in table 4.1. This section will evaluate the agreement between the two approaches and discuss the diļ¬€erences. Preferred pathway Radius Langevin parameter, Ī» Caveat Surface melting > 8a < 40 Shrinkage of Non-melting surfaces Linear melting < 6a > 0, < 200 No pre-melting facets facet induced melting > 8a > 80 Preceded by assymetric faceting Bulk melting - - Not invesigated in this chapter Table 4.1: Conditions for the diļ¬€erent melting pathways as observed in molecular dynamics simulations. The lattice parameter, a is 4.050ĖšAfor FCC aluminium. Note that the conditions on the radius and Langevin parameter are not closely ļ¬tting. In between these boundary values we would often observe a strong competition between two path- ways. Some examples of which are included in ļ¬gure 4.8. The competition between linear and facet induced melting was already shown in ļ¬gure 4.4. The deciding parameter between linear melting and both surface and facet induced melting seems to be the radius, as we found only for the former of these comparisons in the free energy evaluation. From table 4.1, one is tempted to estimate Rsl to be between 6 and 8 lattice spacings. However, we have not been able to identify the nucleation cost of the diļ¬€erent pathways, and closer investigation may reveal a higher critical radius. We have not been able to identify the surface energy as an explicit parameter but have seen the important role of the crystal facets and the Langevin parameter. These ļ¬ndings suggest that for facet induced melting the simple model employed in chapter 2 is not suļ¬ƒcient. Figure 4.8: Two examples of competing pathways, with atoms classiļ¬ed as solid (liquid) shown in copper (blue). Snapshot A show part of the wire surrounded by melt while the other half of the wire attempts to maintain a solid-vapour interface, showing the competition between surface and facet induced melting. Snapshot B shows competition between surface and linear melting as the solid phase becomes fully encapsulated in the melt after a cross-sectional cut.
  • 34. 32 CHAPTER 4. ANALYSIS AND RESULTS
  • 35. Chapter 5 Discussion In this study we investigated three diļ¬€erent pathways through which a nanowire can melt. Chapter 2 compared the pathways by means of a free energy evaluation, calculating the free energy cost and predicting a melting temperature for each wire. This approach made a number of approximations which overlooked possibly important details. Foremost among these was the shape of the nanowire. As we assume a cylindrical shape the faceting seen in the simulations is not taken account of. This is vital in predicting the melting temperature as this depends heavily on the phase interface energy densities. These energy densities depend on the crystal facet at which the interface is located. A more complete investigation may be made when an octagonal cylinder is assumed. The sides of the octogonal can then be taken to be the diļ¬€erent crystal facets and one may acquire an expression for the temperature in terms of the speciļ¬c interface energies of the diļ¬€erent facets. Allowing experimental or simulation parameters to be substituted to predict the melting tem- perature and pathway of a given wire. The cross section may also be allowed to change into an irregular octagonal by letting the wetting surface to melt the sides of non melting surfaces. This would make for signiļ¬cantly more complicated expressions but would model a number of the interesting phenomena encountered in simulation more closely. Another source of discrepancy between theory and simulation originated from ignoring the nucleation cost of the diļ¬€erent pathways. Based on the free energy evaluation in chapter 2 the linear pathway is always preferred below the bulk melting temperature. However, in chapter 4 we certainly saw both pathways. This is likely due to the linear melting transition needing to overcome a greater energy barrier in the nucleation stage during which the simulation is still heated. Thus if this barrier cannot be overcome before reaching a temperature at which surface melting is possible we may see either pathway. Future work may simulate wires with a diļ¬€erently orientated lattice. Of special interest would be a wire with only (110) and (111) facets which should highlight any diļ¬€erence between the two non-melting facets in Aluminium, after the (100) facet was examined here. Another interesting case would be a geometry with only the non-melting surfaces (111) and (100) at the surface, although this is not possible in a cylindrical nanowire. It should also be noted that this study observed multiple wires in which more than one pathway was followed. This co-occurrence of two pathways may be of interest to another study as it changed the stability of the solid signiļ¬cantly, often melting signiļ¬cantly quicker in the boundary areas. An example of co-occurrence of linear and facet induced melting is given in section 4.2.2. 33
  • 36. 34 CHAPTER 5. DISCUSSION Finally future studies could investigate similar behaviour for nanowires consisting of diļ¬€er- ent materials. The authors believe that any FCC metal with two non-melting surfaces should behave qualitatively similar to the aluminium studied here. However, a metal with a diļ¬€erent crystal lattice or an FCC lattice with one or zero non-melting surfaces may behave quite diļ¬€er- ently.
  • 37. Chapter 6 Conclusions Simulations using molecular dynamics were carried out to investigate melting phenomena in metal nanowires. The main focus was on the diļ¬€erent pathways through which a nanowire can melt. In this study surface melting was observed for wires with relatively large radius in canoni- cal simulations. Facet induced melting was observed for similar wires treated micro-canonically. Linear melting was observed for both canonical and micro-canonical systems for wires with a diameter smaller than 12 lattice spacings, where the system ensemble was decided by the Langevin parameter, Ī». Another melting phenomena investigated, albeit less in-depth, was the faceting in melting due to the presence of both wetting and non-melting facets at the nanowire surface. The pres- ence of this state before the system seems to be associated with the melting pathway taken. This state was not signiļ¬cantly present before linear melting, while the facets melt symmetri- cally for surface melting and showed asymmetry for the facet induced melting. An attempt was made to explain and predict the pathways using free energy evaluations for a simple model of a nanowire. This managed to predict the dependence of the melting pathway on the radius of the wire and the importance of the interface energy diļ¬€erence, āˆ†Ī³. However it was discovered that an accurate theory describing these phenomena needs to take the faceting dynamics and the cost of nucleation into account. This work thus identiļ¬ed three diļ¬€erent pathways for the melting transition in a nanowire. It also identiļ¬ed some of the main parameters inļ¬‚uencing the pathway taken by any particular wire. This work is of importance as electronic components reach the nanoscale and are pushed to their thermodynamic limits. Future research directions are outlined in chapter 5. 35
  • 38. 36 CHAPTER 6. CONCLUSIONS
  • 39. Bibliography 1 Y. Kondo and K. Takayanagi, ā€œSynthesis and characterization of helical multi-shell gold nanowires,ā€ Science, vol. 289, no. 5479, pp. 606ā€“608, 2000. 2 A. Xhang, G. Zheng, and C. Lieber, Nanowires, Building Blocks for Nanoscience and Nan- otechnology, vol. 1 of NanoScience and Technology. Springer, 2016. 3 F. A. Lindemann, ā€œUeber die berechnung molekularer eigenfrequenzen,ā€ Phys. Z, vol. 11, pp. 609ā€“612, 1910. 4 P.-L. Sun, S.-P. Wu, and T.-S. Chin, ā€œMelting point depression of tin nanoparticles embedded in a stable alpha-alumina matrix fabricated by ball milling,ā€ Materials Letters, vol. 144, pp. 142 ā€“ 145, 2015. 5 K. M. Ridings, ā€œThe breakup and melting of metal nanowires using molecular dynamics,ā€ University of Auckland, 2015. 6 A. M. Molenbroek and J. W. M. Frenken, ā€œAnharmonicity but absence of surface melting on al(001),ā€ Phys. Rev. B, vol. 50, pp. 11132ā€“11141, Oct 1994. 7 j. W. M. Frenken and J. F. van der Veen, ā€œObservation of surface melting,ā€ Physical review letters, vol. 54, pp. 827ā€“830, 1985. 8 A. M. Rodrguez, G. Bozzolo, and J. Ferrante, ā€œMultilayer relaxation and surface energies of fcc and bcc metals using equivalent crystal theory,ā€ Surface Science, vol. 289, no. 1, pp. 100 ā€“ 126, 1993. 9 L. Vitos, A. Ruban, H. Skriver, and J. Kollr, ā€œThe surface energy of metals,ā€ Surface Science, vol. 411, no. 12, pp. 186 ā€“ 202, 1998. 10 Y. Mishin, D. Farkas, M. J. Mehl, and D. A. Papaconstantopoulos, ā€œInteratomic potentials for monoatomic metals from experimental data and ab initio calculations,ā€ Phys. Rev. B, vol. 59, pp. 3393ā€“3407, Feb 1999. 11 D. Schebarchov and S. C. Hendy, ā€œSuperheating and solid-liquid phase coexistence in nanopar- ticles with nonmelting surfaces,ā€ Phys. Rev. Lett., vol. 96, p. 256101, Jun 2006. 12 D. Schebarchov and S. C. Hendy, ā€œSuperheating and solid-liquid phase coexistence in nanopar- ticles with nonmelting surfaces,ā€ Phys. Rev. Lett., vol. 96, p. 256101, Jun 2006. 13 J. A. Alonso, Structure and Properties of Atomic Nanoclusters. Imperial College Press, 2005. 14 M. Faraday, ā€œOn certain conditions of freezing water,ā€ Journal of the Franklin Institute, vol. 5, pp. 283ā€“284, 1850. 15 L. Hui, F. Pederiva, B. L. Wang, and B. L. Wang, ā€œHow does the nickel nanowire melt?,ā€ Appl. Phys. Lett., vol. 86, 2005. 37
  • 40. 38 BIBLIOGRAPHY 16 E. Fermi, J. Pasta, S. Ulam, and M. Tsingou, ā€œStudies of nonlinear problems,ā€ Los Alamos, 1955. 17 B. Alder and T. Wainwright, ā€œStudies in molecular dynamics. i. general method,ā€ The Journal of Chemical Physics, pp. 459 ā€“ 466, 1959. 18 D. Frenkel and B. Smit, Understanding Molecular Simulations: from algorithms to applica- tions. Computational sciences series, Academic Press, Hartcourt, 2002. 19 J. E. Jones, ā€œOn the determination of molecular ļ¬elds. ii. from the equation of state of a gas,ā€ Proceedings of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, vol. 106, no. 738, pp. 463ā€“477, 1924. 20 M. S. Daw and M. I. Baskes, ā€œSemiempirical, quantum mechanical calculation of hydrogen embrittlement in metals,ā€ Phys. Rev. Lett., vol. 50, pp. 1285ā€“1288, Apr 1983. 21 M. S. Daw and M. I. Baskes, ā€œEmbedded-atom method: Derivation and application to im- purities, surfaces, and other defects in metals,ā€ Phys. Rev. B, vol. 29, pp. 6443ā€“6453, Jun 1984. 22 M. S. Daw, S. M. Foiles, and M. I. Baskes, ā€œThe embedded-atom method: a review of theory and applications,ā€ Materials Science Reports, pp. 251ā€“310, 1993. 23 Y. Mishin, D. Farkas, M. J. Mehl, and D. A. Papaconstantopoulos, ā€œInteratomic potentials for monoatomic metals from experimental data and ab initio calculations,ā€ Phys. Rev. B, vol. 59, pp. 3393ā€“3407, Feb 1999. 24 S. Plimpton, ā€œFast parallel algorithms for short-range molecular dynamics,ā€ Journal of Com- putational Physics, vol. 117, no. 1, pp. 1 ā€“ 19, 1995. 25 T. Schneider and E. Stoll, ā€œMolecular-dynamics study of a three-dimensional one-component model for distortive phase transitions,ā€ Phys. Rev. B, vol. 17, pp. 1302ā€“1322, Feb 1978. 26 I. Snook, The Langevin and Generalised Langevin Approach to the Dynamics of Atomic, Polymeric and Colloidal Systems. Amsterdam: Elsevier, 2007. 27 P. J. Steinhardt, D. R. Nelson, and M. Ronchetti, ā€œBond-orientational order in liquids and glasses,ā€ Phys. Rev. B, vol. 28, pp. 784ā€“805, Jul 1983. 28 A. S. Keys, C. R. Iacovella, and S. C. Glotzer, ā€œCharacterizing structure through shape matching and applications to self-assembly,ā€ Annual Review of Condensed Matter Physics, pp. 263ā€“285, March 2011.