ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chuckDissertation/nanoglass.tex
Revision: 3483
Committed: Tue Jan 13 14:39:50 2009 UTC (15 years, 8 months ago) by chuckv
Content type: application/x-tex
File size: 48122 byte(s)
Log Message:
Chuck's dissertation for PhD Jan 2009

File Contents

# User Rev Content
1 chuckv 3483 %!TEX root = /Users/charles/Documents/chuckDissertation/dissertation.tex
2     \chapter{\label{chap:nanoglass}GLASS FORMATION IN METALLIC NANOPARTICLES}
3    
4     \section{INTRODUCTION}
5    
6     Excitation of the plasmon resonance in metallic nanoparticles has
7     attracted enormous interest in the past several years. This is partly
8     due to the location of the plasmon band in the near IR for particles
9     in a wide range of sizes and geometries. Living tissue is nearly
10     transparent in the near IR, and for this reason, there is an
11     unrealized potential for metallic nanoparticles to be used in both
12     diagnostic and therapeutic settings.\cite{West:2003fk,Hu:2006lr} One
13     of the side effects of absorption of laser radiation at these
14     frequencies is the rapid (sub-picosecond) heating of the electronic
15     degrees of freedom in the metal. This hot electron gas quickly
16     transfers heat to the phonon modes of the particle, resulting in a
17     rapid heating of the lattice of the metal particles. Since metallic
18     nanoparticles have a large surface area to volume ratio, many of the
19     metal atoms are at surface locations and experience relatively weak
20     bonding. This is observable in a lowering of the melting temperatures
21     of these particles when compared with bulk metallic
22     samples.\cite{Buffat:1976yq,Dick:2002qy} One of the side effects of
23     the excitation of small metallic nanoparticles at the plasmon
24     resonance is the facile creation of liquid metal
25     droplets.\cite{Mafune01,HartlandG.V._jp0276092,Link:2000lr,Plech:2003yq,plech:195423,Plech:2007rt}
26    
27     Much of the experimental work on this subject has been carried out in
28     the Hartland, El-Sayed and Plech
29     groups.\cite{HartlandG.V._jp0276092,Hodak:2000rb,Hartland:2003lr,Petrova:2007qy,Link:2000lr,plech:195423,Plech:2007rt}
30     These experiments mostly use the technique of time-resolved optical
31     pump-probe spectroscopy, where a pump laser pulse serves to excite
32     conduction band electrons in the nanoparticle and a following probe
33     laser pulse allows observation of the time evolution of the
34     electron-phonon coupling. Hu and Hartland have observed a direct
35     relation between the size of the nanoparticle and the observed cooling
36     rate using such pump-probe techniques.\cite{Hu:2004lr} Plech {\it et
37     al.} have use pulsed x-ray scattering as a probe to directly access
38     changes to atomic structure following pump
39     excitation.\cite{plech:195423} They further determined that heat
40     transfer in nanoparticles to the surrounding solvent is goverened by
41     interfacial dynamics and not the thermal transport properties of the
42     solvent. This is in agreement with Cahill,\cite{Wilson:2002uq}
43     but opposite to the conclusions in Reference \citen{Hu:2004lr}.
44    
45     Since these experiments are carried out in condensed phase
46     surroundings, the large surface area to volume ratio makes the heat
47     transfer to the surrounding solvent a relatively rapid process. In our
48     recent simulation study of the laser excitation of gold
49     nanoparticles,\cite{VardemanC.F._jp051575r} we observed that the
50     cooling rate for these particles (10$^{11}$-10$^{12}$ K/s) is in
51     excess of the cooling rate required for glass formation in bulk
52     metallic alloys.\cite{Greer:1995qy} Given this fact, it may be
53     possible to use laser excitation to melt, alloy and quench metallic
54     nanoparticles in order to form glassy nanobeads.
55    
56     To study whether or not glass nanobead formation is feasible, we have
57     chosen the bimetallic alloy of Silver (60\%) and Copper (40\%) as a
58     model system because it is an experimentally known glass former, and
59     has been used previously as a theoretical model for glassy
60     dynamics.\cite{Vardeman-II:2001jn} The Hume-Rothery rules suggest that
61     alloys composed of Copper and Silver should be miscible in the solid
62     state, because their lattice constants are within 15\% of each
63     another.\cite{Kittel:1996fk} Experimentally, however Ag-Cu alloys are
64     a well-known exception to this rule and are only miscible in the
65     liquid state given equilibrium conditions.\cite{Massalski:1986rt}
66     Below the eutectic temperature of 779 $^\circ$C and composition
67     (60.1\% Ag, 39.9\% Cu), the solid alloys of Ag and Cu will phase
68     separate into Ag and Cu rich $\alpha$ and $\beta$ phases,
69     respectively.\cite{Banhart:1992sv,Ma:2005fk} This behavior is due to a
70     positive heat of mixing in both the solid and liquid phases. For the
71     one-to-one composition fcc solid solution, $\Delta H_{\rm mix}$ is on
72     the order of +6~kJ/mole.\cite{Ma:2005fk} Non-equilibrium solid
73     solutions may be formed by undercooling, and under these conditions, a
74     compositionally-disordered $\gamma$ fcc phase can be
75     formed.\cite{najafabadi:3144}
76    
77     Metastable alloys composed of Ag-Cu were first reported by Duwez in
78     1960 and were created by using a ``splat quenching'' technique in
79     which a liquid droplet is propelled by a shock wave against a cooled
80     metallic target.\cite{duwez:1136} Because of the small positive
81     $\Delta H_{\rm mix}$, supersaturated crystalline solutions are
82     typically obtained rather than an amorphous phase. Higher $\Delta
83     H_{\rm mix}$ systems, such as Ag-Ni, are immiscible even in liquid
84     states, but they tend to form metastable alloys much more readily than
85     Ag-Cu. If present, the amorphous Ag-Cu phase is usually seen as the
86     minority phase in most experiments. Because of this unique
87     crystalline-amorphous behavior, the Ag-Cu system has been widely
88     studied. Methods for creating such bulk phase structures include splat
89     quenching, vapor deposition, ion beam mixing and mechanical
90     alloying. Both structural \cite{sheng:184203} and
91     dynamic\cite{Vardeman-II:2001jn} computational studies have also been
92     performed on this system.
93    
94     Although bulk Ag-Cu alloys have been studied widely, this alloy has
95     been mostly overlooked in nanoscale materials. The literature on
96     alloyed metallic nanoparticles has dealt with the Ag-Au system, which
97     has the useful property of being miscible on both solid and liquid
98     phases. Nanoparticles of another miscible system, Au-Cu, have been
99     successfully constructed using techniques such as laser
100     ablation,\cite{Malyavantham:2004cu} and the synthetic reduction of
101     metal ions in solution.\cite{Kim:2003lv} Laser induced alloying has
102     been used as a technique for creating Au-Ag alloy particles from
103     core-shell particles.\cite{Hartland:2003lr} To date, attempts at
104     creating Ag-Cu nanoparticles have used ion implantation to embed
105     nanoparticles in a glass matrix.\cite{De:1996ta,Magruder:1994rg} These
106     attempts have been largely unsuccessful in producing mixed alloy
107     nanoparticles, and instead produce phase segregated or core-shell
108     structures.
109    
110     One of the more successful attempts at creating intermixed Ag-Cu
111     nanoparticles used alternate pulsed laser ablation and deposition in
112     an amorphous Al$_2$O$_3$ matrix.\cite{gonzalo:5163} Surface plasmon
113     resonance (SPR) of bimetallic core-shell structures typically show two
114     distinct resonance peaks where mixed particles show a single shifted
115     and broadened resonance.\cite{Hodak:2000rb} The SPR for pure silver
116     occurs at 400 nm and for copper at 570 nm.\cite{HengleinA._jp992950g}
117     On Al$_2$O$_3$ films, these resonances move to 424 nm and 572 nm for
118     the pure metals. For bimetallic nanoparticles with 40\% Ag an
119     absorption peak is seen between 400-550 nm. With increasing Ag
120     content, the SPR shifts towards the blue, with the peaks nearly
121     coincident at a composition of 57\% Ag. Gonzalo {\it et al.} cited the
122     existence of a single broad resonance peak as evidence of an alloyed
123     particle rather than a phase segregated system. However, spectroscopy
124     may not be able to tell the difference between alloyed particles and
125     mixtures of segregated particles. High-resolution electron microscopy
126     has so far been unable to determine whether the mixed nanoparticles
127     were an amorphous phase or a supersaturated crystalline phase.
128    
129     Characterization of glassy behavior by molecular dynamics simulations
130     is typically done using dynamic measurements such as the mean squared
131     displacement, $\langle r^2(t) \rangle$. Liquids exhibit a mean squared
132     displacement that is linear in time (at long times). Glassy materials
133     deviate significantly from this linear behavior at intermediate times,
134     entering a sub-linear regime with a return to linear behavior in the
135     infinite time limit.\cite{Kob:1999fk} However, diffusion in
136     nanoparticles differs significantly from the bulk in that atoms are
137     confined to a roughly spherical volume and cannot explore any region
138     larger than the particle radius ($R$). In these confined geometries,
139     $\langle r^2(t) \rangle$ approaches a limiting value of
140     $3R^2/40$.\cite{ShibataT._ja026764r} This limits the utility of
141     dynamical measures of glass formation when studying nanoparticles.
142    
143     However, glassy materials exhibit strong icosahedral ordering among
144     nearest-neghbors (in contrast with crystalline and liquid-like
145     configurations). Local icosahedral structures are the
146     three-dimensional equivalent of covering a two-dimensional plane with
147     5-sided tiles; they cannot be used to tile space in a periodic
148     fashion, and are therefore an indicator of non-periodic packing in
149     amorphous solids. Steinhart {\it et al.} defined an orientational bond
150     order parameter that is sensitive to icosahedral
151     ordering.\cite{Steinhardt:1983mo} This bond order parameter can
152     therefore be used to characterize glass formation in liquid and solid
153     solutions.\cite{wolde:9932}
154    
155     Theoretical molecular dynamics studies have been performed on the
156     formation of amorphous single component nanoclusters of either
157     gold,\cite{Chen:2004ec,Cleveland:1997jb,Cleveland:1997gu} or
158     nickel,\cite{Gafner:2004bg,Qi:2001nn} by rapid cooling($\thicksim
159     10^{12}-10^{13}$ K/s) from a liquid state. All of these studies found
160     icosahedral ordering in the resulting structures produced by this
161     rapid cooling which can be evidence of the formation of an amorphous
162     structure.\cite{Strandburg:1992qy} The nearest neighbor information
163     was obtained from pair correlation functions, common neighbor analysis
164     and bond order parameters.\cite{Steinhardt:1983mo} It should be noted
165     that these studies used single component systems with cooling rates
166     that are only obtainable in computer simulations and particle sizes
167     less than 20\AA. Single component systems are known to form amorphous
168     states in small clusters,\cite{Breaux:rz} but do not generally form
169     amorphous structures in bulk materials.
170    
171     Since the nanoscale Ag-Cu alloy has been largely unexplored, many
172     interesting questions remain about the formation and properties of
173     such a system. Does the large surface area to volume ratio aid Ag-Cu
174     nanoparticles in rapid cooling and formation of an amorphous state?
175     Nanoparticles have been shown to have a size dependent melting
176     transition ($T_m$),\cite{Buffat:1976yq,Dick:2002qy} so we might expect
177     a similar trend to follow for the glass transition temperature
178     ($T_g$). By analogy, bulk metallic glasses exhibit a correlation
179     between $T_m$ and $T_g$ although such dependence is difficult to
180     establish because of the dependence of $T_g$ on cooling rate and the
181     process by which the glass is formed.\cite{Wang:2003fk} It has also
182     been demonstrated that there is a finite size effect depressing $T_g$
183     in polymer glasses in confined geometries.\cite{Alcoutlabi:2005kx}
184    
185     In the sections below, we describe our modeling of the laser
186     excitation and subsequent cooling of the particles {\it in silico} to
187     mimic real experimental conditions. The simulation parameters have
188     been tuned to the degree possible to match experimental conditions,
189     and we discusss both the icosahedral ordering in the system, as well
190     as the clustering of icosahedral centers that we observed.
191    
192     \section{COMPUTATIONAL METHODOLOGY}
193     \label{nanoglass:sec:details}
194    
195     \subsection{INITIAL GEOMETRIES AND HEATING}
196    
197    
198     Cu-core / Ag-shell and random alloy structures were constructed on an
199     underlying FCC lattice (4.09 {\AA}) at the bulk eutectic composition
200     $\mathrm{Ag}_6\mathrm{Cu}_4$. Both initial geometries were considered
201     although experimental results suggest that the random structure is the
202     most likely structure to be found following
203     synthesis.\cite{Jiang:2005lr,gonzalo:5163} Three different sizes of
204     nanoparticles corresponding to a 20 \AA radius (2382 atoms), 30 {\AA}
205     radius (6603 atoms) and 40 {\AA} radius (15683 atoms) were
206     constructed. These initial structures were relaxed to their
207     equilibrium structures at 20 K for 20 ps and again at 300 K for 100 ps
208     sampling from a Maxwell-Boltzmann distribution at each
209     temperature. All simulations were conducted using the {\sc oopse}
210     molecular dynamics package.\cite{Meineke:2004uq}
211    
212     To mimic the effects of the heating due to laser irradiation, the
213     particles were allowed to melt by sampling velocities from the Maxwell
214     Boltzmann distribution at a temperature of 900 K. The particles were
215     run under microcanonical simulation conditions for 1 ns of simualtion
216     time prior to studying the effects of heat transfer to the solvent.
217     In all cases, center of mass translational and rotational motion of
218     the particles were set to zero before any data collection was
219     undertaken. Structural features (pair distribution functions) were
220     used to verify that the particles were indeed liquid droplets before
221     cooling simulations took place.
222    
223     \subsection{MODELING RANDOM ALLOY AND CORE SHELL PARTICLES IN SOLUTION PHASE ENVIRONMENTS}
224    
225    
226    
227     To approximate the effects of rapid heat transfer to the solvent
228     following a heating at the plasmon resonance, we utilized a
229     methodology in which atoms contained in the outer $4$ {\AA} radius of
230     the nanoparticle evolved under Langevin Dynamics,
231     \begin{equation}
232     m \frac{\partial^2 \vec{x}}{\partial t^2} = F_\textrm{sys}(\vec{x}(t))
233     - 6 \pi a \eta \vec{v}(t) + F_\textrm{ran}
234     \label{eq:langevin}
235     \end{equation}
236     with a solvent friction ($\eta$) approximating the contribution from
237     the solvent and capping agent. Atoms located in the interior of the
238     nanoparticle evolved under Newtonian dynamics. The set-up of our
239     simulations is nearly identical with the ``stochastic boundary
240     molecular dynamics'' ({\sc sbmd}) method that has seen wide use in the
241     protein simulation
242     community.\cite{BROOKS:1985kx,BROOKS:1983uq,BRUNGER:1984fj} A sketch
243     of this setup can be found in Fig. \ref{fig:langevinSketch}. In
244     Eq. (\ref{eq:langevin}) the frictional forces of a spherical atom
245     of radius $a$ depend on the solvent viscosity. The random forces are
246     usually taken as gaussian random variables with zero mean and a
247     variance tied to the solvent viscosity and temperature,
248     \begin{equation}
249     \langle F_\textrm{ran}(t) \cdot F_\textrm{ran} (t')
250     \rangle = 2 k_B T (6 \pi \eta a) \delta(t - t')
251     \label{eq:stochastic}
252     \end{equation}
253     Due to the presence of the capping agent and the lack of details about
254     the atomic-scale interactions between the metallic atoms and the
255     solvent, the effective viscosity is a essentially a free parameter
256     that must be tuned to give experimentally relevant simulations.
257     \begin{figure}[htbp]
258     \centering
259     \includegraphics[width=5in]{images/stochbound.pdf}
260     \caption{Methodology used to mimic the experimental cooling conditions
261     of a hot nanoparticle surrounded by a solvent. Atoms in the core of
262     the particle evolved under Newtonian dynamics, while atoms that were
263     in the outer skin of the particle evolved under Langevin dynamics.
264     The radius of the spherical region operating under Newtonian dynamics,
265     $r_\textrm{Newton}$ was set to be 4 {\AA} smaller than the original
266     radius ($R$) of the liquid droplet.}
267     \label{fig:langevinSketch}
268     \end{figure}
269    
270     The viscosity ($\eta$) can be tuned by comparing the cooling rate that
271     a set of nanoparticles experience with the known cooling rates for
272     similar particles obtained via the laser heating experiments.
273     Essentially, we tune the solvent viscosity until the thermal decay
274     profile matches a heat-transfer model using reasonable values for the
275     interfacial conductance and the thermal conductivity of the solvent.
276    
277     Cooling rates for the experimentally-observed nanoparticles were
278     calculated from the heat transfer equations for a spherical particle
279     embedded in a ambient medium that allows for diffusive heat transport.
280     Following Plech {\it et al.},\cite{plech:195423} we use a heat
281     transfer model that consists of two coupled differential equations
282     in the Laplace domain,
283     \begin{eqnarray}
284     Mc_{P}\cdot(s\cdot T_{p}(s)-T_{0})+4\pi R^{2} G\cdot(T_{p}(s)-T_{f}(r=R,s)=0\\
285     \left(\frac{\partial}{\partial r} T_{f}(r,s)\right)_{r=R} +
286     \frac{G}{K}(T_{p}(s)-T_{f}(r,s) = 0
287     \label{eq:heateqn}
288     \end{eqnarray}
289     where $s$ is the time-conjugate variable in Laplace space. The
290     variables in these equations describe a nanoparticle of radius $R$,
291     mass $M$, and specific heat $c_{p}$ at an initial temperature
292     $T_0$. The surrounding solvent has a thermal profile $T_f(r,t)$,
293     thermal conductivity $K$, density $\rho$, and specific heat $c$. $G$
294     is the interfacial conductance between the nanoparticle and the
295     surrounding solvent, and contains information about heat transfer to
296     the capping agent as well as the direct metal-to-solvent heat loss.
297     The temperature of the nanoparticle as a function of time can then
298     obtained by the inverse Laplace transform,
299     \begin{equation}
300     T_{p}(t)=\frac{2 k R^2 g^2
301     T_0}{\pi}\int_{0}^{\infty}\frac{\exp(-\kappa u^2
302     t/R^2)u^2}{(u^2(1 + R g) - k R g)^2+(u^3 - k R g u)^2}\mathrm{d}u.
303     \label{eq:laplacetransform}
304     \end{equation}
305     For simplicity, we have introduced the thermal diffusivity $\kappa =
306     K/(\rho c)$, and defined $k=4\pi R^3 \rho c /(M c_p)$ and $g = G/K$ in
307     Eq. (\ref{eq:laplacetransform}).
308    
309     Eq. (\ref{eq:laplacetransform}) was solved numerically for the Ag-Cu
310     system using mole-fraction weighted values for $c_p$ and $\rho_p$ of
311     0.295 $(\mathrm{J g^{-1} K^{-1}})$ and $9.826\times 10^6$ $(\mathrm{g
312     m^{-3}})$ respectively. Since most of the laser excitation experiments
313     have been done in aqueous solutions, parameters used for the fluid are
314     $K$ of $0.6$ $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$
315     $(\mathrm{g m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
316    
317     Values for the interfacial conductance have been determined by a
318     number of groups for similar nanoparticles and range from a low
319     $87.5\times 10^{6} (\mathrm{Wm^{-2}K^{-1}})$ to $130\times 10^{6}
320     (\mathrm{Wm^{-2}K^{-1}})$.\cite{Wilson:2002uq} Wilson {\it
321     et al.} worked with Au, Pt, and AuPd nanoparticles and obtained an
322     estimate for the interfacial conductance of $G=130
323     (\mathrm{Wm^{-2}K^{-1}})$.\cite{Wilson:2002uq} Similarly, Plech {\it
324     et al.} reported a value for the interfacial conductance of $G=105\pm
325     15 (\mathrm{Wm^{-2}K^{-1}})$ for Au nanoparticles.\cite{plech:195423}
326    
327     We conducted our simulations at both ends of the range of
328     experimentally-determined values for the interfacial conductance.
329     This allows us to observe both the slowest and fastest heat transfers
330     from the nanoparticle to the solvent that are consistent with
331     experimental observations. For the slowest heat transfer, a value for
332     G of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ was used and for
333     the fastest heat transfer, a value of $117\times 10^{6}$
334     $(\mathrm{Wm^{-2}K^{-1}})$ was used. Based on calculations we have
335     done using raw data from the Hartland group's thermal half-time
336     experiments on Au nanospheres,\cite{HuM._jp020581+} the true G values
337     are probably in the faster regime: $117\times 10^{6}$
338     $(\mathrm{Wm^{-2}K^{-1}})$.
339    
340     The rate of cooling for the nanoparticles in a molecular dynamics
341     simulation can then be tuned by changing the effective solvent
342     viscosity ($\eta$) until the nanoparticle cooling rate matches the
343     cooling rate described by the heat-transfer Eq.
344     (\ref{eq:heateqn}). The effective solvent viscosity (in Pa s) for a G
345     of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $4.2 \times
346     10^{-6}$, $5.0 \times 10^{-6}$, and
347     $5.5 \times 10^{-6}$ for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
348     effective solvent viscosity (again in Pa s) for an interfacial
349     conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $5.7
350     \times 10^{-6}$, $7.2 \times 10^{-6}$, and $7.5 \times 10^{-6}$
351     for 20 {\AA}, 30 {\AA} and 40 {\AA} particles. These viscosities are
352     essentially gas-phase values, a fact which is consistent with the
353     initial temperatures of the particles being well into the
354     super-critical region for the aqueous environment. Gas bubble
355     generation has also been seen experimentally around gold nanoparticles
356     in water.\cite{kotaidis:184702} Instead of a single value for the
357     effective viscosity, a time-dependent parameter might be a better
358     mimic of the cooling vapor layer that surrounds the hot particles.
359     This may also be a contributing factor to the size-dependence of the
360     effective viscosities in our simulations.
361    
362     Cooling traces for each particle size are presented in
363     Fig. \ref{fig:images_cooling_plot}. It should be noted that the
364     Langevin thermostat produces cooling curves that are consistent with
365     Newtonian (single-exponential) cooling, which cannot match the cooling
366     profiles from Eq. (\ref{eq:laplacetransform}) exactly. Fitting the
367     Langevin cooling profiles to a single-exponential produces
368     $\tau=25.576$ ps, $\tau=43.786$ ps, and $\tau=56.621$ ps for the 20,
369     30 and 40 {\AA} nanoparticles and a G of $87.5\times 10^{6}$
370     $(\mathrm{Wm^{-2}K^{-1}})$. For comparison's sake, similar
371     single-exponential fits with an interfacial conductance of G of
372     $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ produced a $\tau=13.391$
373     ps, $\tau=30.426$ ps, $\tau=43.857$ ps for the 20, 30 and 40 {\AA}
374     nanoparticles.
375    
376     \begin{figure}[htbp]
377     \centering
378     \includegraphics[width=5in]{images/cooling_plot.pdf}
379     \caption{Thermal cooling curves obtained from the inverse Laplace
380     transform heat model in Eq. (\ref{eq:laplacetransform}) (solid line) as
381     well as from molecular dynamics simulations (circles). Effective
382     solvent viscosities of 4.2-7.5 $\times 10^{-6}$ Pa s (depending on the
383     radius of the particle) give the best fit to the experimental cooling
384     curves. This viscosity suggests that the nanoparticles in these
385     experiments are surrounded by a vapor layer (which is a reasonable
386     assumptions given the initial temperatures of the particles). }
387     \label{fig:images_cooling_plot}
388     \end{figure}
389    
390     \subsection{POTENIALS FOR CLASSICAL SIMULATIONS OF BIMETALLIC NANOPARTICLES}
391    
392     Several different potential models have been developed that reasonably
393     describe interactions in transition metals. In particular, the
394     Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and
395     Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study
396     a wide range of phenomena in both bulk materials and
397     nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq} Both
398     potentials are based on a model of a metal which treats the nuclei and
399     core electrons as pseudo-atoms embedded in the electron density due to
400     the valence electrons on all of the other atoms in the system. The
401     {\sc sc} potential has a simple form that closely resembles that of
402     the ubiquitous Lennard Jones potential,
403     \begin{equation}
404     \label{eq:SCP1}
405     U_{tot}=\sum _{i}\left[ \frac{1}{2}\sum _{j\neq i}D_{ij}V^{pair}_{ij}(r_{ij})-c_{i}D_{ii}\sqrt{\rho_{i}}\right] ,
406     \end{equation}
407     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
408     \begin{equation}
409     \label{eq:SCP2}
410     V^{pair}_{ij}(r)=\left( \frac{\alpha_{ij}}{r_{ij}}\right)^{n_{ij}}, \rho_{i}=\sum_{j\neq i}\left( \frac{\alpha_{ij}}{r_{ij}}\right) ^{m_{ij}}.
411     \end{equation}
412     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
413     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
414     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
415     the interactions between the valence electrons and the cores of the
416     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
417     scale, $c_i$ scales the attractive portion of the potential relative
418     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
419     that assures a dimensionless form for $\rho$. These parameters are
420     tuned to various experimental properties such as the density, cohesive
421     energy, and elastic moduli for FCC transition metals. The quantum
422     Sutton-Chen ({\sc q-sc}) formulation matches these properties while
423     including zero-point quantum corrections for different transition
424     metals.\cite{PhysRevB.59.3527} This particular parametarization has
425     been shown to reproduce the experimentally available heat of mixing
426     data for both FCC solid solutions of Ag-Cu and the high-temperature
427     liquid.\cite{sheng:184203} In contrast, the {\sc eam} potential does
428     not reproduce the experimentally observed heat of mixing for the
429     liquid alloy.\cite{MURRAY:1984lr} In this work, we have utilized the
430     {\sc q-sc} formulation for our potential energies and forces.
431     Combination rules for the alloy were taken to be the arithmetic
432     average of the atomic parameters with the exception of $c_i$ since its
433     values is only dependent on the identity of the atom where the density
434     is evaluated. For the {\sc q-sc} potential, cutoff distances are
435     traditionally taken to be $2\alpha_{ij}$ and include up to the sixth
436     coordination shell in FCC metals.
437    
438     %\subsection{Sampling single-temperature configurations from a cooling
439     %trajectory}
440    
441     To better understand the structural changes occurring in the
442     nanoparticles throughout the cooling trajectory, configurations were
443     sampled at regular intervals during the cooling trajectory. These
444     configurations were then allowed to evolve under NVE dynamics to
445     sample from the proper distribution in phase space. Fig.
446     \ref{fig:images_cooling_time_traces} illustrates this sampling.
447    
448    
449     \begin{figure}[htbp]
450     \centering
451     \includegraphics[height=3in]{images/cooling_time_traces.pdf}
452     \caption{Illustrative cooling profile for the 40 {\AA}
453     nanoparticle evolving under stochastic boundary conditions
454     corresponding to $G=$$87.5\times 10^{6}$
455     $(\mathrm{Wm^{-2}K^{-1}})$. At temperatures along the cooling
456     trajectory, configurations were sampled and allowed to evolve in the
457     NVE ensemble. These subsequent trajectories were analyzed for
458     structural features associated with bulk glass formation.}
459     \label{fig:images_cooling_time_traces}
460     \end{figure}
461    
462    
463     \begin{figure}[htbp]
464     \centering
465     \includegraphics[width=5in]{images/cross_section_array.jpg}
466     \caption{Cutaway views of 30 \AA\ Ag-Cu nanoparticle structures for
467     random alloy (top) and Cu (core) / Ag (shell) initial conditions
468     (bottom). Shown from left to right are the crystalline, liquid
469     droplet, and final glassy bead configurations.}
470     \label{fig:cross_sections}
471     \end{figure}
472    
473     \section{ANALYSIS}
474    
475     Frank first proposed local icosahedral ordering of atoms as an
476     explanation for supercooled atomic (specifically metallic) liquids,
477     and further showed that a 13-atom icosahedral cluster has a 8.4\%
478     higher binding energy the either a face centered cubic ({\sc fcc}) or
479     hexagonal close-packed ({\sc hcp}) crystal structures.\cite{19521106}
480     Icosahedra also have six five-fold symmetry axes that cannot be
481     extended indefinitely in three dimensions; long-range translational
482     order is therefore incommensurate with local icosahedral ordering.
483     This does not preclude icosahedral clusters from possessing long-range
484     {\it orientational} order. The ``frustrated'' packing of these
485     icosahedral structures into dense clusters has been proposed as a
486     model for glass formation.\cite{19871127} The size of the icosahedral
487     clusters is thought to increase until frustration prevents any further
488     growth.\cite{HOARE:1976fk} Molecular dynamics simulations of a
489     two-component Lennard-Jones glass showed that clusters of face-sharing
490     icosahedra are distributed throughout the
491     material.\cite{PhysRevLett.60.2295} Simulations of freezing of single
492     component metalic nanoclusters have shown a tendency for icosohedral
493     structure formation particularly at the surfaces of these
494     clusters.\cite{Gafner:2004bg,PhysRevLett.89.275502,Chen:2004ec}
495     Experimentally, the splitting (or shoulder) on the second peak of the
496     X-ray structure factor in binary metallic glasses has been attributed
497     to the formation of tetrahedra that share faces of adjoining
498     icosahedra.\cite{Waal:1995lr}
499    
500     Various structural probes have been used to characterize structural
501     order in molecular systems including: common neighbor analysis,
502     Voronoi tesselations, and orientational bond-order
503     parameters.\cite{HoneycuttJ.Dana_j100303a014,Iwamatsu:2007lr,hsu:4974,nose:1803}
504     The method that has been used most extensively for determining local
505     and extended orientational symmetry in condensed phases is the
506     bond-orientational analysis formulated by Steinhart {\it et
507     al.}\cite{Steinhardt:1983mo} In this model, a set of spherical
508     harmonics is associated with each of the near neighbors of a central
509     atom. Neighbors (or ``bonds'') are defined as having a distance from
510     the central atom that is within the first peak in the radial
511     distribution function. The spherical harmonic between a central atom
512     $i$ and a neighboring atom $j$ is
513     \begin{equation}
514     Y_{lm}\left(\vec{r}_{ij}\right)=Y_{lm}\left(\theta(\vec{r}_{ij}),\phi(\vec{r}_{ij})\right)
515     \label{eq:spharm}
516     \end{equation}
517     where, $\{ Y_{lm}(\theta,\phi)\}$ are spherical harmonics, and
518     $\theta(\vec{r})$ and $\phi(\vec{r})$ are the polar and azimuthal
519     angles made by the bond vector $\vec{r}$ with respect to a reference
520     coordinate system. We chose for simplicity the origin as defined by
521     the coordinates for our nanoparticle. (Only even-$l$ spherical
522     harmonics are considered since permutation of a pair of identical
523     particles should not affect the bond-order parameter.) The local
524     environment surrounding atom $i$ can be defined by
525     the average over all neighbors, $N_b(i)$, surrounding that atom,
526     \begin{equation}
527     \bar{q}_{lm}(i) = \frac{1}{N_{b}(i)}\sum_{j=1}^{N_b(i)} Y_{lm}(\vec{r}_{ij}).
528     \label{eq:local_avg_bo}
529     \end{equation}
530     We can further define a global average orientational-bond order over
531     all $\bar{q}_{lm}$ by calculating the average of $\bar{q}_{lm}(i)$
532     over all $N$ particles
533     \begin{equation}
534     \bar{Q}_{lm} = \frac{\sum^{N}_{i=1}N_{b}(i)\bar{q}_{lm}(i)}{\sum^{N}_{i=1}N_{b}(i)}
535     \label{eq:sys_avg_bo}
536     \end{equation}
537     The $\bar{Q}_{lm}$ contained in Eq. (\ref{eq:sys_avg_bo}) is not
538     necessarily invariant under rotations of the arbitrary reference
539     coordinate system. Second- and third-order rotationally invariant
540     combinations, $Q_l$ and $W_l$, can be taken by summing over $m$ values
541     of $\bar{Q}_{lm}$,
542     \begin{equation}
543     Q_{l} = \left( \frac{4\pi}{2 l+1} \sum^{l}_{m=-l} \left| \bar{Q}_{lm} \right|^2\right)^{1/2}
544     \label{eq:sec_ord_inv}
545     \end{equation}
546     and
547     \begin{equation}
548     \hat{W}_{l} = \frac{W_{l}}{\left( \sum^{l}_{m=-l} \left| \bar{Q}_{lm} \right|^2\right)^{3/2}}
549     \label{eq:third_ord_inv}
550     \end{equation}
551     where
552     \begin{equation}
553     W_{l} = \sum_{\substack{m_1,m_2,m_3 \\m_1 + m_2 + m_3 = 0}} \left( \begin{array}{ccc} l & l & l \\ m_1 & m_2 & m_3 \end{array}\right) \\ \times \bar{Q}_{lm_1}\bar{Q}_{lm_2}\bar{Q}_{lm_3}.
554     \label{eq:third_inv}
555     \end{equation}
556     The factor in parentheses in Eq. (\ref{eq:third_inv}) is the Wigner-3$j$
557     symbol, and the sum is over all valid ($|m| \leq l$) values of $m_1$,
558     $m_2$, and $m_3$ which sum to zero.
559    
560     \begin{table}
561     \caption{Values of bond orientational order parameters for
562     simple structures cooresponding to $l=4$ and $l=6$ spherical harmonic
563     functions.\cite{wolde:9932} $Q_4$ and $\hat{W}_4$ have values for {\it
564     individual} icosahedral clusters, but these values are not invariant
565     under rotations of the reference coordinate systems. Similar behavior
566     is observed in the bond-orientational order parameters for individual
567     liquid-like structures.}
568     \begin{center}
569     \begin{tabular}{ccccc}
570     \hline
571     \hline
572     & $Q_4$ & $Q_6$ & $\hat{W}_4$ & $\hat{W}_6$\\
573    
574     fcc & 0.191 & 0.575 & -0.159 & -0.013\\
575    
576     hcp & 0.097 & 0.485 & 0.134 & -0.012\\
577    
578     bcc & 0.036 & 0.511 & 0.159 & 0.013\\
579    
580     sc & 0.764 & 0.354 & 0.159 & 0.013\\
581    
582     Icosahedral & - & 0.663 & - & -0.170\\
583    
584     (liquid) & - & - & - & -\\
585     \hline
586     \end{tabular}
587     \end{center}
588     \label{table:bopval}
589     \end{table}
590    
591     For ideal face-centered cubic ({\sc fcc}), body-centered cubic ({\sc
592     bcc}) and simple cubic ({\sc sc}) as well as hexagonally close-packed
593     ({\sc hcp}) structures, these rotationally invariant bond order
594     parameters have fixed values independent of the choice of coordinate
595     reference frames. For ideal icosahedral structures, the $l=6$
596     invariants, $Q_6$ and $\hat{W}_6$ are also independent of the
597     coordinate system. $Q_4$ and $\hat{W}_4$ will have non-vanishing
598     values for {\it individual} icosahedral clusters, but these values are
599     not invariant under rotations of the reference coordinate systems.
600     Similar behavior is observed in the bond-orientational order
601     parameters for individual liquid-like structures. Additionally, both
602     $Q_6$ and $\hat{W}_6$ are thought to have extreme values for the
603     icosahedral clusters.\cite{Steinhardt:1983mo} This makes the $l=6$
604     bond-orientational order parameters particularly useful in identifying
605     the extent of local icosahedral ordering in condensed phases. For
606     example, a local structure which exhibits $\hat{W}_6$ values near
607     -0.17 is easily identified as an icosahedral cluster and cannot be
608     mistaken for distorted cubic or liquid-like structures.
609    
610     One may use these bond orientational order parameters as an averaged
611     property to obtain the extent of icosahedral ordering in a supercooled
612     liquid or cluster. It is also possible to accumulate information
613     about the {\it distributions} of local bond orientational order
614     parameters, $p(\hat{W}_6)$ and $p(Q_6)$, which provide information
615     about individual atomic sites that are central to local icosahedral
616     structures.
617    
618     The distributions of atomic $Q_6$ and $\hat{W}_6$ values are plotted
619     as a function of temperature for our nanoparticles in Fig.
620     \ref{fig:q6} and \ref{fig:w6}. At high temperatures, the
621     distributions are unstructured and are broadly distributed across the
622     entire range of values. As the particles are cooled, however, there
623     is a dramatic increase in the fraction of atomic sites which have
624     local icosahedral ordering around them. (This corresponds to the
625     sharp peak appearing in Fig. \ref{fig:w6} at $\hat{W}_6=-0.17$ and
626     to the broad shoulder appearing in Fig. \ref{fig:q6} at $Q_6 =
627     0.663$.)
628    
629     \begin{figure}[htbp]
630     \centering
631     \includegraphics[width=5in]{images/w6_stacked_plot.pdf}
632     \caption{Distributions of the bond orientational order parameter
633     ($\hat{W}_6$) at different temperatures. The upper, middle, and lower
634     panels are for 20, 30, and 40 \AA\ particles, respectively. The
635     left-hand column used cooling rates commensurate with a low
636     interfacial conductance ($87.5 \times 10^{6}$
637     $\mathrm{Wm^{-2}K^{-1}}$), while the right-hand column used a more
638     physically reasonable value of $117 \times 10^{6}$
639     $\mathrm{Wm^{-2}K^{-1}}$. The peak at $\hat{W}_6 \approx -0.17$ is
640     due to local icosahedral structures. The different curves in each of
641     the panels indicate the distribution of $\hat{W}_6$ values for samples
642     taken at different times along the cooling trajectory. The initial
643     and final temperatures (in K) are indicated on the plots adjacent to
644     their respective distributions.}
645     \label{fig:w6}
646     \end{figure}
647    
648     \begin{figure}[htbp]
649     \centering
650     \includegraphics[width=5in]{images/q6_stacked_plot.pdf}
651     \caption{Distributions of the bond orientational order parameter
652     ($Q_6$) at different temperatures. The curves in the six panels in
653     this figure were computed at identical conditions to the same panels in
654     figure \ref{fig:w6}.}
655     \label{fig:q6}
656     \end{figure}
657    
658     The probability distributions of local order can be used to generate
659     free energy surfaces using the local orientational ordering as a
660     reaction coordinate. By making the simple statistical equivalence
661     between the free energy and the probabilities of occupying certain
662     states,
663     \begin{equation}
664     g(\hat{W}_6) = - k_B T \ln p(\hat{W}_6),
665     \end{equation}
666     we can obtain a sequence of free energy surfaces (as a function of
667     temperature) for the local ordering around central atoms within our
668     particles. Free energy surfaces for the 40 \AA\ particle at a range
669     of temperatures are shown in Fig. \ref{fig:freeEnergy}. Note that
670     at all temperatures, the liquid-like structures are global minima on
671     the free energy surface, while the local icosahedra appear as local
672     minima once the temperature has fallen below 528 K. As the
673     temperature falls, it is possible for substructures to become trapped
674     in the local icosahedral well, and if the cooling is rapid enough,
675     this trapping leads to vitrification. A similar analysis of the free
676     energy surface for orientational order in bulk glass formers can be
677     found in the work of van~Duijneveldt and
678     Frenkel.\cite{duijneveldt:4655}
679    
680    
681     \begin{figure}[htbp]
682     \centering
683     \includegraphics[width=5in]{images/freeEnergyVsW6.pdf}
684     \caption{Free energy as a function of the orientational order
685     parameter ($\hat{W}_6$) for 40 {\AA} bimetallic nanoparticles as they
686     are cooled from 902 K to 310 K. As the particles cool below 528 K, a
687     local minimum in the free energy surface appears near the perfect
688     icosahedral ordering ($\hat{W}_6 = -0.17$). At all temperatures,
689     liquid-like structures are a global minimum on the free energy
690     surface, but if the cooling rate is fast enough, substructures
691     may become trapped with local icosahedral order, leading to the
692     formation of a glass.}
693     \label{fig:freeEnergy}
694     \end{figure}
695    
696     We have also calculated the fraction of atomic centers which have
697     strong local icosahedral order:
698     \begin{equation}
699     f_\textrm{icos} = \int_{-\infty}^{w_i} p(\hat{W}_6) d \hat{W}_6
700     \label{eq:ficos}
701     \end{equation}
702     where $w_i$ is a cutoff value in $\hat{W}_6$ for atomic centers that
703     are displaying icosahedral environments. We have chosen a (somewhat
704     arbitrary) value of $w_i= -0.15$ for the purposes of this work. A
705     plot of $f_\textrm{icos}(T)$ as a function of temperature of the
706     particles is given in Fig. \ref{fig:ficos}. As the particles cool,
707     the fraction of local icosahedral ordering rises smoothly to a plateau
708     value. The smaller particles (particularly the ones that were cooled
709     in a higher viscosity solvent) show a slightly larger tendency towards
710     icosahedral ordering.
711    
712     \begin{figure}[htbp]
713     \centering
714     \includegraphics[width=5in]{images/fraction_icos.pdf}
715     \caption{Temperautre dependence of the fraction of atoms with local
716     icosahedral ordering, $f_\textrm{icos}(T)$ for 20, 30, and 40 \AA\
717     particles cooled at two different values of the interfacial
718     conductance.}
719     \label{fig:ficos}
720     \end{figure}
721    
722     Since we have atomic-level resolution of the local bond-orientational
723     ordering information, we can also look at the local ordering as a
724     function of the identities of the central atoms. In figure
725     \ref{fig:AgVsCu} we display the distributions of $\hat{W}_6$ values
726     for both the silver and copper atoms, and we note a strong
727     predilection for the copper atoms to be central to icosahedra. This
728     is probably due to local packing competition of the larger silver
729     atoms around the copper, which would tend to favor icosahedral
730     structures over the more densely packed cubic structures.
731    
732     \begin{figure}[htbp]
733     \centering
734     \includegraphics[width=5in]{images/w6_stacked_bytype_plot.pdf}
735     \caption{Distributions of the bond orientational order parameter
736     ($\hat{W}_6$) for the two different elements present in the
737     nanoparticles. This distribution was taken from the fully-cooled 40
738     \AA\ nanoparticle. Local icosahedral ordering around copper atoms is
739     much more prevalent than around silver atoms.}
740     \label{fig:AgVsCu}
741     \end{figure}
742    
743     The locations of these icosahedral centers are not uniformly
744     distrubted throughout the particles. In Fig. \ref{fig:icoscluster}
745     we show snapshots of the centers of the local icosahedra (i.e. any
746     atom which exhibits a local bond orientational order parameter
747     $\hat{W}_6 < -0.15$). At high temperatures, the icosahedral centers
748     are transitory, existing only for a few fs before being reabsorbed
749     into the liquid droplet. As the particle cools, these centers become
750     fixed at certain locations, and additional icosahedra develop
751     throughout the particle, clustering around the sites where the
752     structures originated. There is a strong preference for icosahedral
753     ordering near the surface of the particles. Identification of these
754     structures by the type of atom shows that the silver-centered
755     icosahedra are evident only at the surface of the particles.
756    
757     \begin{figure}[htbp]
758     \centering
759     \begin{tabular}{c c c}
760     \includegraphics[width=2.1in]{images/Cu_Ag_random_30A_liq_icosonly.pdf}
761     \includegraphics[width=2.1in]{images/Cu_Ag_random_30A__0007_icosonly.pdf}
762     \includegraphics[width=2.1in]{images/Cu_Ag_random_30A_glass_icosonly.pdf}
763     \end{tabular}
764     \caption{Centers of local icosahedral order ($\hat{W}_6<0.15$) at 900
765     K, 471 K and 315 K for the 30 \AA\ nanoparticle cooled with an
766     interfacial conductance $G = 87.5 \times 10^{6}$
767     $\mathrm{Wm^{-2}K^{-1}}$. Silver atoms (blue) exhibit icosahedral
768     order at the surface of the nanoparticle while copper icosahedral
769     centers (green) are distributed throughout the nanoparticle. The
770     icosahedral centers appear to cluster together and these clusters
771     increase in size with decreasing temperature.}
772     \label{fig:icoscluster}
773     \end{figure}
774    
775     In contrast with the silver ordering behavior, the copper atoms which
776     have local icosahedral ordering are distributed more evenly throughout
777     the nanoparticles. Fig. \ref{fig:Surface} shows this tendency as a
778     function of distance from the center of the nanoparticle. Silver,
779     since it has a lower surface free energy than copper, tends to coat
780     the skins of the mixed particles.\cite{Zhu:1997lr} This is true even
781     for bimetallic particles that have been prepared in the Ag (core) / Cu
782     (shell) configuration. Upon forming a liquid droplet, approximately 1
783     monolayer of Ag atoms will rise to the surface of the particles. This
784     can be seen visually in Fig. \ref{fig:cross_sections} as well as in
785     the density plots in the bottom panel of Fig. \ref{fig:Surface}.
786     This observation is consistent with previous experimental and
787     theoretical studies on bimetallic alloys composed of noble
788     metals.\cite{MainardiD.S._la0014306,HuangS.-P._jp0204206,Ramirez-Caballero:2006lr}
789     Bond order parameters for surface atoms are averaged only over the
790     neighboring atoms, so packing constraints that may prevent icosahedral
791     ordering around silver in the bulk are removed near the surface. It
792     would certainly be interesting to see if the relative tendency of
793     silver and copper to form local icosahedral structures in a bulk glass
794     differs from our observations on nanoparticles.
795    
796     \begin{figure}[htbp]
797     \centering
798     \includegraphics[width=5in]{images/dens_fracr_stacked_plot.pdf}
799     \caption{Appearance of icosahedral clusters around central silver atoms
800     is largely due to the presence of these silver atoms at or near the
801     surface of the nanoparticle. The upper panel shows the fraction of
802     icosahedral atoms ($f_\textrm{icos}(r)$ for each of the two metallic
803     atoms as a function of distance from the center of the nanoparticle
804     ($r$). The lower panel shows the radial density of the two
805     constituent metals (relative to the overall density of the
806     nanoparticle). Icosahedral clustering around copper atoms are more
807     evenly distributed throughout the particle, while icosahedral
808     clustering around silver is largely confined to the silver atoms at
809     the surface.}
810     \label{fig:Surface}
811     \end{figure}
812    
813     The methods used by Sheng, He, and Ma to estimate the glass transition
814     temperature, $T_g$, in bulk Ag-Cu alloys involve finding
815     discontinuities in the slope of the average atomic volume, $\langle V
816     \rangle / N$, or enthalpy when plotted against the temperature of the
817     alloy. They obtained a bulk glass transition temperature, $T_g$ = 510
818     K for a quenching rate of $2.5 \times 10^{13}$ K/s.
819    
820     For simulations of nanoparticles, there is no periodic box, and
821     therefore, no easy way to compute the volume exactly. Instead, we
822     estimate the volume of our nanoparticles using Barber {\it et al.}'s
823     very fast quickhull algorithm to obtain the convex hull for the
824     collection of 3-d coordinates of all of atoms at each point in
825     time.~\cite{barber96quickhull,qhull} The convex hull is the smallest convex
826     polyhedron which includes all of the atoms, so the volume of this
827     polyhedron is an excellent estimate of the volume of the nanoparticle.
828     This method of estimating the volume will be problematic if the
829     nanoparticle breaks into pieces (i.e. if the bounding surface becomes
830     concave), but for the relatively short trajectories used in this
831     study, it provides an excellent measure of particle volume as a
832     function of time (and temperature).
833    
834     Using the discontinuity in the slope of the average atomic volume
835     vs. temperature, we arrive at an estimate of $T_g$ that is
836     approximately 488 K. We note that this temperature is somewhat below
837     the onset of icosahedral ordering exhibited in the bond orientational
838     order parameters. It appears that icosahedral ordering sets in while
839     the system is still somewhat fluid, and is locked in place once the
840     temperature falls below $T_g$. We did not observe any dependence of
841     our estimates for $T_g$ on either the nanoparticle size or the value
842     of the interfacial conductance. However, the cooling rates and size
843     ranges we utilized are all sampled from a relatively narrow range, and
844     it is possible that much larger particles would have substantially
845     different values for $T_g$. Our estimates for the glass transition
846     temperatures for all three particle sizes and both interfacial
847     conductance values are shown in table \ref{table:Tg}.
848    
849     \begin{table}
850     \caption{Estimates of the glass transition temperatures $T_g$ for
851     three different sizes of bimetallic Ag$_6$Cu$_4$ nanoparticles cooled
852     under two different values of the interfacial conductance, $G$.}
853     \begin{center}
854     \begin{tabular}{ccccc}
855     \hline
856     \hline
857     Radius (\AA\ ) & Interfacial conductance & Effective cooling rate
858     (K/s $\times 10^{13}$) & & $T_g$ (K) \\
859     20 & 87.5 & 2.4 & 477 \\
860     20 & 117 & 4.5 & 502 \\
861     30 & 87.5 & 1.3 & 491 \\
862     30 & 117 & 1.9 & 493 \\
863     40 & 87.5 & 1.0 & 476 \\
864     40 & 117 & 1.3 & 487 \\
865     \hline
866     \end{tabular}
867     \end{center}
868     \label{table:Tg}
869     \end{table}
870    
871     \section{CONCLUSIONS}
872     \label{metglass:sec:conclusion}
873    
874     Our heat-transfer calculations have utilized the best current
875     estimates of the interfacial heat transfer coefficient (G) from recent
876     experiments. Using reasonable values for thermal conductivity in both
877     the metallic particle and the surrounding solvent, we have obtained
878     cooling rates for laser-heated nanoparticles that are in excess of
879     10$^{13}$ K / s. To test whether or not this cooling rate can form
880     glassy nanoparticles, we have performed a mixed molecular dynamics
881     simulation in which the atoms in contact with the solvent or capping
882     agent are evolved under Langevin dynamics while the remaining atoms
883     are evolved under Newtonian dynamics. The effective solvent viscosity
884     ($\eta$) is a free parameter which we have tuned so that the particles
885     in the simulation follow the same cooling curve as their experimental
886     counterparts. From the local icosahedral ordering around the atoms in
887     the nanoparticles (particularly Copper atoms), we deduce that it is
888     likely that glassy nanobeads are created via laser heating of
889     bimetallic nanoparticles, particularly when the initial temperature of
890     the particles approaches the melting temperature of the bulk metal
891     alloy.
892    
893     Improvements to our calculations would require: 1) explicit treatment
894     of the capping agent and solvent, 2) another radial region to handle
895     the heat transfer to the solvent vapor layer that is likely to form
896     immediately surrounding the hot
897     particle,\cite{Hu:2004lr,kotaidis:184702} and 3) larger particles in
898     the size range most easily studied via laser heating experiments.
899    
900     The local icosahedral ordering we observed in these bimetallic
901     particles is centered almost completely around the copper atoms, and
902     this is likely due to the size mismatch leading to a more efficient
903     packing of 5-membered rings of silver around a central copper atom.
904     This size mismatch should be reflected in bulk calculations, and work
905     is ongoing in our lab to confirm this observation in bulk
906     glass-formers.
907    
908     The physical properties (bulk modulus, frequency of the breathing
909     mode, and density) of glassy nanobeads should be somewhat different
910     from their crystalline counterparts. However, observation of these
911     differences may require single-particle resolution of the ultrafast
912     vibrational spectrum of one particle both before and after the
913     crystallite has been converted into a glassy bead.