ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chuckDissertation/bulkmod.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: 24630 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:bulkmod}BREATHING MODE DYNAMICS AND ELASTIC PROPERTIES OF GOLD NANOPARTICLES}
3    
4     In metallic nanoparticles, the relatively large surface area to volume ratio
5     induces a number of well-known size-dependent phenomena. Notable
6     among these are the depression of the bulk melting
7     temperature,\cite{Buffat:1976yq,el-sayed00,el-sayed01} surface melting
8     transitions, increased room-temperature alloying
9     rates,\cite{ShibataT._ja026764r} changes in the breathing mode
10     frequencies,\cite{delfatti99,henglein99,hartland02a,hartland02c} and
11     rapid heat transfer to the surrounding
12     medium.\cite{HuM._jp020581+,hartland02d}
13    
14     This paper reports on atomic-level simulations of the transient
15     response of metallic nanoparticles to the nearly instantaneous heating
16     undergone when photons are absorbed during ultrafast laser excitation
17     experiments. The time scale for heating (determined by the {\it e-ph}
18     coupling constant) is faster than a single period of the breathing
19     mode for spherical nanoparticles.\cite{Simon2001,HartlandG.V._jp0276092}
20     Hot-electron pressure and direct lattice heating contribute to the
21     thermal excitation of the atomic degrees of freedom, and both
22     mechanisms are rapid enough to coherently excite the breathing mode of
23     the spherical particles.\cite{Hartland00}
24    
25     There are questions posed by the experiments that may be easiest to
26     answer via computer simulation. For example, the dephasing seen
27     following coherent excitation of the breathing mode may be due to
28     inhomogeneous size distributions in the sample, but it may also be due
29     to softening of the breathing mode vibrational frequency following a
30     melting transition. Additionally, there are properties (such as the
31     bulk modulus) that may be nearly impossible to obtain experimentally,
32     but which are relatively easily obtained via simulation techniques.
33    
34     We outline our simulation techniques in section \ref{bulkmod:sec:details}.
35     Results are presented in section \ref{bulkmod:sec:results}. We discuss our
36     results in terms of Lamb's classical theory of elastic spheres in
37     section \ref{bulkmod:sec:discussion}.
38    
39     \section{COMPUTATIONAL DETAILS}
40     \label{bulkmod:sec:details}
41    
42     Spherical Au nanoparticles were created in a standard FCC lattice at
43     four different radii [20{\AA} (1926 atoms), 25{\AA} (3884 atoms),
44     30{\AA} (6602 atoms), and 35{\AA} (10606 atoms)]. To create spherical
45     nanoparticles, a large FCC lattice was built at the normal Au lattice
46     spacing (4.08 \AA) and any atoms outside the target radius were
47     excluded from the simulation.
48    
49     \subsection{SIMULATION METHODOLOGY}
50    
51     Potentials were calculated using the Voter-Chen
52     parameterization~\cite{Voter:87} of the Embedded Atom Method ({\sc
53     eam}), which has been widely used for MD simulations of metallic
54     particles.%\cite{Voter:87,Daw84,FBD86,johnson89,Lu97}
55     Like other transition metal potentials,%\cite{Finnis84,Ercolessi88,Chen90,Qi99,Ercolessi02}
56     {\sc eam} describes bonding in metallic systems by including an
57     attractive interaction which models the stabilization of a positively
58     charged metal core ion in a sea of surrounding valence electrons. A
59     repulsive pair potential describes the interactions of the core ions
60     with each other. The {\sc eam} potential has the form:
61     \begin{eqnarray}
62     V & = & \sum_{i} F_{i}\left[\rho_{i}\right] + \sum_{i} \sum_{j \neq i}
63     \phi_{ij}({\bf r}_{ij}) \\
64     \rho_{i} & = & \sum_{j \neq i} f_{j}({\bf r}_{ij})
65     \end{eqnarray}
66     where $\phi_{ij}$ is the (primarily repulsive) pairwise interaction
67     between atoms $i$ and $j$. $F_{i}[\rho_{i}]$ is an embedding function
68     that determines the energy to embed the positively-charged core, $i$,
69     in the electron density, $\rho_{i}$, due to the valence electrons of
70     the surrounding $j$ atoms, and $f_{j}(r)$ describes the radial
71     dependence of the field due to atom $j$. There is a cutoff distance,
72     $r_{cut}$, which limits the summations in the {\sc eam} equation to
73     the few dozen atoms surrounding atom $i$. In these simulations, a
74     cutoff radius of 10~\AA\ was used.
75    
76     %Mixing rules as outlined by Johnson~\cite{johnson89} were used to
77     %compute the heterogenous pair potential,
78     %
79     %\begin{eqnarray}
80     %\label{eq:johnson}
81     %\phi_{ab}(r)=\frac{1}{2}\left(
82     %\frac{f_{b}(r)}{f_{a}(r)}\phi_{aa}(r)+
83     %\frac{f_{a}(r)}{f_{b}(r)}\phi_{bb}(r)
84     %\right).
85     %\end{eqnarray}
86    
87     Before starting the molecular dynamics runs, a relatively short
88     steepest-descent minimization was performed to relax the lattice in
89     the initial configuration. To facilitate the study of the larger
90     particles, simulations were run in parallel over 16 processors using
91     Plimpton's force-decomposition method.\cite{plimpton93}
92    
93     To mimic the events following the absorption of light in the ultrafast
94     laser heating experiments, we have used a simple two-step process to
95     prepare the simulations. Instantaneous heating of the lattice was
96     performed by sampling atomic velocities from a Maxwell-Boltzmann
97     distribution set to twice the target temperature for the simulation.
98     By equipartition, approximately half of the initial kinetic energy of
99     the system winds up in the potential energy of the system. The system
100     was a allowed a very short (10 fs) evolution period with the new
101     velocities.
102    
103     Following this excitation step, the particles evolved under
104     Nos\'{e}-Hoover NVT dynamics~\cite{hoover85} for 40 ps. Given the
105     mass of the constituent metal atoms, time steps of 5 fs give excellent
106     energy conservation in standard NVE integrators, so the same time step
107     was used in the NVT simulations. Target temperatures for these
108     particles spanned the range from 300 K to 1350 K in 50 K intervals.
109     Five independent samples were run for each particle and temperature.
110     The results presented below are averaged properties for each of the
111     five independent samples.
112    
113     \subsection{ANALYSIS}
114    
115     Of primary interest when comparing our simulations to experiments is
116     the dynamics of the low-frequency breathing mode of the particles. To
117     study this motion, we need access to accurate measures of both the
118     volume and surface area as a function of time. Throughout the
119     simulations, we monitored both quantities using Barber {\it et al.}'s
120     very fast quickhull algorithm to obtain the convex hull for the
121     collection of 3-d coordinates of all of atoms at each point in
122     time.~\cite{barber96quickhull,qhull} The convex hull is the smallest convex
123     polyhedron which includes all of the atoms, so the volume of this
124     polyhedron is an excellent estimate of the volume of the nanoparticle.
125     The convex hull estimate of the volume will be problematic if the
126     nanoparticle breaks into pieces (i.e. if the bounding surface becomes
127     concave), but for the relatively short trajectories of this study, it
128     provides an excellent measure of particle volume as a function of
129     time.
130    
131     The bulk modulus, which is the inverse of the compressibility,
132     \begin{equation}
133     K = \frac{1}{\kappa} = - V \left(\frac{\partial P}{\partial
134     V}\right)_{T}
135     \label{eq:Kdef}
136     \end{equation}
137     can be related to quantities that are relatively easily accessible
138     from our molecular dynamics simulations. We present here four
139     different approaches for estimating the bulk modulus directly from
140     basic or derived quantities:
141    
142     \begin{itemize}
143     \item The traditional ``Energetic'' approach
144    
145     Using basic thermodynamics and one of the Maxwell relations, we write
146     \begin{equation}
147     P = T\left(\frac{\partial S}{\partial V}\right)_{T} -
148     \left(\frac{\partial U}{\partial V}\right)_{T}.
149     \label{eq:Pdef}
150     \end{equation}
151     It follows that
152     \begin{equation}
153     K = V \left(T \left(\frac{\partial^{2} S}{\partial V^{2}}\right)_{T} -
154     \left(\frac{\partial^{2} U}{\partial V^{2}}\right)_{T} \right).
155     \label{eq:Ksub}
156     \end{equation}
157     The standard practice in solid state physics is to assume the low
158     temperature limit ({\it i.e.} to neglect the entropic term), which
159     means the bulk modulus is usually expressed
160     \begin{equation}
161     K \approx V \left(\frac{\partial^{2} U }{\partial V^{2}}\right)_{T}.
162     \label{eq:Kuse}
163     \end{equation}
164     When the relationship between the total energy $U$ and volume $V$ of
165     the system are available at a fixed temperature (as it is in these
166     simulations), it is a simple matter to compute the bulk modulus from
167     the response of the system to the perturbation of the instantaneous
168     heating. Although this information would, in theory, be available
169     from longer constant temperature simulations, the ranges of volumes
170     and energies explored by a nanoparticle under equilibrium conditions
171     are actually quite small. Instantaneous heating, since it excites
172     coherent oscillations in the breathing mode, allows us to sample a
173     much wider range of volumes (and energies) for the particles. The
174     problem with this method is that it neglects the entropic term near
175     the melting transition, which gives spurious results (negative bulk
176     moduli) at higher temperatures.
177    
178     \item The Linear Response Approach
179    
180     Linear response theory gives us another approach to calculating the
181     bulk modulus. This method relates the low-wavelength fluctuations in
182     the density to the isothermal compressibility,\cite{BernePecora}
183     \begin{equation}
184     \underset{k \rightarrow 0}{\lim} \langle | \delta \rho(\vec{k}) |^2
185     \rangle = \frac{k_B T \rho^2 \kappa}{V}
186     \end{equation}
187     where the frequency dependent density fluctuations are the Fourier
188     transform of the spatial fluctuations,
189     \begin{equation}
190     \delta \rho(\vec{k}) = \underset{V}{\int} e^{i \vec{k}\cdot\vec{r}} \left( \rho(\vec{r}, t) - \langle \rho \rangle \right) dV
191     \end{equation}
192     This approach is essentially equivalent to using the volume
193     fluctuations directly to estimate the bulk modulus,
194     \begin{equation}
195     K = \frac{V k_B T}{\langle \delta V^2 \rangle} = k_B T \frac{\langle V \rangle}{\langle V^2 \rangle - \langle V \rangle^2}
196     \end{equation}
197     It should be noted that in these experiments, the particles are {\it
198     far from equilibrium}, and so a linear response approach will not be
199     the most suitable way to obtain estimates of the bulk modulus.
200    
201     \item The Extended System Approach
202    
203     Since we are performing these simulations in the NVT ensemble using
204     Nos\'e-Hoover thermostatting, the quantity conserved by our integrator
205     ($H_{NVT}$) can be expressed as:
206     \begin{equation}
207     H_{NVT} = U + f k_B T_{ext} \left( \frac{\tau_T^2 \chi(t)^2}{2} +
208     \int_0^t \chi(s) ds \right).
209     \end{equation}
210     Here $f$ is the number of degrees of freedom in the (real) system,
211     $T_{ext}$ is the temperature of the thermostat, $\tau_T$ is the time
212     constant for the thermostat, and $\chi(t)$ is the instantaneous value
213     of the extended system thermostat variable. The extended Hamiltonian
214     for the system, $H_{NVT}$ is, to within a constant, the Helmholtz free
215     energy.\cite{melchionna93} Since the pressure is a simple derivative
216     of the Helmholtz free energy,
217     \begin{equation}
218     P = -\left( \frac{\partial A}{\partial V} \right)_T ,
219     \end{equation}
220     the bulk modulus can be obtained (theoretically) by a quadratic fit of
221     the fluctuations in $H_{NVT}$ against fluctuations in the volume,
222     \begin{equation}
223     K = -V \left( \frac{\partial^2 H_{NVT}}{\partial V^2} \right)_T.
224     \end{equation}
225     However, $H_{NVT}$ is essentially conserved during these simulations,
226     so fitting fluctuations of this quantity to obtain meaningful physical
227     quantities is somewhat suspect. We also note that this method would
228     fail in periodic systems because the volume itself is fixed in
229     periodic NVT simulations.
230    
231     \item The Direct Pressure Approach
232    
233     Our preferred method for estimating the bulk modulus is to compute it
234     {\it directly} from the internal pressure in the nanoparticle. The
235     pressure is obtained via the trace of the pressure tensor,
236     \begin{equation}
237     P = \frac{1}{3}
238     \mathrm{Tr}\left[\overleftrightarrow{\mathsf{P}}\right],
239     \end{equation}
240     which has a kinetic contribution as well as a contribution from the
241     stress tensor ($\overleftrightarrow{\mathsf{W}}$):
242     \begin{equation}
243     \overleftrightarrow{\mathsf{P}} = \frac{1}{V} \left(
244     \sum_{i=1}^{N} m_i \vec{v}_i \otimes \vec{v}_i \right) +
245     \overleftrightarrow{\mathsf{W}}.
246     \end{equation}
247     Here the $\otimes$ symbol represents the {\it outer} product of the
248     velocity vector for atom $i$ to yield a $3 \times 3$ matrix. The
249     virial is computed during the simulation using forces between pairs of
250     particles,
251     \begin{equation}
252     \overleftrightarrow{\mathsf{W}} = \sum_{i} \sum_{j>i} \vec{r}_{ij} \otimes \vec{f}_{ij}
253     \end{equation}
254     During the simulation, we record the internal pressure, $P$, as well as
255     the total energy, $U$, the extended system's hamiltonian, $H_{NVT}$,
256     and the particle coordinates. Once we have calculated the time
257     dependent volume of the nanoparticle using the convex hull, we can use
258     any of these four methods to estimate the bulk modulus.
259     \end{itemize}
260    
261     We find, however, that only the fourth method (the direct pressure
262     approach) gives meaningful results. Bulk moduli for the 35 \AA\
263     particle were computed with the traditional (Energy vs. Volume)
264     approach as well as the direct pressure approach. A comparison of the
265     Bulk Modulus obtained via both methods and are shown in
266     Fig. \ref{fig:Methods} Note that the second derivative fits in the
267     traditional approach can give (in the liquid droplet region) negative
268     curvature, and this results in negative values for the bulk modulus.
269    
270     \begin{figure}[htbp]
271     \centering
272     \includegraphics[height=3in]{images/Methods.pdf}
273     \caption{Comparison of two of the methods for estimating the bulk
274     modulus as a function of temperature for the 35\AA\ particle.}
275     \label{fig:Methods}
276     \end{figure}
277    
278     The Bulk moduli reported in the rest of this paper were computed using
279     the direct pressure method.
280    
281     To study the frequency of the breathing mode, we have calculated the
282     power spectrum for volume ($V$) fluctuations,
283     \begin{equation}
284     \rho_{\Delta V}(\omega) = \int_{-\infty}^{\infty} \langle \Delta V(t)
285     \Delta V(0) \rangle e^{-i \omega t} dt
286     \label{eq:volspect}
287     \end{equation}
288     where $\Delta V(t) = V(t) - \langle V \rangle$. Because the
289     instantaneous heating excites all of the vibrational modes of the
290     particle, the power spectrum will contain contributions from all modes
291     that perturb the total volume of the particle. The lowest frequency
292     peak in the power spectrum should give the frequency (and period) for
293     the breathing mode, and these quantities are most readily compared
294     with the Hartland group experiments.\cite{HartlandG.V._jp0276092} Further
295     analysis of the breathing dynamics follows in section
296     \ref{bulkmod:sec:discussion}.
297    
298     We have also computed the heat capacity for our simulations to verify
299     the location of the melting transition. Calculations of the heat
300     capacity were performed on the non-equilibrium, instantaneous heating
301     simulations, as well as on simulations of nanoparticles that were at
302     equilibrium at the target temperature.
303    
304     \section{RESULTS}
305     \label{bulkmod:sec:results}
306    
307     \subsection{THE BULK MODULUS AND HEAT CAPACITY}
308    
309     The upper panel in Fig. \ref{fig:BmCp} shows the temperature
310     dependence of the Bulk Modulus ($K$). In all samples, there is a
311     dramatic (size-dependent) drop in $K$ at temperatures well below the
312     melting temperature of bulk polycrystalline gold. This drop in $K$
313     coincides with the actual melting transition of the nanoparticles.
314     Surface melting has been confirmed at even lower temperatures using
315     the radial-dependent density, $\rho(r) / \rho$, which shows a merging
316     of the crystalline peaks in the outer layer of the nanoparticle.
317     However, the bulk modulus only has an appreciable drop when the
318     particle melts fully.
319    
320     \begin{figure}[htbp]
321     \centering
322     \includegraphics[height=3in]{images/Stacked_Bulk_modulus_and_Cp.pdf}
323     \caption{The temperature dependence of the bulk modulus (upper panel)
324     and heat capacity (lower panel) for nanoparticles of four different
325     radii. Note that the peak in the heat capacity coincides with the {\em
326     start} of the peak in the bulk modulus.}
327     \label{fig:BmCp}
328     \end{figure}
329    
330    
331     Another feature of these transient (non-equilibrium) calculations is
332     the width of the peak in the heat capacity. Calculation of $C_{p}$
333     from longer equilibrium trajectories should indicate {\it sharper}
334     features in $C_{p}$ for the larger particles. Since we are initiating
335     and observing the melting process itself in these calculation, the
336     smaller particles melt more rapidly, and thus exhibit sharper features
337     in $C_{p}$. Indeed, longer trajectories do show that $T_{m}$ occurs
338     at lower temperatures and with sharper transitions in larger particles
339     than can be observed from transient response calculations.
340     Fig. \ref{fig:Cp2} shows the results of 300 ps simulations which give
341     much sharper and lower temperature melting transitions than those
342     observed in the 40 ps simulations.
343    
344     \begin{figure}[htbp]
345     \centering
346     \includegraphics[height=3in]{images/Cp_vs_T.pdf}
347     \caption{The dependence of the spike in the heat capacity on the
348     length of the simulation. Longer heating-response calculations result
349     in melting transitions that are sharper and lower in temperature than
350     the short-time transient response simulations. Shorter runs don't
351     allow the particles to melt completely.}
352     \label{fig:Cp2}
353     \end{figure}
354    
355    
356    
357     \subsection{BREATHING MODE DYNAMICS}
358    
359     Fig.\ref{fig:VolTime} shows representative samples of the volume
360     vs. time traces for the 20 \AA\ and 35 \AA\ particles at a number of
361     different temperatures. We can clearly see that the period of the
362     breathing mode is dependent on temperature, and that the coherent
363     oscillations of the particles' volume are destroyed after only a few
364     ps in the smaller particles, while they live on for 10-20 ps in the
365     larger particles. The de-coherence is also strongly temperature
366     dependent, with the high temperature samples decohering much more
367     rapidly than lower temperatures.
368    
369     \begin{figure}[htbp]
370     \centering
371     \includegraphics[height=3in]{images/Vol_vs_time.pdf}
372     \caption{Sample Volume traces for the 20 \AA\ and 35 \AA\ particles at a
373     range of temperatures. Note the relatively rapid ($<$ 10 ps)
374     decoherence due to melting in the 20 \AA\ particle as well as the
375     difference between the 1100 K and 1200 K traces in the 35 \AA\
376     particle.}
377     \label{fig:VolTime}
378     \end{figure}
379    
380    
381     Although $V$ vs. $t$ traces can say a great deal, it is more
382     instructive to compute the autocorrelation function for volume
383     fluctuations to give more accurate short-time information. Fig
384     \ref{fig:volcorr} shows representative autocorrelation functions for
385     volume fluctuations. Although many traces exhibit a single frequency
386     with decaying amplitude, a number of the samples show distinct beat
387     patterns indicating the presence of multiple frequency components in
388     the breathing motion of the nanoparticles. In particular, the 20 \AA\
389     particle shows a distinct beat in the volume fluctuations in the 800 K
390     trace.
391    
392    
393    
394     \begin{figure}[htbp]
395     \centering
396     \includegraphics[height=3in]{images/volcorr.pdf}
397     \caption{Volume fluctuation autocorrelation functions for the 20 \AA\
398     and 35 \AA\ particles at a range of temperatures. Successive
399     temperatures have been translated upwards by one unit. Note the beat
400     pattern in the 20 \AA\ particle at 800K.}
401     \label{fig:volcorr}
402     \end{figure}
403    
404    
405     When the power spectrum of the volume autocorrelation functions are
406     analyzed (Eq. (\ref{eq:volspect})), the samples which exhibit beat
407     patterns do indeed show multiple peaks in the power spectrum. We plot
408     the period corresponding to the two lowest frequency peaks in
409     Fig. \ref{fig:Period}. The smaller particles have the most evident
410     splitting, particularly as the temperature rises above the melting
411     points for these particles.
412    
413     \begin{figure}[htbp]
414     \centering
415     \includegraphics[height=3in]{images/Period_vs_T.pdf}
416     \caption{The temperature dependence of the period of the breathing
417     mode for the four different nanoparticles studied in this
418     work.}
419     \label{fig:Period}
420     \end{figure}
421    
422    
423     \section{DISCUSSION}
424     \label{bulkmod:sec:discussion}
425    
426     Lamb's classical theory of elastic spheres~\cite{Lamb1882} provides
427     two possible explanations for the split peak in the vibrational
428     spectrum. The periods of the longitudinal and transverse vibrations
429     in an elastic sphere of radius $R$ are given by:
430     \begin{equation}
431     \tau_{t} = \frac{2 \pi R}{\theta c_{t}}
432     \end{equation}
433     and
434     \begin{equation}
435     \tau_{l} = \frac{2 \pi R}{\eta c_{l}}
436     \end{equation}
437     where $\theta$ and $n$ are obtained from the solutions to the
438     transcendental equations
439     \begin{equation}
440     \tan \theta = \frac{3 \theta}{3 - \theta^{2}}
441     \end{equation}
442     \begin{equation}
443     \tan \eta = \frac{4 \eta}{4 - \eta^{2}\frac{c_{l}^{2}}{c_{t}^{2}}}
444     \end{equation}.
445    
446     $c_{l}$ and $c_{t}$ are the longitudinal and transverse speeds
447     of sound in the material. In an isotropic material, these speeds are
448     simply related to the elastic constants and the density ($\rho$),
449     \begin{equation}
450     c_{l} = \sqrt{c_{11}/\rho}
451     \end{equation}
452     \begin{equation}
453     c_{t} = \sqrt{c_{44}/\rho}.
454     \end{equation}
455    
456     In crystalline materials, the speeds depend on the direction of
457     propagation of the wave relative to the crystal plane.\cite{Kittel:1996fk}
458     For the remainder of our analysis, we assume the nanoparticles are
459     isotropic (which should be valid only above the melting transition).
460     A more detailed analysis of the lower temperature particles would take
461     the crystal lattice into account.
462    
463     If we use the experimental values for the elastic constants for 30
464     \AA\ Au particles at 300K, the low-frequency longitudinal (breathing)
465     mode should have a period of 2.19 ps while the low-frequency
466     transverse (toroidal) mode should have a period of 2.11 ps. Although
467     the actual calculated frequencies in our simulations are off of these
468     values, the difference in the periods (0.08 ps) is approximately half
469     of the splitting observed room-temperature simulations. This,
470     therefore, may be an explanation for the low-temperature splitting in
471     Fig. \ref{fig:Period}.
472    
473     We note that Cerullo {\it et al.} used a similar treatment to obtain
474     the low frequency longitudinal frequencies for crystalline
475     semiconductor nanoparticles,\cite{Cerullo1999} and Simon and Geller
476     have investigated the effects of ensembles of particle size on the
477     Lamb mode using the classical Lamb theory results for isotropic
478     elastic spheres.\cite{Simon2001}
479    
480     \subsection{MELTED AND PARTIALLY-MELTED PARTICLES}
481    
482     Hartland {\it et al.} have extended the Lamb analysis to include
483     surface stress ($\gamma$).\cite{HartlandG.V._jp0276092} In this case, the
484     transcendental equation that must be solved to obtain the
485     low-frequency longitudinal mode is
486     \begin{equation}
487     \eta \cot \eta = 1 - \frac{\eta^{2} c_{l}^{2}}{4 c_{l}^{2} -
488     2 \gamma / (\rho R)}.
489     \end{equation}
490     In ideal liquids, inclusion of the surface stress is vital since the
491     transverse speed of sound ($c_{t}$) vanishes. Interested readers
492     should consult Hartland {\it et al.}'s paper for more details on the
493     extension to liquid-like particles, but the primary result is that the
494     vibrational period of the breathing mode for liquid droplets may be
495     written
496     \begin{equation}
497     \tau = \frac{2 R}{c_{l}(l)}
498     \end{equation}
499     where $c_{l}(l)$ is the longitudinal speed of sound in the liquid.
500     Iida and Guthrie list the speed of sound in liquid Au metal as
501     \begin{equation}
502     c_{l}(l) = 2560 - 0.55 (T - T_{m}) (\mbox{m s}^{-1})
503     \end{equation}
504     where $T_{m}$ is the melting temperature.\cite{Iida1988} A molten 35
505     \AA\ particle just above $T_{m}$ would therefore have a vibrational
506     period of 2.73 ps, and this would be markedly different from the
507     vibrational period just below $T_{m}$ if the melting transition were
508     sharp.
509    
510     We know from our calculations of $C_{p}$ that the complete melting of
511     the particles is {\it not} sharp, and should take longer than the 40
512     ps observation time. There are therefore two explanations which are
513     commensurate with our observations.
514     \begin{enumerate}
515     \item The melting may occur at some time partway through observation
516     of the response to instantaneous heating. The early part
517     of the simulation would then show a higher-frequency breathing mode
518     than would be evident during the latter parts of the simulation.
519     \item The melting may take place by softening the outer layers of the
520     particle first, followed by a melting of the core at higher
521     temperatures. The liquid-like outer layer would then contribute a
522     lower frequency component than the interior of the particle.
523     \end{enumerate}
524    
525     The second of these explanations is consistent with the core-shell
526     melting hypothesis advanced by Hartland {\it et al.} to explain their
527     laser heating experiments.\cite{HartlandG.V._jp0276092} At this stage, our
528     simulations cannot distinguish between the two hypotheses. One
529     possible avenue for future work would be the computation of a
530     radial-dependent order parameter to help evaluate whether the
531     solid-core/liquid-shell structure exists in our simulation.