| 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 | chuckv | 3496 | In this chapter, atomic-level simulations of the transient | 
| 15 | chuckv | 3483 | response of metallic nanoparticles to the nearly instantaneous heating | 
| 16 |  |  | undergone when photons are absorbed during ultrafast laser excitation | 
| 17 | chuckv | 3496 | experiments are discussed. The time scale for heating (determined by the {\it e-ph} | 
| 18 | chuckv | 3483 | 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 | chuckv | 3496 | results in terms of Lamb's classical theory of elastic spheres \cite{Lamb1882} in | 
| 37 | chuckv | 3483 | section \ref{bulkmod:sec:discussion}. | 
| 38 |  |  |  | 
| 39 | chuckv | 3496 | \section{Computational Details} | 
| 40 | chuckv | 3483 | \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 | chuckv | 3496 | \subsection{Simulation Methodology} | 
| 50 | chuckv | 3483 |  | 
| 51 |  |  | Potentials were calculated using the Voter-Chen | 
| 52 |  |  | parameterization~\cite{Voter:87} of the Embedded Atom Method ({\sc | 
| 53 | chuckv | 3496 | eam}) as described in Section \ref{introSec:eam} of this dissertation. | 
| 54 | chuckv | 3483 | Before starting the molecular dynamics runs, a relatively short | 
| 55 |  |  | steepest-descent minimization was performed to relax the lattice in | 
| 56 |  |  | the initial configuration.  To facilitate the study of the larger | 
| 57 |  |  | particles, simulations were run in parallel over 16 processors using | 
| 58 |  |  | Plimpton's force-decomposition method.\cite{plimpton93} | 
| 59 |  |  |  | 
| 60 |  |  | To mimic the events following the absorption of light in the ultrafast | 
| 61 |  |  | laser heating experiments, we have used a simple two-step process to | 
| 62 |  |  | prepare the simulations.  Instantaneous heating of the lattice was | 
| 63 |  |  | performed by sampling atomic velocities from a Maxwell-Boltzmann | 
| 64 |  |  | distribution set to twice the target temperature for the simulation. | 
| 65 |  |  | By equipartition, approximately half of the initial kinetic energy of | 
| 66 |  |  | the system winds up in the potential energy of the system.  The system | 
| 67 |  |  | was a allowed a very short (10 fs) evolution period with the new | 
| 68 |  |  | velocities. | 
| 69 |  |  |  | 
| 70 |  |  | Following this excitation step, the particles evolved under | 
| 71 |  |  | Nos\'{e}-Hoover NVT dynamics~\cite{hoover85} for 40 ps.  Given the | 
| 72 |  |  | mass of the constituent metal atoms, time steps of 5 fs give excellent | 
| 73 |  |  | energy conservation in standard NVE integrators, so the same time step | 
| 74 |  |  | was used in the NVT simulations.  Target temperatures for these | 
| 75 |  |  | particles spanned the range from 300 K to 1350 K in 50 K intervals. | 
| 76 |  |  | Five independent samples were run for each particle and temperature. | 
| 77 |  |  | The results presented below are averaged properties for each of the | 
| 78 |  |  | five independent samples. | 
| 79 |  |  |  | 
| 80 | chuckv | 3496 | \subsection{\label{bulkmod:sec:analysis}Analysis} | 
| 81 | chuckv | 3483 |  | 
| 82 |  |  | Of primary interest when comparing our simulations to experiments is | 
| 83 |  |  | the dynamics of the low-frequency breathing mode of the particles.  To | 
| 84 | chuckv | 3496 | study this motion, access is needed to accurate measures of both the | 
| 85 | chuckv | 3483 | volume and surface area as a function of time.  Throughout the | 
| 86 | chuckv | 3496 | simulations, both quantities were monitored using Barber {\it et al.}'s | 
| 87 | chuckv | 3483 | very fast quickhull algorithm to obtain the convex hull for the | 
| 88 |  |  | collection of 3-d coordinates of all of atoms at each point in | 
| 89 |  |  | time.~\cite{barber96quickhull,qhull} The convex hull is the smallest convex | 
| 90 |  |  | polyhedron which includes all of the atoms, so the volume of this | 
| 91 |  |  | polyhedron is an excellent estimate of the volume of the nanoparticle. | 
| 92 |  |  | The convex hull estimate of the volume will be problematic if the | 
| 93 |  |  | nanoparticle breaks into pieces (i.e. if the bounding surface becomes | 
| 94 |  |  | concave), but for the relatively short trajectories of this study, it | 
| 95 |  |  | provides an excellent measure of particle volume as a function of | 
| 96 |  |  | time. | 
| 97 |  |  |  | 
| 98 |  |  | The bulk modulus, which is the inverse of the compressibility, | 
| 99 |  |  | \begin{equation} | 
| 100 |  |  | K = \frac{1}{\kappa} = - V \left(\frac{\partial P}{\partial | 
| 101 |  |  | V}\right)_{T} | 
| 102 |  |  | \label{eq:Kdef} | 
| 103 |  |  | \end{equation} | 
| 104 |  |  | can be related to quantities that are relatively easily accessible | 
| 105 | chuckv | 3496 | from our molecular dynamics simulations. Four | 
| 106 |  |  | different approaches are presented here for estimating the bulk modulus directly from | 
| 107 | chuckv | 3483 | basic or derived quantities: | 
| 108 |  |  |  | 
| 109 |  |  | \begin{itemize} | 
| 110 |  |  | \item The traditional ``Energetic'' approach | 
| 111 |  |  |  | 
| 112 | chuckv | 3496 | Using basic thermodynamics and one of the Maxwell relations, it can be written that | 
| 113 | chuckv | 3483 | \begin{equation} | 
| 114 |  |  | P = T\left(\frac{\partial S}{\partial V}\right)_{T} - | 
| 115 |  |  | \left(\frac{\partial U}{\partial V}\right)_{T}. | 
| 116 |  |  | \label{eq:Pdef} | 
| 117 |  |  | \end{equation} | 
| 118 |  |  | It follows that | 
| 119 |  |  | \begin{equation} | 
| 120 |  |  | K = V \left(T \left(\frac{\partial^{2} S}{\partial V^{2}}\right)_{T} - | 
| 121 |  |  | \left(\frac{\partial^{2} U}{\partial V^{2}}\right)_{T} \right). | 
| 122 |  |  | \label{eq:Ksub} | 
| 123 |  |  | \end{equation} | 
| 124 |  |  | The standard practice in solid state physics is to assume the low | 
| 125 |  |  | temperature limit ({\it i.e.} to neglect the entropic term), which | 
| 126 |  |  | means the bulk modulus is usually expressed | 
| 127 |  |  | \begin{equation} | 
| 128 |  |  | K \approx V \left(\frac{\partial^{2} U }{\partial V^{2}}\right)_{T}. | 
| 129 |  |  | \label{eq:Kuse} | 
| 130 |  |  | \end{equation} | 
| 131 |  |  | When the relationship between the total energy $U$ and volume $V$ of | 
| 132 |  |  | the system are available at a fixed temperature (as it is in these | 
| 133 |  |  | simulations), it is a simple matter to compute the bulk modulus from | 
| 134 |  |  | the response of the system to the perturbation of the instantaneous | 
| 135 |  |  | heating.  Although this information would, in theory, be available | 
| 136 |  |  | from longer constant temperature simulations, the ranges of volumes | 
| 137 |  |  | and energies explored by a nanoparticle under equilibrium conditions | 
| 138 |  |  | are actually quite small.  Instantaneous heating, since it excites | 
| 139 |  |  | coherent oscillations in the breathing mode, allows us to sample a | 
| 140 |  |  | much wider range of volumes (and energies) for the particles.  The | 
| 141 |  |  | problem with this method is that it neglects the entropic term near | 
| 142 |  |  | the melting transition, which gives spurious results (negative bulk | 
| 143 |  |  | moduli) at higher temperatures. | 
| 144 |  |  |  | 
| 145 |  |  | \item The Linear Response Approach | 
| 146 |  |  |  | 
| 147 |  |  | Linear response theory gives us another approach to calculating the | 
| 148 |  |  | bulk modulus.  This method relates the low-wavelength fluctuations in | 
| 149 |  |  | the density to the isothermal compressibility,\cite{BernePecora} | 
| 150 |  |  | \begin{equation} | 
| 151 |  |  | \underset{k \rightarrow 0}{\lim} \langle | \delta \rho(\vec{k}) |^2 | 
| 152 |  |  | \rangle  =  \frac{k_B T \rho^2 \kappa}{V} | 
| 153 |  |  | \end{equation} | 
| 154 |  |  | where the frequency dependent density fluctuations are the Fourier | 
| 155 |  |  | transform of the spatial fluctuations, | 
| 156 |  |  | \begin{equation} | 
| 157 |  |  | \delta \rho(\vec{k})  =  \underset{V}{\int} e^{i \vec{k}\cdot\vec{r}} \left( \rho(\vec{r}, t) - \langle \rho \rangle \right) dV | 
| 158 |  |  | \end{equation} | 
| 159 |  |  | This approach is essentially equivalent to using the volume | 
| 160 |  |  | fluctuations directly to estimate the bulk modulus, | 
| 161 |  |  | \begin{equation} | 
| 162 |  |  | 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} | 
| 163 |  |  | \end{equation} | 
| 164 |  |  | It should be noted that in these experiments, the particles are {\it | 
| 165 |  |  | far from equilibrium}, and so a linear response approach will not be | 
| 166 |  |  | the most suitable way to obtain estimates of the bulk modulus. | 
| 167 |  |  |  | 
| 168 |  |  | \item The Extended System Approach | 
| 169 |  |  |  | 
| 170 | chuckv | 3496 | Since these simulations are being performed in the NVT ensemble using | 
| 171 | chuckv | 3483 | Nos\'e-Hoover thermostatting, the quantity conserved by our integrator | 
| 172 |  |  | ($H_{NVT}$) can be expressed as: | 
| 173 |  |  | \begin{equation} | 
| 174 |  |  | H_{NVT}  =  U  + f k_B T_{ext} \left( \frac{\tau_T^2 \chi(t)^2}{2} + | 
| 175 |  |  | \int_0^t \chi(s) ds \right). | 
| 176 |  |  | \end{equation} | 
| 177 |  |  | Here $f$ is the number of degrees of freedom in the (real) system, | 
| 178 |  |  | $T_{ext}$ is the temperature of the thermostat, $\tau_T$ is the time | 
| 179 |  |  | constant for the thermostat, and $\chi(t)$ is the instantaneous value | 
| 180 |  |  | of the extended system thermostat variable.  The extended Hamiltonian | 
| 181 |  |  | for the system, $H_{NVT}$ is, to within a constant, the Helmholtz free | 
| 182 |  |  | energy.\cite{melchionna93} Since the pressure is a simple derivative | 
| 183 |  |  | of the Helmholtz free energy, | 
| 184 |  |  | \begin{equation} | 
| 185 |  |  | P  = -\left( \frac{\partial A}{\partial V} \right)_T , | 
| 186 |  |  | \end{equation} | 
| 187 |  |  | the bulk modulus can be obtained (theoretically) by a quadratic fit of | 
| 188 |  |  | the fluctuations in $H_{NVT}$ against fluctuations in the volume, | 
| 189 |  |  | \begin{equation} | 
| 190 |  |  | K =  -V \left( \frac{\partial^2 H_{NVT}}{\partial V^2} \right)_T. | 
| 191 |  |  | \end{equation} | 
| 192 |  |  | However, $H_{NVT}$ is essentially conserved during these simulations, | 
| 193 |  |  | so fitting fluctuations of this quantity to obtain meaningful physical | 
| 194 | chuckv | 3496 | quantities is somewhat suspect.  It should also noted that this method would | 
| 195 | chuckv | 3483 | fail in periodic systems because the volume itself is fixed in | 
| 196 |  |  | periodic NVT simulations. | 
| 197 |  |  |  | 
| 198 |  |  | \item The Direct Pressure Approach | 
| 199 |  |  |  | 
| 200 |  |  | Our preferred method for estimating the bulk modulus is to compute it | 
| 201 |  |  | {\it directly} from the internal pressure in the nanoparticle.  The | 
| 202 |  |  | pressure is obtained via the trace of the pressure tensor, | 
| 203 |  |  | \begin{equation} | 
| 204 |  |  | P  =  \frac{1}{3} | 
| 205 |  |  | \mathrm{Tr}\left[\overleftrightarrow{\mathsf{P}}\right], | 
| 206 |  |  | \end{equation} | 
| 207 |  |  | which has a kinetic contribution as well as a contribution from the | 
| 208 |  |  | stress tensor ($\overleftrightarrow{\mathsf{W}}$): | 
| 209 |  |  | \begin{equation} | 
| 210 |  |  | \overleftrightarrow{\mathsf{P}}  =  \frac{1}{V} \left( | 
| 211 |  |  | \sum_{i=1}^{N} m_i \vec{v}_i \otimes \vec{v}_i  \right) + | 
| 212 |  |  | \overleftrightarrow{\mathsf{W}}. | 
| 213 |  |  | \end{equation} | 
| 214 |  |  | Here the $\otimes$ symbol represents the {\it outer} product of the | 
| 215 |  |  | velocity vector for atom $i$ to yield a $3 \times 3$ matrix. The | 
| 216 |  |  | virial is computed during the simulation using forces between pairs of | 
| 217 |  |  | particles, | 
| 218 |  |  | \begin{equation} | 
| 219 |  |  | \overleftrightarrow{\mathsf{W}}  =  \sum_{i} \sum_{j>i}  \vec{r}_{ij} \otimes \vec{f}_{ij} | 
| 220 |  |  | \end{equation} | 
| 221 | chuckv | 3496 | During the simulation, the internal pressure, $P$, is recorded as well as | 
| 222 | chuckv | 3483 | the total energy, $U$, the extended system's hamiltonian, $H_{NVT}$, | 
| 223 | chuckv | 3496 | and the particle coordinates.  Once the time | 
| 224 |  |  | dependent volume of the nanoparticle has been calculated using the convex hull, | 
| 225 |  |  | any of these four methods can be used to estimate the bulk modulus. | 
| 226 | chuckv | 3483 | \end{itemize} | 
| 227 |  |  |  | 
| 228 | chuckv | 3496 | It is found, however, that only the fourth method (the direct pressure | 
| 229 | chuckv | 3483 | approach) gives meaningful results.  Bulk moduli for the 35 \AA\ | 
| 230 |  |  | particle were computed with the traditional (Energy vs. Volume) | 
| 231 |  |  | approach as well as the direct pressure approach.  A comparison of the | 
| 232 |  |  | Bulk Modulus obtained via both methods and are shown in | 
| 233 |  |  | Fig. \ref{fig:Methods} Note that the second derivative fits in the | 
| 234 |  |  | traditional approach can give (in the liquid droplet region) negative | 
| 235 |  |  | curvature, and this results in negative values for the bulk modulus. | 
| 236 |  |  |  | 
| 237 |  |  | \begin{figure}[htbp] | 
| 238 |  |  | \centering | 
| 239 |  |  | \includegraphics[height=3in]{images/Methods.pdf} | 
| 240 |  |  | \caption{Comparison of two of the methods for estimating the bulk | 
| 241 |  |  | modulus as a function of temperature for the 35\AA\ particle.} | 
| 242 |  |  | \label{fig:Methods} | 
| 243 |  |  | \end{figure} | 
| 244 |  |  |  | 
| 245 |  |  | The Bulk moduli reported in the rest of this paper were computed using | 
| 246 |  |  | the direct pressure method. | 
| 247 |  |  |  | 
| 248 | chuckv | 3496 | To study the frequency of the breathing mode, the | 
| 249 |  |  | power spectrum for volume ($V$) fluctuations have been calculated according to, | 
| 250 | chuckv | 3483 | \begin{equation} | 
| 251 |  |  | \rho_{\Delta V}(\omega) = \int_{-\infty}^{\infty} \langle \Delta V(t) | 
| 252 |  |  | \Delta V(0) \rangle e^{-i \omega t} dt | 
| 253 |  |  | \label{eq:volspect} | 
| 254 |  |  | \end{equation} | 
| 255 |  |  | where $\Delta V(t) = V(t) - \langle V \rangle$.  Because the | 
| 256 |  |  | instantaneous heating excites all of the vibrational modes of the | 
| 257 |  |  | particle, the power spectrum will contain contributions from all modes | 
| 258 |  |  | that perturb the total volume of the particle.  The lowest frequency | 
| 259 |  |  | peak in the power spectrum should give the frequency (and period) for | 
| 260 |  |  | the breathing mode, and these quantities are most readily compared | 
| 261 |  |  | with the Hartland group experiments.\cite{HartlandG.V._jp0276092} Further | 
| 262 |  |  | analysis of the breathing dynamics follows in section | 
| 263 |  |  | \ref{bulkmod:sec:discussion}. | 
| 264 |  |  |  | 
| 265 | chuckv | 3496 | The heat capacity for our simulations has also been computed to verify | 
| 266 | chuckv | 3483 | the location of the melting transition.  Calculations of the heat | 
| 267 |  |  | capacity were performed on the non-equilibrium, instantaneous heating | 
| 268 |  |  | simulations, as well as on simulations of nanoparticles that were at | 
| 269 |  |  | equilibrium at the target temperature. | 
| 270 |  |  |  | 
| 271 | chuckv | 3496 | \section{Results} | 
| 272 | chuckv | 3483 | \label{bulkmod:sec:results} | 
| 273 |  |  |  | 
| 274 | chuckv | 3496 | \subsection{The Bulk Modulus and Heat Capacity} | 
| 275 | chuckv | 3483 |  | 
| 276 |  |  | The upper panel in Fig. \ref{fig:BmCp} shows the temperature | 
| 277 |  |  | dependence of the Bulk Modulus ($K$).  In all samples, there is a | 
| 278 |  |  | dramatic (size-dependent) drop in $K$ at temperatures well below the | 
| 279 |  |  | melting temperature of bulk polycrystalline gold.  This drop in $K$ | 
| 280 |  |  | coincides with the actual melting transition of the nanoparticles. | 
| 281 |  |  | Surface melting has been confirmed at even lower temperatures using | 
| 282 |  |  | the radial-dependent density, $\rho(r) / \rho$, which shows a merging | 
| 283 |  |  | of the crystalline peaks in the outer layer of the nanoparticle. | 
| 284 |  |  | However, the bulk modulus only has an appreciable drop when the | 
| 285 |  |  | particle melts fully. | 
| 286 |  |  |  | 
| 287 |  |  | \begin{figure}[htbp] | 
| 288 |  |  | \centering | 
| 289 |  |  | \includegraphics[height=3in]{images/Stacked_Bulk_modulus_and_Cp.pdf} | 
| 290 |  |  | \caption{The temperature dependence of the bulk modulus (upper panel) | 
| 291 |  |  | and heat capacity (lower panel) for nanoparticles of four different | 
| 292 |  |  | radii. Note that the peak in the heat capacity coincides with the {\em | 
| 293 |  |  | start} of the peak in the bulk modulus.} | 
| 294 |  |  | \label{fig:BmCp} | 
| 295 |  |  | \end{figure} | 
| 296 |  |  |  | 
| 297 |  |  |  | 
| 298 |  |  | Another feature of these transient (non-equilibrium) calculations is | 
| 299 |  |  | the width of the peak in the heat capacity.  Calculation of $C_{p}$ | 
| 300 |  |  | from longer equilibrium trajectories should indicate {\it sharper} | 
| 301 | chuckv | 3496 | features in $C_{p}$ for the larger particles.  Since the melting process itself is being initiated and observed in these calculations, the | 
| 302 | chuckv | 3483 | smaller particles melt more rapidly, and thus exhibit sharper features | 
| 303 |  |  | in $C_{p}$.  Indeed, longer trajectories do show that $T_{m}$ occurs | 
| 304 |  |  | at lower temperatures and with sharper transitions in larger particles | 
| 305 |  |  | than can be observed from transient response calculations. | 
| 306 |  |  | Fig. \ref{fig:Cp2} shows the results of 300 ps simulations which give | 
| 307 |  |  | much sharper and lower temperature melting transitions than those | 
| 308 |  |  | observed in the 40 ps simulations. | 
| 309 |  |  |  | 
| 310 |  |  | \begin{figure}[htbp] | 
| 311 |  |  | \centering | 
| 312 |  |  | \includegraphics[height=3in]{images/Cp_vs_T.pdf} | 
| 313 |  |  | \caption{The dependence of the spike in the heat capacity on the | 
| 314 |  |  | length of the simulation.  Longer heating-response calculations result | 
| 315 |  |  | in melting transitions that are sharper and lower in temperature than | 
| 316 |  |  | the short-time transient response simulations.  Shorter runs don't | 
| 317 |  |  | allow the particles to melt completely.} | 
| 318 |  |  | \label{fig:Cp2} | 
| 319 |  |  | \end{figure} | 
| 320 |  |  |  | 
| 321 |  |  |  | 
| 322 |  |  |  | 
| 323 | chuckv | 3496 | \subsection{Breathing Mode Dynamics} | 
| 324 | chuckv | 3483 |  | 
| 325 |  |  | Fig.\ref{fig:VolTime} shows representative samples of the volume | 
| 326 |  |  | vs. time traces for the 20 \AA\ and 35 \AA\ particles at a number of | 
| 327 | chuckv | 3496 | different temperatures.  It can clearly be seen that the period of the | 
| 328 | chuckv | 3483 | breathing mode is dependent on temperature, and that the coherent | 
| 329 |  |  | oscillations of the particles' volume are destroyed after only a few | 
| 330 |  |  | ps in the smaller particles, while they live on for 10-20 ps in the | 
| 331 |  |  | larger particles.  The de-coherence is also strongly temperature | 
| 332 |  |  | dependent, with the high temperature samples decohering much more | 
| 333 |  |  | rapidly than lower temperatures. | 
| 334 |  |  |  | 
| 335 |  |  | \begin{figure}[htbp] | 
| 336 |  |  | \centering | 
| 337 |  |  | \includegraphics[height=3in]{images/Vol_vs_time.pdf} | 
| 338 |  |  | \caption{Sample Volume traces for the 20 \AA\ and 35 \AA\ particles at a | 
| 339 |  |  | range of temperatures.  Note the relatively rapid ($<$ 10 ps) | 
| 340 |  |  | decoherence due to melting in the 20 \AA\ particle as well as the | 
| 341 |  |  | difference between the 1100 K and 1200 K traces in the 35 \AA\ | 
| 342 |  |  | particle.} | 
| 343 |  |  | \label{fig:VolTime} | 
| 344 |  |  | \end{figure} | 
| 345 |  |  |  | 
| 346 |  |  |  | 
| 347 |  |  | Although $V$ vs. $t$ traces can say a great deal, it is more | 
| 348 |  |  | instructive to compute the autocorrelation function for volume | 
| 349 |  |  | fluctuations to give more accurate short-time information.  Fig | 
| 350 |  |  | \ref{fig:volcorr} shows representative autocorrelation functions for | 
| 351 |  |  | volume fluctuations.  Although many traces exhibit a single frequency | 
| 352 |  |  | with decaying amplitude, a number of the samples show distinct beat | 
| 353 |  |  | patterns indicating the presence of multiple frequency components in | 
| 354 |  |  | the breathing motion of the nanoparticles.  In particular, the 20 \AA\ | 
| 355 |  |  | particle shows a distinct beat in the volume fluctuations in the 800 K | 
| 356 |  |  | trace. | 
| 357 |  |  |  | 
| 358 |  |  |  | 
| 359 |  |  |  | 
| 360 |  |  | \begin{figure}[htbp] | 
| 361 |  |  | \centering | 
| 362 |  |  | \includegraphics[height=3in]{images/volcorr.pdf} | 
| 363 | chuckv | 3496 | \caption{Volume fluctuation autocorrelation functions for the 20 \AA\ (lower panel) | 
| 364 |  |  | and 35 \AA\ (upper panel) particles at a range of temperatures.  Successive | 
| 365 | chuckv | 3483 | temperatures have been translated upwards by one unit.  Note the beat | 
| 366 |  |  | pattern in the 20 \AA\ particle at 800K.} | 
| 367 |  |  | \label{fig:volcorr} | 
| 368 |  |  | \end{figure} | 
| 369 |  |  |  | 
| 370 |  |  |  | 
| 371 |  |  | When the power spectrum of the volume autocorrelation functions are | 
| 372 |  |  | analyzed (Eq. (\ref{eq:volspect})), the samples which exhibit beat | 
| 373 | chuckv | 3496 | patterns do indeed show multiple peaks in the power spectrum. The period corresponding to the two lowest frequency peaks is plotted in | 
| 374 |  |  | Fig. \ref{fig:Period}.  Smaller particles have the most evident | 
| 375 | chuckv | 3483 | splitting, particularly as the temperature rises above the melting | 
| 376 |  |  | points for these particles. | 
| 377 |  |  |  | 
| 378 |  |  | \begin{figure}[htbp] | 
| 379 |  |  | \centering | 
| 380 |  |  | \includegraphics[height=3in]{images/Period_vs_T.pdf} | 
| 381 |  |  | \caption{The temperature dependence of the period of the breathing | 
| 382 |  |  | mode for the four different nanoparticles studied in this | 
| 383 |  |  | work.} | 
| 384 |  |  | \label{fig:Period} | 
| 385 |  |  | \end{figure} | 
| 386 |  |  |  | 
| 387 |  |  |  | 
| 388 | chuckv | 3496 | \section{Discussion} | 
| 389 | chuckv | 3483 | \label{bulkmod:sec:discussion} | 
| 390 |  |  |  | 
| 391 |  |  | Lamb's classical theory of elastic spheres~\cite{Lamb1882} provides | 
| 392 |  |  | two possible explanations for the split peak in the vibrational | 
| 393 |  |  | spectrum.  The periods of the longitudinal and transverse vibrations | 
| 394 |  |  | in an elastic sphere of radius $R$ are given by: | 
| 395 |  |  | \begin{equation} | 
| 396 |  |  | \tau_{t} = \frac{2 \pi R}{\theta c_{t}} | 
| 397 |  |  | \end{equation} | 
| 398 |  |  | and | 
| 399 |  |  | \begin{equation} | 
| 400 |  |  | \tau_{l} = \frac{2 \pi R}{\eta c_{l}} | 
| 401 |  |  | \end{equation} | 
| 402 |  |  | where $\theta$ and $n$ are obtained from the solutions to the | 
| 403 |  |  | transcendental equations | 
| 404 |  |  | \begin{equation} | 
| 405 |  |  | \tan \theta = \frac{3 \theta}{3 - \theta^{2}} | 
| 406 |  |  | \end{equation} | 
| 407 |  |  | \begin{equation} | 
| 408 |  |  | \tan \eta = \frac{4 \eta}{4 - \eta^{2}\frac{c_{l}^{2}}{c_{t}^{2}}} | 
| 409 |  |  | \end{equation}. | 
| 410 |  |  |  | 
| 411 |  |  | $c_{l}$ and $c_{t}$ are the longitudinal and transverse speeds | 
| 412 |  |  | of sound in the material.  In an isotropic material, these speeds are | 
| 413 |  |  | simply related to the elastic constants and the density ($\rho$), | 
| 414 |  |  | \begin{equation} | 
| 415 |  |  | c_{l} = \sqrt{c_{11}/\rho} | 
| 416 |  |  | \end{equation} | 
| 417 |  |  | \begin{equation} | 
| 418 |  |  | c_{t} = \sqrt{c_{44}/\rho}. | 
| 419 |  |  | \end{equation} | 
| 420 |  |  |  | 
| 421 |  |  | In crystalline materials, the speeds depend on the direction of | 
| 422 |  |  | propagation of the wave relative to the crystal plane.\cite{Kittel:1996fk} | 
| 423 | chuckv | 3496 | For the remainder of our analysis, it is assumed that the nanoparticles are | 
| 424 | chuckv | 3483 | isotropic (which should be valid only above the melting transition). | 
| 425 |  |  | A more detailed analysis of the lower temperature particles would take | 
| 426 |  |  | the crystal lattice into account. | 
| 427 |  |  |  | 
| 428 | chuckv | 3496 | Using the experimental values for the elastic constants for 30 | 
| 429 | chuckv | 3483 | \AA\ Au particles at 300K, the low-frequency longitudinal (breathing) | 
| 430 |  |  | mode should have a period of 2.19 ps while the low-frequency | 
| 431 |  |  | transverse (toroidal) mode should have a period of 2.11 ps.  Although | 
| 432 | chuckv | 3496 | the actual calculated frequencies in the simulations are off of these | 
| 433 | chuckv | 3483 | values, the difference in the periods (0.08 ps) is approximately half | 
| 434 |  |  | of the splitting observed room-temperature simulations.  This, | 
| 435 |  |  | therefore, may be an explanation for the low-temperature splitting in | 
| 436 |  |  | Fig. \ref{fig:Period}. | 
| 437 |  |  |  | 
| 438 |  |  | We note that Cerullo {\it et al.} used a similar treatment to obtain | 
| 439 |  |  | the low frequency longitudinal frequencies for crystalline | 
| 440 |  |  | semiconductor nanoparticles,\cite{Cerullo1999} and Simon and Geller | 
| 441 |  |  | have investigated the effects of ensembles of particle size on the | 
| 442 |  |  | Lamb mode using the classical Lamb theory results for isotropic | 
| 443 |  |  | elastic spheres.\cite{Simon2001} | 
| 444 |  |  |  | 
| 445 | chuckv | 3496 | \subsection{Melted and Partially-Melted Particles} | 
| 446 | chuckv | 3483 |  | 
| 447 |  |  | Hartland {\it et al.} have extended the Lamb analysis to include | 
| 448 |  |  | surface stress ($\gamma$).\cite{HartlandG.V._jp0276092} In this case, the | 
| 449 |  |  | transcendental equation that must be solved to obtain the | 
| 450 |  |  | low-frequency longitudinal mode is | 
| 451 |  |  | \begin{equation} | 
| 452 |  |  | \eta \cot \eta = 1 - \frac{\eta^{2} c_{l}^{2}}{4 c_{l}^{2} - | 
| 453 |  |  | 2 \gamma / (\rho R)}. | 
| 454 |  |  | \end{equation} | 
| 455 |  |  | In ideal liquids, inclusion of the surface stress is vital since the | 
| 456 |  |  | transverse speed of sound ($c_{t}$) vanishes.  Interested readers | 
| 457 |  |  | should consult Hartland {\it et al.}'s paper for more details on the | 
| 458 |  |  | extension to liquid-like particles, but the primary result is that the | 
| 459 |  |  | vibrational period of the breathing mode for liquid droplets may be | 
| 460 |  |  | written | 
| 461 |  |  | \begin{equation} | 
| 462 |  |  | \tau = \frac{2 R}{c_{l}(l)} | 
| 463 |  |  | \end{equation} | 
| 464 |  |  | where $c_{l}(l)$ is the longitudinal speed of sound in the liquid. | 
| 465 |  |  | Iida and Guthrie list the speed of sound in liquid Au metal as | 
| 466 |  |  | \begin{equation} | 
| 467 |  |  | c_{l}(l) = 2560 - 0.55 (T - T_{m})  (\mbox{m s}^{-1}) | 
| 468 |  |  | \end{equation} | 
| 469 |  |  | where $T_{m}$ is the melting temperature.\cite{Iida1988} A molten 35 | 
| 470 |  |  | \AA\ particle just above $T_{m}$ would therefore have a vibrational | 
| 471 |  |  | period of 2.73 ps, and this would be markedly different from the | 
| 472 |  |  | vibrational period just below $T_{m}$ if the melting transition were | 
| 473 |  |  | sharp. | 
| 474 |  |  |  | 
| 475 |  |  | We know from our calculations of $C_{p}$ that the complete melting of | 
| 476 |  |  | the particles is {\it not} sharp, and should take longer than the 40 | 
| 477 |  |  | ps observation time.  There are therefore two explanations which are | 
| 478 |  |  | commensurate with our observations. | 
| 479 |  |  | \begin{enumerate} | 
| 480 |  |  | \item The melting may occur at some time partway through observation | 
| 481 |  |  | of the response to instantaneous heating.  The early part | 
| 482 |  |  | of the simulation would then show a higher-frequency breathing mode | 
| 483 |  |  | than would be evident during the latter parts of the simulation. | 
| 484 |  |  | \item The melting may take place by softening the outer layers of the | 
| 485 |  |  | particle first, followed by a melting of the core at higher | 
| 486 |  |  | temperatures.  The liquid-like outer layer would then contribute a | 
| 487 |  |  | lower frequency component than the interior of the particle. | 
| 488 |  |  | \end{enumerate} | 
| 489 |  |  |  | 
| 490 |  |  | The second of these explanations is consistent with the core-shell | 
| 491 |  |  | melting hypothesis advanced by Hartland {\it et al.} to explain their | 
| 492 |  |  | laser heating experiments.\cite{HartlandG.V._jp0276092} At this stage, our | 
| 493 |  |  | simulations cannot distinguish between the two hypotheses.  One | 
| 494 |  |  | possible avenue for future work would be the computation of a | 
| 495 |  |  | radial-dependent order parameter to help evaluate whether the | 
| 496 |  |  | solid-core/liquid-shell structure exists in our simulation. |