| 1 | chuckv | 3226 | %!TEX root = /Users/charles/Desktop/nanoglass/nanoglass.tex | 
| 2 |  |  |  | 
| 3 | gezelter | 3221 | \section{Computational Methodology} | 
| 4 |  |  | \label{sec:details} | 
| 5 |  |  |  | 
| 6 |  |  | \subsection{Initial Geometries and Heating} | 
| 7 |  |  |  | 
| 8 |  |  | Cu-core / Ag-shell and random alloy structures were constructed on an | 
| 9 |  |  | underlying FCC lattice (4.09 {\AA}) at the bulk eutectic composition | 
| 10 | gezelter | 3233 | $\mathrm{Ag}_6\mathrm{Cu}_4$.  Both initial geometries were considered | 
| 11 | gezelter | 3230 | although experimental results suggest that the random structure is the | 
| 12 | gezelter | 3233 | most likely structure to be found following | 
| 13 |  |  | synthesis.\cite{Jiang:2005lr,gonzalo:5163} Three different sizes of | 
| 14 | gezelter | 3230 | nanoparticles corresponding to a 20 \AA radius (1961 atoms), 30 {\AA} | 
| 15 |  |  | radius (6603 atoms) and 40 {\AA} radius (15683 atoms) were | 
| 16 |  |  | constructed.  These initial structures were relaxed to their | 
| 17 |  |  | equilibrium structures at 20 K for 20 ps and again at 300 K for 100 ps | 
| 18 |  |  | sampling from a Maxwell-Boltzmann distribution at each temperature. | 
| 19 | gezelter | 3221 |  | 
| 20 |  |  | To mimic the effects of the heating due to laser irradiation, the | 
| 21 |  |  | particles were allowed to melt by sampling velocities from the Maxwell | 
| 22 |  |  | Boltzmann distribution at a temperature of 900 K.  The particles were | 
| 23 |  |  | run under microcanonical simulation conditions for 1 ns of simualtion | 
| 24 |  |  | time prior to studying the effects of heat transfer to the solvent. | 
| 25 |  |  | In all cases, center of mass translational and rotational motion of | 
| 26 |  |  | the particles were set to zero before any data collection was | 
| 27 |  |  | undertaken.  Structural features (pair distribution functions) were | 
| 28 |  |  | used to verify that the particles were indeed liquid droplets before | 
| 29 |  |  | cooling simulations took place. | 
| 30 |  |  |  | 
| 31 |  |  | \subsection{Modeling random alloy and core shell particles in solution | 
| 32 |  |  | phase environments} | 
| 33 |  |  |  | 
| 34 |  |  | To approximate the effects of rapid heat transfer to the solvent | 
| 35 |  |  | following a heating at the plasmon resonance, we utilized a | 
| 36 |  |  | methodology in which atoms contained in the outer $4$ {\AA} radius of | 
| 37 | gezelter | 3233 | the nanoparticle evolved under Langevin Dynamics, | 
| 38 |  |  | \begin{equation} | 
| 39 |  |  | m \frac{\partial^2 \vec{x}}{\partial t^2} = F_\textrm{sys}(\vec{x}(t)) | 
| 40 |  |  | - 6 \pi a \eta \vec{v}(t)  + F_\textrm{ran} | 
| 41 |  |  | \label{eq:langevin} | 
| 42 |  |  | \end{equation} | 
| 43 |  |  | with a solvent friction ($\eta$) approximating the contribution from | 
| 44 |  |  | the solvent and capping agent.  Atoms located in the interior of the | 
| 45 |  |  | nanoparticle evolved under Newtonian dynamics.  The set-up of our | 
| 46 |  |  | simulations is nearly identical with the ``stochastic boundary | 
| 47 |  |  | molecular dynamics'' ({\sc sbmd}) method that has seen wide use in the | 
| 48 |  |  | protein simulation | 
| 49 | gezelter | 3221 | community.\cite{BROOKS:1985kx,BROOKS:1983uq,BRUNGER:1984fj} A sketch | 
| 50 | gezelter | 3233 | of this setup can be found in Fig. \ref{fig:langevinSketch}.  In | 
| 51 |  |  | equation \ref{eq:langevin} the frictional forces of a spherical atom | 
| 52 |  |  | of radius $a$ depend on the solvent viscosity.  The random forces are | 
| 53 |  |  | usually taken as gaussian random variables with zero mean and a | 
| 54 |  |  | variance tied to the solvent viscosity and temperature, | 
| 55 |  |  | \begin{equation} | 
| 56 |  |  | \langle F_\textrm{ran}(t) \cdot F_\textrm{ran} (t') | 
| 57 |  |  | \rangle = 2 k_B T (6 \pi \eta a) \delta(t - t') | 
| 58 |  |  | \label{eq:stochastic} | 
| 59 | chuckv | 3208 | \end{equation} | 
| 60 | gezelter | 3233 | Due to the presence of the capping agent and the lack of details about | 
| 61 |  |  | the atomic-scale interactions between the metallic atoms and the | 
| 62 |  |  | solvent, the effective viscosity is a essentially a free parameter | 
| 63 |  |  | that must be tuned to give experimentally relevant simulations. | 
| 64 | chuckv | 3222 | \begin{figure}[htbp] | 
| 65 |  |  | \centering | 
| 66 | gezelter | 3242 | \includegraphics[width=5in]{images/stochbound.pdf} | 
| 67 | gezelter | 3230 | \caption{Methodology used to mimic the experimental cooling conditions | 
| 68 |  |  | of a hot nanoparticle surrounded by a solvent.  Atoms in the core of | 
| 69 |  |  | the particle evolved under Newtonian dynamics, while atoms that were | 
| 70 |  |  | in the outer skin of the particle evolved under Langevin dynamics. | 
| 71 | gezelter | 3242 | The radius of the spherical region operating under Newtonian dynamics, | 
| 72 |  |  | $r_\textrm{Newton}$ was set to be 4 {\AA} smaller than the original | 
| 73 |  |  | radius ($R$) of the liquid droplet.} | 
| 74 | chuckv | 3222 | \label{fig:langevinSketch} | 
| 75 |  |  | \end{figure} | 
| 76 | gezelter | 3230 |  | 
| 77 | gezelter | 3221 | The viscosity ($\eta$) can be tuned by comparing the cooling rate that | 
| 78 |  |  | a set of nanoparticles experience with the known cooling rates for | 
| 79 | gezelter | 3230 | similar particles obtained via the laser heating experiments. | 
| 80 | gezelter | 3221 | Essentially, we tune the solvent viscosity until the thermal decay | 
| 81 |  |  | profile matches a heat-transfer model using reasonable values for the | 
| 82 |  |  | interfacial conductance and the thermal conductivity of the solvent. | 
| 83 |  |  |  | 
| 84 |  |  | Cooling rates for the experimentally-observed nanoparticles were | 
| 85 |  |  | calculated from the heat transfer equations for a spherical particle | 
| 86 | gezelter | 3230 | embedded in a ambient medium that allows for diffusive heat transport. | 
| 87 |  |  | Following Plech {\it et al.},\cite{plech:195423} we use a heat | 
| 88 |  |  | transfer model that consists of two coupled differential equations | 
| 89 |  |  | in the Laplace domain, | 
| 90 | chuckv | 3208 | \begin{eqnarray} | 
| 91 | gezelter | 3221 | 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\\ | 
| 92 |  |  | \left(\frac{\partial}{\partial r} T_{f}(r,s)\right)_{r=R} + | 
| 93 |  |  | \frac{G}{K}(T_{p}(s)-T_{f}(r,s) = 0 | 
| 94 |  |  | \label{eq:heateqn} | 
| 95 | chuckv | 3208 | \end{eqnarray} | 
| 96 | gezelter | 3221 | where $s$ is the time-conjugate variable in Laplace space. The | 
| 97 |  |  | variables in these equations describe a nanoparticle of radius $R$, | 
| 98 |  |  | mass $M$, and specific heat $c_{p}$ at an initial temperature | 
| 99 |  |  | $T_0$. The surrounding solvent has a thermal profile $T_f(r,t)$, | 
| 100 |  |  | thermal conductivity $K$, density $\rho$, and specific heat $c$. $G$ | 
| 101 |  |  | is the interfacial conductance between the nanoparticle and the | 
| 102 |  |  | surrounding solvent, and contains information about heat transfer to | 
| 103 |  |  | the capping agent as well as the direct metal-to-solvent heat loss. | 
| 104 |  |  | The temperature of the nanoparticle as a function of time can then | 
| 105 |  |  | obtained by the inverse Laplace transform, | 
| 106 | chuckv | 3208 | \begin{equation} | 
| 107 | gezelter | 3221 | T_{p}(t)=\frac{2 k R^2 g^2 | 
| 108 |  |  | T_0}{\pi}\int_{0}^{\infty}\frac{\exp(-\kappa u^2 | 
| 109 |  |  | t/R^2)u^2}{(u^2(1 + R g) - k R g)^2+(u^3 - k R g u)^2}\mathrm{d}u. | 
| 110 |  |  | \label{eq:laplacetransform} | 
| 111 | chuckv | 3208 | \end{equation} | 
| 112 | gezelter | 3221 | For simplicity, we have introduced the thermal diffusivity $\kappa = | 
| 113 |  |  | K/(\rho c)$,  and defined $k=4\pi R^3 \rho c /(M c_p)$ and $g = G/K$ in | 
| 114 |  |  | Eq. \ref{eq:laplacetransform}. | 
| 115 | chuckv | 3208 |  | 
| 116 | gezelter | 3221 | Eq. \ref{eq:laplacetransform} was solved numerically for the Ag-Cu | 
| 117 |  |  | system using mole-fraction weighted values for $c_p$ and $\rho_p$ of | 
| 118 |  |  | 0.295 $(\mathrm{J g^{-1} K^{-1}})$ and $9.826\times 10^6$ $(\mathrm{g | 
| 119 |  |  | m^{-3}})$ respectively. Since most of the laser excitation experiments | 
| 120 |  |  | have been done in aqueous solutions, parameters used for the fluid are | 
| 121 | gezelter | 3230 | $K$ of $0.6$ $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$ | 
| 122 |  |  | $(\mathrm{g m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$. | 
| 123 | chuckv | 3208 |  | 
| 124 | gezelter | 3221 | Values for the interfacial conductance have been determined by a | 
| 125 |  |  | number of groups for similar nanoparticles and range from a low | 
| 126 |  |  | $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ to $120\times 10^{6}$ | 
| 127 | gezelter | 3230 | $(\mathrm{Wm^{-2}K^{-1}})$.\cite{hartlandPrv2007} Similarly, Plech | 
| 128 |  |  | {\it et al.}  reported a value for the interfacial conductance of | 
| 129 |  |  | $G=105\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ and $G=130\pm 15$ | 
| 130 |  |  | $(\mathrm{Wm^{-2}K^{-1}})$ for Pt | 
| 131 |  |  | nanoparticles.\cite{plech:195423,PhysRevB.66.224301} | 
| 132 | gezelter | 3221 |  | 
| 133 |  |  | We conducted our simulations at both ends of the range of | 
| 134 |  |  | experimentally-determined values for the interfacial conductance. | 
| 135 |  |  | This allows us to observe both the slowest and fastest heat transfers | 
| 136 |  |  | from the nanoparticle to the solvent that are consistent with | 
| 137 |  |  | experimental observations.  For the slowest heat transfer, a value for | 
| 138 |  |  | G of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ was used and for | 
| 139 |  |  | the fastest heat transfer, a value of $117\times 10^{6}$ | 
| 140 |  |  | $(\mathrm{Wm^{-2}K^{-1}})$ was used.  Based on calculations we have | 
| 141 |  |  | done using raw data from the Hartland group's thermal half-time | 
| 142 | gezelter | 3230 | experiments on Au nanospheres, the true G values are probably in the | 
| 143 |  |  | faster regime: $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$. | 
| 144 | gezelter | 3221 |  | 
| 145 |  |  | The rate of cooling for the nanoparticles in a molecular dynamics | 
| 146 |  |  | simulation can then be tuned by changing the effective solvent | 
| 147 |  |  | viscosity ($\eta$) until the nanoparticle cooling rate matches the | 
| 148 |  |  | cooling rate described by the heat-transfer equations | 
| 149 | gezelter | 3247 | (\ref{eq:heateqn}). The effective solvent viscosity (in Pa s) for a G | 
| 150 |  |  | of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $4.2 \times | 
| 151 |  |  | 10^{-6}$, $5.0 \times 10^{-6}$, and | 
| 152 |  |  | $5.5 \times 10^{-6}$ for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The | 
| 153 |  |  | effective solvent viscosity (again in Pa s) for an interfacial | 
| 154 |  |  | conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $5.7 | 
| 155 |  |  | \times 10^{-6}$, $7.2 \times 10^{-6}$, and $7.5 \times 10^{-6}$ | 
| 156 |  |  | for 20 {\AA}, 30 {\AA} and 40 {\AA} particles.  Cooling traces for | 
| 157 |  |  | each particle size are presented in | 
| 158 | gezelter | 3221 | Fig. \ref{fig:images_cooling_plot}. It should be noted that the | 
| 159 |  |  | Langevin thermostat produces cooling curves that are consistent with | 
| 160 |  |  | Newtonian (single-exponential) cooling, which cannot match the cooling | 
| 161 | gezelter | 3230 | profiles from Eq. \ref{eq:laplacetransform} exactly. Fitting the | 
| 162 |  |  | Langevin cooling profiles to a single-exponential produces | 
| 163 |  |  | $\tau=25.576$ ps, $\tau=43.786$ ps, and $\tau=56.621$ ps for the 20, | 
| 164 |  |  | 30 and 40 {\AA} nanoparticles and a G of $87.5\times 10^{6}$ | 
| 165 |  |  | $(\mathrm{Wm^{-2}K^{-1}})$.  For comparison's sake, similar | 
| 166 |  |  | single-exponential fits with an interfacial conductance of G of | 
| 167 |  |  | $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ produced a $\tau=13.391$ | 
| 168 |  |  | ps, $\tau=30.426$ ps, $\tau=43.857$ ps for the 20, 30 and 40 {\AA} | 
| 169 |  |  | nanoparticles. | 
| 170 | gezelter | 3221 |  | 
| 171 | chuckv | 3213 | \begin{figure}[htbp] | 
| 172 | gezelter | 3221 | \centering | 
| 173 | gezelter | 3242 | \includegraphics[width=5in]{images/cooling_plot.pdf} | 
| 174 | gezelter | 3221 | \caption{Thermal cooling curves obtained from the inverse Laplace | 
| 175 |  |  | transform heat model in Eq. \ref{eq:laplacetransform} (solid line) as | 
| 176 |  |  | well as from molecular dynamics simulations (circles).  Effective | 
| 177 | gezelter | 3247 | solvent viscosities of 4.2-7.5 $\times 10^{-6}$ Pa s (depending on the | 
| 178 |  |  | radius of the particle) give the best fit to the experimental cooling | 
| 179 |  |  | curves.  This viscosity suggests that the nanoparticles in these | 
| 180 |  |  | experiments are surrounded by a vapor layer (which is a reasonable | 
| 181 |  |  | assumptions given the initial temperatures of the particles).  } | 
| 182 | gezelter | 3221 | \label{fig:images_cooling_plot} | 
| 183 | chuckv | 3213 | \end{figure} | 
| 184 | chuckv | 3208 |  | 
| 185 | gezelter | 3221 | \subsection{Potentials for classical simulations of bimetallic | 
| 186 |  |  | nanoparticles} | 
| 187 | chuckv | 3208 |  | 
| 188 | gezelter | 3221 | Several different potential models have been developed that reasonably | 
| 189 |  |  | describe interactions in transition metals. In particular, the | 
| 190 |  |  | Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and | 
| 191 |  |  | Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study | 
| 192 |  |  | a wide range of phenomena in both bulk materials and | 
| 193 |  |  | nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r} Both | 
| 194 |  |  | potentials are based on a model of a metal which treats the nuclei and | 
| 195 |  |  | core electrons as pseudo-atoms embedded in the electron density due to | 
| 196 |  |  | the valence electrons on all of the other atoms in the system. The | 
| 197 |  |  | {\sc sc} potential has a simple form that closely resembles that of | 
| 198 |  |  | the ubiquitous Lennard Jones potential, | 
| 199 |  |  | \begin{equation} | 
| 200 |  |  | \label{eq:SCP1} | 
| 201 |  |  | 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] , | 
| 202 |  |  | \end{equation} | 
| 203 |  |  | where $V^{pair}_{ij}$ and $\rho_{i}$ are given by | 
| 204 |  |  | \begin{equation} | 
| 205 |  |  | \label{eq:SCP2} | 
| 206 |  |  | 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}}. | 
| 207 |  |  | \end{equation} | 
| 208 |  |  | $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for | 
| 209 |  |  | interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in | 
| 210 |  |  | Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models | 
| 211 |  |  | the interactions between the valence electrons and the cores of the | 
| 212 |  |  | pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy | 
| 213 |  |  | scale, $c_i$ scales the attractive portion of the potential relative | 
| 214 |  |  | to the repulsive interaction and $\alpha_{ij}$ is a length parameter | 
| 215 |  |  | that assures a dimensionless form for $\rho$. These parameters are | 
| 216 |  |  | tuned to various experimental properties such as the density, cohesive | 
| 217 |  |  | energy, and elastic moduli for FCC transition metals. The quantum | 
| 218 |  |  | Sutton-Chen ({\sc q-sc}) formulation matches these properties while | 
| 219 |  |  | including zero-point quantum corrections for different transition | 
| 220 |  |  | metals.\cite{PhysRevB.59.3527} This particular parametarization has | 
| 221 |  |  | been shown to reproduce the experimentally available heat of mixing | 
| 222 |  |  | data for both FCC solid solutions of Ag-Cu and the high-temperature | 
| 223 |  |  | liquid.\cite{sheng:184203} In contrast, the {\sc eam} potential does | 
| 224 |  |  | not reproduce the experimentally observed heat of mixing for the | 
| 225 |  |  | liquid alloy.\cite{MURRAY:1984lr} Combination rules for the alloy were | 
| 226 |  |  | taken to be the arithmatic average of the atomic parameters with the | 
| 227 |  |  | exception of $c_i$ since its values is only dependent on the identity | 
| 228 |  |  | of the atom where the density is evaluated.  For the {\sc q-sc} | 
| 229 |  |  | potential, cutoff distances are traditionally taken to be | 
| 230 |  |  | $2\alpha_{ij}$ and include up to the sixth coordination shell in FCC | 
| 231 |  |  | metals. | 
| 232 | chuckv | 3213 |  | 
| 233 | chuckv | 3226 | %\subsection{Sampling single-temperature configurations from a cooling | 
| 234 |  |  | %trajectory} | 
| 235 | chuckv | 3213 |  | 
| 236 | gezelter | 3230 | To better understand the structural changes occurring in the | 
| 237 |  |  | nanoparticles throughout the cooling trajectory, configurations were | 
| 238 |  |  | sampled at regular intervals during the cooling trajectory. These | 
| 239 |  |  | configurations were then allowed to evolve under NVE dynamics to | 
| 240 |  |  | sample from the proper distribution in phase space. Figure | 
| 241 |  |  | \ref{fig:images_cooling_time_traces} illustrates this sampling. | 
| 242 | chuckv | 3226 |  | 
| 243 |  |  |  | 
| 244 |  |  | \begin{figure}[htbp] | 
| 245 |  |  | \centering | 
| 246 |  |  | \includegraphics[height=3in]{images/cooling_time_traces.pdf} | 
| 247 | gezelter | 3230 | \caption{Illustrative cooling profile for the 40 {\AA} | 
| 248 |  |  | nanoparticle evolving under stochastic boundary conditions | 
| 249 |  |  | corresponding to $G=$$87.5\times 10^{6}$ | 
| 250 |  |  | $(\mathrm{Wm^{-2}K^{-1}})$. At temperatures along the cooling | 
| 251 |  |  | trajectory, configurations were sampled and allowed to evolve in the | 
| 252 |  |  | NVE ensemble. These subsequent trajectories were analyzed for | 
| 253 |  |  | structural features associated with bulk glass formation.} | 
| 254 | chuckv | 3226 | \label{fig:images_cooling_time_traces} | 
| 255 |  |  | \end{figure} | 
| 256 |  |  |  | 
| 257 |  |  |  | 
| 258 | gezelter | 3233 | \begin{figure}[htbp] | 
| 259 |  |  | \centering | 
| 260 | gezelter | 3242 | \includegraphics[width=5in]{images/cross_section_array.jpg} | 
| 261 | gezelter | 3233 | \caption{Cutaway views of 30 \AA\ Ag-Cu nanoparticle structures for | 
| 262 |  |  | random alloy (top) and Cu (core) / Ag (shell) initial conditions | 
| 263 |  |  | (bottom).  Shown from left to right are the crystalline, liquid | 
| 264 |  |  | droplet, and final glassy bead configurations.} | 
| 265 | gezelter | 3242 | \label{fig:cross_sections} | 
| 266 | gezelter | 3233 | \end{figure} |