ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nanoglass/experimental.tex
Revision: 3230
Committed: Tue Sep 25 19:23:21 2007 UTC (16 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 12993 byte(s)
Log Message:
Edits

File Contents

# Content
1 %!TEX root = /Users/charles/Desktop/nanoglass/nanoglass.tex
2
3 \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 $\mathrm{Ag}_6\mathrm{Cu}_4$. All three compositions were considered
11 although experimental results suggest that the random structure is the
12 most likely composition after
13 synthesis.\cite{Jiang:2005lr,gonzalo:5163} Three different sizes of
14 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
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 the nanoparticle evolved under Langevin Dynamics with a solvent
38 friction approximating the contribution from the solvent and capping
39 agent. Atoms located in the interior of the nanoparticle evolved
40 under Newtonian dynamics. The set-up of our simulations is nearly
41 identical with the ``stochastic boundary molecular dynamics'' ({\sc
42 sbmd}) method that has seen wide use in the protein simulation
43 community.\cite{BROOKS:1985kx,BROOKS:1983uq,BRUNGER:1984fj} A sketch
44 of this setup can be found in Fig. \ref{fig:langevinSketch}. For a
45 spherical atom of radius $a$, the Langevin frictional forces can be
46 determined by Stokes' law
47 \begin{equation}
48 \mathbf{F}_{\mathrm{frictional}}=6\pi a \eta \mathbf{v}
49 \end{equation}
50 where $\eta$ is the effective viscosity of the solvent in which the
51 particle is embedded. Due to the presence of the capping agent and
52 the lack of details about the atomic-scale interactions between the
53 metallic atoms and the solvent, the effective viscosity is a
54 essentially a free parameter that must be tuned to give experimentally
55 relevant simulations.
56 \begin{figure}[htbp]
57 \centering
58 \includegraphics[width=\linewidth]{images/stochbound.pdf}
59 \caption{Methodology used to mimic the experimental cooling conditions
60 of a hot nanoparticle surrounded by a solvent. Atoms in the core of
61 the particle evolved under Newtonian dynamics, while atoms that were
62 in the outer skin of the particle evolved under Langevin dynamics.
63 The radial cutoff between the two dynamical regions was set to 4 {\AA}
64 smaller than the original radius of the liquid droplet.}
65 \label{fig:langevinSketch}
66 \end{figure}
67
68 The viscosity ($\eta$) can be tuned by comparing the cooling rate that
69 a set of nanoparticles experience with the known cooling rates for
70 similar particles obtained via the laser heating experiments.
71 Essentially, we tune the solvent viscosity until the thermal decay
72 profile matches a heat-transfer model using reasonable values for the
73 interfacial conductance and the thermal conductivity of the solvent.
74
75 Cooling rates for the experimentally-observed nanoparticles were
76 calculated from the heat transfer equations for a spherical particle
77 embedded in a ambient medium that allows for diffusive heat transport.
78 Following Plech {\it et al.},\cite{plech:195423} we use a heat
79 transfer model that consists of two coupled differential equations
80 in the Laplace domain,
81 \begin{eqnarray}
82 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\\
83 \left(\frac{\partial}{\partial r} T_{f}(r,s)\right)_{r=R} +
84 \frac{G}{K}(T_{p}(s)-T_{f}(r,s) = 0
85 \label{eq:heateqn}
86 \end{eqnarray}
87 where $s$ is the time-conjugate variable in Laplace space. The
88 variables in these equations describe a nanoparticle of radius $R$,
89 mass $M$, and specific heat $c_{p}$ at an initial temperature
90 $T_0$. The surrounding solvent has a thermal profile $T_f(r,t)$,
91 thermal conductivity $K$, density $\rho$, and specific heat $c$. $G$
92 is the interfacial conductance between the nanoparticle and the
93 surrounding solvent, and contains information about heat transfer to
94 the capping agent as well as the direct metal-to-solvent heat loss.
95 The temperature of the nanoparticle as a function of time can then
96 obtained by the inverse Laplace transform,
97 \begin{equation}
98 T_{p}(t)=\frac{2 k R^2 g^2
99 T_0}{\pi}\int_{0}^{\infty}\frac{\exp(-\kappa u^2
100 t/R^2)u^2}{(u^2(1 + R g) - k R g)^2+(u^3 - k R g u)^2}\mathrm{d}u.
101 \label{eq:laplacetransform}
102 \end{equation}
103 For simplicity, we have introduced the thermal diffusivity $\kappa =
104 K/(\rho c)$, and defined $k=4\pi R^3 \rho c /(M c_p)$ and $g = G/K$ in
105 Eq. \ref{eq:laplacetransform}.
106
107 Eq. \ref{eq:laplacetransform} was solved numerically for the Ag-Cu
108 system using mole-fraction weighted values for $c_p$ and $\rho_p$ of
109 0.295 $(\mathrm{J g^{-1} K^{-1}})$ and $9.826\times 10^6$ $(\mathrm{g
110 m^{-3}})$ respectively. Since most of the laser excitation experiments
111 have been done in aqueous solutions, parameters used for the fluid are
112 $K$ of $0.6$ $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$
113 $(\mathrm{g m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
114
115 Values for the interfacial conductance have been determined by a
116 number of groups for similar nanoparticles and range from a low
117 $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ to $120\times 10^{6}$
118 $(\mathrm{Wm^{-2}K^{-1}})$.\cite{hartlandPrv2007} Similarly, Plech
119 {\it et al.} reported a value for the interfacial conductance of
120 $G=105\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ and $G=130\pm 15$
121 $(\mathrm{Wm^{-2}K^{-1}})$ for Pt
122 nanoparticles.\cite{plech:195423,PhysRevB.66.224301}
123
124 We conducted our simulations at both ends of the range of
125 experimentally-determined values for the interfacial conductance.
126 This allows us to observe both the slowest and fastest heat transfers
127 from the nanoparticle to the solvent that are consistent with
128 experimental observations. For the slowest heat transfer, a value for
129 G of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ was used and for
130 the fastest heat transfer, a value of $117\times 10^{6}$
131 $(\mathrm{Wm^{-2}K^{-1}})$ was used. Based on calculations we have
132 done using raw data from the Hartland group's thermal half-time
133 experiments on Au nanospheres, the true G values are probably in the
134 faster regime: $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$.
135
136 The rate of cooling for the nanoparticles in a molecular dynamics
137 simulation can then be tuned by changing the effective solvent
138 viscosity ($\eta$) until the nanoparticle cooling rate matches the
139 cooling rate described by the heat-transfer equations
140 (\ref{eq:heateqn}). The effective solvent viscosity (in poise) for a G
141 of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.17, 0.20, and
142 0.22 for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
143 effective solvent viscosity (again in poise) for an interfacial
144 conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.23,
145 0.29, and 0.30 for 20 {\AA}, 30 {\AA} and 40 {\AA} particles. Cooling
146 traces for each particle size are presented in
147 Fig. \ref{fig:images_cooling_plot}. It should be noted that the
148 Langevin thermostat produces cooling curves that are consistent with
149 Newtonian (single-exponential) cooling, which cannot match the cooling
150 profiles from Eq. \ref{eq:laplacetransform} exactly. Fitting the
151 Langevin cooling profiles to a single-exponential produces
152 $\tau=25.576$ ps, $\tau=43.786$ ps, and $\tau=56.621$ ps for the 20,
153 30 and 40 {\AA} nanoparticles and a G of $87.5\times 10^{6}$
154 $(\mathrm{Wm^{-2}K^{-1}})$. For comparison's sake, similar
155 single-exponential fits with an interfacial conductance of G of
156 $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ produced a $\tau=13.391$
157 ps, $\tau=30.426$ ps, $\tau=43.857$ ps for the 20, 30 and 40 {\AA}
158 nanoparticles.
159
160 \begin{figure}[htbp]
161 \centering
162 \includegraphics[width=\linewidth]{images/cooling_plot.pdf}
163 \caption{Thermal cooling curves obtained from the inverse Laplace
164 transform heat model in Eq. \ref{eq:laplacetransform} (solid line) as
165 well as from molecular dynamics simulations (circles). Effective
166 solvent viscosities of 0.23-0.30 poise (depending on the radius of the
167 particle) give the best fit to the experimental cooling curves.
168 %Since
169 %this viscosity is substantially in excess of the viscosity of liquid
170 %water, much of the thermal transfer to the surroundings is probably
171 %due to the capping agent.
172 }
173 \label{fig:images_cooling_plot}
174 \end{figure}
175
176 \subsection{Potentials for classical simulations of bimetallic
177 nanoparticles}
178
179 Several different potential models have been developed that reasonably
180 describe interactions in transition metals. In particular, the
181 Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and
182 Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study
183 a wide range of phenomena in both bulk materials and
184 nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r} Both
185 potentials are based on a model of a metal which treats the nuclei and
186 core electrons as pseudo-atoms embedded in the electron density due to
187 the valence electrons on all of the other atoms in the system. The
188 {\sc sc} potential has a simple form that closely resembles that of
189 the ubiquitous Lennard Jones potential,
190 \begin{equation}
191 \label{eq:SCP1}
192 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] ,
193 \end{equation}
194 where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
195 \begin{equation}
196 \label{eq:SCP2}
197 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}}.
198 \end{equation}
199 $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
200 interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
201 Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
202 the interactions between the valence electrons and the cores of the
203 pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
204 scale, $c_i$ scales the attractive portion of the potential relative
205 to the repulsive interaction and $\alpha_{ij}$ is a length parameter
206 that assures a dimensionless form for $\rho$. These parameters are
207 tuned to various experimental properties such as the density, cohesive
208 energy, and elastic moduli for FCC transition metals. The quantum
209 Sutton-Chen ({\sc q-sc}) formulation matches these properties while
210 including zero-point quantum corrections for different transition
211 metals.\cite{PhysRevB.59.3527} This particular parametarization has
212 been shown to reproduce the experimentally available heat of mixing
213 data for both FCC solid solutions of Ag-Cu and the high-temperature
214 liquid.\cite{sheng:184203} In contrast, the {\sc eam} potential does
215 not reproduce the experimentally observed heat of mixing for the
216 liquid alloy.\cite{MURRAY:1984lr} Combination rules for the alloy were
217 taken to be the arithmatic average of the atomic parameters with the
218 exception of $c_i$ since its values is only dependent on the identity
219 of the atom where the density is evaluated. For the {\sc q-sc}
220 potential, cutoff distances are traditionally taken to be
221 $2\alpha_{ij}$ and include up to the sixth coordination shell in FCC
222 metals.
223
224 %\subsection{Sampling single-temperature configurations from a cooling
225 %trajectory}
226
227 To better understand the structural changes occurring in the
228 nanoparticles throughout the cooling trajectory, configurations were
229 sampled at regular intervals during the cooling trajectory. These
230 configurations were then allowed to evolve under NVE dynamics to
231 sample from the proper distribution in phase space. Figure
232 \ref{fig:images_cooling_time_traces} illustrates this sampling.
233
234
235 \begin{figure}[htbp]
236 \centering
237 \includegraphics[height=3in]{images/cooling_time_traces.pdf}
238 \caption{Illustrative cooling profile for the 40 {\AA}
239 nanoparticle evolving under stochastic boundary conditions
240 corresponding to $G=$$87.5\times 10^{6}$
241 $(\mathrm{Wm^{-2}K^{-1}})$. At temperatures along the cooling
242 trajectory, configurations were sampled and allowed to evolve in the
243 NVE ensemble. These subsequent trajectories were analyzed for
244 structural features associated with bulk glass formation.}
245 \label{fig:images_cooling_time_traces}
246 \end{figure}
247
248