ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nanoglass/experimental.tex
(Generate patch)

Comparing trunk/nanoglass/experimental.tex (file contents):
Revision 3208 by chuckv, Thu Aug 23 17:47:31 2007 UTC vs.
Revision 3233 by gezelter, Thu Sep 27 18:45:55 2007 UTC

# Line 1 | Line 1
1 < \input{header.tex}
2 < \section{Simulation Method} \label{sec:details} Several different potential models have been developed that reasonably describe interactions in transition metals. In particular, the Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study a wide range of phenomena in both bulk materials and nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r} Both potentials are based on a model of a metal which treats the nuclei and core electrons as pseudo-atoms embedded in the electron density due to the valence electrons on all of the other atoms in the system. The {\sc sc} potential has a simple form that closely resembles that of the ubiquitous Lennard Jones potential,
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 > $\mathrm{Ag}_6\mathrm{Cu}_4$.  Both initial geometries were considered
11 > although experimental results suggest that the random structure is the
12 > most likely structure to be found following
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,
38   \begin{equation}
39 <        \label{eq:SCP1} 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] ,
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 < where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
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 > community.\cite{BROOKS:1985kx,BROOKS:1983uq,BRUNGER:1984fj} A sketch
50 > 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 <        \label{eq:SCP2} 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}}.
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   \end{equation}
60 < $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for interactions of the pseudo-atom cores. The $\sqrt{\rho_i}$ term in Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models the interactions between the valence electrons and the cores of the pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy scale, $c_i$ scales the attractive portion of the potential relative to the repulsive interaction and $\alpha_{ij}$ is a length parameter that assures a dimensionless form for V and $\rho$. These parameters are tuned to various experimental properties such as density, cohesive energy, elastic moduli for FCC transition metals. The quantum Sutton-Chen ({\sc q-sc}) formulation matches these properties while including zero-point quantum corrections for different transition metals.\cite{PhysRevB.59.3527} This particular parametarization has been shown to reproduce the experimentally available heat of mixing data for both FCC solid solutions of Ag-Cu and the high-temperature liquid.\cite{sheng:184203} Alternatively, the {\sc eam} potential does not reproduce the experimentally observed heat of mixing for the liquid alloy.\cite{MURRAY:1984lr} Combination rules for the alloy were taken to be the arithmatic average of the atomic parameters with the exception of $c_i$ since its values is only dependent on the identity of the atom where the density is evaluated. Cutoff distances are defined by this potential to be $2\alpha_{ij}$ and include up to the sixth coordination shell in FCC metals.
60 > 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 > \begin{figure}[htbp]
65 > \centering
66 > \includegraphics[width=\linewidth]{images/stochbound.pdf}
67 > \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 > The radial cutoff between the two dynamical regions was set to 4 {\AA}
72 > smaller than the original radius of the liquid droplet.}
73 > \label{fig:langevinSketch}
74 > \end{figure}
75  
76 < Cooling rates for the various sized nanoparticles were calculated from the heat transfer equations for a spherical particle embedded in a ambient medium that allows for diffusive heat transport. These equate to a set of coupled differential equations in the Laplace domain and are given by
76 > The viscosity ($\eta$) can be tuned by comparing the cooling rate that
77 > a set of nanoparticles experience with the known cooling rates for
78 > similar particles obtained via the laser heating experiments.
79 > Essentially, we tune the solvent viscosity until the thermal decay
80 > profile matches a heat-transfer model using reasonable values for the
81 > interfacial conductance and the thermal conductivity of the solvent.
82 >
83 > Cooling rates for the experimentally-observed nanoparticles were
84 > calculated from the heat transfer equations for a spherical particle
85 > embedded in a ambient medium that allows for diffusive heat transport.
86 > Following Plech {\it et al.},\cite{plech:195423} we use a heat
87 > transfer model that consists of two coupled differential equations
88 > in the Laplace domain,
89   \begin{eqnarray}
90 <        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\\
91 <        \left(\frac{
92 <        \partial}{
93 <        \partial r} T_{f}(r,s)\right)_{r=R} + \frac{G}{K}(T_{p}(s)-T_{f}(r,s) = 0 \label{eq:heateqn}
90 > 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\\
91 > \left(\frac{\partial}{\partial r} T_{f}(r,s)\right)_{r=R} +
92 > \frac{G}{K}(T_{p}(s)-T_{f}(r,s) = 0
93 > \label{eq:heateqn}
94   \end{eqnarray}
95 < where s is the time-conjugated variable in Laplace space. The variables in these equations describe a nanoparticle of radius $R$, mass $M$, and specific heat $c_{p}$ at an initial temperature $T_0$. The surrounding solvent at a temperature $T_f(r,t)$ has a thermal conductivity $K$, density $\rho$, and specific heat $c$. $G$ is the interface conductance between the nanoparticle and the surrounding solvent. The temperature of the nanoparticle as a function of time is then obtained by the inverse Laplace transformation
95 > where $s$ is the time-conjugate variable in Laplace space. The
96 > variables in these equations describe a nanoparticle of radius $R$,
97 > mass $M$, and specific heat $c_{p}$ at an initial temperature
98 > $T_0$. The surrounding solvent has a thermal profile $T_f(r,t)$,
99 > thermal conductivity $K$, density $\rho$, and specific heat $c$. $G$
100 > is the interfacial conductance between the nanoparticle and the
101 > surrounding solvent, and contains information about heat transfer to
102 > the capping agent as well as the direct metal-to-solvent heat loss.
103 > The temperature of the nanoparticle as a function of time can then
104 > obtained by the inverse Laplace transform,
105   \begin{equation}
106 <        T_{p}(t)=\frac{2 k R^2 g^2 T_0}{\pi}\int_{0}^{\infty}\frac{\exp(-\kappa u^2 t/R^2)u^2}{(u^2(1+Rg)-kRg)^2+(u^3-kRgu)^2}\mathrm{d}u.
106 > T_{p}(t)=\frac{2 k R^2 g^2
107 > T_0}{\pi}\int_{0}^{\infty}\frac{\exp(-\kappa u^2
108 > t/R^2)u^2}{(u^2(1 + R g) - k R g)^2+(u^3 - k R g u)^2}\mathrm{d}u.
109 > \label{eq:laplacetransform}
110   \end{equation}
111 < For clarity, we introduce the diffusivity $\kappa = K/(pc)$, and define $k=4\pi R^3 pc /(Mc_p)$ and $g = G/K$. This equation is solved numerically for the Ag-Cu system using the mole-fraction weighted values for $c_p$ and $\rho$ of 0.295 $(\mathrm{J g^{-1} K^{-1}})$ and $9.826\times 10^6$ $(\mathrm{g m^{-3}})$ respectively. Parameters used for the fluid are $K$ of $0.6$ $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$ $(\mathrm{g m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$. Interfacial conductance have been determined by a variety of experimental techniques and range from a low $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ to $120\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ in the high end of the heat transfer regime. We have chosen $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ for the value of G to present the slowest heat transfer from the nanoparticle to the solvent consistent with experimental observation.
111 > For simplicity, we have introduced the thermal diffusivity $\kappa =
112 > K/(\rho c)$,  and defined $k=4\pi R^3 \rho c /(M c_p)$ and $g = G/K$ in
113 > Eq. \ref{eq:laplacetransform}.
114  
115 < To approximate laser heating-cooling experiments, we utilized a methodology in which atoms contained in the outer $4\AA$ radius of the nanoparticle evolve under Langevin Dynamics approximating the solvent-nanoparticle interaction and atoms outside of this region evolve under Newtonian Dynamics. For a spherical particle of radius $a$, the Langevin frictional force can be determined by Stokes' law
116 < \begin{equation}
117 < \mathbf{F}_{\mathrm{frictional}}=6\pi a \eta \mathbf{v}
115 > Eq. \ref{eq:laplacetransform} was solved numerically for the Ag-Cu
116 > system using mole-fraction weighted values for $c_p$ and $\rho_p$ of
117 > 0.295 $(\mathrm{J g^{-1} K^{-1}})$ and $9.826\times 10^6$ $(\mathrm{g
118 > m^{-3}})$ respectively. Since most of the laser excitation experiments
119 > have been done in aqueous solutions, parameters used for the fluid are
120 > $K$ of $0.6$ $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$
121 > $(\mathrm{g m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
122 >
123 > Values for the interfacial conductance have been determined by a
124 > number of groups for similar nanoparticles and range from a low
125 > $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ to $120\times 10^{6}$
126 > $(\mathrm{Wm^{-2}K^{-1}})$.\cite{hartlandPrv2007} Similarly, Plech
127 > {\it et al.}  reported a value for the interfacial conductance of
128 > $G=105\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ and $G=130\pm 15$
129 > $(\mathrm{Wm^{-2}K^{-1}})$ for Pt
130 > nanoparticles.\cite{plech:195423,PhysRevB.66.224301}
131 >
132 > We conducted our simulations at both ends of the range of
133 > experimentally-determined values for the interfacial conductance.
134 > This allows us to observe both the slowest and fastest heat transfers
135 > from the nanoparticle to the solvent that are consistent with
136 > experimental observations.  For the slowest heat transfer, a value for
137 > G of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ was used and for
138 > the fastest heat transfer, a value of $117\times 10^{6}$
139 > $(\mathrm{Wm^{-2}K^{-1}})$ was used.  Based on calculations we have
140 > done using raw data from the Hartland group's thermal half-time
141 > experiments on Au nanospheres, the true G values are probably in the
142 > faster regime: $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$.
143 >
144 > The rate of cooling for the nanoparticles in a molecular dynamics
145 > simulation can then be tuned by changing the effective solvent
146 > viscosity ($\eta$) until the nanoparticle cooling rate matches the
147 > cooling rate described by the heat-transfer equations
148 > (\ref{eq:heateqn}). The effective solvent viscosity (in poise) for a G
149 > of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.17, 0.20, and
150 > 0.22 for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
151 > effective solvent viscosity (again in poise) for an interfacial
152 > conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.23,
153 > 0.29, and 0.30 for 20 {\AA}, 30 {\AA} and 40 {\AA} particles.  Cooling
154 > traces for each particle size are presented in
155 > Fig. \ref{fig:images_cooling_plot}. It should be noted that the
156 > Langevin thermostat produces cooling curves that are consistent with
157 > Newtonian (single-exponential) cooling, which cannot match the cooling
158 > profiles from Eq. \ref{eq:laplacetransform} exactly. Fitting the
159 > Langevin cooling profiles to a single-exponential produces
160 > $\tau=25.576$ ps, $\tau=43.786$ ps, and $\tau=56.621$ ps for the 20,
161 > 30 and 40 {\AA} nanoparticles and a G of $87.5\times 10^{6}$
162 > $(\mathrm{Wm^{-2}K^{-1}})$.  For comparison's sake, similar
163 > single-exponential fits with an interfacial conductance of G of
164 > $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ produced a $\tau=13.391$
165 > ps, $\tau=30.426$ ps, $\tau=43.857$ ps for the 20, 30 and 40 {\AA}
166 > nanoparticles.
167 >
168 > \begin{figure}[htbp]
169 > \centering
170 > \includegraphics[width=\linewidth]{images/cooling_plot.pdf}
171 > \caption{Thermal cooling curves obtained from the inverse Laplace
172 > transform heat model in Eq. \ref{eq:laplacetransform} (solid line) as
173 > well as from molecular dynamics simulations (circles).  Effective
174 > solvent viscosities of 0.23-0.30 poise (depending on the radius of the
175 > particle) give the best fit to the experimental cooling curves.
176 > %Since
177 > %this viscosity is substantially in excess of the viscosity of liquid
178 > %water, much of the thermal transfer to the surroundings is probably
179 > %due to the capping agent.
180 > }
181 > \label{fig:images_cooling_plot}
182 > \end{figure}
183 >
184 > \subsection{Potentials for classical simulations of bimetallic
185 > nanoparticles}
186 >
187 > Several different potential models have been developed that reasonably
188 > describe interactions in transition metals. In particular, the
189 > Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and
190 > Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study
191 > a wide range of phenomena in both bulk materials and
192 > nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r} Both
193 > potentials are based on a model of a metal which treats the nuclei and
194 > core electrons as pseudo-atoms embedded in the electron density due to
195 > the valence electrons on all of the other atoms in the system. The
196 > {\sc sc} potential has a simple form that closely resembles that of
197 > the ubiquitous Lennard Jones potential,
198 > \begin{equation}
199 > \label{eq:SCP1}
200 > 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] ,
201   \end{equation}
202 < where $\eta$ is the viscosity of the solvent in which the particle is embedded. The rate of cooling for the nanoparticle can be tuned by changing the solvent viscosity until the nanoparticle cooling rate matches that of the cooling rate as described by the heat-transfer equations (\ref{eq:heateqn}).
202 > where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
203 > \begin{equation}
204 > \label{eq:SCP2}
205 > 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}}.
206 > \end{equation}
207 > $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
208 > interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
209 > Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
210 > the interactions between the valence electrons and the cores of the
211 > pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
212 > scale, $c_i$ scales the attractive portion of the potential relative
213 > to the repulsive interaction and $\alpha_{ij}$ is a length parameter
214 > that assures a dimensionless form for $\rho$. These parameters are
215 > tuned to various experimental properties such as the density, cohesive
216 > energy, and elastic moduli for FCC transition metals. The quantum
217 > Sutton-Chen ({\sc q-sc}) formulation matches these properties while
218 > including zero-point quantum corrections for different transition
219 > metals.\cite{PhysRevB.59.3527} This particular parametarization has
220 > been shown to reproduce the experimentally available heat of mixing
221 > data for both FCC solid solutions of Ag-Cu and the high-temperature
222 > liquid.\cite{sheng:184203} In contrast, the {\sc eam} potential does
223 > not reproduce the experimentally observed heat of mixing for the
224 > liquid alloy.\cite{MURRAY:1984lr} Combination rules for the alloy were
225 > taken to be the arithmatic average of the atomic parameters with the
226 > exception of $c_i$ since its values is only dependent on the identity
227 > of the atom where the density is evaluated.  For the {\sc q-sc}
228 > potential, cutoff distances are traditionally taken to be
229 > $2\alpha_{ij}$ and include up to the sixth coordination shell in FCC
230 > metals.
231  
232 < Core-shell (Cu core) and random initial structures were constructed on an underlying FCC lattice (4.09 \AA) at the bulk eutectic composition $\mathrm{Ag}_6\mathrm{Cu}_4$. Three different sizes corresponding to 20 \AA (1961 atoms), 30 \AA (6603 atoms) and 40 \AA (1683 atoms) in radius. Only the random initial structure was studied for the 40 \AA nanoparticle. These initial structures were relaxed to their equilibrium structures at 20 K for 20 ps and again at 300 K for 100 ps sampling from a Maxwell-Boltzmann distribution at each temperature.
232 > %\subsection{Sampling single-temperature configurations from a cooling
233 > %trajectory}
234  
235 + To better understand the structural changes occurring in the
236 + nanoparticles throughout the cooling trajectory, configurations were
237 + sampled at regular intervals during the cooling trajectory. These
238 + configurations were then allowed to evolve under NVE dynamics to
239 + sample from the proper distribution in phase space. Figure
240 + \ref{fig:images_cooling_time_traces} illustrates this sampling.
241  
242 < \input{footer.tex}
242 >
243 > \begin{figure}[htbp]
244 >        \centering
245 >                \includegraphics[height=3in]{images/cooling_time_traces.pdf}
246 >        \caption{Illustrative cooling profile for the 40 {\AA}
247 > nanoparticle evolving under stochastic boundary conditions
248 > corresponding to $G=$$87.5\times 10^{6}$
249 > $(\mathrm{Wm^{-2}K^{-1}})$. At temperatures along the cooling
250 > trajectory, configurations were sampled and allowed to evolve in the
251 > NVE ensemble. These subsequent trajectories were analyzed for
252 > structural features associated with bulk glass formation.}
253 >        \label{fig:images_cooling_time_traces}
254 > \end{figure}
255 >
256 >
257 > \begin{figure}[htbp]
258 > \centering
259 > \includegraphics[width=\linewidth]{images/cross_section_array.jpg}
260 > \caption{Cutaway views of 30 \AA\ Ag-Cu nanoparticle structures for
261 > random alloy (top) and Cu (core) / Ag (shell) initial conditions
262 > (bottom).  Shown from left to right are the crystalline, liquid
263 > droplet, and final glassy bead configurations.}
264 > \label{fig:q6}
265 > \end{figure}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines