ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nanoglass/experimental.tex
Revision: 3221
Committed: Thu Sep 6 20:44:02 2007 UTC (17 years ago) by gezelter
Content type: application/x-tex
File size: 10515 byte(s)
Log Message:
more writing

File Contents

# User Rev Content
1 gezelter 3221 \section{Computational Methodology}
2     \label{sec:details}
3    
4     \subsection{Initial Geometries and Heating}
5    
6     Cu-core / Ag-shell and random alloy structures were constructed on an
7     underlying FCC lattice (4.09 {\AA}) at the bulk eutectic composition
8     $\mathrm{Ag}_6\mathrm{Cu}_4$. Three different sizes of nanoparticles
9     corresponding to a 20 \AA radius (1961 atoms), 30 {\AA} radius (6603
10     atoms) and 40 {\AA} radius (15683 atoms) were constructed. These
11     initial structures were relaxed to their equilibrium structures at 20
12     K for 20 ps and again at 300 K for 100 ps sampling from a
13     Maxwell-Boltzmann distribution at each temperature.
14    
15     To mimic the effects of the heating due to laser irradiation, the
16     particles were allowed to melt by sampling velocities from the Maxwell
17     Boltzmann distribution at a temperature of 900 K. The particles were
18     run under microcanonical simulation conditions for 1 ns of simualtion
19     time prior to studying the effects of heat transfer to the solvent.
20     In all cases, center of mass translational and rotational motion of
21     the particles were set to zero before any data collection was
22     undertaken. Structural features (pair distribution functions) were
23     used to verify that the particles were indeed liquid droplets before
24     cooling simulations took place.
25    
26     \subsection{Modeling random alloy and core shell particles in solution
27     phase environments}
28    
29     To approximate the effects of rapid heat transfer to the solvent
30     following a heating at the plasmon resonance, we utilized a
31     methodology in which atoms contained in the outer $4$ {\AA} radius of
32     the nanoparticle evolved under Langevin Dynamics with a solvent
33     friction approximating the contribution from the solvent and capping
34     agent. Atoms located in the interior of the nanoparticle evolved
35     under Newtonian dynamics. The set-up of our simulations is nearly
36     identical with the ``stochastic boundary molecular dynamics'' ({\sc
37     sbmd}) method that has seen wide use in the protein simulation
38     community.\cite{BROOKS:1985kx,BROOKS:1983uq,BRUNGER:1984fj} A sketch
39     of this setup can be found in Fig. \ref{fig:langevinSketch}. For a
40     spherical atom of radius $a$, the Langevin frictional forces can be
41     determined by Stokes' law
42     \begin{equation}
43     \mathbf{F}_{\mathrm{frictional}}=6\pi a \eta \mathbf{v}
44 chuckv 3208 \end{equation}
45 gezelter 3221 where $\eta$ is the effective viscosity of the solvent in which the
46     particle is embedded. Due to the presence of the capping agent and
47     the lack of details about the atomic-scale interactions between the
48     metallic atoms and the solvent, the effective viscosity is a
49     essentially a free parameter that must be tuned to give experimentally
50     relevant simulations.
51 chuckv 3208
52 gezelter 3221 The viscosity ($\eta$) can be tuned by comparing the cooling rate that
53     a set of nanoparticles experience with the known cooling rates for
54     those particles obtained via the laser heating experiments.
55     Essentially, we tune the solvent viscosity until the thermal decay
56     profile matches a heat-transfer model using reasonable values for the
57     interfacial conductance and the thermal conductivity of the solvent.
58    
59     Cooling rates for the experimentally-observed nanoparticles were
60     calculated from the heat transfer equations for a spherical particle
61     embedded in a ambient medium that allows for diffusive heat
62     transport. The heat transfer model is a set of two coupled
63     differential equations in the Laplace domain,
64 chuckv 3208 \begin{eqnarray}
65 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\\
66     \left(\frac{\partial}{\partial r} T_{f}(r,s)\right)_{r=R} +
67     \frac{G}{K}(T_{p}(s)-T_{f}(r,s) = 0
68     \label{eq:heateqn}
69 chuckv 3208 \end{eqnarray}
70 gezelter 3221 where $s$ is the time-conjugate variable in Laplace space. The
71     variables in these equations describe a nanoparticle of radius $R$,
72     mass $M$, and specific heat $c_{p}$ at an initial temperature
73     $T_0$. The surrounding solvent has a thermal profile $T_f(r,t)$,
74     thermal conductivity $K$, density $\rho$, and specific heat $c$. $G$
75     is the interfacial conductance between the nanoparticle and the
76     surrounding solvent, and contains information about heat transfer to
77     the capping agent as well as the direct metal-to-solvent heat loss.
78     The temperature of the nanoparticle as a function of time can then
79     obtained by the inverse Laplace transform,
80 chuckv 3208 \begin{equation}
81 gezelter 3221 T_{p}(t)=\frac{2 k R^2 g^2
82     T_0}{\pi}\int_{0}^{\infty}\frac{\exp(-\kappa u^2
83     t/R^2)u^2}{(u^2(1 + R g) - k R g)^2+(u^3 - k R g u)^2}\mathrm{d}u.
84     \label{eq:laplacetransform}
85 chuckv 3208 \end{equation}
86 gezelter 3221 For simplicity, we have introduced the thermal diffusivity $\kappa =
87     K/(\rho c)$, and defined $k=4\pi R^3 \rho c /(M c_p)$ and $g = G/K$ in
88     Eq. \ref{eq:laplacetransform}.
89 chuckv 3208
90 gezelter 3221 Eq. \ref{eq:laplacetransform} was solved numerically for the Ag-Cu
91     system using mole-fraction weighted values for $c_p$ and $\rho_p$ of
92     0.295 $(\mathrm{J g^{-1} K^{-1}})$ and $9.826\times 10^6$ $(\mathrm{g
93     m^{-3}})$ respectively. Since most of the laser excitation experiments
94     have been done in aqueous solutions, parameters used for the fluid are
95     $K$ of $0.6$ $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$ $(\mathrm{g
96     m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
97 chuckv 3208
98 gezelter 3221 Values for the interfacial conductance have been determined by a
99     number of groups for similar nanoparticles and range from a low
100     $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ to $120\times 10^{6}$
101     $(\mathrm{Wm^{-2}K^{-1}})$.\cite{XXX}
102    
103     We conducted our simulations at both ends of the range of
104     experimentally-determined values for the interfacial conductance.
105     This allows us to observe both the slowest and fastest heat transfers
106     from the nanoparticle to the solvent that are consistent with
107     experimental observations. For the slowest heat transfer, a value for
108     G of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ was used and for
109     the fastest heat transfer, a value of $117\times 10^{6}$
110     $(\mathrm{Wm^{-2}K^{-1}})$ was used. Based on calculations we have
111     done using raw data from the Hartland group's thermal half-time
112     experiments on Au nanospheres, we believe that the true G values are
113     closer to the faster regime: $117\times 10^{6}$
114     $(\mathrm{Wm^{-2}K^{-1}})$.
115    
116     The rate of cooling for the nanoparticles in a molecular dynamics
117     simulation can then be tuned by changing the effective solvent
118     viscosity ($\eta$) until the nanoparticle cooling rate matches the
119     cooling rate described by the heat-transfer equations
120     (\ref{eq:heateqn}). The effective solvent viscosity (in poise) for a G
121     of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.17, 0.20, and
122     0.22 for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
123     effective solvent viscosity (again in poise) for an interfacial
124     conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.23,
125     0.29, and 0.30 for 20 {\AA}, 30 {\AA} and 40 {\AA} particles. Cooling
126     traces for each particle size are presented in
127     Fig. \ref{fig:images_cooling_plot}. It should be noted that the
128     Langevin thermostat produces cooling curves that are consistent with
129     Newtonian (single-exponential) cooling, which cannot match the cooling
130     profiles from Eq. \ref{eq:laplacetransform} exactly.
131    
132 chuckv 3213 \begin{figure}[htbp]
133 gezelter 3221 \centering
134     \includegraphics[width=\linewidth]{images/cooling_plot.pdf}
135     \caption{Thermal cooling curves obtained from the inverse Laplace
136     transform heat model in Eq. \ref{eq:laplacetransform} (solid line) as
137     well as from molecular dynamics simulations (circles). Effective
138     solvent viscosities of 0.23-0.30 poise (depending on the radius of the
139     particle) give the best fit to the experimental cooling curves. Since
140     this viscosity is substantially in excess of the viscosity of liquid
141     water, much of the thermal transfer to the surroundings is probably
142     due to the capping agent.}
143     \label{fig:images_cooling_plot}
144 chuckv 3213 \end{figure}
145 chuckv 3208
146 gezelter 3221 \subsection{Potentials for classical simulations of bimetallic
147     nanoparticles}
148 chuckv 3208
149 gezelter 3221 Several different potential models have been developed that reasonably
150     describe interactions in transition metals. In particular, the
151     Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and
152     Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study
153     a wide range of phenomena in both bulk materials and
154     nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r} Both
155     potentials are based on a model of a metal which treats the nuclei and
156     core electrons as pseudo-atoms embedded in the electron density due to
157     the valence electrons on all of the other atoms in the system. The
158     {\sc sc} potential has a simple form that closely resembles that of
159     the ubiquitous Lennard Jones potential,
160     \begin{equation}
161     \label{eq:SCP1}
162     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] ,
163     \end{equation}
164     where $V^{pair}_{ij}$ and $\rho_{i}$ are given by
165     \begin{equation}
166     \label{eq:SCP2}
167     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}}.
168     \end{equation}
169     $V^{pair}_{ij}$ is a repulsive pairwise potential that accounts for
170     interactions between the pseudoatom cores. The $\sqrt{\rho_i}$ term in
171     Eq. (\ref{eq:SCP1}) is an attractive many-body potential that models
172     the interactions between the valence electrons and the cores of the
173     pseudo-atoms. $D_{ij}$, $D_{ii}$ set the appropriate overall energy
174     scale, $c_i$ scales the attractive portion of the potential relative
175     to the repulsive interaction and $\alpha_{ij}$ is a length parameter
176     that assures a dimensionless form for $\rho$. These parameters are
177     tuned to various experimental properties such as the density, cohesive
178     energy, and elastic moduli for FCC transition metals. The quantum
179     Sutton-Chen ({\sc q-sc}) formulation matches these properties while
180     including zero-point quantum corrections for different transition
181     metals.\cite{PhysRevB.59.3527} This particular parametarization has
182     been shown to reproduce the experimentally available heat of mixing
183     data for both FCC solid solutions of Ag-Cu and the high-temperature
184     liquid.\cite{sheng:184203} In contrast, the {\sc eam} potential does
185     not reproduce the experimentally observed heat of mixing for the
186     liquid alloy.\cite{MURRAY:1984lr} Combination rules for the alloy were
187     taken to be the arithmatic average of the atomic parameters with the
188     exception of $c_i$ since its values is only dependent on the identity
189     of the atom where the density is evaluated. For the {\sc q-sc}
190     potential, cutoff distances are traditionally taken to be
191     $2\alpha_{ij}$ and include up to the sixth coordination shell in FCC
192     metals.
193 chuckv 3213
194 gezelter 3221 \subsection{Sampling single-temperature configurations from a cooling
195     trajectory}
196 chuckv 3213
197 gezelter 3221 ffdsafjdksalfdsa