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 3222 by chuckv, Tue Sep 11 15:23:24 2007 UTC vs.
Revision 3242 by gezelter, Tue Oct 2 22:01:48 2007 UTC

# Line 1 | Line 1
1 + %!TEX root = /Users/charles/Desktop/nanoglass/nanoglass.tex
2 +
3   \section{Computational Methodology}
4   \label{sec:details}
5  
# Line 5 | Line 7 | $\mathrm{Ag}_6\mathrm{Cu}_4$.  Three different sizes o
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 < $\mathrm{Ag}_6\mathrm{Cu}_4$.  Three different sizes of nanoparticles
11 < corresponding to a 20 \AA radius (1961 atoms), 30 {\AA} radius (6603
12 < atoms) and 40 {\AA} radius (15683 atoms) were constructed.  These
13 < initial structures were relaxed to their equilibrium structures at 20
14 < K for 20 ps and again at 300 K for 100 ps sampling from a
15 < Maxwell-Boltzmann distribution at each temperature.
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
# Line 29 | Line 34 | the nanoparticle evolved under Langevin Dynamics with
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
37 > 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   community.\cite{BROOKS:1985kx,BROOKS:1983uq,BRUNGER:1984fj} A sketch
50 < of this setup can be found in Fig. \ref{fig:langevinSketch}.  For a
51 < spherical atom of radius $a$, the Langevin frictional forces can be
52 < determined by Stokes' law
53 < \begin{equation}
54 < \mathbf{F}_{\mathrm{frictional}}=6\pi a \eta \mathbf{v}
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 > \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 < where $\eta$ is the effective viscosity of the solvent in which the
61 < particle is embedded.  Due to the presence of the capping agent and
62 < the lack of details about the atomic-scale interactions between the
63 < 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.
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 for nanoparticle cooling. Equations of motion for metal atoms contained in the outer 4 {\AA} were determined by Langevins' Equations of motion. Metal atoms outside this region were allowed to evolve under Newtonian dynamics.}
66 > \includegraphics[width=5in]{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 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   \label{fig:langevinSketch}
75   \end{figure}
76 +
77   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 < those particles obtained via the laser heating experiments.
79 > similar particles obtained via the laser heating experiments.
80   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 < embedded in a ambient medium that allows for diffusive heat
87 < transport. The heat transfer model is a set of two coupled
88 < differential equations in the Laplace domain,
86 > 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   \begin{eqnarray}
91   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} +
# Line 97 | Line 118 | $K$ of $0.6$  $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1
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 < $K$ of $0.6$  $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$ $(\mathrm{g
122 < m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
121 > $K$ 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  
124   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 < $(\mathrm{Wm^{-2}K^{-1}})$.\cite{hartlandPrv2007} Plech {\it et al.} reported a value for the interfacial conductance of $G=105\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ and
128 < $G=130\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ for Pt nanoparticles.\cite{plech:195423,PhysRevB.66.224301}
127 > $(\mathrm{Wm^{-2}K^{-1}})$.\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  
133   We conducted our simulations at both ends of the range of
134   experimentally-determined values for the interfacial conductance.
# Line 115 | Line 139 | experiments on Au nanospheres, we believe that the tru
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 < experiments on Au nanospheres, we believe that the true G values are
143 < closer to the faster regime: $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$.
142 > 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  
121
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
# Line 133 | Line 156 | profiles from Eq. \ref{eq:laplacetransform} exactly.
156   Fig. \ref{fig:images_cooling_plot}. It should be noted that the
157   Langevin thermostat produces cooling curves that are consistent with
158   Newtonian (single-exponential) cooling, which cannot match the cooling
159 < profiles from Eq. \ref{eq:laplacetransform} exactly.
159 > profiles from Eq. \ref{eq:laplacetransform} exactly. Fitting the
160 > Langevin cooling profiles to a single-exponential produces
161 > $\tau=25.576$ ps, $\tau=43.786$ ps, and $\tau=56.621$ ps for the 20,
162 > 30 and 40 {\AA} nanoparticles and a G of $87.5\times 10^{6}$
163 > $(\mathrm{Wm^{-2}K^{-1}})$.  For comparison's sake, similar
164 > single-exponential fits with an interfacial conductance of G of
165 > $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ produced a $\tau=13.391$
166 > ps, $\tau=30.426$ ps, $\tau=43.857$ ps for the 20, 30 and 40 {\AA}
167 > nanoparticles.
168  
169   \begin{figure}[htbp]
170   \centering
171 < \includegraphics[width=\linewidth]{images/cooling_plot.pdf}
171 > \includegraphics[width=5in]{images/cooling_plot.pdf}
172   \caption{Thermal cooling curves obtained from the inverse Laplace
173   transform heat model in Eq. \ref{eq:laplacetransform} (solid line) as
174   well as from molecular dynamics simulations (circles).  Effective
175   solvent viscosities of 0.23-0.30 poise (depending on the radius of the
176 < particle) give the best fit to the experimental cooling curves.  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.}
176 > particle) give the best fit to the experimental cooling curves.
177 > %Since
178 > %this viscosity is substantially in excess of the viscosity of liquid
179 > %water, much of the thermal transfer to the surroundings is probably
180 > %due to the capping agent.
181 > }
182   \label{fig:images_cooling_plot}
183   \end{figure}
184  
# Line 197 | Line 230 | metals.
230   $2\alpha_{ij}$ and include up to the sixth coordination shell in FCC
231   metals.
232  
233 < \subsection{Sampling single-temperature configurations from a cooling
234 < trajectory}
233 > %\subsection{Sampling single-temperature configurations from a cooling
234 > %trajectory}
235  
236 < ffdsafjdksalfdsa
236 > 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 >
243 >
244 > \begin{figure}[htbp]
245 >        \centering
246 >                \includegraphics[height=3in]{images/cooling_time_traces.pdf}
247 >        \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 >        \label{fig:images_cooling_time_traces}
255 > \end{figure}
256 >
257 >
258 > \begin{figure}[htbp]
259 > \centering
260 > \includegraphics[width=5in]{images/cross_section_array.jpg}
261 > \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 > \label{fig:cross_sections}
266 > \end{figure}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines