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 3228 by chuckv, Fri Sep 21 21:31:25 2007 UTC vs.
Revision 3233 by gezelter, Thu Sep 27 18:45:55 2007 UTC

# Line 7 | Line 7 | $\mathrm{Ag}_6\mathrm{Cu}_4$.  All three compositions
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 although experimental results suggest that the random structure is the most likely composition after synthesis.\cite{Jiang:2005lr,gonzalo:5163} 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 31 | 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
51 < essentially a free parameter that must be tuned to give experimentally
52 < 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.}
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   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 < those particles obtained via the laser heating experiments.
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
86 < transport. The heat transfer model is a set of two coupled
87 < differential equations in the Laplace domain,
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{\partial}{\partial r} T_{f}(r,s)\right)_{r=R} +
# Line 99 | Line 117 | $K$ of $0.6$  $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1
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 $1.0\times10^6$ $(\mathrm{g
121 < m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
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} Plech {\it et al.} reported a value for the interfacial conductance of $G=105\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ and
127 < $G=130\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ for Pt nanoparticles.\cite{plech:195423,PhysRevB.66.224301}
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.
# Line 117 | Line 138 | experiments on Au nanospheres, we believe that the tru
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, we believe that the true G values are
142 < closer to the faster regime: $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$.
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  
123
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
# Line 135 | Line 155 | profiles from Eq. \ref{eq:laplacetransform} exactly. F
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 Langevin cooling profiles to a single-exponential produces $\tau=25.576$ ps, $\tau=43.786$ ps, and $\tau=56.621$ ps for the 20, 30 and 40 {\AA} nanoparticles and a G of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$. The faster cooling G of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ produced a $\tau=13.391$ ps, $\tau=30.426$ ps, $\tau=43.857$ ps for the 20, 30 and 40 {\AA} nanoparticles.
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
# Line 144 | Line 172 | particle) give the best fit to the experimental coolin
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.  Since
176 < this viscosity is substantially in excess of the viscosity of liquid
177 < water, much of the thermal transfer to the surroundings is probably
178 < due to the capping agent.}
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  
# Line 202 | Line 232 | To better understand the structural changes occurring
232   %\subsection{Sampling single-temperature configurations from a cooling
233   %trajectory}
234  
235 < To better understand the structural changes occurring in the nanoparticles throughout the cooling trajectory, configurations were sampled at temperatures throughout the cooling trajectory. These configurations were then allowed to evolve under NVE dynamics to sample from the proper distribution in phase space. Figure \ref{fig:images_cooling_time_traces} illustrates this sampling.
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  
243   \begin{figure}[htbp]
244          \centering
245                  \includegraphics[height=3in]{images/cooling_time_traces.pdf}
246 <        \caption{Illustrative cooling profile for the 40 {\AA} nanoparticle evolving under stochastic boundary conditions corresponding to $G=$$87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$. At temperatures along the cooling trajectory, configurations were sampled and allowed to evolve in the NVE ensemble. These subsequent trajectories were analyzed for structural features associated with bulk glass formation.}
246 >        \caption{Illustrative cooling profile for the 40 {\AA}
247 > nanoparticle evolving under stochastic boundary conditions
248 > corresponding to $G=$$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