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 3226 by chuckv, Wed Sep 19 16:53:58 2007 UTC vs.
Revision 3259 by gezelter, Fri Oct 12 21:21:04 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
19 > temperature. All simulations were conducted using the {\sc oopse}
20 > molecular dynamics package.\cite{Meineke:2004uq}
21  
22   To mimic the effects of the heating due to laser irradiation, the
23   particles were allowed to melt by sampling velocities from the Maxwell
# Line 31 | Line 36 | the nanoparticle evolved under Langevin Dynamics with
36   To approximate the effects of rapid heat transfer to the solvent
37   following a heating at the plasmon resonance, we utilized a
38   methodology in which atoms contained in the outer $4$ {\AA} radius of
39 < the nanoparticle evolved under Langevin Dynamics with a solvent
40 < friction approximating the contribution from the solvent and capping
41 < agent.  Atoms located in the interior of the nanoparticle evolved
42 < under Newtonian dynamics.  The set-up of our simulations is nearly
43 < identical with the ``stochastic boundary molecular dynamics'' ({\sc
44 < sbmd}) method that has seen wide use in the protein simulation
39 > the nanoparticle evolved under Langevin Dynamics,
40 > \begin{equation}
41 > m \frac{\partial^2 \vec{x}}{\partial t^2} = F_\textrm{sys}(\vec{x}(t))
42 > - 6 \pi a \eta \vec{v}(t)  + F_\textrm{ran}
43 > \label{eq:langevin}
44 > \end{equation}
45 > with a solvent friction ($\eta$) approximating the contribution from
46 > the solvent and capping agent.  Atoms located in the interior of the
47 > nanoparticle evolved under Newtonian dynamics.  The set-up of our
48 > simulations is nearly identical with the ``stochastic boundary
49 > molecular dynamics'' ({\sc sbmd}) method that has seen wide use in the
50 > protein simulation
51   community.\cite{BROOKS:1985kx,BROOKS:1983uq,BRUNGER:1984fj} A sketch
52 < of this setup can be found in Fig. \ref{fig:langevinSketch}.  For a
53 < spherical atom of radius $a$, the Langevin frictional forces can be
54 < determined by Stokes' law
55 < \begin{equation}
56 < \mathbf{F}_{\mathrm{frictional}}=6\pi a \eta \mathbf{v}
52 > of this setup can be found in Fig. \ref{fig:langevinSketch}.  In
53 > equation \ref{eq:langevin} the frictional forces of a spherical atom
54 > of radius $a$ depend on the solvent viscosity.  The random forces are
55 > usually taken as gaussian random variables with zero mean and a
56 > variance tied to the solvent viscosity and temperature,
57 > \begin{equation}
58 > \langle F_\textrm{ran}(t) \cdot F_\textrm{ran} (t')
59 > \rangle = 2 k_B T (6 \pi \eta a) \delta(t - t')
60 > \label{eq:stochastic}
61   \end{equation}
62 < where $\eta$ is the effective viscosity of the solvent in which the
63 < particle is embedded.  Due to the presence of the capping agent and
64 < the lack of details about the atomic-scale interactions between the
65 < 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.
62 > Due to the presence of the capping agent and the lack of details about
63 > the atomic-scale interactions between the metallic atoms and the
64 > solvent, the effective viscosity is a essentially a free parameter
65 > that must be tuned to give experimentally relevant simulations.
66   \begin{figure}[htbp]
67   \centering
68 < \includegraphics[width=\linewidth]{images/stochbound.pdf}
69 < \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.}
68 > \includegraphics[width=5in]{images/stochbound.pdf}
69 > \caption{Methodology used to mimic the experimental cooling conditions
70 > of a hot nanoparticle surrounded by a solvent.  Atoms in the core of
71 > the particle evolved under Newtonian dynamics, while atoms that were
72 > in the outer skin of the particle evolved under Langevin dynamics.
73 > The radius of the spherical region operating under Newtonian dynamics,
74 > $r_\textrm{Newton}$ was set to be 4 {\AA} smaller than the original
75 > radius ($R$) of the liquid droplet.}
76   \label{fig:langevinSketch}
77   \end{figure}
78 +
79   The viscosity ($\eta$) can be tuned by comparing the cooling rate that
80   a set of nanoparticles experience with the known cooling rates for
81 < those particles obtained via the laser heating experiments.
81 > similar particles obtained via the laser heating experiments.
82   Essentially, we tune the solvent viscosity until the thermal decay
83   profile matches a heat-transfer model using reasonable values for the
84   interfacial conductance and the thermal conductivity of the solvent.
85  
86   Cooling rates for the experimentally-observed nanoparticles were
87   calculated from the heat transfer equations for a spherical particle
88 < embedded in a ambient medium that allows for diffusive heat
89 < transport. The heat transfer model is a set of two coupled
90 < differential equations in the Laplace domain,
88 > embedded in a ambient medium that allows for diffusive heat transport.
89 > Following Plech {\it et al.},\cite{plech:195423} we use a heat
90 > transfer model that consists of two coupled differential equations
91 > in the Laplace domain,
92   \begin{eqnarray}
93   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\\
94   \left(\frac{\partial}{\partial r} T_{f}(r,s)\right)_{r=R} +
# Line 99 | Line 120 | $K$ of $0.6$  $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1
120   0.295 $(\mathrm{J g^{-1} K^{-1}})$ and $9.826\times 10^6$ $(\mathrm{g
121   m^{-3}})$ respectively. Since most of the laser excitation experiments
122   have been done in aqueous solutions, parameters used for the fluid are
123 < $K$ of $0.6$  $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$ $(\mathrm{g
124 < m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
123 > $K$ of $0.6$ $(\mathrm{Wm^{-1}K^{-1}})$, $\rho$ of $1.0\times10^6$
124 > $(\mathrm{g m^{-3}})$ and $c$ of $4.184$ $(\mathrm{J g^{-1} K^{-1}})$.
125  
126   Values for the interfacial conductance have been determined by a
127   number of groups for similar nanoparticles and range from a low
128 < $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ to $120\times 10^{6}$
129 < $(\mathrm{Wm^{-2}K^{-1}})$.\cite{hartlandPrv2007} Plech {\it et al.} reported a value for the interfacial conductance of $G=105\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ and
130 < $G=130\pm 15$ $(\mathrm{Wm^{-2}K^{-1}})$ for Pt nanoparticles.\cite{plech:195423,PhysRevB.66.224301}
128 > $87.5\times 10^{6} (\mathrm{Wm^{-2}K^{-1}})$ to $130\times 10^{6}
129 > (\mathrm{Wm^{-2}K^{-1}})$.\cite{XXXHartland,Wilson:2002uq} Wilson {\it
130 > et al.}  worked with Au, Pt, and AuPd nanoparticles and obtained an
131 > estimate for the interfacial conductance of $G=130
132 > (\mathrm{Wm^{-2}K^{-1}})$.\cite{Wilson:2002uq} Similarly, Plech {\it
133 > et al.}  reported a value for the interfacial conductance of $G=105\pm
134 > 15 (\mathrm{Wm^{-2}K^{-1}})$ for Au nanoparticles.\cite{plech:195423}
135  
136   We conducted our simulations at both ends of the range of
137   experimentally-determined values for the interfacial conductance.
# Line 117 | Line 142 | experiments on Au nanospheres, we believe that the tru
142   the fastest heat transfer, a value of $117\times 10^{6}$
143   $(\mathrm{Wm^{-2}K^{-1}})$ was used.  Based on calculations we have
144   done using raw data from the Hartland group's thermal half-time
145 < experiments on Au nanospheres, we believe that the true G values are
146 < closer to the faster regime: $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$.
145 > experiments on Au nanospheres,\cite{HuM._jp020581+} the true G values
146 > are probably in the faster regime: $117\times 10^{6}$
147 > $(\mathrm{Wm^{-2}K^{-1}})$.
148  
123
149   The rate of cooling for the nanoparticles in a molecular dynamics
150   simulation can then be tuned by changing the effective solvent
151   viscosity ($\eta$) until the nanoparticle cooling rate matches the
152   cooling rate described by the heat-transfer equations
153 < (\ref{eq:heateqn}). The effective solvent viscosity (in poise) for a G
154 < of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.17, 0.20, and
155 < 0.22 for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
156 < effective solvent viscosity (again in poise) for an interfacial
157 < conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.23,
158 < 0.29, and 0.30 for 20 {\AA}, 30 {\AA} and 40 {\AA} particles.  Cooling
159 < traces for each particle size are presented in
153 > (\ref{eq:heateqn}). The effective solvent viscosity (in Pa s) for a G
154 > of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $4.2 \times
155 > 10^{-6}$, $5.0 \times 10^{-6}$, and
156 > $5.5 \times 10^{-6}$ for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
157 > effective solvent viscosity (again in Pa s) for an interfacial
158 > conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $5.7
159 > \times 10^{-6}$, $7.2 \times 10^{-6}$, and $7.5 \times 10^{-6}$
160 > for 20 {\AA}, 30 {\AA} and 40 {\AA} particles.  These viscosities are
161 > essentially gas-phase values, a fact which is consistent with the
162 > initial temperatures of the particles being well into the
163 > super-critical region for the aqueous environment.  Gas bubble
164 > generation has also been seen experimentally around gold nanoparticles
165 > in water.\cite{kotaidis:184702} Instead of a single value for the
166 > effective viscosity, a time-dependent parameter might be a better
167 > mimic of the cooling vapor layer that surrounds the hot particles.
168 > This may also be a contributing factor to the size-dependence of the
169 > effective viscosities in our simulations.
170 >
171 > Cooling traces for each particle size are presented in
172   Fig. \ref{fig:images_cooling_plot}. It should be noted that the
173   Langevin thermostat produces cooling curves that are consistent with
174   Newtonian (single-exponential) cooling, which cannot match the cooling
175 < profiles from Eq. \ref{eq:laplacetransform} exactly.
175 > profiles from Eq. \ref{eq:laplacetransform} exactly. Fitting the
176 > Langevin cooling profiles to a single-exponential produces
177 > $\tau=25.576$ ps, $\tau=43.786$ ps, and $\tau=56.621$ ps for the 20,
178 > 30 and 40 {\AA} nanoparticles and a G of $87.5\times 10^{6}$
179 > $(\mathrm{Wm^{-2}K^{-1}})$.  For comparison's sake, similar
180 > single-exponential fits with an interfacial conductance of G of
181 > $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ produced a $\tau=13.391$
182 > ps, $\tau=30.426$ ps, $\tau=43.857$ ps for the 20, 30 and 40 {\AA}
183 > nanoparticles.
184  
185   \begin{figure}[htbp]
186   \centering
187 < \includegraphics[width=\linewidth]{images/cooling_plot.pdf}
187 > \includegraphics[width=5in]{images/cooling_plot.pdf}
188   \caption{Thermal cooling curves obtained from the inverse Laplace
189   transform heat model in Eq. \ref{eq:laplacetransform} (solid line) as
190   well as from molecular dynamics simulations (circles).  Effective
191 < solvent viscosities of 0.23-0.30 poise (depending on the radius of the
192 < particle) give the best fit to the experimental cooling curves.  Since
193 < this viscosity is substantially in excess of the viscosity of liquid
194 < water, much of the thermal transfer to the surroundings is probably
195 < due to the capping agent.}
191 > solvent viscosities of 4.2-7.5 $\times 10^{-6}$ Pa s (depending on the
192 > radius of the particle) give the best fit to the experimental cooling
193 > curves.  This viscosity suggests that the nanoparticles in these
194 > experiments are surrounded by a vapor layer (which is a reasonable
195 > assumptions given the initial temperatures of the particles).  }
196   \label{fig:images_cooling_plot}
197   \end{figure}
198  
# Line 159 | Line 204 | nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026
204   Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and
205   Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study
206   a wide range of phenomena in both bulk materials and
207 < nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r} Both
207 > nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq} Both
208   potentials are based on a model of a metal which treats the nuclei and
209   core electrons as pseudo-atoms embedded in the electron density due to
210   the valence electrons on all of the other atoms in the system. The
# Line 191 | Line 236 | liquid alloy.\cite{MURRAY:1984lr} Combination rules fo
236   data for both FCC solid solutions of Ag-Cu and the high-temperature
237   liquid.\cite{sheng:184203} In contrast, the {\sc eam} potential does
238   not reproduce the experimentally observed heat of mixing for the
239 < liquid alloy.\cite{MURRAY:1984lr} Combination rules for the alloy were
240 < taken to be the arithmatic average of the atomic parameters with the
241 < exception of $c_i$ since its values is only dependent on the identity
242 < of the atom where the density is evaluated.  For the {\sc q-sc}
243 < potential, cutoff distances are traditionally taken to be
244 < $2\alpha_{ij}$ and include up to the sixth coordination shell in FCC
245 < metals.
239 > liquid alloy.\cite{MURRAY:1984lr} In this work, we have utilized the
240 > {\sc q-sc} formulation for our potential energies and forces.
241 > Combination rules for the alloy were taken to be the arithmetic
242 > average of the atomic parameters with the exception of $c_i$ since its
243 > values is only dependent on the identity of the atom where the density
244 > is evaluated.  For the {\sc q-sc} potential, cutoff distances are
245 > traditionally taken to be $2\alpha_{ij}$ and include up to the sixth
246 > coordination shell in FCC metals.
247  
248   %\subsection{Sampling single-temperature configurations from a cooling
249   %trajectory}
250  
251 < 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.
251 > To better understand the structural changes occurring in the
252 > nanoparticles throughout the cooling trajectory, configurations were
253 > sampled at regular intervals during the cooling trajectory. These
254 > configurations were then allowed to evolve under NVE dynamics to
255 > sample from the proper distribution in phase space. Figure
256 > \ref{fig:images_cooling_time_traces} illustrates this sampling.
257  
258  
259   \begin{figure}[htbp]
260          \centering
261                  \includegraphics[height=3in]{images/cooling_time_traces.pdf}
262 <        \caption{Illustrative cooling profile for the 40 {\AA} nanoparticle evolving under stochastic boundary conditions corresponding to $G=$$87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$. At temperatures along the cooling trajectory, configurations were sampled and allowed to evolve in the NVE ensemble. These subsequent trajectories were analyzed for structural features associated with bulk glass formation.}
262 >        \caption{Illustrative cooling profile for the 40 {\AA}
263 > nanoparticle evolving under stochastic boundary conditions
264 > corresponding to $G=$$87.5\times 10^{6}$
265 > $(\mathrm{Wm^{-2}K^{-1}})$. At temperatures along the cooling
266 > trajectory, configurations were sampled and allowed to evolve in the
267 > NVE ensemble. These subsequent trajectories were analyzed for
268 > structural features associated with bulk glass formation.}
269          \label{fig:images_cooling_time_traces}
270   \end{figure}
271  
272  
273 + \begin{figure}[htbp]
274 + \centering
275 + \includegraphics[width=5in]{images/cross_section_array.jpg}
276 + \caption{Cutaway views of 30 \AA\ Ag-Cu nanoparticle structures for
277 + random alloy (top) and Cu (core) / Ag (shell) initial conditions
278 + (bottom).  Shown from left to right are the crystalline, liquid
279 + droplet, and final glassy bead configurations.}
280 + \label{fig:cross_sections}
281 + \end{figure}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines