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 3230 by gezelter, Tue Sep 25 19:23:21 2007 UTC vs.
Revision 3256 by chuckv, Thu Oct 11 18:53:11 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
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 composition after
13 < synthesis.\cite{Jiang:2005lr,gonzalo:5163}  Three different sizes of
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.
18 > sampling from a Maxwell-Boltzmann distribution at each temperature. All simulations were conducted using the {\sc OOPSE} molecular dynamics package.\cite{Meineke:2004uq}
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 34 | 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
54 < essentially a free parameter that must be tuned to give experimentally
55 < 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}
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 radial cutoff between the two dynamical regions was set to 4 {\AA}
72 < smaller than the original radius of the liquid droplet.}
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  
# Line 115 | Line 124 | $(\mathrm{Wm^{-2}K^{-1}})$.\cite{hartlandPrv2007} Simi
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} Similarly, Plech
127 > $(\mathrm{Wm^{-2}K^{-1}})$.\cite{Wilson:2002uq} 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
# Line 137 | Line 146 | cooling rate described by the heat-transfer equations
146   simulation can then be tuned by changing the effective solvent
147   viscosity ($\eta$) until the nanoparticle cooling rate matches the
148   cooling rate described by the heat-transfer equations
149 < (\ref{eq:heateqn}). The effective solvent viscosity (in poise) for a G
150 < of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.17, 0.20, and
151 < 0.22 for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
152 < effective solvent viscosity (again in poise) for an interfacial
153 < conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is 0.23,
154 < 0.29, and 0.30 for 20 {\AA}, 30 {\AA} and 40 {\AA} particles.  Cooling
155 < traces for each particle size are presented in
149 > (\ref{eq:heateqn}). The effective solvent viscosity (in Pa s) for a G
150 > of $87.5\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $4.2 \times
151 > 10^{-6}$, $5.0 \times 10^{-6}$, and
152 > $5.5 \times 10^{-6}$ for 20 {\AA}, 30 {\AA}, and 40 {\AA} particles, respectively. The
153 > effective solvent viscosity (again in Pa s) for an interfacial
154 > conductance of $117\times 10^{6}$ $(\mathrm{Wm^{-2}K^{-1}})$ is $5.7
155 > \times 10^{-6}$, $7.2 \times 10^{-6}$, and $7.5 \times 10^{-6}$
156 > for 20 {\AA}, 30 {\AA} and 40 {\AA} particles.  Cooling traces for
157 > each particle size are presented in
158   Fig. \ref{fig:images_cooling_plot}. It should be noted that the
159   Langevin thermostat produces cooling curves that are consistent with
160   Newtonian (single-exponential) cooling, which cannot match the cooling
# Line 159 | Line 170 | nanoparticles.
170  
171   \begin{figure}[htbp]
172   \centering
173 < \includegraphics[width=\linewidth]{images/cooling_plot.pdf}
173 > \includegraphics[width=5in]{images/cooling_plot.pdf}
174   \caption{Thermal cooling curves obtained from the inverse Laplace
175   transform heat model in Eq. \ref{eq:laplacetransform} (solid line) as
176   well as from molecular dynamics simulations (circles).  Effective
177 < solvent viscosities of 0.23-0.30 poise (depending on the radius of the
178 < particle) give the best fit to the experimental cooling curves.
179 < %Since
180 < %this viscosity is substantially in excess of the viscosity of liquid
181 < %water, much of the thermal transfer to the surroundings is probably
171 < %due to the capping agent.
172 < }
177 > solvent viscosities of 4.2-7.5 $\times 10^{-6}$ Pa s (depending on the
178 > radius of the particle) give the best fit to the experimental cooling
179 > curves.  This viscosity suggests that the nanoparticles in these
180 > experiments are surrounded by a vapor layer (which is a reasonable
181 > assumptions given the initial temperatures of the particles).  }
182   \label{fig:images_cooling_plot}
183   \end{figure}
184  
# Line 181 | Line 190 | nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026
190   Embedded Atom Model ({\sc eam})~\cite{PhysRevB.33.7983} and
191   Sutton-Chen ({\sc sc})~\cite{Chen90} potential have been used to study
192   a wide range of phenomena in both bulk materials and
193 < nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r} Both
193 > nanoparticles.\cite{Vardeman-II:2001jn,ShibataT._ja026764r,Sankaranarayanan:2005lr,Chui:2003fk,Wang:2005qy,Medasani:2007uq} Both
194   potentials are based on a model of a metal which treats the nuclei and
195   core electrons as pseudo-atoms embedded in the electron density due to
196   the valence electrons on all of the other atoms in the system. The
# Line 246 | Line 255 | structural features associated with bulk glass formati
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