ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/chuckDissertation/dissertation.lof
Revision: 3483
Committed: Tue Jan 13 14:39:50 2009 UTC (15 years, 5 months ago) by chuckv
File size: 9930 byte(s)
Log Message:
Chuck's dissertation for PhD Jan 2009

File Contents

# Content
1 \addvspace {10\p@ }
2 \contentsline {figure}{\numberline {1.1}{\ignorespaces General form of the pair-pair interaction energy between molecule where $-\epsilon $ is the minimum energy, $\sigma $ is the close interaction distance and $R_m$ is the minimum energy seperation.}}{5}
3 \vskip 0.80em
4 \contentsline {figure}{\numberline {1.2}{\ignorespaces FUNCTIONAL FORMS FOR EAM PARAMETERS.}}{14}
5 \vskip 0.80em
6 \addvspace {10\p@ }
7 \contentsline {figure}{\numberline {2.1}{\ignorespaces Radial distribution functions for Ag$_6$Cu$_4$. The appearance of the split second peak at 500 K indicates the onset of a structural change in the super-cooled liquid at temperatures above $T_{g}$.}}{31}
8 \vskip 0.80em
9 \contentsline {figure}{\numberline {2.2}{\ignorespaces Wendt-Abraham parameter ($R_{WA}$) as a function of temperature. $T^{WA}_{g} \approx $ 540 K for a cooling rate of $1.56 \times 10^{11}$ K/s.}}{33}
10 \vskip 0.80em
11 \contentsline {figure}{\numberline {2.3}{\ignorespaces Density of states calculated from quenched normal modes, $\rho _q(\omega )$, and from the Fourier transform of the velocity autocorrelation function $\rho _o(\omega )$.}}{34}
12 \vskip 0.80em
13 \contentsline {figure}{\numberline {2.4}{\ignorespaces Arrhenius plot of the self-diffusion constants indicating significant deviation from Arrhenius behavior at temperatures below 450 K.}}{36}
14 \vskip 0.80em
15 \contentsline {figure}{\numberline {2.5}{\ignorespaces Self-Diffusion constant for the two models under consideration compared to the values computed via standard techniques.}}{37}
16 \vskip 0.80em
17 \contentsline {figure}{\numberline {2.6}{\ignorespaces {\sc ctrw} hopping distance, $\sigma _0(T)$, as a function of temperature assuming $\gamma =1$ and using the cage correlation hopping times.}}{38}
18 \vskip 0.80em
19 \contentsline {figure}{\numberline {2.7}{\ignorespaces Comparison of exponential stretching coefficients, $\beta $ from the cage-correlation function and $\gamma $ from CTRW theory.}}{39}
20 \vskip 0.80em
21 \contentsline {figure}{\numberline {2.8}{\ignorespaces The characteristic hopping time, $\tau _{hop}$, which characterizes the waiting time distribution. The solid circles represent hopping times predicted from {\sc kww} fits to the cage-correlation function. The open triangles represent the characteristic time calculated via the {\sc ctrw} model for $\delimiter "426830A r^{2}(t) \delimiter "526930B $.}}{40}
22 \vskip 0.80em
23 \addvspace {10\p@ }
24 \addvspace {10\p@ }
25 \contentsline {figure}{\numberline {4.1}{\ignorespaces Comparison of two of the methods for estimating the bulk modulus as a function of temperature for the 35\r A\ particle.}}{51}
26 \vskip 0.80em
27 \contentsline {figure}{\numberline {4.2}{\ignorespaces The temperature dependence of the bulk modulus (upper panel) and heat capacity (lower panel) for nanoparticles of four different radii. Note that the peak in the heat capacity coincides with the {\em start} of the peak in the bulk modulus.}}{52}
28 \vskip 0.80em
29 \contentsline {figure}{\numberline {4.3}{\ignorespaces The dependence of the spike in the heat capacity on the length of the simulation. Longer heating-response calculations result in melting transitions that are sharper and lower in temperature than the short-time transient response simulations. Shorter runs don't allow the particles to melt completely.}}{53}
30 \vskip 0.80em
31 \contentsline {figure}{\numberline {4.4}{\ignorespaces Sample Volume traces for the 20 \r A\ and 35 \r A\ particles at a range of temperatures. Note the relatively rapid ($<$ 10 ps) decoherence due to melting in the 20 \r A\ particle as well as the difference between the 1100 K and 1200 K traces in the 35 \r A\ particle.}}{54}
32 \vskip 0.80em
33 \contentsline {figure}{\numberline {4.5}{\ignorespaces Volume fluctuation autocorrelation functions for the 20 \r A\ and 35 \r A\ particles at a range of temperatures. Successive temperatures have been translated upwards by one unit. Note the beat pattern in the 20 \r A\ particle at 800K.}}{55}
34 \vskip 0.80em
35 \contentsline {figure}{\numberline {4.6}{\ignorespaces The temperature dependence of the period of the breathing mode for the four different nanoparticles studied in this work.}}{55}
36 \vskip 0.80em
37 \addvspace {10\p@ }
38 \contentsline {figure}{\numberline {5.1}{\ignorespaces Methodology used to mimic the experimental cooling conditions of a hot nanoparticle surrounded by a solvent. Atoms in the core of the particle evolved under Newtonian dynamics, while atoms that were in the outer skin of the particle evolved under Langevin dynamics. The radius of the spherical region operating under Newtonian dynamics, $r_\textrm {Newton}$ was set to be 4 {\r A} smaller than the original radius ($R$) of the liquid droplet.}}{66}
39 \vskip 0.80em
40 \contentsline {figure}{\numberline {5.2}{\ignorespaces Thermal cooling curves obtained from the inverse Laplace transform heat model in Eq. (5.5\hbox {}) (solid line) as well as from molecular dynamics simulations (circles). Effective solvent viscosities of 4.2-7.5 $\times 10^{-6}$ Pa s (depending on the radius of the particle) give the best fit to the experimental cooling curves. This viscosity suggests that the nanoparticles in these experiments are surrounded by a vapor layer (which is a reasonable assumptions given the initial temperatures of the particles). }}{69}
41 \vskip 0.80em
42 \contentsline {figure}{\numberline {5.3}{\ignorespaces Illustrative cooling profile for the 40 {\r A} 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.}}{71}
43 \vskip 0.80em
44 \contentsline {figure}{\numberline {5.4}{\ignorespaces Cutaway views of 30 \r A\ Ag-Cu nanoparticle structures for random alloy (top) and Cu (core) / Ag (shell) initial conditions (bottom). Shown from left to right are the crystalline, liquid droplet, and final glassy bead configurations.}}{72}
45 \vskip 0.80em
46 \contentsline {figure}{\numberline {5.5}{\ignorespaces Distributions of the bond orientational order parameter ($\mathaccentV {hat}05E{W}_6$) at different temperatures. The upper, middle, and lower panels are for 20, 30, and 40 \r A\ particles, respectively. The left-hand column used cooling rates commensurate with a low interfacial conductance ($87.5 \times 10^{6}$ $\mathrm {Wm^{-2}K^{-1}}$), while the right-hand column used a more physically reasonable value of $117 \times 10^{6}$ $\mathrm {Wm^{-2}K^{-1}}$. The peak at $\mathaccentV {hat}05E{W}_6 \approx -0.17$ is due to local icosahedral structures. The different curves in each of the panels indicate the distribution of $\mathaccentV {hat}05E{W}_6$ values for samples taken at different times along the cooling trajectory. The initial and final temperatures (in K) are indicated on the plots adjacent to their respective distributions.}}{76}
47 \vskip 0.80em
48 \contentsline {figure}{\numberline {5.6}{\ignorespaces Distributions of the bond orientational order parameter ($Q_6$) at different temperatures. The curves in the six panels in this figure were computed at identical conditions to the same panels in figure 5.5\hbox {}.}}{77}
49 \vskip 0.80em
50 \contentsline {figure}{\numberline {5.7}{\ignorespaces Free energy as a function of the orientational order parameter ($\mathaccentV {hat}05E{W}_6$) for 40 {\r A} bimetallic nanoparticles as they are cooled from 902 K to 310 K. As the particles cool below 528 K, a local minimum in the free energy surface appears near the perfect icosahedral ordering ($\mathaccentV {hat}05E{W}_6 = -0.17$). At all temperatures, liquid-like structures are a global minimum on the free energy surface, but if the cooling rate is fast enough, substructures may become trapped with local icosahedral order, leading to the formation of a glass.}}{79}
51 \vskip 0.80em
52 \contentsline {figure}{\numberline {5.8}{\ignorespaces Temperautre dependence of the fraction of atoms with local icosahedral ordering, $f_\textrm {icos}(T)$ for 20, 30, and 40 \r A\ particles cooled at two different values of the interfacial conductance.}}{80}
53 \vskip 0.80em
54 \contentsline {figure}{\numberline {5.9}{\ignorespaces Distributions of the bond orientational order parameter ($\mathaccentV {hat}05E{W}_6$) for the two different elements present in the nanoparticles. This distribution was taken from the fully-cooled 40 \r A\ nanoparticle. Local icosahedral ordering around copper atoms is much more prevalent than around silver atoms.}}{81}
55 \vskip 0.80em
56 \contentsline {figure}{\numberline {5.10}{\ignorespaces Centers of local icosahedral order ($\mathaccentV {hat}05E{W}_6<0.15$) at 900 K, 471 K and 315 K for the 30 \r A\ nanoparticle cooled with an interfacial conductance $G = 87.5 \times 10^{6}$ $\mathrm {Wm^{-2}K^{-1}}$. Silver atoms (blue) exhibit icosahedral order at the surface of the nanoparticle while copper icosahedral centers (green) are distributed throughout the nanoparticle. The icosahedral centers appear to cluster together and these clusters increase in size with decreasing temperature.}}{82}
57 \vskip 0.80em
58 \contentsline {figure}{\numberline {5.11}{\ignorespaces Appearance of icosahedral clusters around central silver atoms is largely due to the presence of these silver atoms at or near the surface of the nanoparticle. The upper panel shows the fraction of icosahedral atoms ($f_\textrm {icos}(r)$ for each of the two metallic atoms as a function of distance from the center of the nanoparticle ($r$). The lower panel shows the radial density of the two constituent metals (relative to the overall density of the nanoparticle). Icosahedral clustering around copper atoms are more evenly distributed throughout the particle, while icosahedral clustering around silver is largely confined to the silver atoms at the surface.}}{83}
59 \vskip 0.80em