ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/nanoglass/analysis.tex
Revision: 3279
Committed: Thu Nov 29 21:22:08 2007 UTC (16 years, 9 months ago) by gezelter
Content type: application/x-tex
File size: 19451 byte(s)
Log Message:
Adding information about Tg

File Contents

# User Rev Content
1 chuckv 3226 %!TEX root = /Users/charles/Desktop/nanoglass/nanoglass.tex
2    
3 chuckv 3213 \section{Analysis}
4    
5 gezelter 3242 Frank first proposed local icosahedral ordering of atoms as an
6     explanation for supercooled atomic (specifically metallic) liquids,
7     and further showed that a 13-atom icosahedral cluster has a 8.4\%
8     higher binding energy the either a face centered cubic ({\sc fcc}) or
9     hexagonal close-packed ({\sc hcp}) crystal structures.\cite{19521106}
10     Icosahedra also have six five-fold symmetry axes that cannot be
11 gezelter 3259 extended indefinitely in three dimensions; long-range translational
12     order is therefore incommensurate with local icosahedral ordering.
13 gezelter 3242 This does not preclude icosahedral clusters from possessing long-range
14     {\it orientational} order. The ``frustrated'' packing of these
15     icosahedral structures into dense clusters has been proposed as a
16     model for glass formation.\cite{19871127} The size of the icosahedral
17     clusters is thought to increase until frustration prevents any further
18     growth.\cite{HOARE:1976fk} Molecular dynamics simulations of a
19     two-component Lennard-Jones glass showed that clusters of face-sharing
20     icosahedra are distributed throughout the
21 gezelter 3230 material.\cite{PhysRevLett.60.2295} Simulations of freezing of single
22     component metalic nanoclusters have shown a tendency for icosohedral
23     structure formation particularly at the surfaces of these
24 chuckv 3254 clusters.\cite{Gafner:2004bg,PhysRevLett.89.275502,Chen:2004ec}
25 gezelter 3230 Experimentally, the splitting (or shoulder) on the second peak of the
26     X-ray structure factor in binary metallic glasses has been attributed
27     to the formation of tetrahedra that share faces of adjoining
28     icosahedra.\cite{Waal:1995lr}
29 chuckv 3226
30 gezelter 3230 Various structural probes have been used to characterize structural
31 gezelter 3242 order in molecular systems including: common neighbor analysis,
32     Voronoi tesselations, and orientational bond-order
33 gezelter 3230 parameters.\cite{HoneycuttJ.Dana_j100303a014,Iwamatsu:2007lr,hsu:4974,nose:1803}
34 gezelter 3242 The method that has been used most extensively for determining local
35     and extended orientational symmetry in condensed phases is the
36     bond-orientational analysis formulated by Steinhart {\it et
37     al.}\cite{Steinhardt:1983mo} In this model, a set of spherical
38 gezelter 3230 harmonics is associated with each of the near neighbors of a central
39     atom. Neighbors (or ``bonds'') are defined as having a distance from
40     the central atom that is within the first peak in the radial
41     distribution function. The spherical harmonic between a central atom
42     $i$ and a neighboring atom $j$ is
43 chuckv 3213 \begin{equation}
44 gezelter 3230 Y_{lm}\left(\vec{r}_{ij}\right)=Y_{lm}\left(\theta(\vec{r}_{ij}),\phi(\vec{r}_{ij})\right)
45     \label{eq:spharm}
46 chuckv 3213 \end{equation}
47 gezelter 3230 where, $\{ Y_{lm}(\theta,\phi)\}$ are spherical harmonics, and
48     $\theta(\vec{r})$ and $\phi(\vec{r})$ are the polar and azimuthal
49     angles made by the bond vector $\vec{r}$ with respect to a reference
50     coordinate system. We chose for simplicity the origin as defined by
51     the coordinates for our nanoparticle. (Only even-$l$ spherical
52     harmonics are considered since permutation of a pair of identical
53     particles should not affect the bond-order parameter.) The local
54     environment surrounding atom $i$ can be defined by
55     the average over all neighbors, $N_b(i)$, surrounding that atom,
56 chuckv 3213 \begin{equation}
57 gezelter 3230 \bar{q}_{lm}(i) = \frac{1}{N_{b}(i)}\sum_{j=1}^{N_b(i)} Y_{lm}(\vec{r}_{ij}).
58     \label{eq:local_avg_bo}
59 chuckv 3213 \end{equation}
60 gezelter 3230 We can further define a global average orientational-bond order over
61     all $\bar{q}_{lm}$ by calculating the average of $\bar{q}_{lm}(i)$
62     over all $N$ particles
63 chuckv 3213 \begin{equation}
64 gezelter 3230 \bar{Q}_{lm} = \frac{\sum^{N}_{i=1}N_{b}(i)\bar{q}_{lm}(i)}{\sum^{N}_{i=1}N_{b}(i)}
65     \label{eq:sys_avg_bo}
66 chuckv 3213 \end{equation}
67 chuckv 3261 The $\bar{Q}_{lm}$ contained in Eq. (\ref{eq:sys_avg_bo}) is not
68 gezelter 3233 necessarily invariant under rotations of the arbitrary reference
69     coordinate system. Second- and third-order rotationally invariant
70     combinations, $Q_l$ and $W_l$, can be taken by summing over $m$ values
71     of $\bar{Q}_{lm}$,
72 chuckv 3222 \begin{equation}
73 gezelter 3230 Q_{l} = \left( \frac{4\pi}{2 l+1} \sum^{l}_{m=-l} \left| \bar{Q}_{lm} \right|^2\right)^{1/2}
74     \label{eq:sec_ord_inv}
75 chuckv 3222 \end{equation}
76     and
77     \begin{equation}
78     \hat{W}_{l} = \frac{W_{l}}{\left( \sum^{l}_{m=-l} \left| \bar{Q}_{lm} \right|^2\right)^{3/2}}
79     \label{eq:third_ord_inv}
80     \end{equation}
81 gezelter 3230 where
82 chuckv 3222 \begin{equation}
83 gezelter 3230 W_{l} = \sum_{\substack{m_1,m_2,m_3 \\m_1 + m_2 + m_3 = 0}} \left( \begin{array}{ccc} l & l & l \\ m_1 & m_2 & m_3 \end{array}\right) \\ \times \bar{Q}_{lm_1}\bar{Q}_{lm_2}\bar{Q}_{lm_3}.
84 chuckv 3222 \label{eq:third_inv}
85     \end{equation}
86 chuckv 3261 The factor in parentheses in Eq. (\ref{eq:third_inv}) is the Wigner-3$j$
87 gezelter 3233 symbol, and the sum is over all valid ($|m| \leq l$) values of $m_1$,
88     $m_2$, and $m_3$ which sum to zero.
89 chuckv 3226
90     \begin{table}
91 gezelter 3230 \caption{Values of bond orientational order parameters for
92     simple structures cooresponding to $l=4$ and $l=6$ spherical harmonic
93     functions.\cite{wolde:9932} $Q_4$ and $\hat{W}_4$ have values for {\it
94     individual} icosahedral clusters, but these values are not invariant
95     under rotations of the reference coordinate systems. Similar behavior
96     is observed in the bond-orientational order parameters for individual
97     liquid-like structures.}
98 chuckv 3226 \begin{center}
99     \begin{tabular}{ccccc}
100     \hline
101     \hline
102     & $Q_4$ & $Q_6$ & $\hat{W}_4$ & $\hat{W}_6$\\
103    
104     fcc & 0.191 & 0.575 & -0.159 & -0.013\\
105    
106     hcp & 0.097 & 0.485 & 0.134 & -0.012\\
107    
108     bcc & 0.036 & 0.511 & 0.159 & 0.013\\
109    
110     sc & 0.764 & 0.354 & 0.159 & 0.013\\
111    
112 gezelter 3230 Icosahedral & - & 0.663 & - & -0.170\\
113 chuckv 3226
114 gezelter 3230 (liquid) & - & - & - & -\\
115 chuckv 3226 \hline
116     \end{tabular}
117     \end{center}
118     \label{table:bopval}
119     \end{table}
120    
121 gezelter 3230 For ideal face-centered cubic ({\sc fcc}), body-centered cubic ({\sc
122     bcc}) and simple cubic ({\sc sc}) as well as hexagonally close-packed
123     ({\sc hcp}) structures, these rotationally invariant bond order
124     parameters have fixed values independent of the choice of coordinate
125     reference frames. For ideal icosahedral structures, the $l=6$
126     invariants, $Q_6$ and $\hat{W}_6$ are also independent of the
127     coordinate system. $Q_4$ and $\hat{W}_4$ will have non-vanishing
128     values for {\it individual} icosahedral clusters, but these values are
129     not invariant under rotations of the reference coordinate systems.
130     Similar behavior is observed in the bond-orientational order
131 gezelter 3259 parameters for individual liquid-like structures. Additionally, both
132     $Q_6$ and $\hat{W}_6$ are thought to have extreme values for the
133     icosahedral clusters.\cite{Steinhardt:1983mo} This makes the $l=6$
134     bond-orientational order parameters particularly useful in identifying
135     the extent of local icosahedral ordering in condensed phases. For
136     example, a local structure which exhibits $\hat{W}_6$ values near
137     -0.17 is easily identified as an icosahedral cluster and cannot be
138     mistaken for distorted cubic or liquid-like structures.
139 gezelter 3230
140 gezelter 3233 One may use these bond orientational order parameters as an averaged
141     property to obtain the extent of icosahedral ordering in a supercooled
142     liquid or cluster. It is also possible to accumulate information
143     about the {\it distributions} of local bond orientational order
144     parameters, $p(\hat{W}_6)$ and $p(Q_6)$, which provide information
145     about individual atomic sites that are central to local icosahedral
146     structures.
147 gezelter 3230
148 gezelter 3233 The distributions of atomic $Q_6$ and $\hat{W}_6$ values are plotted
149 chuckv 3261 as a function of temperature for our nanoparticles in Fig.
150 gezelter 3233 \ref{fig:q6} and \ref{fig:w6}. At high temperatures, the
151     distributions are unstructured and are broadly distributed across the
152     entire range of values. As the particles are cooled, however, there
153     is a dramatic increase in the fraction of atomic sites which have
154     local icosahedral ordering around them. (This corresponds to the
155 chuckv 3261 sharp peak appearing in Fig. \ref{fig:w6} at $\hat{W}_6=-0.17$ and
156     to the broad shoulder appearing in Fig. \ref{fig:q6} at $Q_6 =
157 gezelter 3233 0.663$.)
158    
159     \begin{figure}[htbp]
160     \centering
161 gezelter 3242 \includegraphics[width=5in]{images/w6_stacked_plot.pdf}
162 gezelter 3233 \caption{Distributions of the bond orientational order parameter
163     ($\hat{W}_6$) at different temperatures. The upper, middle, and lower
164     panels are for 20, 30, and 40 \AA\ particles, respectively. The
165     left-hand column used cooling rates commensurate with a low
166     interfacial conductance ($87.5 \times 10^{6}$
167     $\mathrm{Wm^{-2}K^{-1}}$), while the right-hand column used a more
168     physically reasonable value of $117 \times 10^{6}$
169     $\mathrm{Wm^{-2}K^{-1}}$. The peak at $\hat{W}_6 \approx -0.17$ is
170 gezelter 3239 due to local icosahedral structures. The different curves in each of
171     the panels indicate the distribution of $\hat{W}_6$ values for samples
172     taken at different times along the cooling trajectory. The initial
173     and final temperatures (in K) are indicated on the plots adjacent to
174     their respective distributions.}
175 gezelter 3233 \label{fig:w6}
176     \end{figure}
177    
178     \begin{figure}[htbp]
179     \centering
180 gezelter 3242 \includegraphics[width=5in]{images/q6_stacked_plot.pdf}
181 gezelter 3233 \caption{Distributions of the bond orientational order parameter
182     ($Q_6$) at different temperatures. The curves in the six panels in
183     this figure were computed at identical conditions to the same panels in
184     figure \ref{fig:w6}.}
185     \label{fig:q6}
186     \end{figure}
187    
188 gezelter 3252 The probability distributions of local order can be used to generate
189     free energy surfaces using the local orientational ordering as a
190     reaction coordinate. By making the simple statistical equivalence
191     between the free energy and the probabilities of occupying certain
192     states,
193     \begin{equation}
194     g(\hat{W}_6) = - k_B T \ln p(\hat{W}_6),
195     \end{equation}
196     we can obtain a sequence of free energy surfaces (as a function of
197     temperature) for the local ordering around central atoms within our
198     particles. Free energy surfaces for the 40 \AA\ particle at a range
199 chuckv 3261 of temperatures are shown in Fig. \ref{fig:freeEnergy}. Note that
200 gezelter 3252 at all temperatures, the liquid-like structures are global minima on
201     the free energy surface, while the local icosahedra appear as local
202     minima once the temperature has fallen below 528 K. As the
203     temperature falls, it is possible for substructures to become trapped
204     in the local icosahedral well, and if the cooling is rapid enough,
205     this trapping leads to vitrification. A similar analysis of the free
206     energy surface for orientational order in bulk glass formers can be
207     found in the work of van~Duijneveldt and
208 gezelter 3279 Frenkel.\cite{duijneveldt:4655}
209 gezelter 3252
210 gezelter 3279
211 gezelter 3252 \begin{figure}[htbp]
212     \centering
213     \includegraphics[width=5in]{images/freeEnergyVsW6.pdf}
214     \caption{Free energy as a function of the orientational order
215 chuckv 3261 parameter ($\hat{W}_6$) for 40 {\AA} bimetallic nanoparticles as they
216 gezelter 3252 are cooled from 902 K to 310 K. As the particles cool below 528 K, a
217     local minimum in the free energy surface appears near the perfect
218     icosahedral ordering ($\hat{W}_6 = -0.17$). At all temperatures,
219     liquid-like structures are a global minimum on the free energy
220     surface, but if the cooling rate is fast enough, substructures
221     may become trapped with local icosahedral order, leading to the
222     formation of a glass.}
223     \label{fig:freeEnergy}
224     \end{figure}
225    
226 gezelter 3242 We have also calculated the fraction of atomic centers which have
227     strong local icosahedral order:
228 gezelter 3233 \begin{equation}
229     f_\textrm{icos} = \int_{-\infty}^{w_i} p(\hat{W}_6) d \hat{W}_6
230     \label{eq:ficos}
231     \end{equation}
232     where $w_i$ is a cutoff value in $\hat{W}_6$ for atomic centers that
233     are displaying icosahedral environments. We have chosen a (somewhat
234     arbitrary) value of $w_i= -0.15$ for the purposes of this work. A
235     plot of $f_\textrm{icos}(T)$ as a function of temperature of the
236 chuckv 3261 particles is given in Fig. \ref{fig:ficos}. As the particles cool,
237 gezelter 3233 the fraction of local icosahedral ordering rises smoothly to a plateau
238 gezelter 3259 value. The smaller particles (particularly the ones that were cooled
239     in a higher viscosity solvent) show a slightly larger tendency towards
240 gezelter 3242 icosahedral ordering.
241 gezelter 3233
242     \begin{figure}[htbp]
243     \centering
244 gezelter 3242 \includegraphics[width=5in]{images/fraction_icos.pdf}
245 gezelter 3233 \caption{Temperautre dependence of the fraction of atoms with local
246     icosahedral ordering, $f_\textrm{icos}(T)$ for 20, 30, and 40 \AA\
247     particles cooled at two different values of the interfacial
248     conductance.}
249 gezelter 3242 \label{fig:ficos}
250 gezelter 3233 \end{figure}
251    
252     Since we have atomic-level resolution of the local bond-orientational
253     ordering information, we can also look at the local ordering as a
254     function of the identities of the central atoms. In figure
255     \ref{fig:AgVsCu} we display the distributions of $\hat{W}_6$ values
256     for both the silver and copper atoms, and we note a strong
257 gezelter 3242 predilection for the copper atoms to be central to icosahedra. This
258     is probably due to local packing competition of the larger silver
259     atoms around the copper, which would tend to favor icosahedral
260     structures over the more densely packed cubic structures.
261 gezelter 3233
262     \begin{figure}[htbp]
263     \centering
264 gezelter 3242 \includegraphics[width=5in]{images/w6_stacked_bytype_plot.pdf}
265 gezelter 3233 \caption{Distributions of the bond orientational order parameter
266     ($\hat{W}_6$) for the two different elements present in the
267     nanoparticles. This distribution was taken from the fully-cooled 40
268     \AA\ nanoparticle. Local icosahedral ordering around copper atoms is
269     much more prevalent than around silver atoms.}
270 gezelter 3242 \label{fig:AgVsCu}
271 gezelter 3233 \end{figure}
272 gezelter 3242
273 gezelter 3252 The locations of these icosahedral centers are not uniformly
274 chuckv 3261 distrubted throughout the particles. In Fig. \ref{fig:icoscluster}
275 gezelter 3252 we show snapshots of the centers of the local icosahedra (i.e. any
276     atom which exhibits a local bond orientational order parameter
277     $\hat{W}_6 < -0.15$). At high temperatures, the icosahedral centers
278     are transitory, existing only for a few fs before being reabsorbed
279     into the liquid droplet. As the particle cools, these centers become
280     fixed at certain locations, and additional icosahedra develop
281     throughout the particle, clustering around the sites where the
282     structures originated. There is a strong preference for icosahedral
283     ordering near the surface of the particles. Identification of these
284     structures by the type of atom shows that the silver-centered
285     icosahedra are evident only at the surface of the particles.
286    
287     \begin{figure}[htbp]
288     \centering
289     \begin{tabular}{c c c}
290     \includegraphics[width=2.1in]{images/Cu_Ag_random_30A_liq_icosonly.pdf}
291     \includegraphics[width=2.1in]{images/Cu_Ag_random_30A__0007_icosonly.pdf}
292     \includegraphics[width=2.1in]{images/Cu_Ag_random_30A_glass_icosonly.pdf}
293     \end{tabular}
294     \caption{Centers of local icosahedral order ($\hat{W}_6<0.15$) at 900
295     K, 471 K and 315 K for the 30 \AA\ nanoparticle cooled with an
296     interfacial conductance $G = 87.5 \times 10^{6}$
297     $\mathrm{Wm^{-2}K^{-1}}$. Silver atoms (blue) exhibit icosahedral
298     order at the surface of the nanoparticle while copper icosahedral
299     centers (green) are distributed throughout the nanoparticle. The
300     icosahedral centers appear to cluster together and these clusters
301     increase in size with decreasing temperature.}
302     \label{fig:icoscluster}
303     \end{figure}
304    
305 gezelter 3259 In contrast with the silver ordering behavior, the copper atoms which
306     have local icosahedral ordering are distributed more evenly throughout
307 chuckv 3261 the nanoparticles. Fig. \ref{fig:Surface} shows this tendency as a
308 gezelter 3259 function of distance from the center of the nanoparticle. Silver,
309     since it has a lower surface free energy than copper, tends to coat
310     the skins of the mixed particles.\cite{Zhu:1997lr} This is true even
311     for bimetallic particles that have been prepared in the Ag (core) / Cu
312     (shell) configuration. Upon forming a liquid droplet, approximately 1
313     monolayer of Ag atoms will rise to the surface of the particles. This
314 chuckv 3261 can be seen visually in Fig. \ref{fig:cross_sections} as well as in
315     the density plots in the bottom panel of Fig. \ref{fig:Surface}.
316 gezelter 3259 This observation is consistent with previous experimental and
317     theoretical studies on bimetallic alloys composed of noble
318 gezelter 3247 metals.\cite{MainardiD.S._la0014306,HuangS.-P._jp0204206,Ramirez-Caballero:2006lr}
319     Bond order parameters for surface atoms are averaged only over the
320     neighboring atoms, so packing constraints that may prevent icosahedral
321     ordering around silver in the bulk are removed near the surface. It
322     would certainly be interesting to see if the relative tendency of
323     silver and copper to form local icosahedral structures in a bulk glass
324     differs from our observations on nanoparticles.
325    
326     \begin{figure}[htbp]
327     \centering
328     \includegraphics[width=5in]{images/dens_fracr_stacked_plot.pdf}
329     \caption{Appearance of icosahedral clusters around central silver atoms
330     is largely due to the presence of these silver atoms at or near the
331     surface of the nanoparticle. The upper panel shows the fraction of
332     icosahedral atoms ($f_\textrm{icos}(r)$ for each of the two metallic
333     atoms as a function of distance from the center of the nanoparticle
334     ($r$). The lower panel shows the radial density of the two
335     constituent metals (relative to the overall density of the
336     nanoparticle). Icosahedral clustering around copper atoms are more
337     evenly distributed throughout the particle, while icosahedral
338     clustering around silver is largely confined to the silver atoms at
339     the surface.}
340     \label{fig:Surface}
341     \end{figure}
342 gezelter 3279
343     The methods used by Sheng, He, and Ma to estimate the glass transition
344     temperature, $T_g$, in bulk Ag-Cu alloys involve finding
345     discontinuities in the slope of the average atomic volume, $\langle V
346     \rangle / N$, or enthalpy when plotted against the temperature of the
347     alloy. They obtained a bulk glass transition temperature, $T_g$ = 510
348     K for a quenching rate of $2.5 \times 10^{13}$ K/s.
349    
350     For simulations of nanoparticles, there is no periodic box, and
351     therefore, no easy way to compute the volume exactly. Instead, we
352     estimate the volume of our nanoparticles using Barber {\it et al.}'s
353     very fast quickhull algorithm to obtain the convex hull for the
354     collection of 3-d coordinates of all of atoms at each point in
355     time.~\cite{Barber96,qhull} The convex hull is the smallest convex
356     polyhedron which includes all of the atoms, so the volume of this
357     polyhedron is an excellent estimate of the volume of the nanoparticle.
358     This method of estimating the volume will be problematic if the
359     nanoparticle breaks into pieces (i.e. if the bounding surface becomes
360     concave), but for the relatively short trajectories used in this
361     study, it provides an excellent measure of particle volume as a
362     function of time (and temperature).
363    
364     Using the discontinuity in the slope of the average atomic volume
365     vs. temperature, we arrive at an estimate of $T_g$ that is
366     approximately 490 K. We note that this temperature is somewhat below
367     the onset of icosahedral ordering exhibited in the bond orientational
368     order parameters. It appears that icosahedral ordering sets in while
369     the system is still somewhat fluid, and is locked in place once the
370     temperature falls below $T_g$. We did not observe any dependence of
371     our estimates for $T_g$ on either the nanoparticle size or the value
372     of the interfacial conductance. However, the cooling rates and size
373     ranges we utilized are all sampled from a relatively narrow range, and
374     it is possible that much larger particles would have substantially
375     different values for $T_g$. Our estimates for the glass transition
376     temperatures for all three particle sizes and both interfacial
377     conductance values are shown in table \ref{table:Tg}.
378    
379     \begin{table}
380     \caption{Estimates of the glass transition temperatures $T_g$ for
381     three different sizes of bimetallic Ag$_6$Cu$_4$ nanoparticles cooled
382     under two different values of the interfacial conductance, $G$.}
383     \begin{center}
384     \begin{tabular}{ccccc}
385     \hline
386     \hline
387     Radius (\AA\ ) & Interfacial conductance & Effective cooling rate
388     (K/s) & & $T_g$ (K) \\
389     20 & 87.5 & & 477 \\
390     20 & 117 & & 502 \\
391     30 & 87.5 & & 512 \\
392     30 & 117 & & 493 \\
393     40 & 87.5 & & 476 \\
394     40 & 117 & & 487 \\
395     \hline
396     \end{tabular}
397     \end{center}
398     \label{table:Tg}
399     \end{table}