ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3820
Committed: Mon Dec 17 18:45:57 2012 UTC (11 years, 6 months ago) by gezelter
Content type: application/x-tex
Original Path: trunk/COonPt/firstTry.tex
File size: 23808 byte(s)
Log Message:
formatting

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{times}
5 \usepackage{mathptm}
6 \usepackage{setspace}
7 \usepackage{float}
8 \usepackage{caption}
9
10 %\usepackage{tabularx}
11 \usepackage{graphicx}
12 \usepackage{multirow}
13 %\usepackage{booktabs}
14 %\usepackage{bibentry}
15 %\usepackage{mathrsfs}
16 \usepackage[square, comma, sort&compress]{natbib}
17 \usepackage{url}
18 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
19 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
20 9.0in \textwidth 6.5in \brokenpenalty=10000
21
22 % double space list of tables and figures
23 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
24 \setlength{\abovecaptionskip}{20 pt}
25 \setlength{\belowcaptionskip}{30 pt}
26
27 \bibpunct{}{}{,}{s}{}{;}
28 \bibliographystyle{achemso}
29
30 \begin{document}
31
32
33 %%
34 %Introduction
35 % Experimental observations
36 % Previous work on Pt, CO, etc.
37 %
38 %Simulation Methodology
39 % FF (fits and parameters)
40 % MD (setup, equilibration, collection)
41 %
42 % Analysis of trajectories!!!
43 %Discussion
44 % CO preferences for specific locales
45 % CO-CO interactions
46 % Differences between Au & Pt
47 % Causes of 2_layer reordering in Pt
48 %Summary
49 %%
50
51 %Title
52 \title{Molecular Dynamics simulations of the surface reconstructions
53 of Pt(557) and Au(557) under exposure to CO}
54
55 \author{Joseph R. Michalka, Patrick W. McIntyre and J. Daniel
56 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
57 Department of Chemistry and Biochemistry,\\
58 University of Notre Dame\\
59 Notre Dame, Indiana 46556}
60
61 %Date
62 \date{Dec 15, 2012}
63
64 %authors
65
66 % make the title
67 \maketitle
68
69 \begin{doublespace}
70
71 \begin{abstract}
72
73 \end{abstract}
74
75 \newpage
76
77
78 \section{Introduction}
79 % Importance: catalytically active metals are important
80 % Sub: Knowledge of how their surface structure affects their ability to catalytically facilitate certain reactions is growing, but is more reactionary than predictive
81 % Sub: Designing catalysis is the future, and will play an important role in numerous processes (ones that are currently seen to be impractical, or at least inefficient)
82 % Theory can explore temperatures and pressures which are difficult to work with in experiments
83 % Sub: Also, easier to observe what is going on and provide reasons and explanations
84 %
85
86 Industrial catalysts usually consist of small particles exposing
87 different atomic terminations that exhibit a high concentration of
88 step, kink sites, and vacancies at the edges of the facets. These
89 sites are thought to be the locations of catalytic
90 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
91 significant evidence to demonstrate that solid surfaces are often
92 structurally, compositionally, and chemically {\it modified} by
93 reactants under operating conditions.\cite{Tao2008,Tao:2010,Tao2011}
94 The coupling between surface oxidation state and catalytic activity
95 for CO oxidation on Pt, for instance, is widely
96 documented.\cite{Ertl08,Hendriksen:2002} Despite the well-documented
97 role of these effects on reactivity, the ability to capture or predict
98 them in atomistic models is currently somewhat limited. While these
99 effects are perhaps unsurprising on the highly disperse, multi-faceted
100 nanoscale particles that characterize industrial catalysts, they are
101 manifest even on ordered, well-defined surfaces. The Pt(557) surface,
102 for example, exhibits substantial and reversible restructuring under
103 exposure to moderate pressures of carbon monoxide.\cite{Tao:2010}
104
105 This work is part of an ongoing effort to understand the causes,
106 mechanisms and timescales for surface restructuring using molecular
107 simulation methods. Since the dynamics of the process is of
108 particular interest, we utilize classical molecular dynamic methods
109 with force fields that represent a compromise between chemical
110 accuracy and the computational efficiency necessary to observe the
111 process of interest.
112
113 Since restructuring occurs as a result of specific interactions of the catalyst
114 with adsorbates, two metals systems exposed to the same adsorbate, CO,
115 were examined in this work. The Pt(557) surface has already been shown to
116 reconstruct under certain conditions. The Au(557) surface, because of gold's
117 weaker interaction with CO, is less likely to undergo such a large reconstruction.
118 %Platinum molecular dynamics
119 %gold molecular dynamics
120
121
122
123 \section{Simulation Methods}
124 The challenge in modeling any solid/gas interface problem is the
125 development of a sufficiently general yet computationally tractable
126 model of the chemical interactions between the surface atoms and
127 adsorbates. Since the interfaces involved are quite large (10$^3$ -
128 10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
129 molecular dynamics
130 (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
131 methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
132 mechanical potential energy surfaces remain out of reach.
133 Additionally, the ``bonds'' between metal atoms at a surface are
134 typically not well represented in terms of classical pairwise
135 interactions in the same way that bonds in a molecular material are,
136 nor are they captured by simple non-directional interactions like the
137 Coulomb potential. For this work, we have been using classical
138 molecular dynamics with potential energy surfaces that are
139 specifically tuned for transition metals. In particular, we use the
140 EAM potential for Au-Au and Pt-Pt interactions, while modeling the CO
141 using a model developed by Straub and Karplus for studying
142 photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and Pt-CO
143 cross interactions were parameterized as part of this work.
144
145 \subsection{Metal-metal interactions}
146 Many of the potentials used for classical simulation of transition
147 metals are based on a non-pairwise additive functional of the local
148 electron density. The embedded atom method (EAM) is perhaps the best
149 known of these
150 methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
151 but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
152 the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
153 parameter sets. The glue model of Ercolessi {\it et al.} is among the
154 fastest of these density functional approaches.\cite{Ercolessi88} In
155 all of these models, atoms are conceptualized as a positively charged
156 core with a radially-decaying valence electron distribution. To
157 calculate the energy for embedding the core at a particular location,
158 the electron density due to the valence electrons at all of the other
159 atomic sites is computed at atom $i$'s location,
160 \begin{equation*}
161 \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
162 \end{equation*}
163 Here, $\rho_j(r_{ij})$ is the function that describes the distance
164 dependence of the valence electron distribution of atom $j$. The
165 contribution to the potential that comes from placing atom $i$ at that
166 location is then
167 \begin{equation*}
168 V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
169 \end{equation*}
170 where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
171 $\phi_{ij}(r_{ij})$ is an pairwise term that is meant to represent the
172 overlap of the two positively charged cores.
173
174 The {\it modified} embedded atom method (MEAM) adds angular terms to
175 the electron density functions and an angular screening factor to the
176 pairwise interaction between two
177 atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
178 MEAM has become widely used to simulate systems in which angular
179 interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
180 metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
181 MEAM presents significant additional computational costs, however.
182
183 The EAM, Finnis-Sinclair, MEAM, and the Quantum Sutton-Chen potentials
184 have all been widely used by the materials simulation community for
185 simulations of bulk and nanoparticle
186 properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
187 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
188 fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
189 propagation,\cite{BECQUART:1993rg} and alloying
190 dynamics.\cite{Shibata:2002hh} All of these potentials have their
191 strengths and weaknesses. One of the strengths common to all of the
192 methods is the relatively large library of metals for which these
193 potentials have been
194 parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
195
196 \subsection{CO}
197 Since one explanation for the strong surface CO repulsion on metals is
198 the large linear quadrupole moment of carbon monoxide, the model
199 chosen for this molecule exhibits this property in an efficient
200 manner. We used a model first proposed by Karplus and Straub to study
201 the photodissociation of CO from myoglobin.\cite{Straub} The Straub and
202 Karplus model is a rigid three site model which places a massless M
203 site at the center of mass along the CO bond. The geometry used along
204 with the interaction parameters are reproduced in Table 1. The effective
205 dipole moment, calculated from the assigned charges, is still
206 small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
207 to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
208 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
209 %CO Table
210 \begin{table}[H]
211 \caption{Positions, $\sigma$, $\epsilon$ and charges for CO geometry
212 and self-interactions\cite{Straub}. Distances are in \AA~, energies are
213 in kcal/mol, and charges are in $e$.}
214 \centering
215 \begin{tabular}{| c | c | ccc |}
216 \hline
217 \multicolumn{5}{|c|}{\textbf{Self-Interactions}}\\
218 \hline
219 & {\it z} & $\sigma$ & $\epsilon$ & q\\
220 \hline
221 \textbf{C} & -0.6457 & 0.0262 & 3.83 & -0.75 \\
222 \textbf{O} & 0.4843 & 0.1591 & 3.12 & -0.85 \\
223 \textbf{M} & 0.0 & - & - & 1.6 \\
224 \hline
225 \end{tabular}
226 \end{table}
227
228 \subsection{Cross-Interactions}
229
230 One hurdle that must be overcome in classical molecular simulations
231 is the proper parameterization of the potential interactions present
232 in the system. Since the adsorption of CO onto a platinum surface has been
233 the focus of many experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
234 and theoretical studies \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
235 there is a large amount of data in the literature to fit too. We started with parameters
236 reported by Korzeniewski et al. \cite{Pons:1986} and then
237 modified them to ensure that the Pt-CO interaction favored
238 an atop binding position for the CO upon the Pt surface. This
239 constraint led to the binding energies being on the higher side
240 of reported values. Following the method laid out by Korzeniewski,
241 the Pt-C interaction was fit to a strong Lennard-Jones 12-6
242 interaction to mimic binding, while the Pt-O interaction
243 was parameterized to a Morse potential with a large $r_o$
244 to contribute a weak repulsion. The resultant potential-energy
245 surface suitably recovers the calculated Pt-C bond length ( 1.6\AA)\cite{Beurden:2002ys} and affinity
246 for the atop binding position.\cite{Deshlahra:2012, Hopster:1978}
247
248 %where did you actually get the functionals for citation?
249 %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
250 %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
251 The Au-C and Au-O cross-interactions were fit using Lennard-Jones and
252 Morse potentials, respectively, to reproduce Au-CO binding energies.
253
254 The fits were refined against gas-surface calculations using DFT with
255 a periodic supercell plane-wave basis approach, as implemented in the
256 {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores are
257 described with the projector augmented-wave (PAW)
258 method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
259 included to an energy cutoff of 20 Ry. Electronic energies are
260 computed with the PBE implementation of the generalized gradient
261 approximation (GGA) for gold, carbon, and oxygen that was constructed
262 by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
263 Ionic relaxations were performed until the energy difference between
264 subsequent steps was less than 0.0001 eV. In testing the CO-Au
265 interaction, Au(111) supercells were constructed of four layers of 4
266 Au x 2 Au surface planes and separated from vertical images by six
267 layers of vacuum space. The surface atoms were all allowed to relax.
268 Supercell calculations were performed nonspin-polarized, and energies
269 were converged to within 0.03 meV per Au atom with a 4 x 4 x 4
270 Monkhorst-Pack\cite{Monkhorst:1976,PhysRevB.13.5188} {\bf k}-point
271 sampling of the first Brillouin zone. The relaxed gold slab was then
272 used in numerous single point calculations with CO at various heights
273 (and angles relative to the surface) to allow fitting of the empirical
274 force field.
275
276 %Hint at future work
277 The fit parameter sets employed in this work are shown in Table 2 and their
278 reproduction of the binding energies are displayed in Table 3. Currently,
279 charge transfer is not being treated in this system, however, that is a goal
280 for future work as the effect has been seen to affect binding energies and
281 binding site preferences. \cite{Deshlahra:2012}
282
283
284
285
286 \subsection{Construction and Equilibration of 557 Metal interfaces}
287
288 Our model systems are composed of approximately 4000 metal atoms
289 cut along the 557 plane so that they are periodic in the {\it x} and {\it y}
290 directions exposing the 557 plane in the {\it z} direction. Runs at various
291 temperatures ranging from 300~K to 1200~K were started with the intent
292 of viewing relative stability of the surface when CO was not present in the
293 system. Owing to the different melting points (1337~K for Au and 2045~K for Pt),
294 the bare crystal systems were initially run in the Canonical ensemble at
295 800~K and 1000~K respectively for 100 ps. Various amounts of CO were
296 placed in the vacuum region, which upon full adsorption to the surface
297 corresponded to 5\%, 25\%, 33\%, and 50\% coverages. Because of the
298 high temperature and the difference in binding energies, the platinum systems
299 very rarely had CO that was not adsorbed to the surface whereas the gold systems
300 often had a substantial minority of CO away from the surface.
301 These systems were again allowed to reach thermal equilibrium before being run in the
302 microcanonical ensemble. All of the systems examined in this work were
303 run for at least 40 ns. A subset that were undergoing interesting effects
304 have been allowed to continue running with one system approaching 200 ns.
305 All simulations were run using the open source molecular dynamics package, OpenMD. \cite{Ewald, OOPSE}
306
307
308
309
310
311
312 %\subsection{System}
313 %Once equilibration was reached, the systems were exposed to various sub-monolayer coverage of CO: $0, \frac{1}{10}, \frac{1}{4}, \frac{1}{3},\frac{1}{2}$. The CO was started many \AA~above the surface with random velocity and rotational velocity vectors sampling from a Gaussian distribution centered on the temperature of the equilibrated metal block. Full adsorption occurred over the period of approximately 10 ps for Pt, while the binding energy between Au and CO is smaller and led to an incomplete adsorption. The metal-metal interactions were treated using the Embedded Atom Method while the Pt-CO and Au-CO interactions were fit to experimental data and quantum calculations. The raised temperature helped shorten the length of the simulations by allowing the activation barrier of reconstruction to be more easily overcome. A few runs at lower temperatures showed the very beginnings of reconstructions, but their simulation lengths limited their usefulness.
314
315
316 %Table of Parameters
317 %Pt Parameter Set 9
318 %Au Parameter Set 35
319 \begin{table}[H]
320 \caption{Best fit parameters for metal-adsorbate interactions. Distances are in \AA~and energies are in kcal/mol}
321 \centering
322 \begin{tabular}{| c | cc | c | ccc |}
323 \hline
324 \multicolumn{7}{| c |}{\textbf{Cross-Interactions} }\\
325 \hline
326 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ \\
327 \hline
328 \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
329 \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
330
331 \hline
332 \end{tabular}
333 \end{table}
334
335 %Table of energies
336 \begin{table}[H]
337 \caption{Adsorption energies in eV}
338 \centering
339 \begin{tabular}{| c | cc |}
340 \hline
341 & Calc. & Exp. \\
342 \hline
343 \textbf{Pt-CO} & -1.9 & -1.4~\cite{Kelemen:1979}-- -1.9~\cite{Yeo} \\
344 \textbf{Au-CO} & -0.39 & -0.40~\cite{TPD_Gold} \\
345 \hline
346 \end{tabular}
347 \end{table}
348
349
350
351
352
353
354 % Just results, leave discussion for discussion section
355 \section{Results}
356 \subsection{Diffusion}
357 An ideal metal surface displaying a low-energy facet, a (111) face for
358 instance, is unlikely to experience much surface diffusion because of
359 the large energy barrier associated with atoms 'lifting' from the top
360 layer to then be able to explore the surface. Rougher surfaces, those
361 that already contain numerous adatoms, step edges, and kinks, should
362 have concomitantly higher surface diffusion rates. Tao et al. showed
363 that the platinum 557 surface undergoes two separate reconstructions
364 upon CO adsorption. \cite{Tao:2010} The first reconstruction involves a
365 doubling of the step edge height which is accomplished by a doubling
366 of the plateau length. The second reconstruction led to the formation of
367 triangular motifs stretching across the lengthened plateaus.
368
369 As shown in Figure 2, over a period of approximately 100 ns, the surface
370 has reconstructed from a 557 surface by doubling the step height and
371 step length. Focusing on only the platinum, or gold, atoms that were
372 deemed mobile on the surface, an analysis of the surface diffusion was
373 performed. A particle was considered mobile once it had traveled more
374 than 2~\AA between snapshots. This immediately eliminates all of the
375 bulk metal and greatly limits the number of surface atoms examined.
376 Since diffusion on a surface is strongly affected by overcoming energy
377 barriers, the diffusion parallel to the step edge axis was determined
378 separately from the diffusion perpendicular to the step edge. The results
379 at various coverages on both platinum and gold are shown in Table 4.
380
381 %While an ideal metallic surface is unlikely to experience much surface diffusion, high-index surfaces have large numbers of low-coordinated atoms which have a much easier time overcoming the energetic barriers limiting diffusion, leading to easier surface reconstructions. Surface movement was divided between the parallel ($\parallel$) and perpendicular ($\perp$) directions relative to the step edge. We were then able to calculate diffusion constants as a function of CO coverage. As can be seen in Table 4, the presence and amount of CO directly affects the diffusion constants of surface platinum atoms. The presence of two 50\% coverage systems is to show how the diffusion process is affected by time. The majority of the systems were run for approximately 50 ns while the half monolayer system has been running continuously. The lowered diffusion constant at longer run times will be examined in-depth in the discussion section.
382
383 \begin{figure}[H]
384 \includegraphics[scale=0.6]{DiffusionComparison_error.png}
385 \caption{Diffusion parallel to the step edge will always be higher than that perpendicular to the edge because of the lower energy barrier associated with going from approximately 7 nearest neighbors to 5, as compared to the 3 of an adatom. Additionally, the observed maximum and subsequent decrease for the Pt system suggests that the CO self-interactions are playing a significant role with regards to movement of the platinum atoms around and more importantly across the surface. }
386 \end{figure}
387
388 %Table of Diffusion Constants
389 %Add gold?M
390 \begin{table}[H]
391 \caption{Platinum and gold diffusion constants parallel and perpendicular to the 557 step edge. As the coverage increases, the diffusion constants parallel and perpendicular to the step edge both initially increase and then decrease slightly. Units are \AA\textsuperscript{2}/ns}
392 \centering
393 \begin{tabular}{| c | cc | cc | c |}
394 \hline
395 \textbf{System Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & \textbf{Time (ns)}\\
396 \hline
397 &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} & \\
398 \hline
399 50\% & 4.32 $\pm$ 0.02 & 1.185 $\pm$ 0.008 & 1.72 $\pm$ 0.02 & 0.455 $\pm$ 0.006 & 40 \\
400 33\% & 5.18 $\pm$ 0.03 & 1.999 $\pm$ 0.005 & 1.95 $\pm$ 0.02 & 0.337 $\pm$ 0.004 & 40 \\
401 25\% & 5.01 $\pm$ 0.02 & 1.574 $\pm$ 0.004 & 1.26 $\pm$ 0.03 & 0.377 $\pm$ 0.006 & 40 \\
402 5\% & 3.61 $\pm$ 0.02 & 0.355 $\pm$ 0.002 & 1.84 $\pm$ 0.03 & 0.169 $\pm$ 0.004 & 40 \\
403 0\% & 3.27 $\pm$ 0.02 & 0.147 $\pm$ 0.004 & 1.50 $\pm$ 0.02 & 0.194 $\pm$ 0.002 & 40 \\
404 \hline
405 \end{tabular}
406 \end{table}
407
408
409
410 %Discussion
411 \section{Discussion}
412 Comparing the results from simulation to those reported previously by Tao et al. the similarities in the platinum and CO system are quite strong. As shown in figure 1, the simulated platinum system under a CO atmosphere will restructure slightly by doubling the terrace heights. The restructuring appears to occur slowly, one to two platinum atoms at a time. Looking at individual snapshots, these adatoms tend to either rise on top of the plateau or break away from the step edge and then diffuse perpendicularly to the step direction until reaching another step edge. This combination of growth and decay of the step edges appears to be in somewhat of a state of dynamic equilibrium. However, once two previously separated edges meet as shown in figure 1.B, this point tends to act as a focus or growth point for the rest of the edge to meet up, akin to that of a zipper. From the handful of cases where a double layer was formed during the simulation, measuring from the initial appearance of a growth point, the double layer tends to be fully formed within $\sim$~35 ns.
413
414 \subsection{Diffusion}
415 As shown in the results section, the diffusion parallel to the step edge tends to be much faster than that perpendicular to the step edge. Additionally, the coverage of CO appears to play a slight role in relative rates of diffusion, as shown in Table 4. Thus, the bottleneck of the double layer formation appears to be the initial formation of this growth point, which seems to be somewhat of a stochastic event. Once it appears, parallel diffusion, along the now slightly angled step edge, will allow for a faster formation of the double layer than if the entire process were dependent on only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the more likely a growth point is to be formed. One driving force behind this reconstruction appears to be the lowering of surface energy that occurs by doubling the terrace widths. (I'm not really proving this... I have the surface flatness to show it, but surface energy?)
416 \\
417 \\
418 %Evolution of surface
419 \begin{figure}[H]
420 \includegraphics[scale=0.5]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
421 \caption{Four snapshots of the $\frac{1}{2}$ monolayer system at various times a) 258 ps b) 19 ns c) 31.2 ns and d) 86.1 ns. Slight disruption of the surface occurs fairly quickly. However, the doubling of the layers seems to be very dependent on the initial linking of two separate step edges. The focal point in b, appears to be a growth spot for the rest of the double layer.}
422 \end{figure}
423
424
425
426
427 %Peaks!
428 \begin{figure}[H]
429 \includegraphics[scale=0.25]{doublePeaks_noCO.png}
430 \caption{}
431 \end{figure}
432 \section{Conclusion}
433
434
435 \section{Acknowledgments}
436 Support for this project was provided by the National Science
437 Foundation under grant CHE-0848243 and by the Center for Sustainable
438 Energy at Notre Dame (cSEND). Computational time was provided by the
439 Center for Research Computing (CRC) at the University of Notre Dame.
440
441 \newpage
442 \bibliography{firstTryBibliography}
443 \end{doublespace}
444 \end{document}