ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
(Generate patch)

Comparing trunk/COonPt/firstTry.tex (file contents):
Revision 3826 by gezelter, Wed Dec 19 21:37:51 2012 UTC vs.
Revision 3881 by jmichalk, Tue Mar 19 18:08:24 2013 UTC

# Line 1 | Line 1
1 < \documentclass[11pt]{article}
1 > \documentclass[journal = jpccck, manuscript = article]{achemso}
2 > \setkeys{acs}{usetitle = true}
3 > \usepackage{achemso}
4 > \usepackage{caption}
5 > \usepackage{float}
6 > \usepackage{geometry}
7 > \usepackage{natbib}
8 > \usepackage{setspace}
9 > \usepackage{xkeyval}
10 > %%%%%%%%%%%%%%%%%%%%%%%
11   \usepackage{amsmath}
12   \usepackage{amssymb}
13   \usepackage{times}
# Line 6 | Line 15
15   \usepackage{setspace}
16   \usepackage{endfloat}
17   \usepackage{caption}
18 < %\usepackage{tabularx}
18 > \usepackage{tabularx}
19 > \usepackage{longtable}
20   \usepackage{graphicx}
21   \usepackage{multirow}
22 < %\usepackage{booktabs}
23 < %\usepackage{bibentry}
24 < %\usepackage{mathrsfs}
25 < \usepackage[square, comma, sort&compress]{natbib}
22 > \usepackage{multicol}
23 > \usepackage{epstopdf}
24 >
25 > \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
26 > % \usepackage[square, comma, sort&compress]{natbib}
27   \usepackage{url}
28   \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
29   \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
30 < 9.0in \textwidth 6.5in \brokenpenalty=10000
30 > 9.0in \textwidth 6.5in \brokenpenalty=1110000
31  
32   % double space list of tables and figures
33   %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
34   \setlength{\abovecaptionskip}{20 pt}
35   \setlength{\belowcaptionskip}{30 pt}
36 + % \bibpunct{}{}{,}{s}{}{;}
37  
38 < \bibpunct{}{}{,}{s}{}{;}
39 < \bibliographystyle{achemso}
38 > %\citestyle{nature}
39 > % \bibliographystyle{achemso}
40  
41 < \begin{document}
41 > \title{Molecular Dynamics simulations of the surface reconstructions
42 >  of Pt(557) and Au(557) under exposure to CO}
43  
44 + \author{Joseph R. Michalka}
45 + \author{Patrick W. McIntyre}
46 + \author{J. Daniel Gezelter}
47 + \email{gezelter@nd.edu}
48 + \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
49 +  Department of Chemistry and Biochemistry\\ University of Notre
50 +  Dame\\ Notre Dame, Indiana 46556}
51  
52 + \keywords{}
53 +
54 + \begin{document}
55 +
56 +
57   %%
58   %Introduction
59   %       Experimental observations
# Line 47 | Line 72
72   %Summary
73   %%
74  
50 %Title
51 \title{Molecular Dynamics simulations of the surface reconstructions
52  of Pt(557) and Au(557) under exposure to CO}
75  
54 \author{Joseph R. Michalka, Patrick W. McIntyre and J. Daniel
55 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
56 Department of Chemistry and Biochemistry,\\
57 University of Notre Dame\\
58 Notre Dame, Indiana 46556}
59
60 %Date
61 \date{Dec 15, 2012}
62
63 %authors
64
65 % make the title
66 \maketitle
67
68 \begin{doublespace}
69
76   \begin{abstract}
77 + We examine surface reconstructions of Pt and Au(557) under
78 + various CO coverages using molecular dynamics in order to
79 + explore possible mechanisms for any observed reconstructions
80 + and their dynamics. The metal-CO interactions were parameterized
81 + as part of this work so that an efficient large-scale treatment of
82 + this system could be undertaken. The large difference in binding
83 + strengths of the metal-CO interactions was found to play a significant
84 + role with regards to step-edge stability and adatom diffusion. A
85 + small correlation between coverage and the diffusion constant
86 + was also determined. The energetics of CO adsorbed to the surface
87 + is sufficient to explain the reconstructions observed on the Pt
88 + systems and the lack  of reconstruction of the Au systems.
89  
90 +
91 + The mechanism and dynamics of surface reconstructions of Pt(557)
92 + and Au(557) exposed to various coverages of carbon monoxide (CO)
93 + were investigated using molecular dynamics simulations. Metal-CO
94 + interactions were parameterized from experimental data and plane-wave
95 + Density Functional Theory (DFT) calculations.  The large difference in
96 + binding strengths of the Pt-CO and Au-CO interactions was found to play
97 + a significant role in step-edge stability and adatom diffusion constants.
98 + The energetics of CO adsorbed to the surface is sufficient to explain the
99 + step-doubling reconstruction observed on Pt(557) and the lack of such
100 + a reconstruction on the Au(557) surface.
101   \end{abstract}
102  
103   \newpage
# Line 100 | Line 129 | This work an effort to understand the mechanism and ti
129   reversible restructuring under exposure to moderate pressures of
130   carbon monoxide.\cite{Tao:2010}
131  
132 < This work an effort to understand the mechanism and timescale for
132 > This work is an investigation into the mechanism and timescale for the Pt(557) \& Au(557)
133   surface restructuring using molecular simulations.  Since the dynamics
134 < of the process is of particular interest, we utilize classical force
134 > of the process are of particular interest, we employ classical force
135   fields that represent a compromise between chemical accuracy and the
136 < computational efficiency necessary to observe the process of interest.
136 > computational efficiency necessary to simulate the process of interest.
137 > Since restructuring typically occurs as a result of specific interactions of the
138 > catalyst with adsorbates, in this work, two metal systems exposed
139 > to carbon monoxide were examined. The Pt(557) surface has already been shown
140 > to undergo a large scale reconstruction under certain conditions.\cite{Tao:2010}
141 > The Au(557) surface, because of a weaker interaction with CO, is less
142 > likely to undergo this kind of reconstruction. However, Peters {\it et al}.\cite{Peters:2000}
143 > and Piccolo {\it et al}.\cite{Piccolo:2004} have both observed CO-induced
144 > reconstruction of a Au(111) surface. Peters {\it et al}. saw a relaxation to the
145 > 22 x $\sqrt{3}$ cell. They argued that only a few Au atoms
146 > become adatoms, limiting the stress of this reconstruction, while
147 > allowing the rest to relax and approach the ideal (111)
148 > configuration. They did not see the usual herringbone pattern on Au(111) being greatly
149 > affected by this relaxation. Piccolo {\it et al}. on the other hand, did see a
150 > disruption of the herringbone pattern as CO was adsorbed to the
151 > surface. Both groups suggested that the preference CO shows for
152 > low-coordinated Au atoms was the primary driving force for the reconstruction.
153  
154 < Since restructuring occurs as a result of specific interactions of the
110 < catalyst with adsorbates, two metal systems exposed to carbon monoxide
111 < were examined in this work. The Pt(557) surface has already been shown
112 < to reconstruct under certain conditions. The Au(557) surface, because
113 < of a weaker interaction with CO, is less likely to undergo this kind
114 < of reconstruction.  MORE HERE ON PT AND AU PREVIOUS WORK.
154 >
155  
156   %Platinum molecular dynamics
157   %gold molecular dynamics
158  
159   \section{Simulation Methods}
160 < The challenge in modeling any solid/gas interface problem is the
160 > The challenge in modeling any solid/gas interface is the
161   development of a sufficiently general yet computationally tractable
162   model of the chemical interactions between the surface atoms and
163   adsorbates.  Since the interfaces involved are quite large (10$^3$ -
164 < 10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
164 > 10$^4$ atoms) and respond slowly to perturbations, {\it ab initio}
165   molecular dynamics
166   (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
167   methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
# Line 133 | Line 173 | Au-Au and Pt-Pt interactions, while modeling the CO us
173   Coulomb potential.  For this work, we have used classical molecular
174   dynamics with potential energy surfaces that are specifically tuned
175   for transition metals.  In particular, we used the EAM potential for
176 < Au-Au and Pt-Pt interactions, while modeling the CO using a rigid
176 > Au-Au and Pt-Pt interactions.\cite{Foiles86} The CO was modeled using a rigid
177   three-site model developed by Straub and Karplus for studying
178   photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
179   Pt-CO cross interactions were parameterized as part of this work.
# Line 146 | Line 186 | parameter sets. The glue model of Ercolessi {\it et al
186   methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
187   but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
188   the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
189 < parameter sets. The glue model of Ercolessi {\it et al.} is among the
190 < fastest of these density functional approaches.\cite{Ercolessi88} In
191 < all of these models, atoms are conceptualized as a positively charged
189 > parameter sets. The glue model of Ercolessi {\it et al}.\cite{Ercolessi88} is among the
190 > fastest of these density functional approaches. In
191 > all of these models, atoms are treated as a positively charged
192   core with a radially-decaying valence electron distribution. To
193   calculate the energy for embedding the core at a particular location,
194   the electron density due to the valence electrons at all of the other
# Line 164 | Line 204 | $\phi_{ij}(r_{ij})$ is an pairwise term that is meant
204   V_i =  F[ \bar{\rho}_i ]  + \sum_{j \neq i} \phi_{ij}(r_{ij})
205   \end{equation*}
206   where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
207 < $\phi_{ij}(r_{ij})$ is an pairwise term that is meant to represent the
208 < overlap of the two positively charged cores.  
207 > $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
208 > repulsive overlap of the two positively charged cores.  
209  
210   % The {\it modified} embedded atom method (MEAM) adds angular terms to
211   % the electron density functions and an angular screening factor to the
# Line 176 | Line 216 | The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen
216   % metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
217   % MEAM presents significant additional computational costs, however.
218  
219 < The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen potentials
219 > The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
220   have all been widely used by the materials simulation community for
221   simulations of bulk and nanoparticle
222   properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
223   melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
224   fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
225   propagation,\cite{BECQUART:1993rg} and alloying
226 < dynamics.\cite{Shibata:2002hh} All of these potentials have their
227 < strengths and weaknesses.  One of the strengths common to all of the
228 < methods is the relatively large library of metals for which these
229 < potentials have been
230 < parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}  
226 > dynamics.\cite{Shibata:2002hh} One of EAM's strengths
227 > is its sensitivity to small changes in structure. This arises
228 > because interactions
229 > up to the third nearest neighbor were taken into account in the parameterization.\cite{Voter95a}
230 > Comparing that to the glue model of Ercolessi {\it et al}.\cite{Ercolessi88}
231 > which is only parameterized up to the nearest-neighbor
232 > interactions, EAM is a suitable choice for systems where
233 > the bulk properties are of secondary importance to low-index
234 > surface structures. Additionally, the similarity of EAM's functional
235 > treatment of the embedding energy to standard density functional
236 > theory (DFT) makes fitting DFT-derived cross potentials with adsorbates somewhat easier.
237 > \cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}  
238  
239 +
240 +
241 +
242   \subsection{Carbon Monoxide model}
243 < Since previous explanations for the surface rearrangements center on
244 < the large linear quadrupole moment of carbon monoxide, the model
245 < chosen for this molecule exhibits this property in an efficient
246 < manner. We used a model first proposed by Karplus and Straub to study
247 < the photodissociation of CO from myoglobin.\cite{Straub} The Straub
248 < and Karplus model is a rigid linear three site model which places a
249 < massless (M) site at the center of mass along the CO bond.  The
250 < geometry and interaction parameters are reproduced in Table 1. The
251 < effective dipole moment, calculated from the assigned charges, is
252 < still small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is
253 < close to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
243 > Previous explanations for the surface rearrangements center on
244 > the large linear quadrupole moment of carbon monoxide.\cite{Tao:2010}  
245 > We used a model first proposed by Karplus and Straub to study
246 > the photodissociation of CO from myoglobin because it reproduces
247 > the quadrupole moment well.\cite{Straub} The Straub and
248 > Karplus model treats CO as a rigid three site molecule with a massless M
249 > site at the molecular center of mass. The geometry and interaction
250 > parameters are reproduced in Table~\ref{tab:CO}. The effective
251 > dipole moment, calculated from the assigned charges, is still
252 > small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
253 > to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
254   mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
255   %CO Table
256   \begin{table}[H]
257    \caption{Positions, Lennard-Jones parameters ($\sigma$ and
258      $\epsilon$), and charges for the CO-CO
259 <    interactions borrowed from Ref. \bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA~, energies are
259 >    interactions in Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
260      in kcal/mol, and charges are in atomic units.}
261   \centering
262   \begin{tabular}{| c | c | ccc |}
263   \hline
264   &  {\it z} & $\sigma$ & $\epsilon$ & q\\
265   \hline
266 < \textbf{C} & -0.6457 &  0.0262  & 3.83   &   -0.75 \\
267 < \textbf{O} &  0.4843 &   0.1591 &   3.12 &   -0.85 \\
266 > \textbf{C} & -0.6457 &  3.83 & 0.0262   &   -0.75 \\
267 > \textbf{O} &  0.4843 &  3.12 &  0.1591  &   -0.85 \\
268   \textbf{M} & 0.0 & -  &  -  &    1.6 \\
269   \hline
270   \end{tabular}
271 + \label{tab:CO}
272   \end{table}
273  
274   \subsection{Cross-Interactions between the metals and carbon monoxide}
275  
276 < Since the adsorption of CO onto a platinum surface has been the focus
276 > Since the adsorption of CO onto a Pt surface has been the focus
277   of much experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
278   and theoretical work
279   \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
280   there is a significant amount of data on adsorption energies for CO on
281 < clean metal surfaces. Parameters reported by Korzeniewski {\it et
282 <  al.}\cite{Pons:1986} were a starting point for our fits, which were
281 > clean metal surfaces. An earlier model by Korzeniewski {\it et
282 >  al.}\cite{Pons:1986} served as a starting point for our fits. The parameters were
283   modified to ensure that the Pt-CO interaction favored the atop binding
284 < position on Pt(111). This resulting binding energies are on the higher
285 < side of the experimentally-reported values. Following Korzeniewski
286 < {\it et al.},\cite{Pons:1986} the Pt-C interaction was fit to a deep
287 < Lennard-Jones interaction to mimic strong, but short-ranged partial
284 > position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}.
285 > The modified parameters yield binding energies that are slightly higher
286 > than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
287 > {\it et al}.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
288 > Lennard-Jones interaction to mimic strong, but short-ranged, partial
289   binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
290 < Pt-O interaction was parameterized to a Morse potential with a large
291 < range parameter ($r_o$).  In most cases, this contributes a weak
290 > Pt-O interaction was modeled with a Morse potential with a large
291 > equilibrium distance, ($r_o$).  These choices ensure that the C is preferred
292 > over O as the surface-binding atom. In most geometries, the Pt-O parameterization contributes a weak
293   repulsion which favors the atop site.  The resulting potential-energy
294   surface suitably recovers the calculated Pt-C separation length
295   (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
# Line 245 | Line 298 | The Au-C and Au-O cross-interactions were fit using Le
298   %where did you actually get the functionals for citation?
299   %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
300   %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
301 < The Au-C and Au-O cross-interactions were fit using Lennard-Jones and
301 > The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
302   Morse potentials, respectively, to reproduce Au-CO binding energies.
303 <
304 < The fits were refined against gas-surface DFT calculations with a
303 > The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
304 > Adsorption energies were obtained from gas-surface DFT calculations with a
305   periodic supercell plane-wave basis approach, as implemented in the
306 < {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores are
306 > {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores were
307   described with the projector augmented-wave (PAW)
308   method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
309   included to an energy cutoff of 20 Ry. Electronic energies are
310   computed with the PBE implementation of the generalized gradient
311   approximation (GGA) for gold, carbon, and oxygen that was constructed
312   by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
313 < Ionic relaxations were performed until the energy difference between
261 < subsequent steps was less than $10^{-8}$ Ry.  In testing the CO-Au
262 < interaction, Au(111) supercells were constructed of four layers of 4
313 > In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
314   Au x 2 Au surface planes and separated from vertical images by six
315 < layers of vacuum space. The surface atoms were all allowed to relax.
316 < Supercell calculations were performed nonspin-polarized with a 4 x 4 x
317 < 4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
318 < zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxed gold slab was
315 > layers of vacuum space. The surface atoms were all allowed to relax
316 > before CO was added to the system. Electronic relaxations were
317 > performed until the energy difference between subsequent steps
318 > was less than $10^{-8}$ Ry.   Nonspin-polarized supercell calculations
319 > were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
320 > zone.\cite{Monkhorst:1976} The relaxed gold slab was
321   then used in numerous single point calculations with CO at various
322   heights (and angles relative to the surface) to allow fitting of the
323   empirical force field.
324  
325   %Hint at future work
326 < The parameters employed in this work are shown in Table 2 and the
327 < binding energies on the 111 surfaces are displayed in Table 3.  To
328 < speed up the computations, charge transfer and polarization are not
329 < being treated in this model, although these effects are likely to
330 < affect binding energies and binding site
278 < preferences.\cite{Deshlahra:2012}
326 > The parameters employed for the metal-CO cross-interactions in this work
327 > are shown in Table~\ref{tab:co_parameters} and the binding energies on the
328 > (111) surfaces are displayed in Table~\ref{tab:co_energies}.  Charge transfer
329 > and polarization are neglected in this model, although these effects could have
330 > an effect on  binding energies and binding site preferences.
331  
332   %Table  of Parameters
333   %Pt Parameter Set 9
334   %Au Parameter Set 35
335   \begin{table}[H]
336 <  \caption{Best fit parameters for metal-CO cross-interactions.   Metal-C
337 <    interactions are modeled with Lennard-Jones potential, while the
338 <    (mostly-repulsive) metal-O interactions were fit to Morse
336 >  \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
337 >    interactions are modeled with Lennard-Jones potentials. While the
338 >    metal-O interactions were fit to Morse
339      potentials.  Distances are given in \AA~and energies in kcal/mol. }
340   \centering
341   \begin{tabular}{| c | cc | c | ccc |}
# Line 295 | Line 347 | preferences.\cite{Deshlahra:2012}
347  
348   \hline
349   \end{tabular}
350 + \label{tab:co_parameters}
351   \end{table}
352  
353   %Table of energies
354   \begin{table}[H]
355 <  \caption{Adsorption energies for CO on M(111) using the potentials
356 <    described in this work.  All values are in eV}
355 >  \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
356 >    described in this work.  All values are in eV.}
357   \centering
358   \begin{tabular}{| c | cc |}
359    \hline
# Line 309 | Line 362 | preferences.\cite{Deshlahra:2012}
362    \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
363    (Ref. \protect\cite{Kelemen:1979}) \\
364   & &  -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
365 <  \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,}  (Ref. \protect\cite{TPD_Gold}) \\
365 >  \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,}  (Ref. \protect\cite{TPDGold}) \\
366    \hline
367   \end{tabular}
368 + \label{tab:co_energies}
369   \end{table}
370  
371   \subsection{Pt(557) and Au(557) metal interfaces}
372 <
373 < Our model systems are composed of 3888 Pt atoms and XXXX Au atoms in a
374 < FCC crystal that have been cut along the 557 plane so that they are
375 < periodic in the {\it x} and {\it y} directions, and have been rotated
376 < to expose two parallel 557 cuts along the positive and negative {\it
377 <  z}-axis.  Simulations of the bare metal interfaces at temperatures
378 < ranging from 300~K to 1200~K were done to observe the relative
372 > Our Pt system is an orthorhombic periodic box of dimensions
373 > 54.482~x~50.046~x~120.88~\AA~while our Au system has
374 > dimensions of 57.4~x~51.9285~x~100~\AA. The metal slabs
375 > are 9 and 8 atoms deep respectively, corresponding to a slab
376 > thickness of $\sim$21~\AA~ for Pt and $\sim$19~\AA~for Au.
377 > The systems are arranged in a FCC crystal that have been cut
378 > along the (557) plane so that they are periodic in the {\it x} and
379 > {\it y} directions, and have been oriented to expose two aligned
380 > (557) cuts along the extended {\it z}-axis.  Simulations of the
381 > bare metal interfaces at temperatures ranging from 300~K to
382 > 1200~K were performed to confirm the relative
383   stability of the surfaces without a CO overlayer.  
384  
385 < The different bulk (and surface) melting temperatures (1337~K for Au
386 < and 2045~K for Pt) suggest that the reconstruction may happen at
387 < different temperatures for the two metals.  To copy experimental
330 < conditions for the CO-exposed surfaces, the bare surfaces were
385 > The different bulk melting temperatures predicted by EAM (1345~$\pm$~10~K for Au\cite{Au:melting}
386 > and $\sim$~2045~K for Pt\cite{Pt:melting}) suggest that any possible reconstruction should happen at
387 > different temperatures for the two metals.  The bare Au and Pt surfaces were
388   initially run in the canonical (NVT) ensemble at 800~K and 1000~K
389 < respectively for 100 ps.  Each surface was exposed to a range of CO
389 > respectively for 100 ps. The two surfaces were relatively stable at these
390 > temperatures when no CO was present, but experienced increased surface
391 > mobility on addition of CO. Each surface was then dosed with different concentrations of CO
392   that was initially placed in the vacuum region.  Upon full adsorption,
393 < these amounts correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
394 < coverage.  Because of the difference in binding energies, the platinum
395 < systems very rarely had CO that was not bound to the surface, while
396 < the gold surfaces often had a significant CO population in the gas
393 > these concentrations correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
394 > coverage. Higher coverages resulted in the formation of a double layer of CO,
395 > which introduces artifacts that are not relevant to (557) reconstruction.
396 > Because of the difference in binding energies, nearly all of the CO was bound to the Pt surface, while
397 > the Au surfaces often had a significant CO population in the gas
398   phase.  These systems were allowed to reach thermal equilibrium (over
399 < 5 ns) before being shifted to the microcanonical (NVE) ensemble for
400 < data collection. All of the systems examined had at least 40 ns in the
401 < data collection stage, although simulation times for some of the
402 < systems exceeded 200ns.  All simulations were run using the open
403 < source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE,OpenMD}
399 > 5~ns) before being run in the microcanonical (NVE) ensemble for
400 > data collection. All of the systems examined had at least 40~ns in the
401 > data collection stage, although simulation times for some Pt of the
402 > systems exceeded 200~ns.  Simulations were carried out using the open
403 > source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
404  
405 < % Just results, leave discussion for discussion section
405 >
406 >
407 >
408 > % RESULTS
409 > %
410   \section{Results}
411 < Tao {\it et al.} showed experimentally that the Pt(557) surface
412 < undergoes two separate reconstructions upon CO
413 < adsorption.\cite{Tao:2010} The first reconstruction involves a
414 < doubling of the step edge height which is accomplished by a doubling
415 < of the plateau length. The second reconstruction led to the formation
416 < of triangular clusters that arrange themselves along the lengthened
417 < plateaus.
411 > \subsection{Structural remodeling}
412 > The bare metal surfaces experienced minor roughening of the
413 > step-edge because of the elevated temperatures, but the (557)
414 > face was stable throughout the simulations. The surface of both
415 > systems, upon dosage of CO, began to undergo extensive remodeling
416 > that was not observed in the bare systems. Reconstructions of
417 > the Au systems were limited to breakup of the step-edges and
418 > some step wandering. The lower coverage Pt systems experienced
419 > similar restructuring but to a greater extent. The 50\% coverage
420 > Pt system was unique among our simulations in that it formed
421 > well-defined and stable double layers through step coalescence,
422 > similar to results reported by Tao {\it et al}.\cite{Tao:2010}
423  
355 The primary observation and results of our simulation is that the
356 presence of CO overlayer on Pt(557) causes the same kind of
357 reconstruction observed experimentally. The 6-atom 111 facets
358 initially become disordered, and after 20-40 ns, a double-layer (with
359 a 2-atom step between terraces) forms.  However, we did not observe
360 the triangular cluster formation that was observed at longer times in
361 the experiments.  Without the CO present on the Pt(557) surface, there
362 was some disorder at the step edges, but no significant restructuring
363 was observed.
424  
425 < In these simulations, the Au(557) surface did not exhibit any
426 < significant restructuring either with or without the presence of a CO
427 < overlayer.
425 > \subsubsection{Step wandering}
426 > The 0\% coverage surfaces for both metals showed minimal
427 > step-wandering at their respective temperatures. As the CO
428 > coverage increased however, the mobility of the surface atoms,
429 > described through adatom diffusion and step-edge wandering,
430 > also increased.  Except for the 50\% Pt system where step
431 > coalescence occurred, the step-edges in the other simulations
432 > preferred to keep nearly the same distance between steps as in
433 > the original (557) lattice, $\sim$13\AA~for Pt and $\sim$14\AA~for Au.
434 > Previous work by Williams {\it et al}.\cite{Williams:1991, Williams:1994}
435 > highlights the repulsion that exists between step-edges even
436 > when no direct interactions are present in the system. This
437 > repulsion is caused by an entropic barrier that arises from
438 > the fact that steps cannot cross over one another. This entropic
439 > repulsion does not completely define the interactions between
440 > steps, however, so it is possible to observe step coalescence
441 > on some surfaces.\cite{Williams:1991} The presence and
442 > concentration of adsorbates, as shown in this work, can
443 > affect step-step interactions, potentially leading to a new
444 > surface structure as the thermodynamic equilibrium.
445  
446 < \subsection{Transport of surface metal atoms}
447 < An ideal metal surface displaying a low energy (111) face is unlikely
448 < to experience much surface diffusion because of the large vacancy
449 < formation energy for atoms at the surface.  This implies that
450 < significant energy must be expended to lift an atom out of the flat
451 < face so it can migrate on the surface.  Rougher surfaces and those
452 < that already contain numerous adatoms, step edges, and kinks, are
453 < expected to have higher surface diffusion rates.  Metal atoms that are
454 < mobile on the surface were observed to leave and then rejoin step
455 < edges or other formations. They may travel together or as isolated
456 < atoms.  The primary challenge of quantifying the overall surface
457 < mobility is in defining ``mobile'' vs. ``static'' atoms.
446 > \subsubsection{Double layers}
447 > Tao {\it et al}.\cite{Tao:2010} have shown experimentally that the Pt(557) surface
448 > undergoes two separate reconstructions upon CO adsorption.
449 > The first involves a doubling of the step height and plateau length.
450 > Similar behavior has been seen on a number of surfaces
451 > at varying conditions, including Ni(977) and Si(111).\cite{Williams:1994,Williams:1991,Pearl}
452 > Of the two systems we examined, the Pt system showed a greater
453 > propensity for reconstruction  
454 > because of the larger surface mobility and the greater extent of step wandering.
455 > The amount of reconstruction was strongly correlated to the amount of CO
456 > adsorbed upon the surface.  This appears to be related to the
457 > effect that adsorbate coverage has on edge breakup and on the
458 > surface diffusion of metal adatoms. Only the 50\% Pt surface underwent the
459 > doubling seen by Tao {\it et al}.\cite{Tao:2010} within the time scales studied here.
460 > Over a longer time scale (150~ns) two more double layers formed
461 > on this surface. Although double layer formation did not occur
462 > in the other Pt systems, they exhibited more step-wandering and
463 > roughening compared to their Au counterparts. The
464 > 50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
465 > various times along the simulation showing the evolution of a double layer step-edge.
466  
467 < A particle was considered mobile once it had traveled more than 2~\AA~
468 < between saved configurations (XX ps). Restricting the transport
469 < calculations to only mobile atoms eliminates all of the bulk metal as
470 < well as any surface atoms that remain fixed for a significant length
471 < of time.  Since diffusion on a surface is strongly affected by local
387 < structures, the diffusion parallel to the step edges was determined
388 < separately from the diffusion perpendicular to these edges.  The
389 < parallel and perpendicular diffusion constants (determined using
390 < linear fits to the mean squared displacement) are shown in figure \ref{fig:diff}.
467 > The second reconstruction observed by
468 > Tao {\it et al}.\cite{Tao:2010} involved the formation of triangular clusters that stretched
469 > across the plateau between two step-edges. Neither metal, within
470 > the 40~ns time scale or the extended simulation time of 150~ns for
471 > the 50\% Pt system, experienced this reconstruction.
472  
473 < %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.
473 > %Evolution of surface
474 > \begin{figure}[H]
475 > \includegraphics[width=\linewidth]{EPS_ProgressionOfDoubleLayerFormation.pdf}
476 > \caption{The Pt(557) / 50\% CO system at a sequence of times after
477 >  initial exposure to the CO: (a) 258~ps, (b) 19~ns, (c) 31.2~ns, and
478 >  (d) 86.1~ns. Disruption of the (557) step-edges occurs quickly.  The
479 >  doubling of the layers appears only after two adjacent step-edges
480 >  touch.  The circled spot in (b) nucleated the growth of the double
481 >  step observed in the later configurations.}
482 >  \label{fig:reconstruct}
483 > \end{figure}
484  
485 + \subsection{Dynamics}
486 + Previous experimental work by Pearl and Sibener\cite{Pearl},
487 + using STM, has been able to capture the coalescence of steps
488 + on Ni(977). The time scale of the image acquisition, $\sim$70~s/image,
489 + provides an upper bound for the time required for the doubling
490 + to occur. By utilizing Molecular Dynamics we are able to probe
491 + the dynamics of these reconstructions at elevated temperatures
492 + and in this section we provide data on the timescales for transport
493 + properties, e.g. diffusion and layer formation time.
494 +
495 +
496 + \subsubsection{Transport of surface metal atoms}
497 + %forcedSystems/stepSeparation
498 + The wandering of a step-edge is a cooperative effect
499 + arising from the individual movements of the atoms making up the steps. An ideal metal surface
500 + displaying a low index facet, (111) or (100), is unlikely to experience
501 + much surface diffusion because of the large energetic barrier that must
502 + be overcome to lift an atom out of the surface. The presence of step-edges and other surface features
503 + on higher-index facets provides a lower energy source for mobile metal atoms.
504 + Single-atom break-away from a step-edge on a clean surface still imposes an
505 + energetic penalty around $\sim$~45 kcal/mol, but this is easier than lifting
506 + the same metal atom vertically out of the surface,  \textgreater~60 kcal/mol.
507 + The penalty lowers significantly when CO is present in sufficient quantities
508 + on the surface. For certain distributions of CO, see Discussion, the penalty can fall to as low as
509 + $\sim$~20 kcal/mol. Once an adatom exists on the surface, the barrier for
510 + diffusion is negligible (\textless~4 kcal/mol for a Pt adatom). These adatoms are then
511 + able to explore the terrace before rejoining either their original step-edge or
512 + becoming a part of a different edge. It is an energetically unfavorable process with a high barrier for an atom
513 + to traverse to a separate terrace although the presence of CO can lower the
514 + energy barrier required to lift or lower an adatom. By tracking the mobility of individual
515 + metal atoms on the Pt and Au surfaces we were able to determine the relative
516 + diffusion constants, as well as how varying coverages of CO affect the diffusion. Close
517 + observation of the mobile metal atoms showed that they were typically in
518 + equilibrium with the step-edges.
519 + At times, their motion was concerted and two or more adatoms would be
520 + observed moving together across the surfaces.
521 +
522 + A particle was considered ``mobile'' once it had traveled more than 2~\AA~
523 + between saved configurations of the system (typically 10-100 ps). A mobile atom
524 + would typically travel much greater distances than this, but the 2~\AA~cutoff
525 + was used to prevent swamping the diffusion data with the in-place vibrational
526 + movement of buried atoms. Diffusion on a surface is strongly affected by
527 + local structures and in this work, the presence of single and double layer
528 + step-edges causes the diffusion parallel to the step-edges to be larger than
529 + the diffusion perpendicular to these edges. Parallel and perpendicular
530 + diffusion constants are shown in Figure \ref{fig:diff}.
531 +
532 + %Diffusion graph
533   \begin{figure}[H]
534 < \includegraphics[scale=0.6]{DiffusionComparison_error.png}
534 > \includegraphics[width=\linewidth]{Portrait_DiffusionComparison_1.pdf}
535   \caption{Diffusion constants for mobile surface atoms along directions
536    parallel ($\mathbf{D}_{\parallel}$) and perpendicular
537 <  ($\mathbf{D}_{\perp}$) to the 557 step edges as a function of CO
538 <  surface coverage.  Diffusion parallel to the step edge is higher
537 >  ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
538 >  surface coverage.  Diffusion parallel to the step-edge is higher
539    than that perpendicular to the edge because of the lower energy
540 <  barrier associated with going from approximately 7 nearest neighbors
541 <  to 5, as compared to the 3 of an adatom. Additionally, the observed
542 <  maximum and subsequent decrease for the Pt system suggests that the
543 <  CO self-interactions are playing a significant role with regards to
544 <  movement of the platinum atoms around and more importantly across
406 <  the surface. }
540 >  barrier associated with traversing along the edge as compared to
541 >  completely breaking away. The two reported diffusion constants for
542 >  the 50\% Pt system arise from different sample sets. The lower values
543 >  correspond to the same 40~ns amount that all of the other systems were
544 >  examined at, while the larger values correspond to a 20~ns period }
545   \label{fig:diff}
546   \end{figure}
547  
548 < %Table of Diffusion Constants
549 < %Add gold?M
550 < % \begin{table}[H]
551 < %   \caption{}
552 < %   \centering
553 < % \begin{tabular}{| c | cc | cc | }
554 < %   \hline
555 < %   &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
556 < %   \hline
557 < %   \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$  \\
558 < %   \hline
559 < %   50\% & 4.32(2) & 1.185 $\pm$ 0.008 & 1.72 $\pm$ 0.02 & 0.455 $\pm$ 0.006 \\
560 < %   33\% & 5.18 $\pm$ 0.03 & 1.999 $\pm$ 0.005 & 1.95 $\pm$ 0.02 & 0.337 $\pm$ 0.004  \\
561 < %   25\% & 5.01 $\pm$ 0.02 & 1.574 $\pm$ 0.004 & 1.26 $\pm$ 0.03 & 0.377 $\pm$ 0.006 \\
424 < %   5\%   & 3.61 $\pm$ 0.02 & 0.355 $\pm$ 0.002 & 1.84 $\pm$ 0.03 & 0.169 $\pm$ 0.004 \\
425 < %   0\%   & 3.27 $\pm$ 0.02 & 0.147 $\pm$ 0.004 & 1.50 $\pm$ 0.02 & 0.194 $\pm$ 0.002  \\
426 < %   \hline
427 < % \end{tabular}
428 < % \end{table}
548 > The weaker Au-CO interaction is evident in the weak CO-coverage
549 > dependance of Au diffusion. This weak interaction leads to lower
550 > observed coverages when compared to dosage amounts. This further
551 > limits the effect the CO can have on surface diffusion. The correlation
552 > between coverage and Pt diffusion rates shows a near linear relationship
553 > at the earliest times in the simulations. Following double layer formation,
554 > however, there is a precipitous drop in adatom diffusion. As the double
555 > layer forms, many atoms that had been tracked for mobility data have
556 > now been buried resulting in a smaller reported diffusion constant. A
557 > secondary effect of higher coverages is CO-CO cross interactions that
558 > lower the effective mobility of the Pt adatoms that are bound to each CO.
559 > This effect would become evident only at higher coverages. A detailed
560 > account of Pt adatom energetics follows in the Discussion.
561 >
562  
563 + \subsubsection{Dynamics of double layer formation}
564 + The increased diffusion on Pt at the higher CO coverages is the primary
565 + contributor to double layer formation. However, this is not a complete
566 + explanation -- the 33\%~Pt system has higher diffusion constants, but
567 + did not show any signs of edge doubling in 40~ns. On the 50\%~Pt
568 + system, one double layer formed within the first 40~ns of simulation time,
569 + while two more were formed as the system was allowed to run for an
570 + additional 110~ns (150~ns total). This suggests that this reconstruction
571 + is a rapid process and that the previously mentioned upper bound is a
572 + very large overestimate.\cite{Williams:1991,Pearl} In this system the first
573 + appearance of a double layer appears at 19~ns into the simulation.
574 + Within 12~ns of this nucleation event, nearly half of the step has formed
575 + the double layer and by 86~ns the complete layer has flattened out.
576 + From the appearance of the first nucleation event to the first observed
577 + double layer, the process took $\sim$20~ns. Another $\sim$40~ns was
578 + necessary for the layer to completely straighten. The other two layers in
579 + this simulation formed over periods of 22~ns and 42~ns respectively.
580 + A possible explanation for this rapid reconstruction is the elevated
581 + temperatures under which our systems were simulated. The process
582 + would almost certainly take longer at lower temperatures. Additionally,
583 + our measured times for completion of the doubling after the appearance
584 + of a nucleation site are likely affected by our periodic boxes. A longer
585 + step-edge will likely take longer to ``zipper''.
586 +
587 +
588   %Discussion
589   \section{Discussion}
590 + We have shown that a classical potential model is able to model the
591 + initial reconstruction of the Pt(557) surface upon CO adsorption as
592 + shown by Tao {\it et al}.\cite{Tao:2010}. More importantly, we were
593 + able to observe features of the dynamic processes necessary for
594 + this reconstruction. Here we discuss the features of the model that
595 + give rise to the observed dynamical properties of the (557) reconstruction.
596  
597 < Mechanism for restructuring
597 > \subsection{Diffusion}
598 > The perpendicular diffusion constant
599 > appears to be the most important indicator of double layer
600 > formation. As highlighted in Figure \ref{fig:reconstruct}, the
601 > formation of the double layer did not begin until a nucleation
602 > site appeared. And as mentioned by Williams {\it et al}.\cite{Williams:1991, Williams:1994},
603 > the inability for edges to cross leads to an effective edge-edge repulsion that
604 > must be overcome to allow step coalescence.
605 > A greater $\textbf{D}_\perp$ implies more step-wandering
606 > and a larger chance for the stochastic meeting of two edges
607 > to create a nucleation point. Parallel diffusion along the step-edge can help ``zipper'' up a nascent double
608 > layer. This helps explain why the time scale for formation after
609 > the appearance of a nucleation site was rapid, while the initial
610 > appearance of the nucleation site was unpredictable.
611  
612 < There are a number of possible mechanisms to explain the role of
613 < adsorbed CO in restructuring the Pt surface. Quadrupolar repulsion
614 < between adjacent CO molecules adsorbed on the surface is one
615 < possibility.  However, the quadrupole-quadrupole interaction is
616 < short-ranged and is attractive for some orientations.  If the CO
617 < molecules are locked in a specific orientation relative to each other,
618 < this explanation gains some weight.  
612 > \subsection{Mechanism for restructuring}
613 > Since the Au surface showed no large scale restructuring in any of
614 > our simulations, our discussion will focus on the 50\% Pt-CO system
615 > which did exhibit doubling. A
616 > number of possible mechanisms exist to explain the role of adsorbed
617 > CO in restructuring the Pt surface. Quadrupolar repulsion between
618 > adjacent CO molecules adsorbed on the surface is one possibility.  
619 > However, the quadrupole-quadrupole interaction is short-ranged and
620 > is attractive for some orientations.  If the CO molecules are ``locked'' in
621 > a specific orientation relative to each other, through atop adsorption for
622 > example, this explanation would gain credence. The calculated energetic repulsion
623 > between two CO molecules located a distance of 2.77~\AA~apart
624 > (nearest-neighbor distance of Pt) and both in a vertical orientation,
625 > is 8.62 kcal/mol. Moving the CO to the second nearest-neighbor distance
626 > of 4.8~\AA~drops the repulsion to nearly 0. Allowing the CO to rotate away
627 > from a purely vertical orientation also lowers the repulsion. When the
628 > carbons are locked at a distance of 2.77~\AA, a minimum of 6.2 kcal/mol is
629 > reached when the angle between the 2 CO is $\sim$24\textsuperscript{o}.
630 > The calculated barrier for surface diffusion of a Pt adatom is only 4 kcal/mol, so
631 > repulsion between adjacent CO molecules bound to Pt could increase the surface
632 > diffusion. However, the residence time of CO on Pt suggests that these
633 > molecules are extremely mobile, with diffusion constants 40 to 2500 times
634 > larger than surface Pt atoms. This mobility suggests that the CO molecules jump
635 > between different Pt atoms throughout the simulation, but will stay bound for
636 > significant periods of time.
637  
638 < Another possible mechanism for the restructuring is in the
639 < destabilization of strong Pt-Pt interactions by CO adsorbed on surface
640 < Pt atoms.  This could have the effect of increasing surface mobility
641 < of these atoms.  
638 > A different interpretation of the above mechanism, taking into account the large
639 > mobility of the CO, looks at how instantaneous and short-lived configurations of
640 > CO on the surface can destabilize Pt-Pt interactions leading to increased step-edge
641 > breakup and diffusion. On the bare Pt(557) surface the barrier to completely detach
642 > an edge atom is $\sim$43~kcal/mol, as is shown in configuration (a) in Figures
643 > \ref{fig:SketchGraphic} \& \ref{fig:SketchEnergies}. For certain configurations, cases
644 > (e), (g), and (h), the barrier can be lowered to $\sim$23~kcal/mole. In these instances,
645 > it becomes quite energetically favorable to roughen the edge by introducing a small
646 > separation of 0.5 to 1.0~\AA. This roughening becomes immediately obvious in
647 > simulations with significant CO populations. The roughening is present to a lesser extent
648 > on lower coverage surfaces and even on the bare surfaces, although in these cases it is likely
649 > due to stochastic vibrational processes that squeeze out step-edge atoms. The mechanism
650 > of step-edge breakup suggested by these energy curves is one of the most difficult
651 > processes, a complete break-away from the step-edge in one unbroken movement.
652 > Easier multistep mechanisms likely exist where an adatom moves laterally on the surface
653 > after being ejected so it ends up alongside the ledge. This provides the atom with 5 nearest
654 > neighbors, which while lower than the 7 if it had stayed a part of the step-edge, is higher
655 > than the 3 it could maintain located on the terrace. In this proposed mechanism, the CO
656 > quadrupolar repulsion is still playing a primary role, but for its importance in roughening
657 > the step-edge, rather than maintaining long-term bonds with a single Pt atom which is not
658 > born out by their mobility data. The requirement for a large density of CO on the surface
659 > for some of the more favorable suggested configurations in Figure \ref{fig:SketchGraphic}
660 > correspond well with the increased mobility seen on higher coverage surfaces.
661  
662 < Comparing the results from simulation to those reported previously by
663 < Tao et al. the similarities in the platinum and CO system are quite
664 < strong. As shown in figure, the simulated platinum system under a CO
665 < atmosphere will restructure slightly by doubling the terrace
666 < heights. The restructuring appears to occur slowly, one to two
667 < platinum atoms at a time. Looking at individual snapshots, these
668 < adatoms tend to either rise on top of the plateau or break away from
669 < the step edge and then diffuse perpendicularly to the step direction
670 < until reaching another step edge. This combination of growth and decay
671 < of the step edges appears to be in somewhat of a state of dynamic
458 < equilibrium. However, once two previously separated edges meet as
459 < shown in figure 1.B, this point tends to act as a focus or growth
460 < point for the rest of the edge to meet up, akin to that of a
461 < zipper. From the handful of cases where a double layer was formed
462 < during the simulation, measuring from the initial appearance of a
463 < growth point, the double layer tends to be fully formed within
464 < $\sim$~35 ns.
662 > %Sketch graphic of different configurations
663 > \begin{figure}[H]
664 > \includegraphics[width=0.8\linewidth, height=0.8\textheight]{COpathsSketch.pdf}
665 > \caption{The dark grey atoms refer to the upper ledge, while the white atoms are
666 > the lower terrace. The blue highlighted atoms had a CO in a vertical atop position
667 > upon them. These are a sampling of the configurations examined to gain a more
668 > complete understanding of the effects CO has on surface diffusion and edge breakup.
669 > Energies associated with each configuration are displayed in Figure \ref{fig:SketchEnergies}.}
670 > \label{fig:SketchGraphic}
671 > \end{figure}
672  
673 < \subsection{Diffusion}
467 < 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?)
468 < \\
469 < \\
470 < %Evolution of surface
673 > %energy graph corresponding to sketch graphic
674   \begin{figure}[H]
675 < \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
676 < \caption{The Pt(557) / 50\% CO system at a sequence of times after
677 <  initial exposure to the CO: (a) 258 ps, (b) 19 ns, (c) 31.2 ns, and
678 <  (d) 86.1 ns. Disruption of the 557 step edges occurs quickly.  The
679 <  doubling of the layers appears only after two adjacent step edges
680 <  touch.  The circled spot in (b) nucleated the growth of the double
681 <  step observed in the later configurations.}
675 > \includegraphics[width=\linewidth]{Portrait_SeparationComparison.pdf}
676 > \caption{The energy curves directly correspond to the labeled model
677 > surface in Figure \ref{fig:SketchGraphic}. All energy curves are relative
678 > to their initial configuration so the energy of a and h do not have the
679 > same zero value. As is seen, certain arrangements of CO can lower
680 > the energetic barrier that must be overcome to create an adatom.
681 > However, it is the highest coverages where these higher-energy
682 > configurations of CO will be more likely. }
683 > \label{fig:SketchEnergies}
684   \end{figure}
685  
686 + While configurations of CO on the surface are able to increase diffusion,
687 + this does not immediately provide an explanation for the formation of double
688 + layers. If adatoms were constrained to their terrace then doubling would be
689 + much less likely to occur. Nucleation sites could still potentially form, but there
690 + would not be enough atoms to finish the doubling. For a non-simulated metal surface, where the
691 + step lengths can be assumed to be infinite relative to atomic sizes, local doubling would be possible, but in
692 + our simulations with our periodic treatment of the system, the system is not large enough to experience this effect.
693 + Thus, there must be a mechanism that explains how adatoms are able to move
694 + amongst terraces. Figure \ref{fig:lambda} shows points along a reaction coordinate
695 + where an adatom along the step-edge with an adsorbed CO ``burrows'' into the
696 + edge displacing an atom onto the higher terrace. This mechanism was chosen
697 + because of similar events that were observed during the simulations. The barrier
698 + heights we obtained are only approximations because we constrained the movement
699 + of the highlighted atoms along a specific concerted path. The calculated $\Delta E$'s
700 + are provide a strong energetic support for this modeled lifting mechanism. When CO is not present and
701 + this reaction coordinate is followed, the $\Delta E > 3$~kcal/mol. The example shown
702 + in the figure, where CO is present in the atop position, has a $\Delta E < -15$~kcal/mol.
703 + While the barrier height is comparable for both cases, there is nearly a 20~kcal/mol
704 + difference in energies and makes the process energetically favorable.
705  
706 < %Peaks!
706 > %lambda progression of Pt -> shoving its way into the step
707   \begin{figure}[H]
708 < \includegraphics[width=\linewidth]{doublePeaks_noCO.png}
709 < \caption{}
708 > \includegraphics[width=\linewidth]{EPS_rxnCoord.pdf}
709 > \caption{ Various points along a reaction coordinate are displayed in the figure.
710 > The mechanism of edge traversal is examined in the presence of CO. The approximate
711 > barrier for the displayed process is 20~kcal/mol. However, the $\Delta E$ of this process
712 > is -15~kcal/mol making it an energetically favorable process.}
713 > \label{fig:lambda}
714   \end{figure}
715 +
716 + The mechanism for doubling on this surface appears to require the cooperation of at least
717 + these two described processes. For complete doubling of a layer to occur there must
718 + be the equivalent removal of a separate terrace. For those atoms to ``disappear'' from
719 + that terrace they must either rise up on the ledge above them or drop to the ledge below
720 + them. The presence of CO helps with the energetics of both of these situations. There must be sufficient
721 + breakage of the step-edge to increase the concentration of adatoms on the surface and
722 + these adatoms must then undergo the burrowing highlighted above or some comparable
723 + mechanism to traverse the step-edge. Over time, these mechanisms working in concert
724 + lead to the formation of a double layer.
725 +
726 + \subsection{CO Removal and double layer stability}
727 + Once a double layer had formed on the 50\%~Pt system it
728 + remained for the rest of the simulation time with minimal
729 + movement. There were configurations that showed small
730 + wells or peaks forming, but typically within a few nanoseconds
731 + the feature would smooth away. Within our simulation time,
732 + the formation of the double layer was irreversible and a double
733 + layer was never observed to split back into two single layer
734 + step-edges while CO was present. To further gauge the effect
735 + CO had on this system, additional simulations were run starting
736 + from a late configuration of the 50\%~Pt system that had formed
737 + double layers. These simulations then had their CO removed.
738 + The double layer breaks rapidly in these simulations, already
739 + showing a well-defined splitting after 100~ps. Configurations of
740 + this system are shown in Figure \ref{fig:breaking}. The coloring
741 + of the top and bottom layers helps to exhibit how much mixing
742 + the edges experience as they split. These systems were only
743 + examined briefly, 10~ns, and within that time despite the initial
744 + rapid splitting, the edges only moved another few \AA~apart.
745 + It is possible with longer simulation times that the
746 + (557) lattice could be recovered as seen by Tao {\it et al}.\cite{Tao:2010}
747 +
748 +
749 +
750 + %breaking of the double layer upon removal of CO
751 + \begin{figure}[H]
752 + \includegraphics[width=\linewidth]{EPS_doubleLayerBreaking.pdf}
753 + \caption{(A)  0~ps, (B) 100~ps, (C) 1~ns, after the removal of CO. The presence of the CO
754 + helped maintain the stability of the double layer and its microfaceting of the double layer
755 + into a (111) configuration. This microfacet immediately reverts to the original (100) step
756 + edge which is a hallmark of the (557) surface. The separation is not a simple sliding apart, rather
757 + there is a mixing of the lower and upper atoms at the edge.}
758 + \label{fig:breaking}
759 + \end{figure}
760 +
761 +
762 +
763 +
764 + %Peaks!
765 + %\begin{figure}[H]
766 + %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
767 + %\caption{At the initial formation of this double layer  ( $\sim$ 37 ns) there is a degree
768 + %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
769 + %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
770 + %\label{fig:peaks}
771 + %\end{figure}
772 +
773 +
774 + %Don't think I need this
775 + %clean surface...
776 + %\begin{figure}[H]
777 + %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
778 + %\caption{}
779 +
780 + %\end{figure}
781 + %\label{fig:clean}
782 +
783 +
784   \section{Conclusion}
785 + The strength of the Pt-CO binding interaction as well as the large
786 + quadrupolar repulsion between CO molecules are sufficient to
787 + explain the observed increase in surface mobility and the resultant
788 + reconstructions at the highest simulated coverage. The weaker
789 + Au-CO interaction results in lower diffusion constants, less step-wandering,
790 + and a lack of the double layer reconstruction. An in-depth examination
791 + of the energetics shows the important role CO plays in increasing
792 + step-breakup and in facilitating edge traversal which are both
793 + necessary for double layer formation.
794  
795  
796 < \section{Acknowledgments}
796 >
797 > %Things I am not ready to remove yet
798 >
799 > %Table of Diffusion Constants
800 > %Add gold?M
801 > % \begin{table}[H]
802 > %   \caption{}
803 > %   \centering
804 > % \begin{tabular}{| c | cc | cc | }
805 > %   \hline
806 > %   &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
807 > %   \hline
808 > %   \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$  \\
809 > %   \hline
810 > %   50\% & 4.32(2) & 1.185(8)  & 1.72(2) & 0.455(6) \\
811 > %   33\% & 5.18(3)  & 1.999(5)  & 1.95(2) & 0.337(4)   \\
812 > %   25\% & 5.01(2)  & 1.574(4)  & 1.26(3) & 0.377(6) \\
813 > %   5\%   & 3.61(2)  & 0.355(2)  & 1.84(3)  & 0.169(4)  \\
814 > %   0\%   & 3.27(2)  & 0.147(4)  & 1.50(2)  & 0.194(2)   \\
815 > %   \hline
816 > % \end{tabular}
817 > % \end{table}
818 >
819 > \begin{acknowledgement}
820   Support for this project was provided by the National Science
821   Foundation under grant CHE-0848243 and by the Center for Sustainable
822   Energy at Notre Dame (cSEND). Computational time was provided by the
823   Center for Research Computing (CRC) at the University of Notre Dame.
824 <
824 > \end{acknowledgement}
825   \newpage
826   \bibliography{firstTryBibliography}
827 < \end{doublespace}
827 > %\end{doublespace}
828 >
829 > \begin{tocentry}
830 > %\includegraphics[height=3.5cm]{timelapse}
831 > \end{tocentry}
832 >
833   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines