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 3870 by jmichalk, Fri Mar 8 22:06:22 2013 UTC vs.
trunk/COonPt/COonPtAu.tex (file contents), Revision 3894 by jmichalk, Wed Jun 12 19:30:39 2013 UTC

# Line 1 | Line 1
1 < \documentclass[11pt]{article}
2 < \usepackage{amsmath}
3 < \usepackage{amssymb}
4 < \usepackage{times}
5 < \usepackage{mathptm}
6 < \usepackage{setspace}
7 < \usepackage{endfloat}
8 < \usepackage{caption}
9 < %\usepackage{tabularx}
10 < \usepackage{graphicx}
1 > \documentclass[journal = jpccck, manuscript = article]{achemso}
2 > \setkeys{acs}{usetitle = true}
3 > \usepackage{achemso}
4 > \usepackage{natbib}
5   \usepackage{multirow}
6 < %\usepackage{booktabs}
7 < %\usepackage{bibentry}
8 < %\usepackage{mathrsfs}
9 < \usepackage[square, comma, sort&compress]{natbib}
6 > \usepackage{wrapfig}
7 > \usepackage{fixltx2e}
8 > %\mciteErrorOnUnknownfalse
9 >
10 > \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
11   \usepackage{url}
17 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
18 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
19 9.0in \textwidth 6.5in \brokenpenalty=10000
12  
13 < % double space list of tables and figures
14 < %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
23 < \setlength{\abovecaptionskip}{20 pt}
24 < \setlength{\belowcaptionskip}{30 pt}
13 > \title{Molecular Dynamics simulations of the surface reconstructions
14 >  of Pt(557) and Au(557) under exposure to CO}
15  
16 < \bibpunct{}{}{,}{s}{}{;}
17 < \bibliographystyle{achemso}
16 > \author{Joseph R. Michalka}
17 > \author{Patrick W. McIntyre}
18 > \author{J. Daniel Gezelter}
19 > \email{gezelter@nd.edu}
20 > \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
21 >  Department of Chemistry and Biochemistry\\ University of Notre
22 >  Dame\\ Notre Dame, Indiana 46556}
23  
24 + \keywords{}
25 +
26   \begin{document}
27  
28 <
28 >
29   %%
30   %Introduction
31   %       Experimental observations
# Line 47 | Line 44
44   %Summary
45   %%
46  
50 %Title
51 \title{Molecular Dynamics simulations of the surface reconstructions
52  of Pt(557) and Au(557) under exposure to CO}
47  
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{Mar 5, 2013}
62
63 %authors
64
65 % make the title
66 \maketitle
67
68 \begin{doublespace}
69
48   \begin{abstract}
49 < We examine surface reconstructions of Pt and Au(557) under
50 < various CO coverages using molecular dynamics in order to
51 < explore possible mechanisms for any observed reconstructions
52 < and their dynamics. The metal-CO interactions were parameterized
53 < as part of this work so that an efficient large-scale treatment of
54 < this system could be undertaken. The large difference in binding
55 < strengths of the metal-CO interactions was found to play a significant
56 < role with regards to step-edge stability and adatom diffusion. A
57 < small correlation between coverage and the diffusion constant
58 < was also determined. The energetics of CO adsorbed to the surface
59 < is sufficient to explain the reconstructions observed on the Pt
60 < systems and the lack  of reconstruction of the Au systems.
61 <
49 >  The mechanism and dynamics of surface reconstructions of Pt(557) and
50 >  Au(557) exposed to various coverages of carbon monoxide (CO) were
51 >  investigated using molecular dynamics simulations.  Metal-CO
52 >  interactions were parameterized from experimental data and
53 >  plane-wave Density Functional Theory (DFT) calculations.  The large
54 >  difference in binding strengths of the Pt-CO and Au-CO interactions
55 >  was found to play a significant role in step-edge stability and
56 >  adatom diffusion constants.  Various mechanisms for CO-mediated step
57 >  wandering and step doubling were investigated on the Pt(557)
58 >  surface.  We find that the energetics of CO adsorbed to the surface
59 >  can explain the step-doubling reconstruction observed on Pt(557) and
60 >  the lack of such a reconstruction on the Au(557) surface.  However,
61 >  more complicated reconstructions into triangular clusters that have
62 >  been seen in recent experiments were not observed in these
63 >  simulations.
64   \end{abstract}
65  
66   \newpage
# Line 112 | Line 92 | This work is an attempt to understand the mechanism an
92   reversible restructuring under exposure to moderate pressures of
93   carbon monoxide.\cite{Tao:2010}
94  
95 < This work is an attempt to understand the mechanism and timescale for
96 < surface restructuring by using molecular simulations.  Since the dynamics
97 < of the process are of particular interest, we employ classical force
98 < fields that represent a compromise between chemical accuracy and the
99 < computational efficiency necessary to simulate the process of interest.
100 < Since restructuring typically occurs as a result of specific interactions of the
101 < catalyst with adsorbates, in this work, two metal systems exposed
102 < to carbon monoxide were examined. The Pt(557) surface has already been shown
103 < to undergo a large scale reconstruction under certain conditions.\cite{Tao:2010}
104 < The Au(557) surface, because of a weaker interaction with CO, is seen as less
105 < likely to undergo this kind of reconstruction. However, Peters et al.\cite{Peters:2000}
106 < and Piccolo et al.\cite{Piccolo:2004} have both observed CO induced
107 < reconstruction of a Au(111) surface. Peters et al. saw a relaxing of the
108 < 22 x $\sqrt{3}$ cell. They argued that a very small number of Au atoms
109 < would become adatoms, limiting the stress of this reconstruction while
110 < allowing the rest of the row to relax and approach the ideal (111)
111 < configuration. They did not see the ``herringbone'' pattern being greatly
112 < affected by this relaxation. Piccolo et al. on the other hand, did see a
113 < disruption of the ``herringbone'' pattern as CO was adsorbed to the
114 < surface. Both groups suggested that the preference CO shows for
115 < low-coordinated Au particles was the primary driving force for these reconstructions.
116 <
117 <
95 > This work is an investigation into the mechanism and timescale for the
96 > Pt(557) \& Au(557) surface restructuring using molecular simulation.
97 > Since the dynamics of the process are of particular interest, we
98 > employ classical force fields that represent a compromise between
99 > chemical accuracy and the computational efficiency necessary to
100 > simulate the process of interest.  Since restructuring typically
101 > occurs as a result of specific interactions of the catalyst with
102 > adsorbates, in this work, two metal systems exposed to carbon monoxide
103 > were examined. The Pt(557) surface has already been shown to undergo a
104 > large scale reconstruction under certain conditions.\cite{Tao:2010}
105 > The Au(557) surface, because of weaker interactions with CO, is less
106 > likely to undergo this kind of reconstruction. However, Peters {\it et
107 >  al}.\cite{Peters:2000} and Piccolo {\it et al}.\cite{Piccolo:2004}
108 > have both observed CO-induced modification of reconstructions to the
109 > Au(111) surface. Peters {\it et al}. observed the Au(111)-($22 \times
110 > \sqrt{3}$) ``herringbone'' reconstruction relaxing slightly under CO
111 > adsorption. They argued that only a few Au atoms become adatoms,
112 > limiting the stress of this reconstruction, while allowing the rest to
113 > relax and approach the ideal (111) configuration.  Piccolo {\it et
114 >  al}. on the other hand, saw a more significant disruption of the
115 > Au(111)-($22 \times \sqrt{3}$) herringbone pattern as CO adsorbed on
116 > the surface. Both groups suggested that the preference CO shows for
117 > low-coordinated Au atoms was the primary driving force for the
118 > relaxation.  Although the Au(111) reconstruction was not the primary
119 > goal of our work, the classical models we have fit may be of future
120 > use in simulating this reconstruction.
121  
122   %Platinum molecular dynamics
123   %gold molecular dynamics
124  
125   \section{Simulation Methods}
126 < The challenge in modeling any solid/gas interface is the
127 < development of a sufficiently general yet computationally tractable
128 < model of the chemical interactions between the surface atoms and
129 < adsorbates.  Since the interfaces involved are quite large (10$^3$ -
130 < 10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
126 > The challenge in modeling any solid/gas interface is the development
127 > of a sufficiently general yet computationally tractable model of the
128 > chemical interactions between the surface atoms and adsorbates.  Since
129 > the interfaces involved are quite large (10$^3$ - 10$^4$ atoms), have
130 > many electrons, and respond slowly to perturbations, {\it ab initio}
131   molecular dynamics
132   (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
133   methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
# Line 156 | Line 139 | Au-Au and Pt-Pt interactions\cite{EAM}. The CO was mod
139   Coulomb potential.  For this work, we have used classical molecular
140   dynamics with potential energy surfaces that are specifically tuned
141   for transition metals.  In particular, we used the EAM potential for
142 < Au-Au and Pt-Pt interactions\cite{EAM}. The CO was modeled using a rigid
143 < three-site model developed by Straub and Karplus for studying
142 > Au-Au and Pt-Pt interactions.\cite{Foiles86} The CO was modeled using
143 > a rigid three-site model developed by Straub and Karplus for studying
144   photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
145   Pt-CO cross interactions were parameterized as part of this work.
146    
# Line 169 | Line 152 | parameter sets. The glue model of Ercolessi et al. is
152   methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
153   but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
154   the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
155 < parameter sets. The glue model of Ercolessi et al. is among the
156 < fastest of these density functional approaches.\cite{Ercolessi88} In
157 < all of these models, atoms are conceptualized as a positively charged
158 < core with a radially-decaying valence electron distribution. To
159 < calculate the energy for embedding the core at a particular location,
160 < the electron density due to the valence electrons at all of the other
161 < atomic sites is computed at atom $i$'s location,
155 > parameter sets. The glue model of Ercolessi {\it et
156 >  al}.\cite{Ercolessi88} is among the fastest of these density
157 > functional approaches. In all of these models, atoms are treated as a
158 > positively charged core with a radially-decaying valence electron
159 > distribution. To calculate the energy for embedding the core at a
160 > particular location, the electron density due to the valence electrons
161 > at all of the other atomic sites is computed at atom $i$'s location,
162   \begin{equation*}
163   \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
164   \end{equation*}
# Line 202 | Line 185 | properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007
185   The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
186   have all been widely used by the materials simulation community for
187   simulations of bulk and nanoparticle
188 < properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
188 > properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq,mishin99:_inter}
189   melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
190 < fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
191 < propagation,\cite{BECQUART:1993rg} and alloying
192 < dynamics.\cite{Shibata:2002hh} One of EAM's strengths
193 < is its sensitivity to small changes in structure. This arises
194 < from the original parameterization, where the interactions
195 < up to the third nearest-neighbor were taken into account.\cite{Voter95a}
196 < Comparing that to the glue model of Ercolessi et al.\cite{Ercolessi88}
197 < which only parameterized up to the nearest-neighbor
198 < interactions, EAM is a suitable choice for systems where
199 < the bulk properties are of secondary importance to low-index
200 < surface structures. Additionally, the similarity of EAMs functional
201 < treatment of the embedding energy to standard density functional
202 < theory (DFT) approaches gives EAM, and conclusions derived, a firm theoretical footing.
203 < \cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}  
190 > fracture,\cite{Shastry:1996qg,Shastry:1998dx,mishin01:cu} crack
191 > propagation,\cite{BECQUART:1993rg,Rifkin1992} and alloying
192 > dynamics.\cite{Shibata:2002hh,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
193 > One of EAM's strengths is its sensitivity to small changes in
194 > structure. This is due to the inclusion of up to the third nearest
195 > neighbor interactions during fitting of the parameters.\cite{Voter95a}
196 > In comparison, the glue model of Ercolessi {\it et
197 >  al}.\cite{Ercolessi88} was only parameterized to include
198 > nearest-neighbor interactions, EAM is a suitable choice for systems
199 > where the bulk properties are of secondary importance to low-index
200 > surface structures. Additionally, the similarity of EAM's functional
201 > treatment of the embedding energy to standard density functional
202 > theory (DFT) makes fitting DFT-derived cross potentials with
203 > adsorbates somewhat easier.
204  
222
223
224
205   \subsection{Carbon Monoxide model}
206 < Previous explanations for the surface rearrangements center on
207 < the large linear quadrupole moment of carbon monoxide.\cite{Tao:2010}  
208 < We used a model first proposed by Karplus and Straub to study
209 < the photodissociation of CO from myoglobin because it reproduces
210 < the quadrupole moment well.\cite{Straub} The Straub and
211 < Karplus model, treats CO as a rigid three site molecule with a massless M
212 < site at the molecular center of mass. The geometry and interaction
213 < parameters are reproduced in Table~\ref{tab:CO}. The effective
214 < dipole moment, calculated from the assigned charges, is still
215 < small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
216 < to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
206 > Previous explanations for the surface rearrangements center on the
207 > large linear quadrupole moment of carbon monoxide.\cite{Tao:2010} We
208 > used a model first proposed by Karplus and Straub to study the
209 > photodissociation of CO from myoglobin because it reproduces the
210 > quadrupole moment well.\cite{Straub} The Straub and Karplus model
211 > treats CO as a rigid three site molecule with a massless
212 > charge-carrying ``M'' site at the center of mass. The geometry and
213 > interaction parameters are reproduced in Table~\ref{tab:CO}. The
214 > effective dipole moment, calculated from the assigned charges, is
215 > still small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is
216 > close to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
217   mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
218   %CO Table
219   \begin{table}[H]
220    \caption{Positions, Lennard-Jones parameters ($\sigma$ and
221 <    $\epsilon$), and charges for the CO-CO
222 <    interactions in Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
223 <    in kcal/mol, and charges are in atomic units.}
221 >    $\epsilon$), and charges for CO-CO
222 >    interactions. Distances are in \AA, energies are
223 >    in kcal/mol, and charges are in atomic units.  The CO model
224 >    from Ref.\bibpunct{}{}{,}{n}{}{,}
225 >    \protect\cite{Straub} was used without modification.}
226   \centering
227   \begin{tabular}{| c | c | ccc |}
228   \hline
# Line 267 | Line 249 | et al.,\cite{Pons:1986} the Pt-C interaction was fit t
249   position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}.
250   The modified parameters yield binding energies that are slightly higher
251   than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
252 < et al.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
253 < Lennard-Jones interaction to mimic strong, but short-ranged partial
252 > {\it et al}.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
253 > Lennard-Jones interaction to mimic strong, but short-ranged, partial
254   binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
255   Pt-O interaction was modeled with a Morse potential with a large
256   equilibrium distance, ($r_o$).  These choices ensure that the C is preferred
257 < over O as the surface-binding atom. In most cases, the Pt-O parameterization contributes a weak
257 > over O as the surface-binding atom. In most geometries, the Pt-O parameterization contributes a weak
258   repulsion which favors the atop site.  The resulting potential-energy
259   surface suitably recovers the calculated Pt-C separation length
260   (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
# Line 286 | Line 268 | periodic supercell plane-wave basis approach, as imple
268   The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
269   Adsorption energies were obtained from gas-surface DFT calculations with a
270   periodic supercell plane-wave basis approach, as implemented in the
271 < {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores were
271 > Quantum ESPRESSO package.\cite{QE-2009} Electron cores were
272   described with the projector augmented-wave (PAW)
273   method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
274   included to an energy cutoff of 20 Ry. Electronic energies are
# Line 300 | Line 282 | zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxe
282   performed until the energy difference between subsequent steps
283   was less than $10^{-8}$ Ry.   Nonspin-polarized supercell calculations
284   were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
285 < zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxed gold slab was
285 > zone.\cite{Monkhorst:1976} The relaxed gold slab was
286   then used in numerous single point calculations with CO at various
287   heights (and angles relative to the surface) to allow fitting of the
288   empirical force field.
# Line 309 | Line 291 | and polarization are neglected in this model, although
291   The parameters employed for the metal-CO cross-interactions in this work
292   are shown in Table~\ref{tab:co_parameters} and the binding energies on the
293   (111) surfaces are displayed in Table~\ref{tab:co_energies}.  Charge transfer
294 < and polarization are neglected in this model, although these effects are likely to
295 < affect binding energies and binding site preferences, and will be addressed in
314 < a future work.\cite{Deshlahra:2012,StreitzMintmire:1994}
294 > and polarization are neglected in this model, although these effects could have
295 > an effect on binding energies and binding site preferences.
296  
297   %Table  of Parameters
298   %Pt Parameter Set 9
299   %Au Parameter Set 35
300   \begin{table}[H]
301 <  \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
302 <    interactions are modeled with Lennard-Jones potentials. While the
303 <    metal-O interactions were fit to Morse
301 >  \caption{Parameters for the metal-CO cross-interactions. Metal-C
302 >    interactions are modeled with Lennard-Jones potentials, while the
303 >    metal-O interactions were fit to broad Morse
304      potentials.  Distances are given in \AA~and energies in kcal/mol. }
305   \centering
306   \begin{tabular}{| c | cc | c | ccc |}
# Line 343 | Line 324 | a future work.\cite{Deshlahra:2012,StreitzMintmire:199
324    \hline
325    & Calculated & Experimental \\
326    \hline
327 <  \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
327 >  \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.81} & -1.4 \bibpunct{}{}{,}{n}{}{,}
328    (Ref. \protect\cite{Kelemen:1979}) \\
329   & &  -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
330 <  \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,}  (Ref. \protect\cite{TPD_Gold}) \\
330 >  \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,}  (Ref. \protect\cite{TPDGold}) \\
331    \hline
332   \end{tabular}
333   \label{tab:co_energies}
334   \end{table}
335  
336 +
337 + \subsection{Forcefield validation}
338 + The CO-Pt cross interactions were compared directly to DFT results
339 + found in the supporting information of Tao {\it et al.}
340 + \cite{Tao:2010}, while the CO-Au results are interpreted on their own.
341 + These calculations are estimates of the stabilization
342 + energy provided to double-layer reconstructions of the perfect (557)
343 + surface by an overlayer of CO molecules in a $c (2 \times 4)$ pattern.
344 + To make the comparison, metal slabs of both Pt and Au that were five atoms thick and
345 + which displayed a (557) facet were constructed.  Double-layer
346 + (reconstructed) systems were created using six atomic layers where
347 + enough of a layer was removed from both exposed (557) facets to create
348 + the double step.  In all cases, the metal slabs contained 480 atoms
349 + and were minimized using steepest descent under the EAM force
350 + field. Both the bare metal slabs and slabs with 50\% carbon monoxide
351 + coverage (arranged in the $c (2 \times 4)$ pattern) were used.  The
352 + systems are periodic along and perpendicular to the step-edge axes
353 + with a large vacuum above the displayed (557) facet.
354 +
355 + Energies calculated using our force field for the various systems are
356 + displayed in Table ~\ref{tab:steps}.  The relative energies are calculated
357 + as $E_{relative} = E_{system} - E_{M-557-S} - N_{CO}*E_{M-CO}$,
358 + where $E_{M-CO}$ is -1.8 eV for CO-Pt and -0.39 eV for CO-Au. Our
359 + calculated CO-Pt minimum is actually at -1.83 eV at a distance of 1.53~\AA,
360 + which was obtained from single-atom liftoffs from a Pt(111) surface. The
361 + arrangement of CO on the single and double steps however, leads to a
362 + slight displacement from the minimum. For a 1 ps run at 3 K, the single
363 + step Pt-CO average bond length was 1.60~\AA, and for the double step,
364 + the bond length was 1.58~\AA. This slight increase is likely due to small
365 + electrostatic interactions among the CO and the non-ideality of the surface.
366 +
367 + For platinum, the bare double layer is less stable then the original single
368 + (557) step by about 0.25 kcal/mole per Pt atom. However, addition of carbon
369 + monoxide to the double step system provides a greater amount of stabilization
370 + when compared to single step system with CO on the order of 230 kcal/mole
371 + for this system size. The absolute difference is minimal, but this result is in
372 + qualitative agreement with DFT calculations in Tao {\it et al.}\cite{Tao:2010},
373 + who also showed that the addition of CO leads to a reversal in stability.
374 +
375 + The gold systems show a smaller energy difference between the clean
376 + single and double layers when compared to platinum. Upon addition of
377 + CO however, the single step surface becomes much more stable. These
378 + results, while helpful, need to be tempered by the weaker binding energy
379 + of CO to Au. From our simulations we see that at the elevated temperatures
380 + we are running at, it is difficult for the gold systems to maintain > than 25\%
381 + coverage, despite their being enough CO in the system.
382 +
383 + %Table of single step double step calculations
384 + \begin{table}[H]
385 +  \caption{Minimized single point energies of (S)ingle and (D)ouble
386 +    steps.  The addition of CO in a 50\% $c(2 \times 4)$ coverage acts as a
387 +    stabilizing presence and suggests a driving force for the observed
388 +    reconstruction on the highest coverage Pt system. All energies are
389 +    in kcal/mol.}
390 + \centering
391 + \begin{tabular}{| c | c | c | c | c | c |}
392 + \hline
393 + \textbf{Step} & \textbf{N}\textsubscript{M} & \textbf{N\textsubscript{CO}} & \textbf{Relative Energy} & \textbf{$\Delta$E/M} & \textbf{$\Delta$E/CO} \\
394 + \hline
395 + Pt(557)-S & 480 & 0 & 0 & 0 & - \\
396 + Pt(557)-D & 480 & 0 & 119.788 & 0.2495 & -\\
397 + Pt(557)-S & 480 & 40 & -109.734 & -0.2286 & -2.743\\
398 + Pt(557)-D & 480 & 48 & -110.039 & -0.2292 & -2.292\\
399 + \hline
400 + \hline
401 + Au(557)-S & 480 & 0 & 0 & 0 & - \\
402 + Au(557)-D & 480 & 0 & 83.853 & 0.1747 & - \\
403 + Au(557)-S & 480 & 40 & -253.604 & -0.5283 & -6.340\\
404 + Au(557)-D & 480 & 48 & -156.150 & -0.3253 & -3.253 \\
405 + \hline
406 + \end{tabular}
407 + \label{tab:steps}
408 + \end{table}
409 +
410 +
411   \subsection{Pt(557) and Au(557) metal interfaces}
412 < Our Pt system has dimensions of 18~x~24~x~9 in a box of size
413 < 54.482~x~50.046~x~120.88~\AA while our Au system has
414 < dimensions of 18~x~24~x~8 in a box of size 57.4~x~51.9285~x~100~\AA.
412 > Our Pt system is an orthorhombic periodic box of dimensions
413 > 54.482~x~50.046~x~120.88~\AA~while our Au system has
414 > dimensions of 57.4~x~51.9285~x~100~\AA. The metal slabs
415 > are 9 and 8 atoms deep respectively, corresponding to a slab
416 > thickness of $\sim$21~\AA~ for Pt and $\sim$19~\AA~for Au.
417   The systems are arranged in a FCC crystal that have been cut
418   along the (557) plane so that they are periodic in the {\it x} and
419   {\it y} directions, and have been oriented to expose two aligned
420   (557) cuts along the extended {\it z}-axis.  Simulations of the
421   bare metal interfaces at temperatures ranging from 300~K to
422 < 1200~K were performed to observe the relative
422 > 1200~K were performed to confirm the relative
423   stability of the surfaces without a CO overlayer.  
424  
425 < The different bulk melting temperatures (1337~K for Au
426 < and 2045~K for Pt) suggest that any possible reconstruction should happen at
427 < different temperatures for the two metals.  The bare Au and Pt surfaces were
428 < initially run in the canonical (NVT) ensemble at 800~K and 1000~K
429 < respectively for 100 ps. The two surfaces were relatively stable at these
430 < temperatures when no CO was present, but experienced increased surface
431 < mobility on addition of CO. Each surface was then dosed with different concentrations of CO
432 < that was initially placed in the vacuum region.  Upon full adsorption,
433 < these concentrations correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
434 < coverage. Higher coverages resulted in CO double layer formation, which introduces artifacts that are not relevant to (557) reconstruction.
435 < Because of the difference in binding energies, nearly all of the CO was bound to the Pt surface, while
436 < the Au surfaces often had a significant CO population in the gas
437 < phase.  These systems were allowed to reach thermal equilibrium (over
438 < 5 ns) before being run in the microcanonical (NVE) ensemble for
439 < data collection. All of the systems examined had at least 40 ns in the
440 < data collection stage, although simulation times for some of the
441 < systems exceeded 200~ns.  Simulations were run using the open
442 < source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
425 > The different bulk melting temperatures predicted by EAM
426 > (1345~$\pm$~10~K for Au\cite{Au:melting} and $\sim$~2045~K for
427 > Pt\cite{Pt:melting}) suggest that any reconstructions should happen at
428 > different temperatures for the two metals.  The bare Au and Pt
429 > surfaces were initially run in the canonical (NVT) ensemble at 800~K
430 > and 1000~K respectively for 100 ps. The two surfaces were relatively
431 > stable at these temperatures when no CO was present, but experienced
432 > increased surface mobility on addition of CO. Each surface was then
433 > dosed with different concentrations of CO that was initially placed in
434 > the vacuum region.  Upon full adsorption, these concentrations
435 > correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface coverage. Higher
436 > coverages resulted in the formation of a double layer of CO, which
437 > introduces artifacts that are not relevant to (557) reconstruction.
438 > Because of the difference in binding energies, nearly all of the CO
439 > was bound to the Pt surface, while the Au surfaces often had a
440 > significant CO population in the gas phase.  These systems were
441 > allowed to reach thermal equilibrium (over 5~ns) before being run in
442 > the microcanonical (NVE) ensemble for data collection. All of the
443 > systems examined had at least 40~ns in the data collection stage,
444 > although simulation times for some Pt of the systems exceeded 200~ns.
445 > Simulations were carried out using the open source molecular dynamics
446 > package, OpenMD.\cite{Ewald,OOPSE,openmd}
447  
448 < % Just results, leave discussion for discussion section
449 < % structure
450 < %       Pt: step wandering, double layers, no triangular motifs
389 < %       Au: step wandering, no double layers
390 < % dynamics
391 < %       diffusion
392 < %       time scale, formation, breakage
448 >
449 > % RESULTS
450 > %
451   \section{Results}
452   \subsection{Structural remodeling}
453 < Tao et al. have shown experimentally that the Pt(557) surface
454 < undergoes two separate reconstructions upon CO
455 < adsorption.\cite{Tao:2010} The first involves a doubling of
456 < the step height and plateau length. Similar behavior has been
457 < seen to occur on numerous surfaces at varying conditions: Ni(977), Si(111).
458 < \cite{Williams:1994,Williams:1991,Pearl} Of the two systems
459 < we examined, the Pt system showed a larger amount of
460 < reconstruction when compared to the Au system. The amount
461 < of reconstruction is correlated to the amount of CO
462 < adsorbed upon the surface.  This appears to be related to the
463 < effect that adsorbate coverage has on edge breakup and on the surface
406 < diffusion of metal adatoms. While both systems displayed step-edge
407 < wandering, only the Pt surface underwent the doubling seen by
408 < Tao et al. within the time scales studied here.  
409 < Only the 50\% coverage Pt system exhibited
410 < a complete doubling in the time scales we
411 < were able to monitor. Over longer periods (150~ns) two more double layers formed on this interface.
412 < Although double layer formation did not occur in the other Pt systems, they show
413 < more lateral movement of the step-edges
414 < compared to their Au counterparts. The 50\% Pt system is highlighted
415 < in Figure \ref{fig:reconstruct} at various times along the simulation
416 < showing the evolution of a step-edge.
453 > The bare metal surfaces experienced minor roughening of the step-edge
454 > because of the elevated temperatures, but the (557) face was stable
455 > throughout the simulations. The surfaces of both systems, upon dosage
456 > of CO, began to undergo extensive remodeling that was not observed in
457 > the bare systems. Reconstructions of the Au systems were limited to
458 > breakup of the step-edges and some step wandering. The lower coverage
459 > Pt systems experienced similar step edge wandering but to a greater
460 > extent. The 50\% coverage Pt system was unique among our simulations
461 > in that it formed well-defined and stable double layers through step
462 > coalescence, similar to results reported by Tao {\it et
463 >  al}.\cite{Tao:2010}
464  
465 < The second reconstruction on the Pt(557) surface observed by
466 < Tao involved the formation of triangular clusters that stretched
467 < across the plateau between two step-edges. Neither system, within
468 < the 40~ns time scale, experienced this reconstruction.
465 > \subsubsection{Step wandering}
466 > The bare surfaces for both metals showed minimal step-wandering at
467 > their respective temperatures. As the CO coverage increased however,
468 > the mobility of the surface atoms, described through adatom diffusion
469 > and step-edge wandering, also increased.  Except for the 50\% Pt
470 > system where step coalescence occurred, the step-edges in the other
471 > simulations preferred to keep nearly the same distance between steps
472 > as in the original (557) lattice, $\sim$13\AA~for Pt and
473 > $\sim$14\AA~for Au.  Previous work by Williams {\it et
474 >  al}.\cite{Williams:1991, Williams:1994} highlights the repulsion
475 > that exists between step-edges even when no direct interactions are
476 > present in the system. This repulsion is caused by an entropic barrier
477 > that arises from the fact that steps cannot cross over one
478 > another. This entropic repulsion does not completely define the
479 > interactions between steps, however, so it is possible to observe step
480 > coalescence on some surfaces.\cite{Williams:1991} The presence and
481 > concentration of adsorbates, as shown in this work, can affect
482 > step-step interactions, potentially leading to a new surface structure
483 > as the thermodynamic equilibrium.
484  
485 < \subsection{Dynamics}
486 < Previous atomistic simulations of stepped surfaces were largely
487 < concerned with the energetics and structures at different conditions
488 < \cite{Williams:1991,Williams:1994}. Consequently, the most common
489 < technique has been Monte Carlo. Monte Carlo gives an efficient
490 < sampling of the equilibrium thermodynamic landscape at the expense
491 < of ignoring the dynamics of the system. Previous work by Pearl and
492 < Sibener\cite{Pearl}, using STM, has been able to show the coalescing
493 < of steps on Ni(977). The time scale of the image acquisition,
494 < $\sim$70 s/image provides an upper bound for the time required for
495 < the doubling to occur. In this section we give data on dynamic and
496 < transport properties, e.g. diffusion, layer formation time, etc.
497 <
498 <
499 < \subsubsection{Transport of surface metal atoms}
500 < %forcedSystems/stepSeparation
501 < The movement or wandering of a step-edge is a cooperative effect
502 < arising from the individual movements, primarily through surface
503 < diffusion, of the atoms making up the steps An ideal metal surface
504 < displaying a low index facet, (111) or (100) is unlikely to experience
505 < much surface diffusion because of the large energetic barrier that must
444 < be overcome to lift an atom out of the surface. The presence of step-edges
445 < on higher-index surfaces provide a source for mobile metal atoms.
446 < Breaking away from the step-edge on a clean surface still imposes an
447 < energetic penalty around $\sim$~40 kcal/mol, but is much less than lifting
448 < the same metal atom vertically out of the surface,  \textgreater~60 kcal/mol.
449 < The penalty lowers significantly when CO is present in sufficient quantities
450 < on the surface. For certain distributions of CO, the penalty can be as low as
451 < $\sim$~20 kcal/mol. Once an adatom exists on the surface, the barrier for
452 < diffusion is negligible ( \textless~4 kcal/mol) and these adatoms are well
453 < able to explore the terrace before rejoining either the original step-edge or becoming a part
454 < of a different edge. Atoms traversing separate terraces is a more difficult
455 < process, but can be overcome through a joining and lifting stage which is
456 < examined in the discussion section. By tracking the mobility of individual
457 < metal atoms on the Pt and Au surfaces we were able to determine the relative
458 < diffusion constants, as well as how varying coverages of CO affect the diffusion. Close
459 < observation of the mobile metal atoms showed that they were typically in
460 < equilibrium with the step-edges, dynamically breaking apart and rejoining the edges.
461 < At times, their motion was concerted and two or more adatoms would be
462 < observed moving together across the surfaces. The primary challenge in
463 < quantifying the overall surface mobility was in defining ``mobile" vs. ``static" atoms.
485 > \subsubsection{Double layers}
486 > Tao {\it et al}.\cite{Tao:2010} have shown experimentally that the
487 > Pt(557) surface undergoes two separate reconstructions upon CO
488 > adsorption.  The first involves a doubling of the step height and
489 > plateau length.  Similar behavior has been seen on a number of
490 > surfaces at varying conditions, including Ni(977) and
491 > Si(111).\cite{Williams:1994,Williams:1991,Pearl} Of the two systems we
492 > examined, the Pt system showed a greater propensity for reconstruction
493 > because of the larger surface mobility and the greater extent of step
494 > wandering.  The amount of reconstruction was strongly correlated to
495 > the amount of CO adsorbed upon the surface.  This appears to be
496 > related to the effect that adsorbate coverage has on edge breakup and
497 > on the surface diffusion of metal adatoms. Only the 50\% Pt surface
498 > underwent the doubling seen by Tao {\it et al}.\cite{Tao:2010} within
499 > the time scales studied here.  Over a longer time scale (150~ns) two
500 > more double layers formed on this surface. Although double layer
501 > formation did not occur in the other Pt systems, they exhibited more
502 > step-wandering and roughening compared to their Au counterparts. The
503 > 50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
504 > various times along the simulation showing the evolution of a double
505 > layer step-edge.
506  
507 < A particle was considered mobile once it had traveled more than 2~\AA~
508 < between saved configurations of the system (typically 10-100 ps). An atom that was
509 < truly mobile would typically travel much greater distances than this, but the 2~\AA~ cutoff
510 < was to prevent swamping the diffusion data with the in-place vibrational
511 < movement of buried atoms. Diffusion on  a surface is strongly affected by
470 < local structures and in this work, the presence of single and double layer
471 < step-edges causes the diffusion parallel to the step-edges to be different
472 < from the diffusion perpendicular to these edges. Parallel and perpendicular
473 < diffusion constants are shown in Figure \ref{fig:diff}.
507 > The second reconstruction observed by Tao {\it et al}.\cite{Tao:2010}
508 > involved the formation of triangular clusters that stretched across
509 > the plateau between two step-edges. Neither of the simulated metal
510 > interfaces, within the 40~ns time scale or the extended time of 150~ns
511 > for the 50\% Pt system, experienced this reconstruction.
512  
475 \subsubsection{Double layer formation dynamics}
476 The increased amounts of diffusion on Pt at the higher CO coverages plays a primary role in the formation of the double layers observed on Pt. However, this is not a complete explanation as seen by the 33\% Pt system which has higher diffusion constants but did not show any signs of undergoing the doubling. This difference will be explored more fully in the discussion. On the 50\% Pt system, three separate layers were formed over the extended run time of this system. Previous experimental work has given some insight into the upper bounds of the time required for step coalescing.\cite{Williams:1991,Pearl} In this system, as seen in Figure \ref{fig:reconstruct}, the first appearance of a double layer, a nodal site, appears at 19 ns into the simulation. Within 12 ns, nearly half of the step has formed the double layer and by 86 ns, the complete layer has been smoothed. The double layer could be considered ``complete" by 37 ns but is a bit rough or wavy. From the appearance of the first node to the first observed double layer, ignoring roughening, the process took $\sim$20 ns. Another $\sim$40 ns was necessary for the layer to completely straighten. The other two layers in this simulation form over a period of 22 ns and 42 ns respectively. Comparing this to the upper bounds of the image scan, it is likely that aspects of this reconstruction occur very quickly. A possible explanation for this rapid reconstruction is the elevated temperatures our systems were run at. It is likely that the process would take longer at lower temperatures and is an area of exploration for future work.
477
513   %Evolution of surface
514   \begin{figure}[H]
515 < \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
516 < \caption{The Pt(557) / 50\% CO system at a sequence of times after
517 <  initial exposure to the CO: (a) 258 ps, (b) 19 ns, (c) 31.2 ns, and
518 <  (d) 86.1 ns. Disruption of the (557) step-edges occurs quickly.  The
515 > \includegraphics[width=\linewidth]{EPS_ProgressionOfDoubleLayerFormation}
516 > \caption{The Pt(557) / 50\% CO interface upon exposure to the CO: (a)
517 >  258~ps, (b) 19~ns, (c) 31.2~ns, and (d) 86.1~ns after
518 >  exposure. Disruption of the (557) step-edges occurs quickly.  The
519    doubling of the layers appears only after two adjacent step-edges
520    touch.  The circled spot in (b) nucleated the growth of the double
521    step observed in the later configurations.}
522    \label{fig:reconstruct}
523   \end{figure}
524  
525 + \subsection{Dynamics}
526 + Previous experimental work by Pearl and Sibener\cite{Pearl}, using
527 + STM, has been able to capture the coalescence of steps on Ni(977). The
528 + time scale of the image acquisition, $\sim$70~s/image, provides an
529 + upper bound for the time required for the doubling to occur. By
530 + utilizing Molecular Dynamics we are able to probe the dynamics of
531 + these reconstructions at elevated temperatures and in this section we
532 + provide data on the timescales for transport properties,
533 + e.g. diffusion and layer formation time.
534 +
535 +
536 + \subsubsection{Transport of surface metal atoms}
537 + %forcedSystems/stepSeparation
538 +
539 + The wandering of a step-edge is a cooperative effect arising from the
540 + individual movements of the atoms making up the steps. An ideal metal
541 + surface displaying a low index facet, (111) or (100), is unlikely to
542 + experience much surface diffusion because of the large energetic
543 + barrier that must be overcome to lift an atom out of the surface. The
544 + presence of step-edges and other surface features on higher-index
545 + facets provides a lower energy source for mobile metal atoms.  Using
546 + our potential model, single-atom break-away from a step-edge on a
547 + clean surface still imposes an energetic penalty around
548 + $\sim$~45~kcal/mol, but this is certainly easier than lifting the same
549 + metal atom vertically out of the surface, \textgreater~60~kcal/mol.
550 + The penalty lowers significantly when CO is present in sufficient
551 + quantities on the surface. For certain distributions of CO, the
552 + energetic penalty can fall to as low as $\sim$~20~kcal/mol. The
553 + configurations that create these lower barriers are detailed in the
554 + discussion section below.
555 +
556 + Once an adatom exists on the surface, the barrier for diffusion is
557 + negligible (\textless~4~kcal/mol for a Pt adatom). These adatoms are
558 + then able to explore the terrace before rejoining either their
559 + original step-edge or becoming a part of a different edge. It is an
560 + energetically unfavorable process with a high barrier for an atom to
561 + traverse to a separate terrace although the presence of CO can lower
562 + the energy barrier required to lift or lower an adatom. By tracking
563 + the mobility of individual metal atoms on the Pt and Au surfaces we
564 + were able to determine the relative diffusion constants, as well as
565 + how varying coverages of CO affect the diffusion. Close observation of
566 + the mobile metal atoms showed that they were typically in equilibrium
567 + with the step-edges.  At times, their motion was concerted, and two or
568 + more adatoms would be observed moving together across the surfaces.
569 +
570 + A particle was considered ``mobile'' once it had traveled more than
571 + 2~\AA~ between saved configurations of the system (typically 10-100
572 + ps). A mobile atom would typically travel much greater distances than
573 + this, but the 2~\AA~cutoff was used to prevent swamping the diffusion
574 + data with the in-place vibrational movement of buried atoms. Diffusion
575 + on a surface is strongly affected by local structures and the presence
576 + of single and double layer step-edges causes the diffusion parallel to
577 + the step-edges to be larger than the diffusion perpendicular to these
578 + edges. Parallel and perpendicular diffusion constants are shown in
579 + Figure \ref{fig:diff}.  Diffusion parallel to the step-edge is higher
580 + than diffusion perpendicular to the edge because of the lower energy
581 + barrier associated with sliding along an edge compared to breaking
582 + away to form an isolated adatom.
583 +
584 + %Diffusion graph
585   \begin{figure}[H]
586 < \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade.pdf}
586 > \includegraphics[width=\linewidth]{Portrait_DiffusionComparison_1}
587   \caption{Diffusion constants for mobile surface atoms along directions
588    parallel ($\mathbf{D}_{\parallel}$) and perpendicular
589    ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
590 <  surface coverage.  Diffusion parallel to the step-edge is higher
591 <  than that perpendicular to the edge because of the lower energy
592 <  barrier associated with traversing along the edge as compared to
593 <  completely breaking away. Additionally, the observed
499 <  maximum and subsequent decrease for the Pt system suggests that the
500 <  CO self-interactions are playing a significant role with regards to
501 <  movement of the Pt atoms around and across the surface. }
590 >  surface coverage.  The two reported diffusion constants for the 50\%
591 >  Pt system correspond to a 20~ns period before the formation of the
592 >  double layer (upper points), and to the full 40~ns sampling period
593 >  (lower points).}
594   \label{fig:diff}
595   \end{figure}
596  
597 + The weaker Au-CO interaction is evident in the weak CO-coverage
598 + dependance of Au diffusion. This weak interaction leads to lower
599 + observed coverages when compared to dosage amounts. This further
600 + limits the effect the CO can have on surface diffusion. The correlation
601 + between coverage and Pt diffusion rates shows a near linear relationship
602 + at the earliest times in the simulations. Following double layer formation,
603 + however, there is a precipitous drop in adatom diffusion. As the double
604 + layer forms, many atoms that had been tracked for mobility data have
605 + now been buried, resulting in a smaller reported diffusion constant. A
606 + secondary effect of higher coverages is CO-CO cross interactions that
607 + lower the effective mobility of the Pt adatoms that are bound to each CO.
608 + This effect would become evident only at higher coverages. A detailed
609 + account of Pt adatom energetics follows in the Discussion.
610 +
611 + \subsubsection{Dynamics of double layer formation}
612 + The increased diffusion on Pt at the higher CO coverages is the primary
613 + contributor to double layer formation. However, this is not a complete
614 + explanation -- the 33\%~Pt system has higher diffusion constants, but
615 + did not show any signs of edge doubling in 40~ns. On the 50\%~Pt
616 + system, one double layer formed within the first 40~ns of simulation time,
617 + while two more were formed as the system was allowed to run for an
618 + additional 110~ns (150~ns total). This suggests that this reconstruction
619 + is a rapid process and that the previously mentioned upper bound is a
620 + very large overestimate.\cite{Williams:1991,Pearl} In this system the first
621 + appearance of a double layer appears at 19~ns into the simulation.
622 + Within 12~ns of this nucleation event, nearly half of the step has formed
623 + the double layer and by 86~ns the complete layer has flattened out.
624 + From the appearance of the first nucleation event to the first observed
625 + double layer, the process took $\sim$20~ns. Another $\sim$40~ns was
626 + necessary for the layer to completely straighten. The other two layers in
627 + this simulation formed over periods of 22~ns and 42~ns respectively.
628 + A possible explanation for this rapid reconstruction is the elevated
629 + temperatures under which our systems were simulated. The process
630 + would almost certainly take longer at lower temperatures. Additionally,
631 + our measured times for completion of the doubling after the appearance
632 + of a nucleation site are likely affected by our periodic boxes. A longer
633 + step-edge will likely take longer to ``zipper''.
634  
635  
507
636   %Discussion
637   \section{Discussion}
638 < In this paper we have shown that we were able to accurately model the initial reconstruction of the
639 < Pt(557) surface upon CO adsorption as shown by Tao et al. \cite{Tao:2010}. More importantly, we
640 < were able to observe the dynamic processes necessary for this reconstruction.
638 > We have shown that a classical potential is able to model the initial
639 > reconstruction of the Pt(557) surface upon CO adsorption, and have
640 > reproduced the double layer structure observed by Tao {\it et
641 >  al}.\cite{Tao:2010}. Additionally, this reconstruction appears to be
642 > rapid -- occurring within 100 ns of the initial exposure to CO.  Here
643 > we discuss the features of the classical potential that are
644 > contributing to the stability and speed of the Pt(557) reconstruction.
645  
646 < \subsection{Mechanism for restructuring}
647 < Since the Au surface showed no large scale restructuring throughout
648 < our simulation time our discussion will focus on the 50\% Pt-CO system
649 < which did undergo the doubling featured in Figure \ref{fig:reconstruct}.
650 < Comparing the results from this simulation to those reported previously by
651 < Tao et al.\cite{Tao:2010} the similarities in the Pt-CO system are quite
652 < strong. As shown in Figure \ref{fig:reconstruct}, the simulated Pt
653 < system exposed to a large dosage of CO will restructure by doubling the terrace
654 < widths and step heights. The restructuring occurs in a piecemeal fashion, one to two Pt atoms at a time and as such is a fairly stochastic event.
655 < Looking at individual configurations of the system, the adatoms either
656 < break away from the step-edge and stay on the lower terrace or they lift
657 < up onto the higher terrace. Once ``free'', they will diffuse on the terrace
658 < until reaching another step-edge or rejoining their original edge.  
659 < This combination of growth and decay of the step-edges is in a state of
528 < dynamic equilibrium. However, once two previously separated edges
529 < meet as shown in Figure 1.B, this meeting point tends to act as a focus
530 < or growth point for the rest of the edge to meet up, akin to that of a zipper.
531 < From the handful of cases where a double layer was formed during the
532 < simulation, measuring from the initial appearance of a growth point, the
533 < double layer tends to be fully formed within $\sim$35 ns.
646 > \subsection{Diffusion}
647 > The perpendicular diffusion constant appears to be the most important
648 > indicator of double layer formation. As highlighted in Figure
649 > \ref{fig:reconstruct}, the formation of the double layer did not begin
650 > until a nucleation site appeared.  Williams {\it et
651 >  al}.\cite{Williams:1991,Williams:1994} cite an effective edge-edge
652 > repulsion arising from the inability of edge crossing.  This repulsion
653 > must be overcome to allow step coalescence.  A larger
654 > $\textbf{D}_\perp$ value implies more step-wandering and a larger
655 > chance for the stochastic meeting of two edges to create a nucleation
656 > point.  Diffusion parallel to the step-edge can help ``zipper'' up a
657 > nascent double layer. This helps explain the rapid time scale for
658 > double layer completion after the appearance of a nucleation site, while
659 > the initial appearance of the nucleation site was unpredictable.
660  
661 < A number of possible mechanisms exist to explain the role of adsorbed
662 < CO in restructuring the Pt surface. Quadrupolar repulsion between adjacent
663 < CO molecules adsorbed on the surface is one likely possibility.  However,
664 < the quadrupole-quadrupole interaction is short-ranged and is attractive for
665 < some orientations.  If the CO molecules are ``locked'' in a specific orientation
666 < relative to each other, through atop adsorption for example, this explanation
667 < gains some weight.  The energetic repulsion between two CO located a
668 < distance of 2.77~\AA~apart (nearest-neighbor distance of Pt) with both in
669 < a  vertical orientation is 8.62 kcal/mol. Moving the CO apart to the second
670 < nearest-neighbor distance of 4.8~\AA~or 5.54~\AA~drops the repulsion to
671 < nearly 0 kcal/mol. Allowing the CO's to leave a purely vertical orientation
672 < also quickly drops the repulsion, a minimum of 6.2 kcal/mol is reached at $\sim$24 degrees between the 2 CO when the carbons are locked at a distance of 2.77 \AA apart.
673 < As mentioned above, the energy barrier for surface diffusion
674 < of a Pt adatom is only 4 kcal/mol. So this repulsion between CO can help
675 < increase the surface diffusion. However, the residence time of CO on Pt was
676 < examined and while the majority of the CO is on or near the surface throughout
677 < the run, it is extremely mobile. This mobility suggests that the CO are more
678 < likely to shift their positions without necessarily dragging the Pt along with them.
661 > \subsection{Mechanism for restructuring}
662 > Since the Au surface showed no large scale restructuring in any of our
663 > simulations, our discussion will focus on the 50\% Pt-CO system which
664 > did exhibit doubling. A number of possible mechanisms exist to explain
665 > the role of adsorbed CO in restructuring the Pt surface. Quadrupolar
666 > repulsion between adjacent CO molecules adsorbed on the surface is one
667 > possibility.  However, the quadrupole-quadrupole interaction is
668 > short-ranged and is attractive for some orientations.  If the CO
669 > molecules are ``locked'' in a vertical orientation, through atop
670 > adsorption for example, this explanation would gain credence. Within
671 > the framework of our classical potential, the calculated energetic
672 > repulsion between two CO molecules located a distance of
673 > 2.77~\AA~apart (nearest-neighbor distance of Pt) and both in a
674 > vertical orientation, is 8.62 kcal/mol. Moving the CO to the second
675 > nearest-neighbor distance of 4.8~\AA~drops the repulsion to nearly
676 > 0. Allowing the CO to rotate away from a purely vertical orientation
677 > also lowers the repulsion. When the carbons are locked at a distance
678 > of 2.77~\AA, a minimum of 6.2 kcal/mol is reached when the angle
679 > between the 2 CO is $\sim$24\textsuperscript{o}.  The calculated
680 > barrier for surface diffusion of a Pt adatom is only 4 kcal/mol, so
681 > repulsion between adjacent CO molecules bound to Pt could indeed
682 > increase the surface diffusion. However, the residence time of CO on
683 > Pt suggests that the CO molecules are extremely mobile, with diffusion
684 > constants 40 to 2500 times larger than surface Pt atoms. This mobility
685 > suggests that the CO molecules jump between different Pt atoms
686 > throughout the simulation.  However, they do stay bound to individual
687 > Pt atoms for long enough to modify the local energy landscape for the
688 > mobile adatoms.
689  
690 < Another possible and more likely mechanism for the restructuring is in the
691 < destabilization of strong Pt-Pt interactions by CO adsorbed on surface
692 < Pt atoms.  This would then have the effect of increasing surface mobility
693 < of these atoms.  To test this hypothesis, numerous configurations of
694 < CO in varying quantities were arranged on the higher and lower plateaus
695 < around a step on a otherwise clean Pt(557) surface. One representative
696 < configuration is displayed in Figure \ref{fig:lambda}. Single or concerted movement
697 < of Pt atoms was then examined to determine possible barriers. Because
698 < the movement was forced along a pre-defined reaction coordinate that may differ
699 < from the true minimum of this path, only the beginning and ending energies
700 < are displayed in Table \ref{tab:energies}. These values suggest that the presence of CO at suitable
701 < locations can lead to lowered barriers for Pt breaking apart from the step-edge.
702 < Additionally, as highlighted in Figure \ref{fig:lambda}, the presence of CO makes the
703 < burrowing and lifting of adatoms favorable, whereas without CO, the process is neutral
704 < in terms of energetics.
690 > A different interpretation of the above mechanism which takes the
691 > large mobility of the CO into account, would be in the destabilization
692 > of Pt-Pt interactions due to bound CO.  Destabilizing Pt-Pt bonds at
693 > the edges could lead to increased step-edge breakup and diffusion. On
694 > the bare Pt(557) surface the barrier to completely detach an edge atom
695 > is $\sim$43~kcal/mol, as is shown in configuration (a) in Figures
696 > \ref{fig:SketchGraphic} \& \ref{fig:SketchEnergies}. For certain
697 > configurations, cases (e), (g), and (h), the barrier can be lowered to
698 > $\sim$23~kcal/mol by the presence of bound CO molecules. In these
699 > instances, it becomes energetically favorable to roughen the edge by
700 > introducing a small separation of 0.5 to 1.0~\AA. This roughening
701 > becomes immediately obvious in simulations with significant CO
702 > populations. The roughening is present to a lesser extent on surfaces
703 > with lower CO coverage (and even on the bare surfaces), although in
704 > these cases it is likely due to random fluctuations that squeeze out
705 > step-edge atoms. Step-edge breakup by direct single-atom translations
706 > (as suggested by these energy curves) is probably a worst-case
707 > scenario.  Multistep mechanisms in which an adatom moves laterally on
708 > the surface after being ejected would be more energetically favorable.
709 > This would leave the adatom alongside the ledge, providing it with
710 > five nearest neighbors.  While fewer than the seven neighbors it had
711 > as part of the step-edge, it keeps more Pt neighbors than the three
712 > neighbors an isolated adatom has on the terrace. In this proposed
713 > mechanism, the CO quadrupolar repulsion still plays a role in the
714 > initial roughening of the step-edge, but not in any long-term bonds
715 > with individual Pt atoms.  Higher CO coverages create more
716 > opportunities for the crowded CO configurations shown in Figure
717 > \ref{fig:SketchGraphic}, and this is likely to cause an increased
718 > propensity for step-edge breakup.
719  
720 + %Sketch graphic of different configurations
721 + \begin{figure}[H]
722 + \includegraphics[width=\linewidth]{COpaths}
723 + \caption{Configurations used to investigate the mechanism of step-edge
724 +  breakup on Pt(557). In each case, the central (starred) atom was
725 +  pulled directly across the surface away from the step edge.  The Pt
726 +  atoms on the upper terrace are colored dark grey, while those on the
727 +  lower terrace are in white.  In each of these configurations, some
728 +  of the atoms (highlighted in blue) had CO molecules bound in the
729 +  vertical atop position.  The energies of these configurations as a
730 +  function of central atom displacement are displayed in Figure
731 +  \ref{fig:SketchEnergies}.}
732 + \label{fig:SketchGraphic}
733 + \end{figure}
734 +
735 + %energy graph corresponding to sketch graphic
736 + \begin{figure}[H]
737 + \includegraphics[width=\linewidth]{Portrait_SeparationComparison}
738 + \caption{Energies for displacing a single edge atom perpendicular to
739 +  the step edge as a function of atomic displacement. Each of the
740 +  energy curves corresponds to one of the labeled configurations in
741 +  Figure \ref{fig:SketchGraphic}, and the energies are referenced to
742 +  the unperturbed step-edge.  Certain arrangements of bound CO
743 +  (notably configurations g and h) can lower the energetic barrier for
744 +  creating an adatom relative to the bare surface (configuration a).}
745 + \label{fig:SketchEnergies}
746 + \end{figure}
747 +
748 + While configurations of CO on the surface are able to increase
749 + diffusion and the likelihood of edge wandering, this does not provide
750 + a complete explanation for the formation of double layers. If adatoms
751 + were constrained to their original terraces then doubling could not
752 + occur.  A mechanism for vertical displacement of adatoms at the
753 + step-edge is required to explain the doubling.
754 +
755 + We have discovered one possible mechanism for a CO-mediated vertical
756 + displacement of Pt atoms at the step edge. Figure \ref{fig:lambda}
757 + shows four points along a reaction coordinate in which a CO-bound
758 + adatom along the step-edge ``burrows'' into the edge and displaces the
759 + original edge atom onto the higher terrace.  A number of events
760 + similar to this mechanism were observed during the simulations.  We
761 + predict an energetic barrier of 20~kcal/mol for this process (in which
762 + the displaced edge atom follows a curvilinear path into an adjacent
763 + 3-fold hollow site).  The barrier heights we obtain for this reaction
764 + coordinate are approximate because the exact path is unknown, but the
765 + calculated energy barriers would be easily accessible at operating
766 + conditions.  Additionally, this mechanism is exothermic, with a final
767 + energy 15~kcal/mol below the original $\lambda = 0$ configuration.
768 + When CO is not present and this reaction coordinate is followed, the
769 + process is endothermic by 3~kcal/mol.  The difference in the relative
770 + energies for the $\lambda=0$ and $\lambda=1$ case when CO is present
771 + provides strong support for CO-mediated Pt-Pt interactions giving rise
772 + to the doubling reconstruction.
773 +
774   %lambda progression of Pt -> shoving its way into the step
775   \begin{figure}[H]
776 < \includegraphics[width=\linewidth]{lambdaProgression_atopCO.png}
777 < \caption{A model system of the Pt(557) surface was used as the framework
778 < for exploring energy barriers along a reaction coordinate. Various numbers,
779 < placements, and rotations of CO were examined as they affect Pt movement.
780 < The coordinate displayed in this Figure was a representative run. As shown
781 < in Table \ref{tab:rxcoord}, relative to the energy of the system at 0\%, there
782 < is a slight decrease upon insertion of the Pt atom into the step-edge along
579 < with the resultant lifting of the other Pt atom when CO is present at certain positions.}
776 > \includegraphics[width=\linewidth]{EPS_rxnCoord}
777 > \caption{Points along a possible reaction coordinate for CO-mediated
778 >  edge doubling. Here, a CO-bound adatom burrows into an established
779 >  step edge and displaces an edge atom onto the upper terrace along a
780 >  curvilinear path.  The approximate barrier for the process is
781 >  20~kcal/mol, and the complete process is exothermic by 15~kcal/mol
782 >  in the presence of CO, but is endothermic by 3~kcal/mol without CO.}
783   \label{fig:lambda}
784   \end{figure}
785  
786 + The mechanism for doubling on the Pt(557) surface appears to require
787 + the cooperation of at least two distinct processes. For complete
788 + doubling of a layer to occur there must be a breakup of one
789 + terrace. These atoms must then ``disappear'' from that terrace, either
790 + by travelling to the terraces above or below their original levels.
791 + The presence of CO helps explain mechanisms for both of these
792 + situations. There must be sufficient breakage of the step-edge to
793 + increase the concentration of adatoms on the surface and these adatoms
794 + must then undergo the burrowing highlighted above (or a comparable
795 + mechanism) to create the double layer.  With sufficient time, these
796 + mechanisms working in concert lead to the formation of a double layer.
797  
798 + \subsection{CO Removal and double layer stability}
799 + Once the double layers had formed on the 50\%~Pt system, they remained
800 + stable for the rest of the simulation time with minimal movement.
801 + Random fluctuations that involved small clusters or divots were
802 + observed, but these features typically healed within a few
803 + nanoseconds.  Within our simulations, the formation of the double
804 + layer appeared to be irreversible and a double layer was never
805 + observed to split back into two single layer step-edges while CO was
806 + present.
807  
808 < \subsection{Diffusion}
809 < As shown in the results section, the diffusion parallel to the step-edge tends to be
810 < much larger than that perpendicular to the step-edge, likely because of the dynamic
811 < equilibrium that is established between the step-edge and adatom interface. The coverage
812 < of CO also appears to play a slight role in relative rates of diffusion, as shown in Figure \ref{fig:diff}.
813 < The
814 < Thus, the bottleneck of the double layer formation appears to be the initial formation
815 < of this growth point, which seems to be somewhat of a stochastic event. Once it
816 < appears, parallel diffusion, along the now slightly angled step-edge, will allow for
817 < a faster formation of the double layer than if the entire process were dependent on
818 < only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the
819 < more likely a growth point is to be formed.
597 < \\
808 > To further gauge the effect CO has on this surface, additional
809 > simulations were run starting from a late configuration of the 50\%~Pt
810 > system that had already formed double layers. These simulations then
811 > had their CO molecules suddenly removed.  The double layer broke apart
812 > rapidly in these simulations, showing a well-defined edge-splitting
813 > after 100~ps. Configurations of this system are shown in Figure
814 > \ref{fig:breaking}. The coloring of the top and bottom layers helps to
815 > show how much mixing the edges experience as they split. These systems
816 > were only examined for 10~ns, and within that time despite the initial
817 > rapid splitting, the edges only moved another few \AA~apart. It is
818 > possible that with longer simulation times, the (557) surface recovery
819 > observed by Tao {\it et al}.\cite{Tao:2010} could also be recovered.
820  
599
821   %breaking of the double layer upon removal of CO
822   \begin{figure}[H]
823 < \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
824 < %:
825 < \caption{(A)  0 ps, (B) 100 ps, (C) 1 ns, after the removal of CO. The presence of the CO
826 < helped maintain the stability of the double layer and upon removal the two layers break
827 < and begin separating. The separation is not a simple pulling apart however, rather
828 < there is a mixing of the lower and upper atoms at the edge.}
823 > \includegraphics[width=\linewidth]{EPS_doubleLayerBreaking}
824 > \caption{Behavior of an established (111) double step after removal of
825 >  the adsorbed CO: (A) 0~ps, (B) 100~ps, and (C) 1~ns after the
826 >  removal of CO.  Nearly immediately after the CO is removed, the
827 >  step edge reforms in a (100) configuration, which is also the step
828 >  type seen on clean (557) surfaces. The step separation involves
829 >  significant mixing of the lower and upper atoms at the edge.}
830   \label{fig:breaking}
831   \end{figure}
832  
833  
612
613
834   %Peaks!
835 < \begin{figure}[H]
836 < \includegraphics[width=\linewidth]{doublePeaks_noCO.png}
837 < \caption{At the initial formation of this double layer  ( $\sim$ 37 ns) there is a degree
838 < of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
839 < aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
840 < \label{fig:peaks}
841 < \end{figure}
835 > %\begin{figure}[H]
836 > %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
837 > %\caption{At the initial formation of this double layer  ( $\sim$ 37 ns) there is a degree
838 > %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
839 > %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
840 > %\label{fig:peaks}
841 > %\end{figure}
842  
843  
844   %Don't think I need this
845   %clean surface...
846   %\begin{figure}[H]
847 < %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
847 > %\includegraphics[width=\linewidth]{557_300K_cleanPDF}
848   %\caption{}
849  
850   %\end{figure}
# Line 632 | Line 852 | In this work we have shown the reconstruction of the P
852  
853  
854   \section{Conclusion}
855 < In this work we have shown the reconstruction of the Pt(557) crystalline surface upon adsorption of CO in less than a $\mu s$. Only the highest coverage Pt system showed this initial reconstruction similar to that seen previously. The strong interaction between Pt and CO and the limited interaction between Au and CO helps explain the differences between the two systems.
855 > The strength and directionality of the Pt-CO binding interaction, as
856 > well as the large quadrupolar repulsion between atop-bound CO
857 > molecules, help to explain the observed increase in surface mobility
858 > of Pt(557) and the resultant reconstruction into a double-layer
859 > configuration at the highest simulated CO-coverages.  The weaker Au-CO
860 > interaction results in significantly lower adataom diffusion
861 > constants, less step-wandering, and a lack of the double layer
862 > reconstruction on the Au(557) surface.
863  
864 + An in-depth examination of the energetics shows the important role CO
865 + plays in increasing step-breakup and in facilitating edge traversal
866 + which are both necessary for double layer formation.
867 +
868   %Things I am not ready to remove yet
869  
870   %Table of Diffusion Constants
# Line 656 | Line 887 | In this work we have shown the reconstruction of the P
887   % \end{tabular}
888   % \end{table}
889  
890 < \section{Acknowledgments}
891 < Support for this project was provided by the National Science
892 < Foundation under grant CHE-0848243 and by the Center for Sustainable
893 < Energy at Notre Dame (cSEND). Computational time was provided by the
894 < Center for Research Computing (CRC) at the University of Notre Dame.
895 <
890 > \begin{acknowledgement}
891 >  We gratefully acknowledge conversations with Dr. William
892 >  F. Schneider and Dr. Feng Tao.  Support for this project was
893 >  provided by the National Science Foundation under grant CHE-0848243
894 >  and by the Center for Sustainable Energy at Notre Dame
895 >  (cSEND). Computational time was provided by the Center for Research
896 >  Computing (CRC) at the University of Notre Dame.
897 > \end{acknowledgement}
898   \newpage
899 < \bibliography{firstTryBibliography}
900 < \end{doublespace}
899 > \bibstyle{achemso}
900 > \bibliography{COonPtAu}
901 > %\end{doublespace}
902 >
903 > \begin{tocentry}
904 > \begin{wrapfigure}{l}{0.5\textwidth}
905 > \begin{center}
906 > \includegraphics[width=\linewidth]{TOC_doubleLayer}
907 > \end{center}
908 > \end{wrapfigure}
909 > A reconstructed Pt(557) surface after 86~ns exposure to a half a
910 > monolayer of CO.  The double layer that forms is a result of
911 > CO-mediated step-edge wandering as well as a burrowing mechanism that
912 > helps lift edge atoms onto an upper terrace.
913 > \end{tocentry}
914 >
915   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines