ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3890
Committed: Tue Jun 4 20:22:12 2013 UTC (11 years, 1 month ago) by jmichalk
Content type: application/x-tex
File size: 47350 byte(s)
Log Message:
new table

File Contents

# User Rev Content
1 gezelter 3875 \documentclass[journal = jpccck, manuscript = article]{achemso}
2     \setkeys{acs}{usetitle = true}
3     \usepackage{achemso}
4     \usepackage{natbib}
5 gezelter 3808 \usepackage{multirow}
6 jmichalk 3885 \usepackage{wrapfig}
7 jmichalk 3889 \usepackage{fixltx2e}
8 gezelter 3887 %\mciteErrorOnUnknownfalse
9 gezelter 3875
10     \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
11 gezelter 3808 \usepackage{url}
12 jmichalk 3802
13 gezelter 3875 \title{Molecular Dynamics simulations of the surface reconstructions
14     of Pt(557) and Au(557) under exposure to CO}
15    
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 gezelter 3808 \begin{document}
27    
28 gezelter 3875
29 jmichalk 3802 %%
30     %Introduction
31     % Experimental observations
32     % Previous work on Pt, CO, etc.
33     %
34     %Simulation Methodology
35     % FF (fits and parameters)
36     % MD (setup, equilibration, collection)
37     %
38     % Analysis of trajectories!!!
39     %Discussion
40     % CO preferences for specific locales
41     % CO-CO interactions
42     % Differences between Au & Pt
43     % Causes of 2_layer reordering in Pt
44     %Summary
45     %%
46    
47 gezelter 3818
48 gezelter 3808 \begin{abstract}
49 gezelter 3882 The mechanism and dynamics of surface reconstructions of Pt(557) and
50     Au(557) exposed to various coverages of carbon monoxide (CO) were
51 gezelter 3887 investigated using molecular dynamics simulations. Metal-CO
52 gezelter 3882 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 gezelter 3887 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 gezelter 3808 \end{abstract}
65 jmichalk 3802
66 gezelter 3808 \newpage
67    
68    
69 jmichalk 3802 \section{Introduction}
70     % Importance: catalytically active metals are important
71     % Sub: Knowledge of how their surface structure affects their ability to catalytically facilitate certain reactions is growing, but is more reactionary than predictive
72     % Sub: Designing catalysis is the future, and will play an important role in numerous processes (ones that are currently seen to be impractical, or at least inefficient)
73     % Theory can explore temperatures and pressures which are difficult to work with in experiments
74     % Sub: Also, easier to observe what is going on and provide reasons and explanations
75     %
76    
77 gezelter 3826 Industrial catalysts usually consist of small particles that exhibit a
78     high concentration of steps, kink sites, and vacancies at the edges of
79     the facets. These sites are thought to be the locations of catalytic
80 gezelter 3808 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
81 gezelter 3826 significant evidence that solid surfaces are often structurally,
82     compositionally, and chemically modified by reactants under operating
83     conditions.\cite{Tao2008,Tao:2010,Tao2011} The coupling between
84     surface oxidation states and catalytic activity for CO oxidation on
85     Pt, for instance, is widely documented.\cite{Ertl08,Hendriksen:2002}
86     Despite the well-documented role of these effects on reactivity, the
87     ability to capture or predict them in atomistic models is somewhat
88     limited. While these effects are perhaps unsurprising on the highly
89     disperse, multi-faceted nanoscale particles that characterize
90     industrial catalysts, they are manifest even on ordered, well-defined
91     surfaces. The Pt(557) surface, for example, exhibits substantial and
92     reversible restructuring under exposure to moderate pressures of
93     carbon monoxide.\cite{Tao:2010}
94 jmichalk 3802
95 gezelter 3887 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 gezelter 3826
122 jmichalk 3811 %Platinum molecular dynamics
123     %gold molecular dynamics
124 jmichalk 3802
125     \section{Simulation Methods}
126 gezelter 3887 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 gezelter 3808 molecular dynamics
132     (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
133     methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
134     mechanical potential energy surfaces remain out of reach.
135     Additionally, the ``bonds'' between metal atoms at a surface are
136     typically not well represented in terms of classical pairwise
137     interactions in the same way that bonds in a molecular material are,
138     nor are they captured by simple non-directional interactions like the
139 gezelter 3826 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 gezelter 3887 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 gezelter 3826 photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
145     Pt-CO cross interactions were parameterized as part of this work.
146 gezelter 3808
147     \subsection{Metal-metal interactions}
148 gezelter 3826 Many of the potentials used for modeling transition metals are based
149     on a non-pairwise additive functional of the local electron
150     density. The embedded atom method (EAM) is perhaps the best known of
151     these
152 gezelter 3808 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 gezelter 3887 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 gezelter 3808 \begin{equation*}
163     \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
164     \end{equation*}
165     Here, $\rho_j(r_{ij})$ is the function that describes the distance
166     dependence of the valence electron distribution of atom $j$. The
167     contribution to the potential that comes from placing atom $i$ at that
168     location is then
169     \begin{equation*}
170     V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
171     \end{equation*}
172     where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
173 jmichalk 3866 $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
174     repulsive overlap of the two positively charged cores.
175 jmichalk 3807
176 gezelter 3826 % The {\it modified} embedded atom method (MEAM) adds angular terms to
177     % the electron density functions and an angular screening factor to the
178     % pairwise interaction between two
179     % atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
180     % MEAM has become widely used to simulate systems in which angular
181     % interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
182     % metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
183     % MEAM presents significant additional computational costs, however.
184 jmichalk 3807
185 jmichalk 3866 The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
186 gezelter 3808 have all been widely used by the materials simulation community for
187     simulations of bulk and nanoparticle
188 jmichalk 3885 properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq,mishin99:_inter}
189 gezelter 3808 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
190 jmichalk 3885 fracture,\cite{Shastry:1996qg,Shastry:1998dx,mishin01:cu} crack
191     propagation,\cite{BECQUART:1993rg,Rifkin1992} and alloying
192 gezelter 3887 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 gezelter 3808
205 gezelter 3826 \subsection{Carbon Monoxide model}
206 gezelter 3887 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 jmichalk 3812 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
218 jmichalk 3802 %CO Table
219     \begin{table}[H]
220 gezelter 3826 \caption{Positions, Lennard-Jones parameters ($\sigma$ and
221 gezelter 3887 $\epsilon$), 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 jmichalk 3802 \centering
227 jmichalk 3810 \begin{tabular}{| c | c | ccc |}
228 jmichalk 3802 \hline
229 jmichalk 3814 & {\it z} & $\sigma$ & $\epsilon$ & q\\
230 jmichalk 3802 \hline
231 jmichalk 3869 \textbf{C} & -0.6457 & 3.83 & 0.0262 & -0.75 \\
232     \textbf{O} & 0.4843 & 3.12 & 0.1591 & -0.85 \\
233 jmichalk 3814 \textbf{M} & 0.0 & - & - & 1.6 \\
234 jmichalk 3802 \hline
235     \end{tabular}
236 jmichalk 3866 \label{tab:CO}
237 jmichalk 3802 \end{table}
238 gezelter 3808
239 gezelter 3826 \subsection{Cross-Interactions between the metals and carbon monoxide}
240 jmichalk 3802
241 jmichalk 3867 Since the adsorption of CO onto a Pt surface has been the focus
242 gezelter 3826 of much experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
243     and theoretical work
244     \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
245     there is a significant amount of data on adsorption energies for CO on
246 jmichalk 3869 clean metal surfaces. An earlier model by Korzeniewski {\it et
247     al.}\cite{Pons:1986} served as a starting point for our fits. The parameters were
248 gezelter 3826 modified to ensure that the Pt-CO interaction favored the atop binding
249 jmichalk 3869 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 jmichalk 3866 than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
252 jmichalk 3878 {\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 gezelter 3826 binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
255 jmichalk 3869 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 jmichalk 3878 over O as the surface-binding atom. In most geometries, the Pt-O parameterization contributes a weak
258 gezelter 3826 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
261     position.\cite{Deshlahra:2012, Hopster:1978}
262 jmichalk 3811
263 jmichalk 3812 %where did you actually get the functionals for citation?
264     %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
265     %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
266 jmichalk 3866 The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
267 gezelter 3818 Morse potentials, respectively, to reproduce Au-CO binding energies.
268 jmichalk 3869 The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
269 jmichalk 3866 Adsorption energies were obtained from gas-surface DFT calculations with a
270 gezelter 3826 periodic supercell plane-wave basis approach, as implemented in the
271 gezelter 3887 Quantum ESPRESSO package.\cite{QE-2009} Electron cores were
272 gezelter 3818 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
275     computed with the PBE implementation of the generalized gradient
276     approximation (GGA) for gold, carbon, and oxygen that was constructed
277     by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
278 jmichalk 3866 In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
279 gezelter 3818 Au x 2 Au surface planes and separated from vertical images by six
280 jmichalk 3866 layers of vacuum space. The surface atoms were all allowed to relax
281     before CO was added to the system. Electronic relaxations were
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 gezelter 3875 zone.\cite{Monkhorst:1976} The relaxed gold slab was
286 gezelter 3826 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.
289 gezelter 3818
290 jmichalk 3812 %Hint at future work
291 jmichalk 3866 The parameters employed for the metal-CO cross-interactions in this work
292 jmichalk 3869 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 jmichalk 3878 and polarization are neglected in this model, although these effects could have
295 gezelter 3887 an effect on binding energies and binding site preferences.
296 jmichalk 3811
297 jmichalk 3802 %Table of Parameters
298     %Pt Parameter Set 9
299     %Au Parameter Set 35
300     \begin{table}[H]
301 gezelter 3887 \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 gezelter 3826 potentials. Distances are given in \AA~and energies in kcal/mol. }
305 jmichalk 3802 \centering
306     \begin{tabular}{| c | cc | c | ccc |}
307     \hline
308 gezelter 3826 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
309 jmichalk 3802 \hline
310     \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
311     \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
312    
313     \hline
314     \end{tabular}
315 jmichalk 3866 \label{tab:co_parameters}
316 jmichalk 3802 \end{table}
317    
318     %Table of energies
319     \begin{table}[H]
320 jmichalk 3869 \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
321 jmichalk 3867 described in this work. All values are in eV.}
322 jmichalk 3802 \centering
323     \begin{tabular}{| c | cc |}
324 gezelter 3826 \hline
325     & Calculated & Experimental \\
326     \hline
327 jmichalk 3890 \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.84} & -1.4 \bibpunct{}{}{,}{n}{}{,}
328 gezelter 3826 (Ref. \protect\cite{Kelemen:1979}) \\
329     & & -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
330 gezelter 3875 \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{TPDGold}) \\
331 gezelter 3826 \hline
332 jmichalk 3802 \end{tabular}
333 jmichalk 3866 \label{tab:co_energies}
334 jmichalk 3802 \end{table}
335    
336 jmichalk 3889
337     \subsection{Validation of forcefield selections}
338     By calculating minimum energies for commensurate systems of
339     single and double layer Pt and Au systems with 0 and 50\% coverages
340     (arranged in a c(2x4) pattern), our forcefield selections were able to be
341     indirectly compared to results shown in the supporting information of Tao
342     {\it et al.} \cite{Tao:2010}. Five layer thick systems, displaying a 557 facet
343     were constructed, each composed of 480 metal atoms. Double layers systems
344     were constructed from six layer thick systems where an entire layer was
345     removed from both displayed facets to create a double step. By design, the
346     double step system also contains 480 atoms, five layers thick, so energy
347     comparisons between the arrangements can be made directly. The positions
348     of the atoms were allowed to relax, along with the box sizes, before a
349     minimum energy was calculated. Carbon monoxide, equivalent to 50\%
350     coverage on one side of the metal system was added in a c(2x4) arrangement
351     and again allowed to relax before a minimum energy was calculated.
352    
353     Energies for the various systems are displayed in Table ~\ref{tab:steps}. Examining
354     the Pt systems first, it is apparent that the double layer system is slightly less stable
355     then the original single step. However, upon addition of carbon monoxide, the
356     stability is reversed and the double layer system becomes more stable. This result
357     is in agreement with DFT calculations in Tao {\it et al.}\cite{Tao:2010}, who also show
358     that the addition of CO leads to a reversal in the most stable system. While our
359     results agree qualitatively, quantitatively, they are approximately an order of magnitude
360     different. Looking at additional stability per atom in kcal/mol, the DFT calculations suggest
361     an increased stability of 0.1 kcal/mol per Pt atom, whereas we are seeing closer to a 0.4 kcal/mol
362     increase in stability per Pt atom.
363    
364     The gold systems show a much smaller energy difference between the single and double
365     systems, likely arising from their lower energy per atom values. Additionally, the weaker
366     binding of CO to Au is evidenced by the much smaller energy change between the two systems,
367     when compared to the Pt results. This limited change helps explain our lack of any reconstruction
368     on the Au systems.
369    
370    
371     %Table of single step double step calculations
372     \begin{table}[H]
373 jmichalk 3890 \caption{Minimized single point energies of unit cell crystals displaying (S)ingle or (D)double steps. Systems are periodic along and perpendicular to the step-edge axes with a large vacuum above the displayed 557 facet. The relative energies are calculated as $E_{relative} = E_{system} - E_{M-557-S} - N_{CO}\Delta E_{CO-M}$ , where $E_{CO-M}$ is -1.84 eV for Pt-CO and -0.39 eV for Pt-CO. The addition of CO in a 50\% c(2x4) coverage acts as a stabilizing presence and suggests a driving force for the observed reconstruction on the highest coverage Pt system. All energies are in kcal/mol.}
374 jmichalk 3889 \centering
375 jmichalk 3890 \begin{tabular}{| c | c | c | c | c | c |}
376 jmichalk 3889 \hline
377 jmichalk 3890 \textbf{Step} & \textbf{N}\textsubscript{M} & \textbf{N\textsubscript{CO}} & \textbf{Relative Energy} & \textbf{$\Delta$E/M} & \textbf{$\Delta$E/CO} \\
378 jmichalk 3889 \hline
379 jmichalk 3890 Pt(557)-S & 480 & 0 & 0 & 0 & - \\
380     Pt(557)-D & 480 & 0 & 114.783 & 0.239 & -\\
381     Pt(557)-S & 480 & 40 & -124.546 & -0.259 & -3.114\\
382     Pt(557)-D & 480 & 44 & -34.953 & -0.073 & -0.794\\
383 jmichalk 3889 \hline
384     \hline
385 jmichalk 3890 Au(557)-S & 480 & 0 & 0 & 0 & - \\
386     Au(557)-D & 480 & 0 & 79.572 & 0.166 & - \\
387     Au(557)-S & 480 & 40 & -157.199 & -0.327 & -3.930\\
388     Au(557)-D & 480 & 44 & -123.297 & -0.257 & -2.802 \\
389 jmichalk 3889 \hline
390     \end{tabular}
391     \label{tab:steps}
392     \end{table}
393    
394    
395 gezelter 3826 \subsection{Pt(557) and Au(557) metal interfaces}
396 jmichalk 3872 Our Pt system is an orthorhombic periodic box of dimensions
397     54.482~x~50.046~x~120.88~\AA~while our Au system has
398 jmichalk 3878 dimensions of 57.4~x~51.9285~x~100~\AA. The metal slabs
399     are 9 and 8 atoms deep respectively, corresponding to a slab
400     thickness of $\sim$21~\AA~ for Pt and $\sim$19~\AA~for Au.
401 jmichalk 3870 The systems are arranged in a FCC crystal that have been cut
402     along the (557) plane so that they are periodic in the {\it x} and
403     {\it y} directions, and have been oriented to expose two aligned
404     (557) cuts along the extended {\it z}-axis. Simulations of the
405     bare metal interfaces at temperatures ranging from 300~K to
406 jmichalk 3872 1200~K were performed to confirm the relative
407 gezelter 3826 stability of the surfaces without a CO overlayer.
408 jmichalk 3802
409 gezelter 3887 The different bulk melting temperatures predicted by EAM
410     (1345~$\pm$~10~K for Au\cite{Au:melting} and $\sim$~2045~K for
411     Pt\cite{Pt:melting}) suggest that any reconstructions should happen at
412     different temperatures for the two metals. The bare Au and Pt
413     surfaces were initially run in the canonical (NVT) ensemble at 800~K
414     and 1000~K respectively for 100 ps. The two surfaces were relatively
415     stable at these temperatures when no CO was present, but experienced
416     increased surface mobility on addition of CO. Each surface was then
417     dosed with different concentrations of CO that was initially placed in
418     the vacuum region. Upon full adsorption, these concentrations
419     correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface coverage. Higher
420     coverages resulted in the formation of a double layer of CO, which
421     introduces artifacts that are not relevant to (557) reconstruction.
422     Because of the difference in binding energies, nearly all of the CO
423     was bound to the Pt surface, while the Au surfaces often had a
424     significant CO population in the gas phase. These systems were
425     allowed to reach thermal equilibrium (over 5~ns) before being run in
426     the microcanonical (NVE) ensemble for data collection. All of the
427     systems examined had at least 40~ns in the data collection stage,
428     although simulation times for some Pt of the systems exceeded 200~ns.
429     Simulations were carried out using the open source molecular dynamics
430     package, OpenMD.\cite{Ewald,OOPSE,openmd}
431 jmichalk 3802
432 jmichalk 3872
433     % RESULTS
434     %
435 jmichalk 3802 \section{Results}
436 jmichalk 3860 \subsection{Structural remodeling}
437 gezelter 3887 The bare metal surfaces experienced minor roughening of the step-edge
438     because of the elevated temperatures, but the (557) face was stable
439     throughout the simulations. The surfaces of both systems, upon dosage
440     of CO, began to undergo extensive remodeling that was not observed in
441     the bare systems. Reconstructions of the Au systems were limited to
442     breakup of the step-edges and some step wandering. The lower coverage
443     Pt systems experienced similar step edge wandering but to a greater
444     extent. The 50\% coverage Pt system was unique among our simulations
445     in that it formed well-defined and stable double layers through step
446     coalescence, similar to results reported by Tao {\it et
447     al}.\cite{Tao:2010}
448 jmichalk 3872
449 jmichalk 3871 \subsubsection{Step wandering}
450 gezelter 3887 The bare surfaces for both metals showed minimal step-wandering at
451     their respective temperatures. As the CO coverage increased however,
452     the mobility of the surface atoms, described through adatom diffusion
453     and step-edge wandering, also increased. Except for the 50\% Pt
454     system where step coalescence occurred, the step-edges in the other
455     simulations preferred to keep nearly the same distance between steps
456     as in the original (557) lattice, $\sim$13\AA~for Pt and
457     $\sim$14\AA~for Au. Previous work by Williams {\it et
458     al}.\cite{Williams:1991, Williams:1994} highlights the repulsion
459     that exists between step-edges even when no direct interactions are
460     present in the system. This repulsion is caused by an entropic barrier
461     that arises from the fact that steps cannot cross over one
462     another. This entropic repulsion does not completely define the
463     interactions between steps, however, so it is possible to observe step
464     coalescence on some surfaces.\cite{Williams:1991} The presence and
465     concentration of adsorbates, as shown in this work, can affect
466     step-step interactions, potentially leading to a new surface structure
467     as the thermodynamic equilibrium.
468 jmichalk 3872
469 jmichalk 3871 \subsubsection{Double layers}
470 gezelter 3887 Tao {\it et al}.\cite{Tao:2010} have shown experimentally that the
471     Pt(557) surface undergoes two separate reconstructions upon CO
472     adsorption. The first involves a doubling of the step height and
473     plateau length. Similar behavior has been seen on a number of
474     surfaces at varying conditions, including Ni(977) and
475     Si(111).\cite{Williams:1994,Williams:1991,Pearl} Of the two systems we
476     examined, the Pt system showed a greater propensity for reconstruction
477     because of the larger surface mobility and the greater extent of step
478     wandering. The amount of reconstruction was strongly correlated to
479     the amount of CO adsorbed upon the surface. This appears to be
480     related to the effect that adsorbate coverage has on edge breakup and
481     on the surface diffusion of metal adatoms. Only the 50\% Pt surface
482     underwent the doubling seen by Tao {\it et al}.\cite{Tao:2010} within
483     the time scales studied here. Over a longer time scale (150~ns) two
484     more double layers formed on this surface. Although double layer
485     formation did not occur in the other Pt systems, they exhibited more
486     step-wandering and roughening compared to their Au counterparts. The
487     50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
488     various times along the simulation showing the evolution of a double
489     layer step-edge.
490 jmichalk 3802
491 gezelter 3887 The second reconstruction observed by Tao {\it et al}.\cite{Tao:2010}
492     involved the formation of triangular clusters that stretched across
493     the plateau between two step-edges. Neither of the simulated metal
494     interfaces, within the 40~ns time scale or the extended time of 150~ns
495     for the 50\% Pt system, experienced this reconstruction.
496 jmichalk 3817
497 jmichalk 3876 %Evolution of surface
498     \begin{figure}[H]
499 gezelter 3882 \includegraphics[width=\linewidth]{EPS_ProgressionOfDoubleLayerFormation}
500 gezelter 3887 \caption{The Pt(557) / 50\% CO interface upon exposure to the CO: (a)
501     258~ps, (b) 19~ns, (c) 31.2~ns, and (d) 86.1~ns after
502     exposure. Disruption of the (557) step-edges occurs quickly. The
503 jmichalk 3876 doubling of the layers appears only after two adjacent step-edges
504     touch. The circled spot in (b) nucleated the growth of the double
505     step observed in the later configurations.}
506     \label{fig:reconstruct}
507     \end{figure}
508    
509 jmichalk 3860 \subsection{Dynamics}
510 gezelter 3887 Previous experimental work by Pearl and Sibener\cite{Pearl}, using
511     STM, has been able to capture the coalescence of steps on Ni(977). The
512     time scale of the image acquisition, $\sim$70~s/image, provides an
513     upper bound for the time required for the doubling to occur. By
514     utilizing Molecular Dynamics we are able to probe the dynamics of
515     these reconstructions at elevated temperatures and in this section we
516     provide data on the timescales for transport properties,
517     e.g. diffusion and layer formation time.
518 gezelter 3826
519 jmichalk 3867
520 jmichalk 3860 \subsubsection{Transport of surface metal atoms}
521 jmichalk 3862 %forcedSystems/stepSeparation
522 gezelter 3826
523 gezelter 3887 The wandering of a step-edge is a cooperative effect arising from the
524     individual movements of the atoms making up the steps. An ideal metal
525     surface displaying a low index facet, (111) or (100), is unlikely to
526     experience much surface diffusion because of the large energetic
527     barrier that must be overcome to lift an atom out of the surface. The
528     presence of step-edges and other surface features on higher-index
529     facets provides a lower energy source for mobile metal atoms. Using
530     our potential model, single-atom break-away from a step-edge on a
531     clean surface still imposes an energetic penalty around
532     $\sim$~45~kcal/mol, but this is certainly easier than lifting the same
533     metal atom vertically out of the surface, \textgreater~60~kcal/mol.
534     The penalty lowers significantly when CO is present in sufficient
535     quantities on the surface. For certain distributions of CO, the
536     energetic penalty can fall to as low as $\sim$~20~kcal/mol. The
537     configurations that create these lower barriers are detailed in the
538     discussion section below.
539 gezelter 3826
540 gezelter 3887 Once an adatom exists on the surface, the barrier for diffusion is
541     negligible (\textless~4~kcal/mol for a Pt adatom). These adatoms are
542     then able to explore the terrace before rejoining either their
543     original step-edge or becoming a part of a different edge. It is an
544     energetically unfavorable process with a high barrier for an atom to
545     traverse to a separate terrace although the presence of CO can lower
546     the energy barrier required to lift or lower an adatom. By tracking
547     the mobility of individual metal atoms on the Pt and Au surfaces we
548     were able to determine the relative diffusion constants, as well as
549     how varying coverages of CO affect the diffusion. Close observation of
550     the mobile metal atoms showed that they were typically in equilibrium
551     with the step-edges. At times, their motion was concerted, and two or
552     more adatoms would be observed moving together across the surfaces.
553    
554     A particle was considered ``mobile'' once it had traveled more than
555     2~\AA~ between saved configurations of the system (typically 10-100
556     ps). A mobile atom would typically travel much greater distances than
557     this, but the 2~\AA~cutoff was used to prevent swamping the diffusion
558     data with the in-place vibrational movement of buried atoms. Diffusion
559     on a surface is strongly affected by local structures and the presence
560     of single and double layer step-edges causes the diffusion parallel to
561     the step-edges to be larger than the diffusion perpendicular to these
562     edges. Parallel and perpendicular diffusion constants are shown in
563     Figure \ref{fig:diff}. Diffusion parallel to the step-edge is higher
564     than diffusion perpendicular to the edge because of the lower energy
565     barrier associated with sliding along an edge compared to breaking
566     away to form an isolated adatom.
567    
568 jmichalk 3876 %Diffusion graph
569     \begin{figure}[H]
570 gezelter 3882 \includegraphics[width=\linewidth]{Portrait_DiffusionComparison_1}
571 jmichalk 3876 \caption{Diffusion constants for mobile surface atoms along directions
572     parallel ($\mathbf{D}_{\parallel}$) and perpendicular
573     ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
574 gezelter 3887 surface coverage. The two reported diffusion constants for the 50\%
575     Pt system correspond to a 20~ns period before the formation of the
576     double layer (upper points), and to the full 40~ns sampling period
577     (lower points).}
578 jmichalk 3876 \label{fig:diff}
579     \end{figure}
580    
581 jmichalk 3878 The weaker Au-CO interaction is evident in the weak CO-coverage
582     dependance of Au diffusion. This weak interaction leads to lower
583     observed coverages when compared to dosage amounts. This further
584     limits the effect the CO can have on surface diffusion. The correlation
585     between coverage and Pt diffusion rates shows a near linear relationship
586     at the earliest times in the simulations. Following double layer formation,
587     however, there is a precipitous drop in adatom diffusion. As the double
588     layer forms, many atoms that had been tracked for mobility data have
589 gezelter 3887 now been buried, resulting in a smaller reported diffusion constant. A
590 jmichalk 3878 secondary effect of higher coverages is CO-CO cross interactions that
591     lower the effective mobility of the Pt adatoms that are bound to each CO.
592     This effect would become evident only at higher coverages. A detailed
593     account of Pt adatom energetics follows in the Discussion.
594    
595     \subsubsection{Dynamics of double layer formation}
596     The increased diffusion on Pt at the higher CO coverages is the primary
597     contributor to double layer formation. However, this is not a complete
598     explanation -- the 33\%~Pt system has higher diffusion constants, but
599     did not show any signs of edge doubling in 40~ns. On the 50\%~Pt
600     system, one double layer formed within the first 40~ns of simulation time,
601     while two more were formed as the system was allowed to run for an
602     additional 110~ns (150~ns total). This suggests that this reconstruction
603     is a rapid process and that the previously mentioned upper bound is a
604     very large overestimate.\cite{Williams:1991,Pearl} In this system the first
605     appearance of a double layer appears at 19~ns into the simulation.
606     Within 12~ns of this nucleation event, nearly half of the step has formed
607     the double layer and by 86~ns the complete layer has flattened out.
608     From the appearance of the first nucleation event to the first observed
609     double layer, the process took $\sim$20~ns. Another $\sim$40~ns was
610     necessary for the layer to completely straighten. The other two layers in
611     this simulation formed over periods of 22~ns and 42~ns respectively.
612     A possible explanation for this rapid reconstruction is the elevated
613     temperatures under which our systems were simulated. The process
614     would almost certainly take longer at lower temperatures. Additionally,
615     our measured times for completion of the doubling after the appearance
616     of a nucleation site are likely affected by our periodic boxes. A longer
617     step-edge will likely take longer to ``zipper''.
618 jmichalk 3876
619    
620 jmichalk 3878 %Discussion
621     \section{Discussion}
622 gezelter 3882 We have shown that a classical potential is able to model the initial
623     reconstruction of the Pt(557) surface upon CO adsorption, and have
624     reproduced the double layer structure observed by Tao {\it et
625     al}.\cite{Tao:2010}. Additionally, this reconstruction appears to be
626     rapid -- occurring within 100 ns of the initial exposure to CO. Here
627     we discuss the features of the classical potential that are
628     contributing to the stability and speed of the Pt(557) reconstruction.
629 jmichalk 3817
630 jmichalk 3878 \subsection{Diffusion}
631 gezelter 3882 The perpendicular diffusion constant appears to be the most important
632     indicator of double layer formation. As highlighted in Figure
633     \ref{fig:reconstruct}, the formation of the double layer did not begin
634     until a nucleation site appeared. Williams {\it et
635     al}.\cite{Williams:1991,Williams:1994} cite an effective edge-edge
636     repulsion arising from the inability of edge crossing. This repulsion
637     must be overcome to allow step coalescence. A larger
638     $\textbf{D}_\perp$ value implies more step-wandering and a larger
639     chance for the stochastic meeting of two edges to create a nucleation
640     point. Diffusion parallel to the step-edge can help ``zipper'' up a
641     nascent double layer. This helps explain the rapid time scale for
642     double layer completion after the appearance of a nucleation site, while
643     the initial appearance of the nucleation site was unpredictable.
644 jmichalk 3876
645 jmichalk 3878 \subsection{Mechanism for restructuring}
646 gezelter 3882 Since the Au surface showed no large scale restructuring in any of our
647     simulations, our discussion will focus on the 50\% Pt-CO system which
648     did exhibit doubling. A number of possible mechanisms exist to explain
649     the role of adsorbed CO in restructuring the Pt surface. Quadrupolar
650     repulsion between adjacent CO molecules adsorbed on the surface is one
651     possibility. However, the quadrupole-quadrupole interaction is
652     short-ranged and is attractive for some orientations. If the CO
653     molecules are ``locked'' in a vertical orientation, through atop
654 gezelter 3887 adsorption for example, this explanation would gain credence. Within
655     the framework of our classical potential, the calculated energetic
656     repulsion between two CO molecules located a distance of
657     2.77~\AA~apart (nearest-neighbor distance of Pt) and both in a
658     vertical orientation, is 8.62 kcal/mol. Moving the CO to the second
659     nearest-neighbor distance of 4.8~\AA~drops the repulsion to nearly
660     0. Allowing the CO to rotate away from a purely vertical orientation
661     also lowers the repulsion. When the carbons are locked at a distance
662     of 2.77~\AA, a minimum of 6.2 kcal/mol is reached when the angle
663     between the 2 CO is $\sim$24\textsuperscript{o}. The calculated
664 gezelter 3882 barrier for surface diffusion of a Pt adatom is only 4 kcal/mol, so
665 gezelter 3887 repulsion between adjacent CO molecules bound to Pt could indeed
666     increase the surface diffusion. However, the residence time of CO on
667     Pt suggests that the CO molecules are extremely mobile, with diffusion
668     constants 40 to 2500 times larger than surface Pt atoms. This mobility
669     suggests that the CO molecules jump between different Pt atoms
670     throughout the simulation. However, they do stay bound to individual
671     Pt atoms for long enough to modify the local energy landscape for the
672     mobile adatoms.
673 jmichalk 3876
674 gezelter 3882 A different interpretation of the above mechanism which takes the
675     large mobility of the CO into account, would be in the destabilization
676     of Pt-Pt interactions due to bound CO. Destabilizing Pt-Pt bonds at
677     the edges could lead to increased step-edge breakup and diffusion. On
678     the bare Pt(557) surface the barrier to completely detach an edge atom
679     is $\sim$43~kcal/mol, as is shown in configuration (a) in Figures
680     \ref{fig:SketchGraphic} \& \ref{fig:SketchEnergies}. For certain
681     configurations, cases (e), (g), and (h), the barrier can be lowered to
682     $\sim$23~kcal/mol by the presence of bound CO molecules. In these
683     instances, it becomes energetically favorable to roughen the edge by
684     introducing a small separation of 0.5 to 1.0~\AA. This roughening
685     becomes immediately obvious in simulations with significant CO
686     populations. The roughening is present to a lesser extent on surfaces
687     with lower CO coverage (and even on the bare surfaces), although in
688     these cases it is likely due to random fluctuations that squeeze out
689 gezelter 3887 step-edge atoms. Step-edge breakup by direct single-atom translations
690     (as suggested by these energy curves) is probably a worst-case
691     scenario. Multistep mechanisms in which an adatom moves laterally on
692     the surface after being ejected would be more energetically favorable.
693     This would leave the adatom alongside the ledge, providing it with
694     five nearest neighbors. While fewer than the seven neighbors it had
695     as part of the step-edge, it keeps more Pt neighbors than the three
696     neighbors an isolated adatom has on the terrace. In this proposed
697     mechanism, the CO quadrupolar repulsion still plays a role in the
698     initial roughening of the step-edge, but not in any long-term bonds
699     with individual Pt atoms. Higher CO coverages create more
700 gezelter 3882 opportunities for the crowded CO configurations shown in Figure
701     \ref{fig:SketchGraphic}, and this is likely to cause an increased
702     propensity for step-edge breakup.
703 jmichalk 3876
704     %Sketch graphic of different configurations
705 jmichalk 3816 \begin{figure}[H]
706 gezelter 3882 \includegraphics[width=\linewidth]{COpaths}
707     \caption{Configurations used to investigate the mechanism of step-edge
708 gezelter 3887 breakup on Pt(557). In each case, the central (starred) atom was
709 gezelter 3882 pulled directly across the surface away from the step edge. The Pt
710     atoms on the upper terrace are colored dark grey, while those on the
711     lower terrace are in white. In each of these configurations, some
712 gezelter 3887 of the atoms (highlighted in blue) had CO molecules bound in the
713     vertical atop position. The energies of these configurations as a
714 gezelter 3882 function of central atom displacement are displayed in Figure
715     \ref{fig:SketchEnergies}.}
716 jmichalk 3876 \label{fig:SketchGraphic}
717 jmichalk 3862 \end{figure}
718    
719 jmichalk 3876 %energy graph corresponding to sketch graphic
720 jmichalk 3862 \begin{figure}[H]
721 gezelter 3882 \includegraphics[width=\linewidth]{Portrait_SeparationComparison}
722     \caption{Energies for displacing a single edge atom perpendicular to
723     the step edge as a function of atomic displacement. Each of the
724     energy curves corresponds to one of the labeled configurations in
725 gezelter 3887 Figure \ref{fig:SketchGraphic}, and the energies are referenced to
726     the unperturbed step-edge. Certain arrangements of bound CO
727     (notably configurations g and h) can lower the energetic barrier for
728     creating an adatom relative to the bare surface (configuration a).}
729 jmichalk 3876 \label{fig:SketchEnergies}
730 jmichalk 3816 \end{figure}
731    
732 gezelter 3882 While configurations of CO on the surface are able to increase
733     diffusion and the likelihood of edge wandering, this does not provide
734     a complete explanation for the formation of double layers. If adatoms
735     were constrained to their original terraces then doubling could not
736     occur. A mechanism for vertical displacement of adatoms at the
737     step-edge is required to explain the doubling.
738 jmichalk 3802
739 gezelter 3882 We have discovered one possible mechanism for a CO-mediated vertical
740     displacement of Pt atoms at the step edge. Figure \ref{fig:lambda}
741     shows four points along a reaction coordinate in which a CO-bound
742     adatom along the step-edge ``burrows'' into the edge and displaces the
743 gezelter 3887 original edge atom onto the higher terrace. A number of events
744     similar to this mechanism were observed during the simulations. We
745     predict an energetic barrier of 20~kcal/mol for this process (in which
746     the displaced edge atom follows a curvilinear path into an adjacent
747     3-fold hollow site). The barrier heights we obtain for this reaction
748 gezelter 3882 coordinate are approximate because the exact path is unknown, but the
749     calculated energy barriers would be easily accessible at operating
750     conditions. Additionally, this mechanism is exothermic, with a final
751     energy 15~kcal/mol below the original $\lambda = 0$ configuration.
752     When CO is not present and this reaction coordinate is followed, the
753 gezelter 3887 process is endothermic by 3~kcal/mol. The difference in the relative
754 gezelter 3882 energies for the $\lambda=0$ and $\lambda=1$ case when CO is present
755     provides strong support for CO-mediated Pt-Pt interactions giving rise
756 gezelter 3887 to the doubling reconstruction.
757 gezelter 3882
758 jmichalk 3862 %lambda progression of Pt -> shoving its way into the step
759     \begin{figure}[H]
760 gezelter 3882 \includegraphics[width=\linewidth]{EPS_rxnCoord}
761     \caption{Points along a possible reaction coordinate for CO-mediated
762     edge doubling. Here, a CO-bound adatom burrows into an established
763     step edge and displaces an edge atom onto the upper terrace along a
764     curvilinear path. The approximate barrier for the process is
765     20~kcal/mol, and the complete process is exothermic by 15~kcal/mol
766 gezelter 3887 in the presence of CO, but is endothermic by 3~kcal/mol without CO.}
767 jmichalk 3862 \label{fig:lambda}
768     \end{figure}
769    
770 gezelter 3882 The mechanism for doubling on the Pt(557) surface appears to require
771     the cooperation of at least two distinct processes. For complete
772     doubling of a layer to occur there must be a breakup of one
773     terrace. These atoms must then ``disappear'' from that terrace, either
774 gezelter 3887 by travelling to the terraces above or below their original levels.
775 gezelter 3882 The presence of CO helps explain mechanisms for both of these
776     situations. There must be sufficient breakage of the step-edge to
777     increase the concentration of adatoms on the surface and these adatoms
778     must then undergo the burrowing highlighted above (or a comparable
779     mechanism) to create the double layer. With sufficient time, these
780     mechanisms working in concert lead to the formation of a double layer.
781 jmichalk 3879
782 jmichalk 3878 \subsection{CO Removal and double layer stability}
783 gezelter 3887 Once the double layers had formed on the 50\%~Pt system, they remained
784     stable for the rest of the simulation time with minimal movement.
785     Random fluctuations that involved small clusters or divots were
786     observed, but these features typically healed within a few
787     nanoseconds. Within our simulations, the formation of the double
788     layer appeared to be irreversible and a double layer was never
789     observed to split back into two single layer step-edges while CO was
790     present.
791 jmichalk 3862
792 gezelter 3882 To further gauge the effect CO has on this surface, additional
793     simulations were run starting from a late configuration of the 50\%~Pt
794     system that had already formed double layers. These simulations then
795 gezelter 3887 had their CO molecules suddenly removed. The double layer broke apart
796     rapidly in these simulations, showing a well-defined edge-splitting
797     after 100~ps. Configurations of this system are shown in Figure
798 gezelter 3882 \ref{fig:breaking}. The coloring of the top and bottom layers helps to
799 gezelter 3887 show how much mixing the edges experience as they split. These systems
800     were only examined for 10~ns, and within that time despite the initial
801     rapid splitting, the edges only moved another few \AA~apart. It is
802     possible that with longer simulation times, the (557) surface recovery
803     observed by Tao {\it et al}.\cite{Tao:2010} could also be recovered.
804 jmichalk 3862
805     %breaking of the double layer upon removal of CO
806 jmichalk 3802 \begin{figure}[H]
807 gezelter 3882 \includegraphics[width=\linewidth]{EPS_doubleLayerBreaking}
808 gezelter 3887 \caption{Behavior of an established (111) double step after removal of
809     the adsorbed CO: (A) 0~ps, (B) 100~ps, and (C) 1~ns after the
810     removal of CO. Nearly immediately after the CO is removed, the
811     step edge reforms in a (100) configuration, which is also the step
812     type seen on clean (557) surfaces. The step separation involves
813 gezelter 3882 significant mixing of the lower and upper atoms at the edge.}
814 jmichalk 3862 \label{fig:breaking}
815 jmichalk 3802 \end{figure}
816    
817    
818     %Peaks!
819 jmichalk 3872 %\begin{figure}[H]
820     %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
821     %\caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
822     %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
823     %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
824     %\label{fig:peaks}
825     %\end{figure}
826 jmichalk 3862
827 jmichalk 3867
828     %Don't think I need this
829 jmichalk 3862 %clean surface...
830 jmichalk 3867 %\begin{figure}[H]
831 gezelter 3882 %\includegraphics[width=\linewidth]{557_300K_cleanPDF}
832 jmichalk 3867 %\caption{}
833 jmichalk 3862
834 jmichalk 3867 %\end{figure}
835     %\label{fig:clean}
836    
837    
838 jmichalk 3802 \section{Conclusion}
839 gezelter 3882 The strength and directionality of the Pt-CO binding interaction, as
840     well as the large quadrupolar repulsion between atop-bound CO
841     molecules, help to explain the observed increase in surface mobility
842     of Pt(557) and the resultant reconstruction into a double-layer
843     configuration at the highest simulated CO-coverages. The weaker Au-CO
844     interaction results in significantly lower adataom diffusion
845     constants, less step-wandering, and a lack of the double layer
846     reconstruction on the Au(557) surface.
847 jmichalk 3802
848 gezelter 3882 An in-depth examination of the energetics shows the important role CO
849     plays in increasing step-breakup and in facilitating edge traversal
850     which are both necessary for double layer formation.
851 jmichalk 3880
852 jmichalk 3862 %Things I am not ready to remove yet
853    
854     %Table of Diffusion Constants
855     %Add gold?M
856     % \begin{table}[H]
857     % \caption{}
858     % \centering
859     % \begin{tabular}{| c | cc | cc | }
860     % \hline
861     % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
862     % \hline
863     % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
864     % \hline
865     % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
866     % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
867     % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
868     % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
869     % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
870     % \hline
871     % \end{tabular}
872     % \end{table}
873    
874 gezelter 3875 \begin{acknowledgement}
875 gezelter 3882 We gratefully acknowledge conversations with Dr. William
876     F. Schneider and Dr. Feng Tao. Support for this project was
877     provided by the National Science Foundation under grant CHE-0848243
878     and by the Center for Sustainable Energy at Notre Dame
879     (cSEND). Computational time was provided by the Center for Research
880     Computing (CRC) at the University of Notre Dame.
881 gezelter 3875 \end{acknowledgement}
882 gezelter 3808 \newpage
883 gezelter 3887 \bibstyle{achemso}
884     \bibliography{COonPtAu}
885 gezelter 3875 %\end{doublespace}
886    
887     \begin{tocentry}
888 gezelter 3887 \begin{wrapfigure}{l}{0.5\textwidth}
889     \begin{center}
890     \includegraphics[width=\linewidth]{TOC_doubleLayer}
891     \end{center}
892     \end{wrapfigure}
893     A reconstructed Pt(557) surface after 86~ns exposure to a half a
894     monolayer of CO. The double layer that forms is a result of
895     CO-mediated step-edge wandering as well as a burrowing mechanism that
896     helps lift edge atoms onto an upper terrace.
897 gezelter 3875 \end{tocentry}
898    
899 gezelter 3808 \end{document}