ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3896
Committed: Thu Jun 13 18:56:15 2013 UTC (11 years, 2 months ago) by gezelter
Content type: application/x-tex
File size: 48858 byte(s)
Log Message:
Final edits

File Contents

# Content
1 \documentclass[journal = jpccck, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3 \usepackage{achemso}
4 \usepackage{natbib}
5 \usepackage{multirow}
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}
12
13 \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 \begin{document}
27
28
29 %%
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
48 \begin{abstract}
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
67
68
69 \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 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 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
81 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
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 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
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 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{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
147 \subsection{Metal-metal interactions}
148 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 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 {\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*}
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 $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
174 repulsive overlap of the two positively charged cores.
175
176 % 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
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,mishin99:_inter}
189 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
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
205 \subsection{Carbon Monoxide model}
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 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
229 & {\it z} & $\sigma$ & $\epsilon$ & q\\
230 \hline
231 \textbf{C} & -0.6457 & 3.83 & 0.0262 & -0.75 \\
232 \textbf{O} & 0.4843 & 3.12 & 0.1591 & -0.85 \\
233 \textbf{M} & 0.0 & - & - & 1.6 \\
234 \hline
235 \end{tabular}
236 \label{tab:CO}
237 \end{table}
238
239 \subsection{Cross-Interactions between the metals and carbon monoxide}
240
241 Since the adsorption of CO onto a Pt surface has been the focus
242 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 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 modified to ensure that the Pt-CO interaction favored the atop binding
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 {\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 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
261 position.\cite{Deshlahra:2012, Hopster:1978}
262
263 %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 The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
267 Morse potentials, respectively, to reproduce Au-CO binding energies.
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 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
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 In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
279 Au x 2 Au surface planes and separated from vertical images by six
280 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 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.
289
290 %Hint at future work
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 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{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 |}
307 \hline
308 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
309 \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 \label{tab:co_parameters}
316 \end{table}
317
318 %Table of energies
319 \begin{table}[H]
320 \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
321 described in this work. All values are in eV.}
322 \centering
323 \begin{tabular}{| c | cc |}
324 \hline
325 & Calculated & Experimental \\
326 \hline
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{TPDGold}) \\
331 \hline
332 \end{tabular}
333 \label{tab:co_energies}
334 \end{table}
335
336
337 \subsection{Force field validation}
338 The CO-Pt cross interactions were compared directly to DFT results
339 found in the supporting information of reference
340 {\bibpunct{}{}{,}{n}{}{,} \protect\cite{Tao:2010}}. These energies are
341 estimates of the degree of stabilization provided to double-layer
342 reconstructions of the M(557) surface by an overlayer of CO molecules
343 in a $c (2 \times 4)$ pattern. To make the comparison, five atom
344 thick metal slabs of both Pt and Au displaying the (557) facet were
345 constructed. Double-layer (reconstructed) systems were created using
346 six atomic layers where enough of a layer was removed from both
347 exposed (557) facets to create the double step. In all cases, the
348 metal slabs contained 480 atoms and were minimized using steepest
349 descent under the EAM force field. Both the bare metal slabs and slabs
350 with 50\% carbon monoxide coverage (arranged in the $c (2 \times 4)$
351 pattern) were used. The systems are periodic along and perpendicular
352 to the step-edge axes with a large vacuum above the displayed (557)
353 facet.
354
355 Energies computed using our force field are displayed in Table
356 ~\ref{tab:steps}. The relative energies are calculated as
357 $E_{relative} = E_{system} - E_{M(557)-S} - N_{CO}*E_{M-CO}(r)$, where
358 $E_{M(557)-S}$ is the energy of a clean (557) surface. $N_{CO}$ is the
359 number of CO molecules present on the surface. In the $c (2 \times
360 4)$ patterning, the CO molecules relax to an average separation, $r$,
361 from the nearest surface metal atom. $E_{M-CO}(r)$ is taken as the
362 energy of a single CO molecule on a flat M(111) surface at a distance
363 $r$ from a metal atop site. These energies correspond to -1.8 eV for
364 CO-Pt and -0.39 eV for CO-Au.
365
366 One important note is that the $c (2 \times 4)$ patterning on the
367 stepped surfaces yields a slightly larger M-CO separation than one
368 would find on a clean (111) surface. On a clean Pt(111) surface, for
369 example, the optimized geometry has a C-Pt distance of 1.53~\AA
370 (corresponding to a binding energy of -1.83 eV). On the double-layer
371 reconstruction and the single (557) step, the half monolayer optimizes
372 to C-Pt separations of 1.58-1.60~\AA, respectively. Although this
373 difference seems quite small, there are notable consequences for
374 $E_{Pt-CO}(r)$ which then takes values from -1.815 eV to -1.8 eV.
375
376 For platinum, the bare double layer reconstruction is less stable than
377 the bare (557) step by about 0.25 kcal/mol per Pt atom. However,
378 addition of carbon monoxide changes the relative energetics of the two
379 systems. This is a quite dramatic shift, $\Delta\Delta E$ (the change
380 in energy for going from single to double-layer structures upon
381 addition of a CO layer) shifts by -0.5~kcal/mol per Pt atom. This
382 result is in qualitative agreement with the DFT calculations in
383 reference {\bibpunct{}{}{,}{n}{}{,} \protect\cite{Tao:2010}}, which
384 also showed that the addition of CO leads to a reversal in stability.
385
386 The gold systems show a smaller energy difference between the clean
387 single and double layers. Upon addition of CO, the single step surface
388 is much more stable than the double-layer reconstruction. However,
389 the CO-Au binding energy is much weaker, so at operating temperatures,
390 the actual coverage by CO will be much lower than the 50\% coverage
391 afforded by the $c (2 \times 4)$ pattern, so single-point energy
392 comparisons are not as helpful.
393
394 %Table of single step double step calculations
395 \begin{table}[H]
396 \caption{Relative energies (in kcal/mol) of (S)ingle M(557) and
397 (D)ouble-step reconstructions. 50\% coverage by CO in a $c(2
398 \times 4)$ pattern stabilizes the D-reconstructed Pt(557)
399 surface, but leaves the single-step Au(557) as the more stable structure.}
400 \centering
401 \begin{tabular}{| c | c | c | c | c |}
402 \hline
403 Step & $N_{M}$ & $N_{CO}$ & Relative Energy & $\Delta E / N_{M}$ \\
404 \hline
405 Pt(557)-S & 480 & 0 & 0 & 0 \\
406 Pt(557)-D & 480 & 0 & 119.788 & 0.2495 \\
407 Pt(557)-S & 480 & 40 & -109.734 & -0.2286 \\
408 Pt(557)-D & 480 & 48 & -110.039 & -0.2292 \\
409 \hline
410 \hline
411 Au(557)-S & 480 & 0 & 0 & 0 \\
412 Au(557)-D & 480 & 0 & 83.853 & 0.1747 \\
413 Au(557)-S & 480 & 40 & -253.604 & -0.5283 \\
414 Au(557)-D & 480 & 48 & -156.150 & -0.3253 \\
415 \hline
416 \end{tabular}
417 \label{tab:steps}
418 \end{table}
419
420 Qualitatively, our classical force field for the metal-CO cross
421 interactions reproduces the results predicted by DFT studies in
422 reference {\bibpunct{}{}{,}{n}{}{,} \protect\cite{Tao:2010}}. Addition
423 of polarization effects, both in the CO and in the metal surfaces,
424 could make the model significantly more accurate. For example,
425 because of the relatively large fixed charges, the current model will
426 be unable to reproduce coverages in excess of 50\% without forming an
427 inverted CO second layer on the surface. The M-CO cross interactions
428 would also be more accurate if they included the direct interactions
429 between charges on the CO and their image charges inside the metal
430 slab. These polarization effects have been shown to play an important
431 role,\cite{Deshlahra:2012} and would be one way of improving the
432 numerical agreement with quantum mechanical calculations.
433
434 \subsection{Pt(557) and Au(557) metal interfaces}
435 Our Pt system is an orthorhombic periodic box of dimensions
436 54.482~x~50.046~x~120.88~\AA~while our Au system has
437 dimensions of 57.4~x~51.9285~x~100~\AA. The metal slabs
438 are 9 and 8 atoms deep respectively, corresponding to a slab
439 thickness of $\sim$21~\AA~ for Pt and $\sim$19~\AA~for Au.
440 The systems are arranged in a FCC crystal that have been cut
441 along the (557) plane so that they are periodic in the {\it x} and
442 {\it y} directions, and have been oriented to expose two aligned
443 (557) cuts along the extended {\it z}-axis. Simulations of the
444 bare metal interfaces at temperatures ranging from 300~K to
445 1200~K were performed to confirm the relative
446 stability of the surfaces without a CO overlayer.
447
448 The different bulk melting temperatures predicted by EAM
449 (1345~$\pm$~10~K for Au\cite{Au:melting} and $\sim$~2045~K for
450 Pt\cite{Pt:melting}) suggest that any reconstructions should happen at
451 different temperatures for the two metals. The bare Au and Pt
452 surfaces were initially run in the canonical (NVT) ensemble at 800~K
453 and 1000~K respectively for 100 ps. The two surfaces were relatively
454 stable at these temperatures when no CO was present, but experienced
455 increased surface mobility on addition of CO. Each surface was then
456 dosed with different concentrations of CO that was initially placed in
457 the vacuum region. Upon full adsorption, these concentrations
458 correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface coverage. Higher
459 coverages resulted in the formation of a double layer of CO, which
460 introduces artifacts that are not relevant to (557) reconstruction.
461 Because of the difference in binding energies, nearly all of the CO
462 was bound to the Pt surface, while the Au surfaces often had a
463 significant CO population in the gas phase. These systems were
464 allowed to reach thermal equilibrium (over 5~ns) before being run in
465 the microcanonical (NVE) ensemble for data collection. All of the
466 systems examined had at least 40~ns in the data collection stage,
467 although simulation times for some Pt of the systems exceeded 200~ns.
468 Simulations were carried out using the open source molecular dynamics
469 package, OpenMD.\cite{Ewald,OOPSE,openmd}
470
471
472 % RESULTS
473 %
474 \section{Results}
475 \subsection{Structural remodeling}
476 The bare metal surfaces experienced minor roughening of the step-edge
477 because of the elevated temperatures, but the (557) face was stable
478 throughout the simulations. The surfaces of both systems, upon dosage
479 of CO, began to undergo extensive remodeling that was not observed in
480 the bare systems. Reconstructions of the Au systems were limited to
481 breakup of the step-edges and some step wandering. The lower coverage
482 Pt systems experienced similar step edge wandering but to a greater
483 extent. The 50\% coverage Pt system was unique among our simulations
484 in that it formed well-defined and stable double layers through step
485 coalescence, similar to results reported by Tao {\it et
486 al}.\cite{Tao:2010}
487
488 \subsubsection{Step wandering}
489 The bare surfaces for both metals showed minimal step-wandering at
490 their respective temperatures. As the CO coverage increased however,
491 the mobility of the surface atoms, described through adatom diffusion
492 and step-edge wandering, also increased. Except for the 50\% Pt
493 system where step coalescence occurred, the step-edges in the other
494 simulations preferred to keep nearly the same distance between steps
495 as in the original (557) lattice, $\sim$13\AA~for Pt and
496 $\sim$14\AA~for Au. Previous work by Williams {\it et
497 al}.\cite{Williams:1991, Williams:1994} highlights the repulsion
498 that exists between step-edges even when no direct interactions are
499 present in the system. This repulsion is caused by an entropic barrier
500 that arises from the fact that steps cannot cross over one
501 another. This entropic repulsion does not completely define the
502 interactions between steps, however, so it is possible to observe step
503 coalescence on some surfaces.\cite{Williams:1991} The presence and
504 concentration of adsorbates, as shown in this work, can affect
505 step-step interactions, potentially leading to a new surface structure
506 as the thermodynamic equilibrium.
507
508 \subsubsection{Double layers}
509 Tao {\it et al}.\cite{Tao:2010} have shown experimentally that the
510 Pt(557) surface undergoes two separate reconstructions upon CO
511 adsorption. The first involves a doubling of the step height and
512 plateau length. Similar behavior has been seen on a number of
513 surfaces at varying conditions, including Ni(977) and
514 Si(111).\cite{Williams:1994,Williams:1991,Pearl} Of the two systems we
515 examined, the Pt system showed a greater propensity for reconstruction
516 because of the larger surface mobility and the greater extent of step
517 wandering. The amount of reconstruction was strongly correlated to
518 the amount of CO adsorbed upon the surface. This appears to be
519 related to the effect that adsorbate coverage has on edge breakup and
520 on the surface diffusion of metal adatoms. Only the 50\% Pt surface
521 underwent the doubling seen by Tao {\it et al}.\cite{Tao:2010} within
522 the time scales studied here. Over a longer time scale (150~ns) two
523 more double layers formed on this surface. Although double layer
524 formation did not occur in the other Pt systems, they exhibited more
525 step-wandering and roughening compared to their Au counterparts. The
526 50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
527 various times along the simulation showing the evolution of a double
528 layer step-edge.
529
530 The second reconstruction observed by Tao {\it et al}.\cite{Tao:2010}
531 involved the formation of triangular clusters that stretched across
532 the plateau between two step-edges. Neither of the simulated metal
533 interfaces, within the 40~ns time scale or the extended time of 150~ns
534 for the 50\% Pt system, experienced this reconstruction.
535
536 %Evolution of surface
537 \begin{figure}[H]
538 \includegraphics[width=\linewidth]{EPS_ProgressionOfDoubleLayerFormation}
539 \caption{The Pt(557) / 50\% CO interface upon exposure to the CO: (a)
540 258~ps, (b) 19~ns, (c) 31.2~ns, and (d) 86.1~ns after
541 exposure. Disruption of the (557) step-edges occurs quickly. The
542 doubling of the layers appears only after two adjacent step-edges
543 touch. The circled spot in (b) nucleated the growth of the double
544 step observed in the later configurations.}
545 \label{fig:reconstruct}
546 \end{figure}
547
548 \subsection{Dynamics}
549 Previous experimental work by Pearl and Sibener\cite{Pearl}, using
550 STM, has been able to capture the coalescence of steps on Ni(977). The
551 time scale of the image acquisition, $\sim$70~s/image, provides an
552 upper bound for the time required for the doubling to occur. By
553 utilizing Molecular Dynamics we are able to probe the dynamics of
554 these reconstructions at elevated temperatures and in this section we
555 provide data on the timescales for transport properties,
556 e.g. diffusion and layer formation time.
557
558
559 \subsubsection{Transport of surface metal atoms}
560 %forcedSystems/stepSeparation
561
562 The wandering of a step-edge is a cooperative effect arising from the
563 individual movements of the atoms making up the steps. An ideal metal
564 surface displaying a low index facet, (111) or (100), is unlikely to
565 experience much surface diffusion because of the large energetic
566 barrier that must be overcome to lift an atom out of the surface. The
567 presence of step-edges and other surface features on higher-index
568 facets provides a lower energy source for mobile metal atoms. Using
569 our potential model, single-atom break-away from a step-edge on a
570 clean surface still imposes an energetic penalty around
571 $\sim$~45~kcal/mol, but this is certainly easier than lifting the same
572 metal atom vertically out of the surface, \textgreater~60~kcal/mol.
573 The penalty lowers significantly when CO is present in sufficient
574 quantities on the surface. For certain distributions of CO, the
575 energetic penalty can fall to as low as $\sim$~20~kcal/mol. The
576 configurations that create these lower barriers are detailed in the
577 discussion section below.
578
579 Once an adatom exists on the surface, the barrier for diffusion is
580 negligible (\textless~4~kcal/mol for a Pt adatom). These adatoms are
581 then able to explore the terrace before rejoining either their
582 original step-edge or becoming a part of a different edge. It is an
583 energetically unfavorable process with a high barrier for an atom to
584 traverse to a separate terrace although the presence of CO can lower
585 the energy barrier required to lift or lower an adatom. By tracking
586 the mobility of individual metal atoms on the Pt and Au surfaces we
587 were able to determine the relative diffusion constants, as well as
588 how varying coverages of CO affect the diffusion. Close observation of
589 the mobile metal atoms showed that they were typically in equilibrium
590 with the step-edges. At times, their motion was concerted, and two or
591 more adatoms would be observed moving together across the surfaces.
592
593 A particle was considered ``mobile'' once it had traveled more than
594 2~\AA~ between saved configurations of the system (typically 10-100
595 ps). A mobile atom would typically travel much greater distances than
596 this, but the 2~\AA~cutoff was used to prevent swamping the diffusion
597 data with the in-place vibrational movement of buried atoms. Diffusion
598 on a surface is strongly affected by local structures and the presence
599 of single and double layer step-edges causes the diffusion parallel to
600 the step-edges to be larger than the diffusion perpendicular to these
601 edges. Parallel and perpendicular diffusion constants are shown in
602 Figure \ref{fig:diff}. Diffusion parallel to the step-edge is higher
603 than diffusion perpendicular to the edge because of the lower energy
604 barrier associated with sliding along an edge compared to breaking
605 away to form an isolated adatom.
606
607 %Diffusion graph
608 \begin{figure}[H]
609 \includegraphics[width=\linewidth]{Portrait_DiffusionComparison_1}
610 \caption{Diffusion constants for mobile surface atoms along directions
611 parallel ($\mathbf{D}_{\parallel}$) and perpendicular
612 ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
613 surface coverage. The two reported diffusion constants for the 50\%
614 Pt system correspond to a 20~ns period before the formation of the
615 double layer (upper points), and to the full 40~ns sampling period
616 (lower points).}
617 \label{fig:diff}
618 \end{figure}
619
620 The weaker Au-CO interaction is evident in the weak CO-coverage
621 dependance of Au diffusion. This weak interaction leads to lower
622 observed coverages when compared to dosage amounts. This further
623 limits the effect the CO can have on surface diffusion. The correlation
624 between coverage and Pt diffusion rates shows a near linear relationship
625 at the earliest times in the simulations. Following double layer formation,
626 however, there is a precipitous drop in adatom diffusion. As the double
627 layer forms, many atoms that had been tracked for mobility data have
628 now been buried, resulting in a smaller reported diffusion constant. A
629 secondary effect of higher coverages is CO-CO cross interactions that
630 lower the effective mobility of the Pt adatoms that are bound to each CO.
631 This effect would become evident only at higher coverages. A detailed
632 account of Pt adatom energetics follows in the Discussion.
633
634 \subsubsection{Dynamics of double layer formation}
635 The increased diffusion on Pt at the higher CO coverages is the primary
636 contributor to double layer formation. However, this is not a complete
637 explanation -- the 33\%~Pt system has higher diffusion constants, but
638 did not show any signs of edge doubling in 40~ns. On the 50\%~Pt
639 system, one double layer formed within the first 40~ns of simulation time,
640 while two more were formed as the system was allowed to run for an
641 additional 110~ns (150~ns total). This suggests that this reconstruction
642 is a rapid process and that the previously mentioned upper bound is a
643 very large overestimate.\cite{Williams:1991,Pearl} In this system the first
644 appearance of a double layer appears at 19~ns into the simulation.
645 Within 12~ns of this nucleation event, nearly half of the step has formed
646 the double layer and by 86~ns the complete layer has flattened out.
647 From the appearance of the first nucleation event to the first observed
648 double layer, the process took $\sim$20~ns. Another $\sim$40~ns was
649 necessary for the layer to completely straighten. The other two layers in
650 this simulation formed over periods of 22~ns and 42~ns respectively.
651 A possible explanation for this rapid reconstruction is the elevated
652 temperatures under which our systems were simulated. The process
653 would almost certainly take longer at lower temperatures. Additionally,
654 our measured times for completion of the doubling after the appearance
655 of a nucleation site are likely affected by our periodic boxes. A longer
656 step-edge will likely take longer to ``zipper''.
657
658
659 %Discussion
660 \section{Discussion}
661 We have shown that a classical potential is able to model the initial
662 reconstruction of the Pt(557) surface upon CO adsorption, and have
663 reproduced the double layer structure observed by Tao {\it et
664 al}.\cite{Tao:2010}. Additionally, this reconstruction appears to be
665 rapid -- occurring within 100 ns of the initial exposure to CO. Here
666 we discuss the features of the classical potential that are
667 contributing to the stability and speed of the Pt(557) reconstruction.
668
669 \subsection{Diffusion}
670 The perpendicular diffusion constant appears to be the most important
671 indicator of double layer formation. As highlighted in Figure
672 \ref{fig:reconstruct}, the formation of the double layer did not begin
673 until a nucleation site appeared. Williams {\it et
674 al}.\cite{Williams:1991,Williams:1994} cite an effective edge-edge
675 repulsion arising from the inability of edge crossing. This repulsion
676 must be overcome to allow step coalescence. A larger
677 $\textbf{D}_\perp$ value implies more step-wandering and a larger
678 chance for the stochastic meeting of two edges to create a nucleation
679 point. Diffusion parallel to the step-edge can help ``zipper'' up a
680 nascent double layer. This helps explain the rapid time scale for
681 double layer completion after the appearance of a nucleation site, while
682 the initial appearance of the nucleation site was unpredictable.
683
684 \subsection{Mechanism for restructuring}
685 Since the Au surface showed no large scale restructuring in any of our
686 simulations, our discussion will focus on the 50\% Pt-CO system which
687 did exhibit doubling. A number of possible mechanisms exist to explain
688 the role of adsorbed CO in restructuring the Pt surface. Quadrupolar
689 repulsion between adjacent CO molecules adsorbed on the surface is one
690 possibility. However, the quadrupole-quadrupole interaction is
691 short-ranged and is attractive for some orientations. If the CO
692 molecules are ``locked'' in a vertical orientation, through atop
693 adsorption for example, this explanation would gain credence. Within
694 the framework of our classical potential, the calculated energetic
695 repulsion between two CO molecules located a distance of
696 2.77~\AA~apart (nearest-neighbor distance of Pt) and both in a
697 vertical orientation, is 8.62 kcal/mol. Moving the CO to the second
698 nearest-neighbor distance of 4.8~\AA~drops the repulsion to nearly
699 0. Allowing the CO to rotate away from a purely vertical orientation
700 also lowers the repulsion. When the carbons are locked at a distance
701 of 2.77~\AA, a minimum of 6.2 kcal/mol is reached when the angle
702 between the 2 CO is $\sim$24\textsuperscript{o}. The calculated
703 barrier for surface diffusion of a Pt adatom is only 4 kcal/mol, so
704 repulsion between adjacent CO molecules bound to Pt could indeed
705 increase the surface diffusion. However, the residence time of CO on
706 Pt suggests that the CO molecules are extremely mobile, with diffusion
707 constants 40 to 2500 times larger than surface Pt atoms. This mobility
708 suggests that the CO molecules jump between different Pt atoms
709 throughout the simulation. However, they do stay bound to individual
710 Pt atoms for long enough to modify the local energy landscape for the
711 mobile adatoms.
712
713 A different interpretation of the above mechanism which takes the
714 large mobility of the CO into account, would be in the destabilization
715 of Pt-Pt interactions due to bound CO. Destabilizing Pt-Pt bonds at
716 the edges could lead to increased step-edge breakup and diffusion. On
717 the bare Pt(557) surface the barrier to completely detach an edge atom
718 is $\sim$43~kcal/mol, as is shown in configuration (a) in Figures
719 \ref{fig:SketchGraphic} \& \ref{fig:SketchEnergies}. For certain
720 configurations, cases (e), (g), and (h), the barrier can be lowered to
721 $\sim$23~kcal/mol by the presence of bound CO molecules. In these
722 instances, it becomes energetically favorable to roughen the edge by
723 introducing a small separation of 0.5 to 1.0~\AA. This roughening
724 becomes immediately obvious in simulations with significant CO
725 populations. The roughening is present to a lesser extent on surfaces
726 with lower CO coverage (and even on the bare surfaces), although in
727 these cases it is likely due to random fluctuations that squeeze out
728 step-edge atoms. Step-edge breakup by direct single-atom translations
729 (as suggested by these energy curves) is probably a worst-case
730 scenario. Multistep mechanisms in which an adatom moves laterally on
731 the surface after being ejected would be more energetically favorable.
732 This would leave the adatom alongside the ledge, providing it with
733 five nearest neighbors. While fewer than the seven neighbors it had
734 as part of the step-edge, it keeps more Pt neighbors than the three
735 neighbors an isolated adatom has on the terrace. In this proposed
736 mechanism, the CO quadrupolar repulsion still plays a role in the
737 initial roughening of the step-edge, but not in any long-term bonds
738 with individual Pt atoms. Higher CO coverages create more
739 opportunities for the crowded CO configurations shown in Figure
740 \ref{fig:SketchGraphic}, and this is likely to cause an increased
741 propensity for step-edge breakup.
742
743 %Sketch graphic of different configurations
744 \begin{figure}[H]
745 \includegraphics[width=\linewidth]{COpaths}
746 \caption{Configurations used to investigate the mechanism of step-edge
747 breakup on Pt(557). In each case, the central (starred) atom was
748 pulled directly across the surface away from the step edge. The Pt
749 atoms on the upper terrace are colored dark grey, while those on the
750 lower terrace are in white. In each of these configurations, some
751 of the atoms (highlighted in blue) had CO molecules bound in the
752 vertical atop position. The energies of these configurations as a
753 function of central atom displacement are displayed in Figure
754 \ref{fig:SketchEnergies}.}
755 \label{fig:SketchGraphic}
756 \end{figure}
757
758 %energy graph corresponding to sketch graphic
759 \begin{figure}[H]
760 \includegraphics[width=\linewidth]{Portrait_SeparationComparison}
761 \caption{Energies for displacing a single edge atom perpendicular to
762 the step edge as a function of atomic displacement. Each of the
763 energy curves corresponds to one of the labeled configurations in
764 Figure \ref{fig:SketchGraphic}, and the energies are referenced to
765 the unperturbed step-edge. Certain arrangements of bound CO
766 (notably configurations g and h) can lower the energetic barrier for
767 creating an adatom relative to the bare surface (configuration a).}
768 \label{fig:SketchEnergies}
769 \end{figure}
770
771 While configurations of CO on the surface are able to increase
772 diffusion and the likelihood of edge wandering, this does not provide
773 a complete explanation for the formation of double layers. If adatoms
774 were constrained to their original terraces then doubling could not
775 occur. A mechanism for vertical displacement of adatoms at the
776 step-edge is required to explain the doubling.
777
778 We have discovered one possible mechanism for a CO-mediated vertical
779 displacement of Pt atoms at the step edge. Figure \ref{fig:lambda}
780 shows four points along a reaction coordinate in which a CO-bound
781 adatom along the step-edge ``burrows'' into the edge and displaces the
782 original edge atom onto the higher terrace. A number of events
783 similar to this mechanism were observed during the simulations. We
784 predict an energetic barrier of 20~kcal/mol for this process (in which
785 the displaced edge atom follows a curvilinear path into an adjacent
786 3-fold hollow site). The barrier heights we obtain for this reaction
787 coordinate are approximate because the exact path is unknown, but the
788 calculated energy barriers would be easily accessible at operating
789 conditions. Additionally, this mechanism is exothermic, with a final
790 energy 15~kcal/mol below the original $\lambda = 0$ configuration.
791 When CO is not present and this reaction coordinate is followed, the
792 process is endothermic by 3~kcal/mol. The difference in the relative
793 energies for the $\lambda=0$ and $\lambda=1$ case when CO is present
794 provides strong support for CO-mediated Pt-Pt interactions giving rise
795 to the doubling reconstruction.
796
797 %lambda progression of Pt -> shoving its way into the step
798 \begin{figure}[H]
799 \includegraphics[width=\linewidth]{EPS_rxnCoord}
800 \caption{Points along a possible reaction coordinate for CO-mediated
801 edge doubling. Here, a CO-bound adatom burrows into an established
802 step edge and displaces an edge atom onto the upper terrace along a
803 curvilinear path. The approximate barrier for the process is
804 20~kcal/mol, and the complete process is exothermic by 15~kcal/mol
805 in the presence of CO, but is endothermic by 3~kcal/mol without CO.}
806 \label{fig:lambda}
807 \end{figure}
808
809 The mechanism for doubling on the Pt(557) surface appears to require
810 the cooperation of at least two distinct processes. For complete
811 doubling of a layer to occur there must be a breakup of one
812 terrace. These atoms must then ``disappear'' from that terrace, either
813 by travelling to the terraces above or below their original levels.
814 The presence of CO helps explain mechanisms for both of these
815 situations. There must be sufficient breakage of the step-edge to
816 increase the concentration of adatoms on the surface and these adatoms
817 must then undergo the burrowing highlighted above (or a comparable
818 mechanism) to create the double layer. With sufficient time, these
819 mechanisms working in concert lead to the formation of a double layer.
820
821 \subsection{CO Removal and double layer stability}
822 Once the double layers had formed on the 50\%~Pt system, they remained
823 stable for the rest of the simulation time with minimal movement.
824 Random fluctuations that involved small clusters or divots were
825 observed, but these features typically healed within a few
826 nanoseconds. Within our simulations, the formation of the double
827 layer appeared to be irreversible and a double layer was never
828 observed to split back into two single layer step-edges while CO was
829 present.
830
831 To further gauge the effect CO has on this surface, additional
832 simulations were run starting from a late configuration of the 50\%~Pt
833 system that had already formed double layers. These simulations then
834 had their CO molecules suddenly removed. The double layer broke apart
835 rapidly in these simulations, showing a well-defined edge-splitting
836 after 100~ps. Configurations of this system are shown in Figure
837 \ref{fig:breaking}. The coloring of the top and bottom layers helps to
838 show how much mixing the edges experience as they split. These systems
839 were only examined for 10~ns, and within that time despite the initial
840 rapid splitting, the edges only moved another few \AA~apart. It is
841 possible that with longer simulation times, the (557) surface recovery
842 observed by Tao {\it et al}.\cite{Tao:2010} could also be recovered.
843
844 %breaking of the double layer upon removal of CO
845 \begin{figure}[H]
846 \includegraphics[width=\linewidth]{EPS_doubleLayerBreaking}
847 \caption{Behavior of an established (111) double step after removal of
848 the adsorbed CO: (A) 0~ps, (B) 100~ps, and (C) 1~ns after the
849 removal of CO. Nearly immediately after the CO is removed, the
850 step edge reforms in a (100) configuration, which is also the step
851 type seen on clean (557) surfaces. The step separation involves
852 significant mixing of the lower and upper atoms at the edge.}
853 \label{fig:breaking}
854 \end{figure}
855
856
857 %Peaks!
858 %\begin{figure}[H]
859 %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
860 %\caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
861 %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
862 %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
863 %\label{fig:peaks}
864 %\end{figure}
865
866
867 %Don't think I need this
868 %clean surface...
869 %\begin{figure}[H]
870 %\includegraphics[width=\linewidth]{557_300K_cleanPDF}
871 %\caption{}
872
873 %\end{figure}
874 %\label{fig:clean}
875
876
877 \section{Conclusion}
878 The strength and directionality of the Pt-CO binding interaction, as
879 well as the large quadrupolar repulsion between atop-bound CO
880 molecules, help to explain the observed increase in surface mobility
881 of Pt(557) and the resultant reconstruction into a double-layer
882 configuration at the highest simulated CO-coverages. The weaker Au-CO
883 interaction results in significantly lower adataom diffusion
884 constants, less step-wandering, and a lack of the double layer
885 reconstruction on the Au(557) surface.
886
887 An in-depth examination of the energetics shows the important role CO
888 plays in increasing step-breakup and in facilitating edge traversal
889 which are both necessary for double layer formation.
890
891 %Things I am not ready to remove yet
892
893 %Table of Diffusion Constants
894 %Add gold?M
895 % \begin{table}[H]
896 % \caption{}
897 % \centering
898 % \begin{tabular}{| c | cc | cc | }
899 % \hline
900 % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
901 % \hline
902 % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
903 % \hline
904 % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
905 % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
906 % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
907 % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
908 % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
909 % \hline
910 % \end{tabular}
911 % \end{table}
912
913 \begin{acknowledgement}
914 We gratefully acknowledge conversations with Dr. William
915 F. Schneider and Dr. Feng Tao. Support for this project was
916 provided by the National Science Foundation under grant CHE-0848243
917 and by the Center for Sustainable Energy at Notre Dame
918 (cSEND). Computational time was provided by the Center for Research
919 Computing (CRC) at the University of Notre Dame.
920 \end{acknowledgement}
921 \newpage
922 \bibstyle{achemso}
923 \bibliography{COonPtAu}
924 %\end{doublespace}
925
926 \begin{tocentry}
927 \begin{wrapfigure}{l}{0.5\textwidth}
928 \begin{center}
929 \includegraphics[width=\linewidth]{TOC_doubleLayer}
930 \end{center}
931 \end{wrapfigure}
932 A reconstructed Pt(557) surface after 86~ns exposure to a half a
933 monolayer of CO. The double layer that forms is a result of
934 CO-mediated step-edge wandering as well as a burrowing mechanism that
935 helps lift edge atoms onto an upper terrace.
936 \end{tocentry}
937
938 \end{document}