ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3887
Committed: Thu Mar 21 15:28:49 2013 UTC (11 years, 5 months ago) by gezelter
Content type: application/x-tex
File size: 43742 byte(s)
Log Message:
Final version

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 gezelter 3887 %\mciteErrorOnUnknownfalse
8 gezelter 3875
9     \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
10 gezelter 3808 \usepackage{url}
11 jmichalk 3802
12 gezelter 3875 \title{Molecular Dynamics simulations of the surface reconstructions
13     of Pt(557) and Au(557) under exposure to CO}
14    
15     \author{Joseph R. Michalka}
16     \author{Patrick W. McIntyre}
17     \author{J. Daniel Gezelter}
18     \email{gezelter@nd.edu}
19     \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
20     Department of Chemistry and Biochemistry\\ University of Notre
21     Dame\\ Notre Dame, Indiana 46556}
22    
23     \keywords{}
24    
25 gezelter 3808 \begin{document}
26    
27 gezelter 3875
28 jmichalk 3802 %%
29     %Introduction
30     % Experimental observations
31     % Previous work on Pt, CO, etc.
32     %
33     %Simulation Methodology
34     % FF (fits and parameters)
35     % MD (setup, equilibration, collection)
36     %
37     % Analysis of trajectories!!!
38     %Discussion
39     % CO preferences for specific locales
40     % CO-CO interactions
41     % Differences between Au & Pt
42     % Causes of 2_layer reordering in Pt
43     %Summary
44     %%
45    
46 gezelter 3818
47 gezelter 3808 \begin{abstract}
48 gezelter 3882 The mechanism and dynamics of surface reconstructions of Pt(557) and
49     Au(557) exposed to various coverages of carbon monoxide (CO) were
50 gezelter 3887 investigated using molecular dynamics simulations. Metal-CO
51 gezelter 3882 interactions were parameterized from experimental data and
52     plane-wave Density Functional Theory (DFT) calculations. The large
53     difference in binding strengths of the Pt-CO and Au-CO interactions
54     was found to play a significant role in step-edge stability and
55     adatom diffusion constants. Various mechanisms for CO-mediated step
56     wandering and step doubling were investigated on the Pt(557)
57     surface. We find that the energetics of CO adsorbed to the surface
58     can explain the step-doubling reconstruction observed on Pt(557) and
59 gezelter 3887 the lack of such a reconstruction on the Au(557) surface. However,
60     more complicated reconstructions into triangular clusters that have
61     been seen in recent experiments were not observed in these
62     simulations.
63 gezelter 3808 \end{abstract}
64 jmichalk 3802
65 gezelter 3808 \newpage
66    
67    
68 jmichalk 3802 \section{Introduction}
69     % Importance: catalytically active metals are important
70     % Sub: Knowledge of how their surface structure affects their ability to catalytically facilitate certain reactions is growing, but is more reactionary than predictive
71     % 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)
72     % Theory can explore temperatures and pressures which are difficult to work with in experiments
73     % Sub: Also, easier to observe what is going on and provide reasons and explanations
74     %
75    
76 gezelter 3826 Industrial catalysts usually consist of small particles that exhibit a
77     high concentration of steps, kink sites, and vacancies at the edges of
78     the facets. These sites are thought to be the locations of catalytic
79 gezelter 3808 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
80 gezelter 3826 significant evidence that solid surfaces are often structurally,
81     compositionally, and chemically modified by reactants under operating
82     conditions.\cite{Tao2008,Tao:2010,Tao2011} The coupling between
83     surface oxidation states and catalytic activity for CO oxidation on
84     Pt, for instance, is widely documented.\cite{Ertl08,Hendriksen:2002}
85     Despite the well-documented role of these effects on reactivity, the
86     ability to capture or predict them in atomistic models is somewhat
87     limited. While these effects are perhaps unsurprising on the highly
88     disperse, multi-faceted nanoscale particles that characterize
89     industrial catalysts, they are manifest even on ordered, well-defined
90     surfaces. The Pt(557) surface, for example, exhibits substantial and
91     reversible restructuring under exposure to moderate pressures of
92     carbon monoxide.\cite{Tao:2010}
93 jmichalk 3802
94 gezelter 3887 This work is an investigation into the mechanism and timescale for the
95     Pt(557) \& Au(557) surface restructuring using molecular simulation.
96     Since the dynamics of the process are of particular interest, we
97     employ classical force fields that represent a compromise between
98     chemical accuracy and the computational efficiency necessary to
99     simulate the process of interest. Since restructuring typically
100     occurs as a result of specific interactions of the catalyst with
101     adsorbates, in this work, two metal systems exposed to carbon monoxide
102     were examined. The Pt(557) surface has already been shown to undergo a
103     large scale reconstruction under certain conditions.\cite{Tao:2010}
104     The Au(557) surface, because of weaker interactions with CO, is less
105     likely to undergo this kind of reconstruction. However, Peters {\it et
106     al}.\cite{Peters:2000} and Piccolo {\it et al}.\cite{Piccolo:2004}
107     have both observed CO-induced modification of reconstructions to the
108     Au(111) surface. Peters {\it et al}. observed the Au(111)-($22 \times
109     \sqrt{3}$) ``herringbone'' reconstruction relaxing slightly under CO
110     adsorption. They argued that only a few Au atoms become adatoms,
111     limiting the stress of this reconstruction, while allowing the rest to
112     relax and approach the ideal (111) configuration. Piccolo {\it et
113     al}. on the other hand, saw a more significant disruption of the
114     Au(111)-($22 \times \sqrt{3}$) herringbone pattern as CO adsorbed on
115     the surface. Both groups suggested that the preference CO shows for
116     low-coordinated Au atoms was the primary driving force for the
117     relaxation. Although the Au(111) reconstruction was not the primary
118     goal of our work, the classical models we have fit may be of future
119     use in simulating this reconstruction.
120 gezelter 3826
121 jmichalk 3811 %Platinum molecular dynamics
122     %gold molecular dynamics
123 jmichalk 3802
124     \section{Simulation Methods}
125 gezelter 3887 The challenge in modeling any solid/gas interface is the development
126     of a sufficiently general yet computationally tractable model of the
127     chemical interactions between the surface atoms and adsorbates. Since
128     the interfaces involved are quite large (10$^3$ - 10$^4$ atoms), have
129     many electrons, and respond slowly to perturbations, {\it ab initio}
130 gezelter 3808 molecular dynamics
131     (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
132     methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
133     mechanical potential energy surfaces remain out of reach.
134     Additionally, the ``bonds'' between metal atoms at a surface are
135     typically not well represented in terms of classical pairwise
136     interactions in the same way that bonds in a molecular material are,
137     nor are they captured by simple non-directional interactions like the
138 gezelter 3826 Coulomb potential. For this work, we have used classical molecular
139     dynamics with potential energy surfaces that are specifically tuned
140     for transition metals. In particular, we used the EAM potential for
141 gezelter 3887 Au-Au and Pt-Pt interactions.\cite{Foiles86} The CO was modeled using
142     a rigid three-site model developed by Straub and Karplus for studying
143 gezelter 3826 photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
144     Pt-CO cross interactions were parameterized as part of this work.
145 gezelter 3808
146     \subsection{Metal-metal interactions}
147 gezelter 3826 Many of the potentials used for modeling transition metals are based
148     on a non-pairwise additive functional of the local electron
149     density. The embedded atom method (EAM) is perhaps the best known of
150     these
151 gezelter 3808 methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
152     but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
153     the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
154 gezelter 3887 parameter sets. The glue model of Ercolessi {\it et
155     al}.\cite{Ercolessi88} is among the fastest of these density
156     functional approaches. In all of these models, atoms are treated as a
157     positively charged core with a radially-decaying valence electron
158     distribution. To calculate the energy for embedding the core at a
159     particular location, the electron density due to the valence electrons
160     at all of the other atomic sites is computed at atom $i$'s location,
161 gezelter 3808 \begin{equation*}
162     \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
163     \end{equation*}
164     Here, $\rho_j(r_{ij})$ is the function that describes the distance
165     dependence of the valence electron distribution of atom $j$. The
166     contribution to the potential that comes from placing atom $i$ at that
167     location is then
168     \begin{equation*}
169     V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
170     \end{equation*}
171     where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
172 jmichalk 3866 $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
173     repulsive overlap of the two positively charged cores.
174 jmichalk 3807
175 gezelter 3826 % The {\it modified} embedded atom method (MEAM) adds angular terms to
176     % the electron density functions and an angular screening factor to the
177     % pairwise interaction between two
178     % atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
179     % MEAM has become widely used to simulate systems in which angular
180     % interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
181     % metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
182     % MEAM presents significant additional computational costs, however.
183 jmichalk 3807
184 jmichalk 3866 The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
185 gezelter 3808 have all been widely used by the materials simulation community for
186     simulations of bulk and nanoparticle
187 jmichalk 3885 properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq,mishin99:_inter}
188 gezelter 3808 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
189 jmichalk 3885 fracture,\cite{Shastry:1996qg,Shastry:1998dx,mishin01:cu} crack
190     propagation,\cite{BECQUART:1993rg,Rifkin1992} and alloying
191 gezelter 3887 dynamics.\cite{Shibata:2002hh,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
192     One of EAM's strengths is its sensitivity to small changes in
193     structure. This is due to the inclusion of up to the third nearest
194     neighbor interactions during fitting of the parameters.\cite{Voter95a}
195     In comparison, the glue model of Ercolessi {\it et
196     al}.\cite{Ercolessi88} was only parameterized to include
197     nearest-neighbor interactions, EAM is a suitable choice for systems
198     where the bulk properties are of secondary importance to low-index
199     surface structures. Additionally, the similarity of EAM's functional
200     treatment of the embedding energy to standard density functional
201     theory (DFT) makes fitting DFT-derived cross potentials with
202     adsorbates somewhat easier.
203 gezelter 3808
204 gezelter 3826 \subsection{Carbon Monoxide model}
205 gezelter 3887 Previous explanations for the surface rearrangements center on the
206     large linear quadrupole moment of carbon monoxide.\cite{Tao:2010} We
207     used a model first proposed by Karplus and Straub to study the
208     photodissociation of CO from myoglobin because it reproduces the
209     quadrupole moment well.\cite{Straub} The Straub and Karplus model
210     treats CO as a rigid three site molecule with a massless
211     charge-carrying ``M'' site at the center of mass. The geometry and
212     interaction parameters are reproduced in Table~\ref{tab:CO}. The
213     effective dipole moment, calculated from the assigned charges, is
214     still small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is
215     close to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
216 jmichalk 3812 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
217 jmichalk 3802 %CO Table
218     \begin{table}[H]
219 gezelter 3826 \caption{Positions, Lennard-Jones parameters ($\sigma$ and
220 gezelter 3887 $\epsilon$), and charges for CO-CO
221     interactions. Distances are in \AA, energies are
222     in kcal/mol, and charges are in atomic units. The CO model
223     from Ref.\bibpunct{}{}{,}{n}{}{,}
224     \protect\cite{Straub} was used without modification.}
225 jmichalk 3802 \centering
226 jmichalk 3810 \begin{tabular}{| c | c | ccc |}
227 jmichalk 3802 \hline
228 jmichalk 3814 & {\it z} & $\sigma$ & $\epsilon$ & q\\
229 jmichalk 3802 \hline
230 jmichalk 3869 \textbf{C} & -0.6457 & 3.83 & 0.0262 & -0.75 \\
231     \textbf{O} & 0.4843 & 3.12 & 0.1591 & -0.85 \\
232 jmichalk 3814 \textbf{M} & 0.0 & - & - & 1.6 \\
233 jmichalk 3802 \hline
234     \end{tabular}
235 jmichalk 3866 \label{tab:CO}
236 jmichalk 3802 \end{table}
237 gezelter 3808
238 gezelter 3826 \subsection{Cross-Interactions between the metals and carbon monoxide}
239 jmichalk 3802
240 jmichalk 3867 Since the adsorption of CO onto a Pt surface has been the focus
241 gezelter 3826 of much experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
242     and theoretical work
243     \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
244     there is a significant amount of data on adsorption energies for CO on
245 jmichalk 3869 clean metal surfaces. An earlier model by Korzeniewski {\it et
246     al.}\cite{Pons:1986} served as a starting point for our fits. The parameters were
247 gezelter 3826 modified to ensure that the Pt-CO interaction favored the atop binding
248 jmichalk 3869 position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}.
249     The modified parameters yield binding energies that are slightly higher
250 jmichalk 3866 than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
251 jmichalk 3878 {\it et al}.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
252     Lennard-Jones interaction to mimic strong, but short-ranged, partial
253 gezelter 3826 binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
254 jmichalk 3869 Pt-O interaction was modeled with a Morse potential with a large
255     equilibrium distance, ($r_o$). These choices ensure that the C is preferred
256 jmichalk 3878 over O as the surface-binding atom. In most geometries, the Pt-O parameterization contributes a weak
257 gezelter 3826 repulsion which favors the atop site. The resulting potential-energy
258     surface suitably recovers the calculated Pt-C separation length
259     (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
260     position.\cite{Deshlahra:2012, Hopster:1978}
261 jmichalk 3811
262 jmichalk 3812 %where did you actually get the functionals for citation?
263     %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
264     %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
265 jmichalk 3866 The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
266 gezelter 3818 Morse potentials, respectively, to reproduce Au-CO binding energies.
267 jmichalk 3869 The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
268 jmichalk 3866 Adsorption energies were obtained from gas-surface DFT calculations with a
269 gezelter 3826 periodic supercell plane-wave basis approach, as implemented in the
270 gezelter 3887 Quantum ESPRESSO package.\cite{QE-2009} Electron cores were
271 gezelter 3818 described with the projector augmented-wave (PAW)
272     method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
273     included to an energy cutoff of 20 Ry. Electronic energies are
274     computed with the PBE implementation of the generalized gradient
275     approximation (GGA) for gold, carbon, and oxygen that was constructed
276     by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
277 jmichalk 3866 In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
278 gezelter 3818 Au x 2 Au surface planes and separated from vertical images by six
279 jmichalk 3866 layers of vacuum space. The surface atoms were all allowed to relax
280     before CO was added to the system. Electronic relaxations were
281     performed until the energy difference between subsequent steps
282     was less than $10^{-8}$ Ry. Nonspin-polarized supercell calculations
283     were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
284 gezelter 3875 zone.\cite{Monkhorst:1976} The relaxed gold slab was
285 gezelter 3826 then used in numerous single point calculations with CO at various
286     heights (and angles relative to the surface) to allow fitting of the
287     empirical force field.
288 gezelter 3818
289 jmichalk 3812 %Hint at future work
290 jmichalk 3866 The parameters employed for the metal-CO cross-interactions in this work
291 jmichalk 3869 are shown in Table~\ref{tab:co_parameters} and the binding energies on the
292     (111) surfaces are displayed in Table~\ref{tab:co_energies}. Charge transfer
293 jmichalk 3878 and polarization are neglected in this model, although these effects could have
294 gezelter 3887 an effect on binding energies and binding site preferences.
295 jmichalk 3811
296 jmichalk 3802 %Table of Parameters
297     %Pt Parameter Set 9
298     %Au Parameter Set 35
299     \begin{table}[H]
300 gezelter 3887 \caption{Parameters for the metal-CO cross-interactions. Metal-C
301     interactions are modeled with Lennard-Jones potentials, while the
302     metal-O interactions were fit to broad Morse
303 gezelter 3826 potentials. Distances are given in \AA~and energies in kcal/mol. }
304 jmichalk 3802 \centering
305     \begin{tabular}{| c | cc | c | ccc |}
306     \hline
307 gezelter 3826 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
308 jmichalk 3802 \hline
309     \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
310     \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
311    
312     \hline
313     \end{tabular}
314 jmichalk 3866 \label{tab:co_parameters}
315 jmichalk 3802 \end{table}
316    
317     %Table of energies
318     \begin{table}[H]
319 jmichalk 3869 \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
320 jmichalk 3867 described in this work. All values are in eV.}
321 jmichalk 3802 \centering
322     \begin{tabular}{| c | cc |}
323 gezelter 3826 \hline
324     & Calculated & Experimental \\
325     \hline
326     \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
327     (Ref. \protect\cite{Kelemen:1979}) \\
328     & & -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
329 gezelter 3875 \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{TPDGold}) \\
330 gezelter 3826 \hline
331 jmichalk 3802 \end{tabular}
332 jmichalk 3866 \label{tab:co_energies}
333 jmichalk 3802 \end{table}
334    
335 gezelter 3826 \subsection{Pt(557) and Au(557) metal interfaces}
336 jmichalk 3872 Our Pt system is an orthorhombic periodic box of dimensions
337     54.482~x~50.046~x~120.88~\AA~while our Au system has
338 jmichalk 3878 dimensions of 57.4~x~51.9285~x~100~\AA. The metal slabs
339     are 9 and 8 atoms deep respectively, corresponding to a slab
340     thickness of $\sim$21~\AA~ for Pt and $\sim$19~\AA~for Au.
341 jmichalk 3870 The systems are arranged in a FCC crystal that have been cut
342     along the (557) plane so that they are periodic in the {\it x} and
343     {\it y} directions, and have been oriented to expose two aligned
344     (557) cuts along the extended {\it z}-axis. Simulations of the
345     bare metal interfaces at temperatures ranging from 300~K to
346 jmichalk 3872 1200~K were performed to confirm the relative
347 gezelter 3826 stability of the surfaces without a CO overlayer.
348 jmichalk 3802
349 gezelter 3887 The different bulk melting temperatures predicted by EAM
350     (1345~$\pm$~10~K for Au\cite{Au:melting} and $\sim$~2045~K for
351     Pt\cite{Pt:melting}) suggest that any reconstructions should happen at
352     different temperatures for the two metals. The bare Au and Pt
353     surfaces were initially run in the canonical (NVT) ensemble at 800~K
354     and 1000~K respectively for 100 ps. The two surfaces were relatively
355     stable at these temperatures when no CO was present, but experienced
356     increased surface mobility on addition of CO. Each surface was then
357     dosed with different concentrations of CO that was initially placed in
358     the vacuum region. Upon full adsorption, these concentrations
359     correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface coverage. Higher
360     coverages resulted in the formation of a double layer of CO, which
361     introduces artifacts that are not relevant to (557) reconstruction.
362     Because of the difference in binding energies, nearly all of the CO
363     was bound to the Pt surface, while the Au surfaces often had a
364     significant CO population in the gas phase. These systems were
365     allowed to reach thermal equilibrium (over 5~ns) before being run in
366     the microcanonical (NVE) ensemble for data collection. All of the
367     systems examined had at least 40~ns in the data collection stage,
368     although simulation times for some Pt of the systems exceeded 200~ns.
369     Simulations were carried out using the open source molecular dynamics
370     package, OpenMD.\cite{Ewald,OOPSE,openmd}
371 jmichalk 3802
372 jmichalk 3872
373     % RESULTS
374     %
375 jmichalk 3802 \section{Results}
376 jmichalk 3860 \subsection{Structural remodeling}
377 gezelter 3887 The bare metal surfaces experienced minor roughening of the step-edge
378     because of the elevated temperatures, but the (557) face was stable
379     throughout the simulations. The surfaces of both systems, upon dosage
380     of CO, began to undergo extensive remodeling that was not observed in
381     the bare systems. Reconstructions of the Au systems were limited to
382     breakup of the step-edges and some step wandering. The lower coverage
383     Pt systems experienced similar step edge wandering but to a greater
384     extent. The 50\% coverage Pt system was unique among our simulations
385     in that it formed well-defined and stable double layers through step
386     coalescence, similar to results reported by Tao {\it et
387     al}.\cite{Tao:2010}
388 jmichalk 3872
389 jmichalk 3871 \subsubsection{Step wandering}
390 gezelter 3887 The bare surfaces for both metals showed minimal step-wandering at
391     their respective temperatures. As the CO coverage increased however,
392     the mobility of the surface atoms, described through adatom diffusion
393     and step-edge wandering, also increased. Except for the 50\% Pt
394     system where step coalescence occurred, the step-edges in the other
395     simulations preferred to keep nearly the same distance between steps
396     as in the original (557) lattice, $\sim$13\AA~for Pt and
397     $\sim$14\AA~for Au. Previous work by Williams {\it et
398     al}.\cite{Williams:1991, Williams:1994} highlights the repulsion
399     that exists between step-edges even when no direct interactions are
400     present in the system. This repulsion is caused by an entropic barrier
401     that arises from the fact that steps cannot cross over one
402     another. This entropic repulsion does not completely define the
403     interactions between steps, however, so it is possible to observe step
404     coalescence on some surfaces.\cite{Williams:1991} The presence and
405     concentration of adsorbates, as shown in this work, can affect
406     step-step interactions, potentially leading to a new surface structure
407     as the thermodynamic equilibrium.
408 jmichalk 3872
409 jmichalk 3871 \subsubsection{Double layers}
410 gezelter 3887 Tao {\it et al}.\cite{Tao:2010} have shown experimentally that the
411     Pt(557) surface undergoes two separate reconstructions upon CO
412     adsorption. The first involves a doubling of the step height and
413     plateau length. Similar behavior has been seen on a number of
414     surfaces at varying conditions, including Ni(977) and
415     Si(111).\cite{Williams:1994,Williams:1991,Pearl} Of the two systems we
416     examined, the Pt system showed a greater propensity for reconstruction
417     because of the larger surface mobility and the greater extent of step
418     wandering. The amount of reconstruction was strongly correlated to
419     the amount of CO adsorbed upon the surface. This appears to be
420     related to the effect that adsorbate coverage has on edge breakup and
421     on the surface diffusion of metal adatoms. Only the 50\% Pt surface
422     underwent the doubling seen by Tao {\it et al}.\cite{Tao:2010} within
423     the time scales studied here. Over a longer time scale (150~ns) two
424     more double layers formed on this surface. Although double layer
425     formation did not occur in the other Pt systems, they exhibited more
426     step-wandering and roughening compared to their Au counterparts. The
427     50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
428     various times along the simulation showing the evolution of a double
429     layer step-edge.
430 jmichalk 3802
431 gezelter 3887 The second reconstruction observed by Tao {\it et al}.\cite{Tao:2010}
432     involved the formation of triangular clusters that stretched across
433     the plateau between two step-edges. Neither of the simulated metal
434     interfaces, within the 40~ns time scale or the extended time of 150~ns
435     for the 50\% Pt system, experienced this reconstruction.
436 jmichalk 3817
437 jmichalk 3876 %Evolution of surface
438     \begin{figure}[H]
439 gezelter 3882 \includegraphics[width=\linewidth]{EPS_ProgressionOfDoubleLayerFormation}
440 gezelter 3887 \caption{The Pt(557) / 50\% CO interface upon exposure to the CO: (a)
441     258~ps, (b) 19~ns, (c) 31.2~ns, and (d) 86.1~ns after
442     exposure. Disruption of the (557) step-edges occurs quickly. The
443 jmichalk 3876 doubling of the layers appears only after two adjacent step-edges
444     touch. The circled spot in (b) nucleated the growth of the double
445     step observed in the later configurations.}
446     \label{fig:reconstruct}
447     \end{figure}
448    
449 jmichalk 3860 \subsection{Dynamics}
450 gezelter 3887 Previous experimental work by Pearl and Sibener\cite{Pearl}, using
451     STM, has been able to capture the coalescence of steps on Ni(977). The
452     time scale of the image acquisition, $\sim$70~s/image, provides an
453     upper bound for the time required for the doubling to occur. By
454     utilizing Molecular Dynamics we are able to probe the dynamics of
455     these reconstructions at elevated temperatures and in this section we
456     provide data on the timescales for transport properties,
457     e.g. diffusion and layer formation time.
458 gezelter 3826
459 jmichalk 3867
460 jmichalk 3860 \subsubsection{Transport of surface metal atoms}
461 jmichalk 3862 %forcedSystems/stepSeparation
462 gezelter 3826
463 gezelter 3887 The wandering of a step-edge is a cooperative effect arising from the
464     individual movements of the atoms making up the steps. An ideal metal
465     surface displaying a low index facet, (111) or (100), is unlikely to
466     experience much surface diffusion because of the large energetic
467     barrier that must be overcome to lift an atom out of the surface. The
468     presence of step-edges and other surface features on higher-index
469     facets provides a lower energy source for mobile metal atoms. Using
470     our potential model, single-atom break-away from a step-edge on a
471     clean surface still imposes an energetic penalty around
472     $\sim$~45~kcal/mol, but this is certainly easier than lifting the same
473     metal atom vertically out of the surface, \textgreater~60~kcal/mol.
474     The penalty lowers significantly when CO is present in sufficient
475     quantities on the surface. For certain distributions of CO, the
476     energetic penalty can fall to as low as $\sim$~20~kcal/mol. The
477     configurations that create these lower barriers are detailed in the
478     discussion section below.
479 gezelter 3826
480 gezelter 3887 Once an adatom exists on the surface, the barrier for diffusion is
481     negligible (\textless~4~kcal/mol for a Pt adatom). These adatoms are
482     then able to explore the terrace before rejoining either their
483     original step-edge or becoming a part of a different edge. It is an
484     energetically unfavorable process with a high barrier for an atom to
485     traverse to a separate terrace although the presence of CO can lower
486     the energy barrier required to lift or lower an adatom. By tracking
487     the mobility of individual metal atoms on the Pt and Au surfaces we
488     were able to determine the relative diffusion constants, as well as
489     how varying coverages of CO affect the diffusion. Close observation of
490     the mobile metal atoms showed that they were typically in equilibrium
491     with the step-edges. At times, their motion was concerted, and two or
492     more adatoms would be observed moving together across the surfaces.
493    
494     A particle was considered ``mobile'' once it had traveled more than
495     2~\AA~ between saved configurations of the system (typically 10-100
496     ps). A mobile atom would typically travel much greater distances than
497     this, but the 2~\AA~cutoff was used to prevent swamping the diffusion
498     data with the in-place vibrational movement of buried atoms. Diffusion
499     on a surface is strongly affected by local structures and the presence
500     of single and double layer step-edges causes the diffusion parallel to
501     the step-edges to be larger than the diffusion perpendicular to these
502     edges. Parallel and perpendicular diffusion constants are shown in
503     Figure \ref{fig:diff}. Diffusion parallel to the step-edge is higher
504     than diffusion perpendicular to the edge because of the lower energy
505     barrier associated with sliding along an edge compared to breaking
506     away to form an isolated adatom.
507    
508 jmichalk 3876 %Diffusion graph
509     \begin{figure}[H]
510 gezelter 3882 \includegraphics[width=\linewidth]{Portrait_DiffusionComparison_1}
511 jmichalk 3876 \caption{Diffusion constants for mobile surface atoms along directions
512     parallel ($\mathbf{D}_{\parallel}$) and perpendicular
513     ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
514 gezelter 3887 surface coverage. The two reported diffusion constants for the 50\%
515     Pt system correspond to a 20~ns period before the formation of the
516     double layer (upper points), and to the full 40~ns sampling period
517     (lower points).}
518 jmichalk 3876 \label{fig:diff}
519     \end{figure}
520    
521 jmichalk 3878 The weaker Au-CO interaction is evident in the weak CO-coverage
522     dependance of Au diffusion. This weak interaction leads to lower
523     observed coverages when compared to dosage amounts. This further
524     limits the effect the CO can have on surface diffusion. The correlation
525     between coverage and Pt diffusion rates shows a near linear relationship
526     at the earliest times in the simulations. Following double layer formation,
527     however, there is a precipitous drop in adatom diffusion. As the double
528     layer forms, many atoms that had been tracked for mobility data have
529 gezelter 3887 now been buried, resulting in a smaller reported diffusion constant. A
530 jmichalk 3878 secondary effect of higher coverages is CO-CO cross interactions that
531     lower the effective mobility of the Pt adatoms that are bound to each CO.
532     This effect would become evident only at higher coverages. A detailed
533     account of Pt adatom energetics follows in the Discussion.
534    
535     \subsubsection{Dynamics of double layer formation}
536     The increased diffusion on Pt at the higher CO coverages is the primary
537     contributor to double layer formation. However, this is not a complete
538     explanation -- the 33\%~Pt system has higher diffusion constants, but
539     did not show any signs of edge doubling in 40~ns. On the 50\%~Pt
540     system, one double layer formed within the first 40~ns of simulation time,
541     while two more were formed as the system was allowed to run for an
542     additional 110~ns (150~ns total). This suggests that this reconstruction
543     is a rapid process and that the previously mentioned upper bound is a
544     very large overestimate.\cite{Williams:1991,Pearl} In this system the first
545     appearance of a double layer appears at 19~ns into the simulation.
546     Within 12~ns of this nucleation event, nearly half of the step has formed
547     the double layer and by 86~ns the complete layer has flattened out.
548     From the appearance of the first nucleation event to the first observed
549     double layer, the process took $\sim$20~ns. Another $\sim$40~ns was
550     necessary for the layer to completely straighten. The other two layers in
551     this simulation formed over periods of 22~ns and 42~ns respectively.
552     A possible explanation for this rapid reconstruction is the elevated
553     temperatures under which our systems were simulated. The process
554     would almost certainly take longer at lower temperatures. Additionally,
555     our measured times for completion of the doubling after the appearance
556     of a nucleation site are likely affected by our periodic boxes. A longer
557     step-edge will likely take longer to ``zipper''.
558 jmichalk 3876
559    
560 jmichalk 3878 %Discussion
561     \section{Discussion}
562 gezelter 3882 We have shown that a classical potential is able to model the initial
563     reconstruction of the Pt(557) surface upon CO adsorption, and have
564     reproduced the double layer structure observed by Tao {\it et
565     al}.\cite{Tao:2010}. Additionally, this reconstruction appears to be
566     rapid -- occurring within 100 ns of the initial exposure to CO. Here
567     we discuss the features of the classical potential that are
568     contributing to the stability and speed of the Pt(557) reconstruction.
569 jmichalk 3817
570 jmichalk 3878 \subsection{Diffusion}
571 gezelter 3882 The perpendicular diffusion constant appears to be the most important
572     indicator of double layer formation. As highlighted in Figure
573     \ref{fig:reconstruct}, the formation of the double layer did not begin
574     until a nucleation site appeared. Williams {\it et
575     al}.\cite{Williams:1991,Williams:1994} cite an effective edge-edge
576     repulsion arising from the inability of edge crossing. This repulsion
577     must be overcome to allow step coalescence. A larger
578     $\textbf{D}_\perp$ value implies more step-wandering and a larger
579     chance for the stochastic meeting of two edges to create a nucleation
580     point. Diffusion parallel to the step-edge can help ``zipper'' up a
581     nascent double layer. This helps explain the rapid time scale for
582     double layer completion after the appearance of a nucleation site, while
583     the initial appearance of the nucleation site was unpredictable.
584 jmichalk 3876
585 jmichalk 3878 \subsection{Mechanism for restructuring}
586 gezelter 3882 Since the Au surface showed no large scale restructuring in any of our
587     simulations, our discussion will focus on the 50\% Pt-CO system which
588     did exhibit doubling. A number of possible mechanisms exist to explain
589     the role of adsorbed CO in restructuring the Pt surface. Quadrupolar
590     repulsion between adjacent CO molecules adsorbed on the surface is one
591     possibility. However, the quadrupole-quadrupole interaction is
592     short-ranged and is attractive for some orientations. If the CO
593     molecules are ``locked'' in a vertical orientation, through atop
594 gezelter 3887 adsorption for example, this explanation would gain credence. Within
595     the framework of our classical potential, the calculated energetic
596     repulsion between two CO molecules located a distance of
597     2.77~\AA~apart (nearest-neighbor distance of Pt) and both in a
598     vertical orientation, is 8.62 kcal/mol. Moving the CO to the second
599     nearest-neighbor distance of 4.8~\AA~drops the repulsion to nearly
600     0. Allowing the CO to rotate away from a purely vertical orientation
601     also lowers the repulsion. When the carbons are locked at a distance
602     of 2.77~\AA, a minimum of 6.2 kcal/mol is reached when the angle
603     between the 2 CO is $\sim$24\textsuperscript{o}. The calculated
604 gezelter 3882 barrier for surface diffusion of a Pt adatom is only 4 kcal/mol, so
605 gezelter 3887 repulsion between adjacent CO molecules bound to Pt could indeed
606     increase the surface diffusion. However, the residence time of CO on
607     Pt suggests that the CO molecules are extremely mobile, with diffusion
608     constants 40 to 2500 times larger than surface Pt atoms. This mobility
609     suggests that the CO molecules jump between different Pt atoms
610     throughout the simulation. However, they do stay bound to individual
611     Pt atoms for long enough to modify the local energy landscape for the
612     mobile adatoms.
613 jmichalk 3876
614 gezelter 3882 A different interpretation of the above mechanism which takes the
615     large mobility of the CO into account, would be in the destabilization
616     of Pt-Pt interactions due to bound CO. Destabilizing Pt-Pt bonds at
617     the edges could lead to increased step-edge breakup and diffusion. On
618     the bare Pt(557) surface the barrier to completely detach an edge atom
619     is $\sim$43~kcal/mol, as is shown in configuration (a) in Figures
620     \ref{fig:SketchGraphic} \& \ref{fig:SketchEnergies}. For certain
621     configurations, cases (e), (g), and (h), the barrier can be lowered to
622     $\sim$23~kcal/mol by the presence of bound CO molecules. In these
623     instances, it becomes energetically favorable to roughen the edge by
624     introducing a small separation of 0.5 to 1.0~\AA. This roughening
625     becomes immediately obvious in simulations with significant CO
626     populations. The roughening is present to a lesser extent on surfaces
627     with lower CO coverage (and even on the bare surfaces), although in
628     these cases it is likely due to random fluctuations that squeeze out
629 gezelter 3887 step-edge atoms. Step-edge breakup by direct single-atom translations
630     (as suggested by these energy curves) is probably a worst-case
631     scenario. Multistep mechanisms in which an adatom moves laterally on
632     the surface after being ejected would be more energetically favorable.
633     This would leave the adatom alongside the ledge, providing it with
634     five nearest neighbors. While fewer than the seven neighbors it had
635     as part of the step-edge, it keeps more Pt neighbors than the three
636     neighbors an isolated adatom has on the terrace. In this proposed
637     mechanism, the CO quadrupolar repulsion still plays a role in the
638     initial roughening of the step-edge, but not in any long-term bonds
639     with individual Pt atoms. Higher CO coverages create more
640 gezelter 3882 opportunities for the crowded CO configurations shown in Figure
641     \ref{fig:SketchGraphic}, and this is likely to cause an increased
642     propensity for step-edge breakup.
643 jmichalk 3876
644     %Sketch graphic of different configurations
645 jmichalk 3816 \begin{figure}[H]
646 gezelter 3882 \includegraphics[width=\linewidth]{COpaths}
647     \caption{Configurations used to investigate the mechanism of step-edge
648 gezelter 3887 breakup on Pt(557). In each case, the central (starred) atom was
649 gezelter 3882 pulled directly across the surface away from the step edge. The Pt
650     atoms on the upper terrace are colored dark grey, while those on the
651     lower terrace are in white. In each of these configurations, some
652 gezelter 3887 of the atoms (highlighted in blue) had CO molecules bound in the
653     vertical atop position. The energies of these configurations as a
654 gezelter 3882 function of central atom displacement are displayed in Figure
655     \ref{fig:SketchEnergies}.}
656 jmichalk 3876 \label{fig:SketchGraphic}
657 jmichalk 3862 \end{figure}
658    
659 jmichalk 3876 %energy graph corresponding to sketch graphic
660 jmichalk 3862 \begin{figure}[H]
661 gezelter 3882 \includegraphics[width=\linewidth]{Portrait_SeparationComparison}
662     \caption{Energies for displacing a single edge atom perpendicular to
663     the step edge as a function of atomic displacement. Each of the
664     energy curves corresponds to one of the labeled configurations in
665 gezelter 3887 Figure \ref{fig:SketchGraphic}, and the energies are referenced to
666     the unperturbed step-edge. Certain arrangements of bound CO
667     (notably configurations g and h) can lower the energetic barrier for
668     creating an adatom relative to the bare surface (configuration a).}
669 jmichalk 3876 \label{fig:SketchEnergies}
670 jmichalk 3816 \end{figure}
671    
672 gezelter 3882 While configurations of CO on the surface are able to increase
673     diffusion and the likelihood of edge wandering, this does not provide
674     a complete explanation for the formation of double layers. If adatoms
675     were constrained to their original terraces then doubling could not
676     occur. A mechanism for vertical displacement of adatoms at the
677     step-edge is required to explain the doubling.
678 jmichalk 3802
679 gezelter 3882 We have discovered one possible mechanism for a CO-mediated vertical
680     displacement of Pt atoms at the step edge. Figure \ref{fig:lambda}
681     shows four points along a reaction coordinate in which a CO-bound
682     adatom along the step-edge ``burrows'' into the edge and displaces the
683 gezelter 3887 original edge atom onto the higher terrace. A number of events
684     similar to this mechanism were observed during the simulations. We
685     predict an energetic barrier of 20~kcal/mol for this process (in which
686     the displaced edge atom follows a curvilinear path into an adjacent
687     3-fold hollow site). The barrier heights we obtain for this reaction
688 gezelter 3882 coordinate are approximate because the exact path is unknown, but the
689     calculated energy barriers would be easily accessible at operating
690     conditions. Additionally, this mechanism is exothermic, with a final
691     energy 15~kcal/mol below the original $\lambda = 0$ configuration.
692     When CO is not present and this reaction coordinate is followed, the
693 gezelter 3887 process is endothermic by 3~kcal/mol. The difference in the relative
694 gezelter 3882 energies for the $\lambda=0$ and $\lambda=1$ case when CO is present
695     provides strong support for CO-mediated Pt-Pt interactions giving rise
696 gezelter 3887 to the doubling reconstruction.
697 gezelter 3882
698 jmichalk 3862 %lambda progression of Pt -> shoving its way into the step
699     \begin{figure}[H]
700 gezelter 3882 \includegraphics[width=\linewidth]{EPS_rxnCoord}
701     \caption{Points along a possible reaction coordinate for CO-mediated
702     edge doubling. Here, a CO-bound adatom burrows into an established
703     step edge and displaces an edge atom onto the upper terrace along a
704     curvilinear path. The approximate barrier for the process is
705     20~kcal/mol, and the complete process is exothermic by 15~kcal/mol
706 gezelter 3887 in the presence of CO, but is endothermic by 3~kcal/mol without CO.}
707 jmichalk 3862 \label{fig:lambda}
708     \end{figure}
709    
710 gezelter 3882 The mechanism for doubling on the Pt(557) surface appears to require
711     the cooperation of at least two distinct processes. For complete
712     doubling of a layer to occur there must be a breakup of one
713     terrace. These atoms must then ``disappear'' from that terrace, either
714 gezelter 3887 by travelling to the terraces above or below their original levels.
715 gezelter 3882 The presence of CO helps explain mechanisms for both of these
716     situations. There must be sufficient breakage of the step-edge to
717     increase the concentration of adatoms on the surface and these adatoms
718     must then undergo the burrowing highlighted above (or a comparable
719     mechanism) to create the double layer. With sufficient time, these
720     mechanisms working in concert lead to the formation of a double layer.
721 jmichalk 3879
722 jmichalk 3878 \subsection{CO Removal and double layer stability}
723 gezelter 3887 Once the double layers had formed on the 50\%~Pt system, they remained
724     stable for the rest of the simulation time with minimal movement.
725     Random fluctuations that involved small clusters or divots were
726     observed, but these features typically healed within a few
727     nanoseconds. Within our simulations, the formation of the double
728     layer appeared to be irreversible and a double layer was never
729     observed to split back into two single layer step-edges while CO was
730     present.
731 jmichalk 3862
732 gezelter 3882 To further gauge the effect CO has on this surface, additional
733     simulations were run starting from a late configuration of the 50\%~Pt
734     system that had already formed double layers. These simulations then
735 gezelter 3887 had their CO molecules suddenly removed. The double layer broke apart
736     rapidly in these simulations, showing a well-defined edge-splitting
737     after 100~ps. Configurations of this system are shown in Figure
738 gezelter 3882 \ref{fig:breaking}. The coloring of the top and bottom layers helps to
739 gezelter 3887 show how much mixing the edges experience as they split. These systems
740     were only examined for 10~ns, and within that time despite the initial
741     rapid splitting, the edges only moved another few \AA~apart. It is
742     possible that with longer simulation times, the (557) surface recovery
743     observed by Tao {\it et al}.\cite{Tao:2010} could also be recovered.
744 jmichalk 3862
745     %breaking of the double layer upon removal of CO
746 jmichalk 3802 \begin{figure}[H]
747 gezelter 3882 \includegraphics[width=\linewidth]{EPS_doubleLayerBreaking}
748 gezelter 3887 \caption{Behavior of an established (111) double step after removal of
749     the adsorbed CO: (A) 0~ps, (B) 100~ps, and (C) 1~ns after the
750     removal of CO. Nearly immediately after the CO is removed, the
751     step edge reforms in a (100) configuration, which is also the step
752     type seen on clean (557) surfaces. The step separation involves
753 gezelter 3882 significant mixing of the lower and upper atoms at the edge.}
754 jmichalk 3862 \label{fig:breaking}
755 jmichalk 3802 \end{figure}
756    
757    
758     %Peaks!
759 jmichalk 3872 %\begin{figure}[H]
760     %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
761     %\caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
762     %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
763     %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
764     %\label{fig:peaks}
765     %\end{figure}
766 jmichalk 3862
767 jmichalk 3867
768     %Don't think I need this
769 jmichalk 3862 %clean surface...
770 jmichalk 3867 %\begin{figure}[H]
771 gezelter 3882 %\includegraphics[width=\linewidth]{557_300K_cleanPDF}
772 jmichalk 3867 %\caption{}
773 jmichalk 3862
774 jmichalk 3867 %\end{figure}
775     %\label{fig:clean}
776    
777    
778 jmichalk 3802 \section{Conclusion}
779 gezelter 3882 The strength and directionality of the Pt-CO binding interaction, as
780     well as the large quadrupolar repulsion between atop-bound CO
781     molecules, help to explain the observed increase in surface mobility
782     of Pt(557) and the resultant reconstruction into a double-layer
783     configuration at the highest simulated CO-coverages. The weaker Au-CO
784     interaction results in significantly lower adataom diffusion
785     constants, less step-wandering, and a lack of the double layer
786     reconstruction on the Au(557) surface.
787 jmichalk 3802
788 gezelter 3882 An in-depth examination of the energetics shows the important role CO
789     plays in increasing step-breakup and in facilitating edge traversal
790     which are both necessary for double layer formation.
791 jmichalk 3880
792 jmichalk 3862 %Things I am not ready to remove yet
793    
794     %Table of Diffusion Constants
795     %Add gold?M
796     % \begin{table}[H]
797     % \caption{}
798     % \centering
799     % \begin{tabular}{| c | cc | cc | }
800     % \hline
801     % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
802     % \hline
803     % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
804     % \hline
805     % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
806     % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
807     % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
808     % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
809     % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
810     % \hline
811     % \end{tabular}
812     % \end{table}
813    
814 gezelter 3875 \begin{acknowledgement}
815 gezelter 3882 We gratefully acknowledge conversations with Dr. William
816     F. Schneider and Dr. Feng Tao. Support for this project was
817     provided by the National Science Foundation under grant CHE-0848243
818     and by the Center for Sustainable Energy at Notre Dame
819     (cSEND). Computational time was provided by the Center for Research
820     Computing (CRC) at the University of Notre Dame.
821 gezelter 3875 \end{acknowledgement}
822 gezelter 3808 \newpage
823 gezelter 3887 \bibstyle{achemso}
824     \bibliography{COonPtAu}
825 gezelter 3875 %\end{doublespace}
826    
827     \begin{tocentry}
828 gezelter 3887 \begin{wrapfigure}{l}{0.5\textwidth}
829     \begin{center}
830     \includegraphics[width=\linewidth]{TOC_doubleLayer}
831     \end{center}
832     \end{wrapfigure}
833     A reconstructed Pt(557) surface after 86~ns exposure to a half a
834     monolayer of CO. The double layer that forms is a result of
835     CO-mediated step-edge wandering as well as a burrowing mechanism that
836     helps lift edge atoms onto an upper terrace.
837 gezelter 3875 \end{tocentry}
838    
839 gezelter 3808 \end{document}