ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3879
Committed: Mon Mar 18 21:20:12 2013 UTC (11 years, 3 months ago) by jmichalk
Content type: application/x-tex
Original Path: trunk/COonPt/firstTry.tex
File size: 43206 byte(s)
Log Message:
I think the discussion is in a pretty good place right now

File Contents

# Content
1 \documentclass[journal = jpccck, manuscript = article]{achemso}
2 \setkeys{acs}{usetitle = true}
3 \usepackage{achemso}
4 \usepackage{caption}
5 \usepackage{float}
6 \usepackage{geometry}
7 \usepackage{natbib}
8 \usepackage{setspace}
9 \usepackage{xkeyval}
10 %%%%%%%%%%%%%%%%%%%%%%%
11 \usepackage{amsmath}
12 \usepackage{amssymb}
13 \usepackage{times}
14 \usepackage{mathptm}
15 \usepackage{setspace}
16 \usepackage{endfloat}
17 \usepackage{caption}
18 \usepackage{tabularx}
19 \usepackage{longtable}
20 \usepackage{graphicx}
21 \usepackage{multirow}
22 \usepackage{multicol}
23
24 \usepackage[version=3]{mhchem} % this is a great package for formatting chemical reactions
25 % \usepackage[square, comma, sort&compress]{natbib}
26 \usepackage{url}
27 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
28 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
29 9.0in \textwidth 6.5in \brokenpenalty=10000
30
31 % double space list of tables and figures
32 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
33 \setlength{\abovecaptionskip}{20 pt}
34 \setlength{\belowcaptionskip}{30 pt}
35 % \bibpunct{}{}{,}{s}{}{;}
36
37 %\citestyle{nature}
38 % \bibliographystyle{achemso}
39
40 \title{Molecular Dynamics simulations of the surface reconstructions
41 of Pt(557) and Au(557) under exposure to CO}
42
43 \author{Joseph R. Michalka}
44 \author{Patrick W. McIntyre}
45 \author{J. Daniel Gezelter}
46 \email{gezelter@nd.edu}
47 \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
48 Department of Chemistry and Biochemistry\\ University of Notre
49 Dame\\ Notre Dame, Indiana 46556}
50
51 \keywords{}
52
53 \begin{document}
54
55
56 %%
57 %Introduction
58 % Experimental observations
59 % Previous work on Pt, CO, etc.
60 %
61 %Simulation Methodology
62 % FF (fits and parameters)
63 % MD (setup, equilibration, collection)
64 %
65 % Analysis of trajectories!!!
66 %Discussion
67 % CO preferences for specific locales
68 % CO-CO interactions
69 % Differences between Au & Pt
70 % Causes of 2_layer reordering in Pt
71 %Summary
72 %%
73
74
75 \begin{abstract}
76 We examine surface reconstructions of Pt and Au(557) under
77 various CO coverages using molecular dynamics in order to
78 explore possible mechanisms for any observed reconstructions
79 and their dynamics. The metal-CO interactions were parameterized
80 as part of this work so that an efficient large-scale treatment of
81 this system could be undertaken. The large difference in binding
82 strengths of the metal-CO interactions was found to play a significant
83 role with regards to step-edge stability and adatom diffusion. A
84 small correlation between coverage and the diffusion constant
85 was also determined. The energetics of CO adsorbed to the surface
86 is sufficient to explain the reconstructions observed on the Pt
87 systems and the lack of reconstruction of the Au systems.
88
89 \end{abstract}
90
91 \newpage
92
93
94 \section{Introduction}
95 % Importance: catalytically active metals are important
96 % Sub: Knowledge of how their surface structure affects their ability to catalytically facilitate certain reactions is growing, but is more reactionary than predictive
97 % 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)
98 % Theory can explore temperatures and pressures which are difficult to work with in experiments
99 % Sub: Also, easier to observe what is going on and provide reasons and explanations
100 %
101
102 Industrial catalysts usually consist of small particles that exhibit a
103 high concentration of steps, kink sites, and vacancies at the edges of
104 the facets. These sites are thought to be the locations of catalytic
105 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
106 significant evidence that solid surfaces are often structurally,
107 compositionally, and chemically modified by reactants under operating
108 conditions.\cite{Tao2008,Tao:2010,Tao2011} The coupling between
109 surface oxidation states and catalytic activity for CO oxidation on
110 Pt, for instance, is widely documented.\cite{Ertl08,Hendriksen:2002}
111 Despite the well-documented role of these effects on reactivity, the
112 ability to capture or predict them in atomistic models is somewhat
113 limited. While these effects are perhaps unsurprising on the highly
114 disperse, multi-faceted nanoscale particles that characterize
115 industrial catalysts, they are manifest even on ordered, well-defined
116 surfaces. The Pt(557) surface, for example, exhibits substantial and
117 reversible restructuring under exposure to moderate pressures of
118 carbon monoxide.\cite{Tao:2010}
119
120 This work is an investigation into the mechanism and timescale for the Pt(557) \& Au(557)
121 surface restructuring using molecular simulations. Since the dynamics
122 of the process are of particular interest, we employ classical force
123 fields that represent a compromise between chemical accuracy and the
124 computational efficiency necessary to simulate the process of interest.
125 Since restructuring typically occurs as a result of specific interactions of the
126 catalyst with adsorbates, in this work, two metal systems exposed
127 to carbon monoxide were examined. The Pt(557) surface has already been shown
128 to undergo a large scale reconstruction under certain conditions.\cite{Tao:2010}
129 The Au(557) surface, because of a weaker interaction with CO, is less
130 likely to undergo this kind of reconstruction. However, Peters {\it et al}.\cite{Peters:2000}
131 and Piccolo {\it et al}.\cite{Piccolo:2004} have both observed CO-induced
132 reconstruction of a Au(111) surface. Peters {\it et al}. saw a relaxation to the
133 22 x $\sqrt{3}$ cell. They argued that only a few Au atoms
134 become adatoms, limiting the stress of this reconstruction, while
135 allowing the rest to relax and approach the ideal (111)
136 configuration. They did not see the usual herringbone pattern on Au(111) being greatly
137 affected by this relaxation. Piccolo {\it et al}. on the other hand, did see a
138 disruption of the herringbone pattern as CO was adsorbed to the
139 surface. Both groups suggested that the preference CO shows for
140 low-coordinated Au atoms was the primary driving force for the reconstruction.
141
142
143
144 %Platinum molecular dynamics
145 %gold molecular dynamics
146
147 \section{Simulation Methods}
148 The challenge in modeling any solid/gas interface is the
149 development of a sufficiently general yet computationally tractable
150 model of the chemical interactions between the surface atoms and
151 adsorbates. Since the interfaces involved are quite large (10$^3$ -
152 10$^4$ atoms) and respond slowly to perturbations, {\it ab initio}
153 molecular dynamics
154 (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
155 methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
156 mechanical potential energy surfaces remain out of reach.
157 Additionally, the ``bonds'' between metal atoms at a surface are
158 typically not well represented in terms of classical pairwise
159 interactions in the same way that bonds in a molecular material are,
160 nor are they captured by simple non-directional interactions like the
161 Coulomb potential. For this work, we have used classical molecular
162 dynamics with potential energy surfaces that are specifically tuned
163 for transition metals. In particular, we used the EAM potential for
164 Au-Au and Pt-Pt interactions.\cite{EAM} The CO was modeled using a rigid
165 three-site model developed by Straub and Karplus for studying
166 photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
167 Pt-CO cross interactions were parameterized as part of this work.
168
169 \subsection{Metal-metal interactions}
170 Many of the potentials used for modeling transition metals are based
171 on a non-pairwise additive functional of the local electron
172 density. The embedded atom method (EAM) is perhaps the best known of
173 these
174 methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
175 but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
176 the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
177 parameter sets. The glue model of Ercolessi {\it et al}. is among the
178 fastest of these density functional approaches.\cite{Ercolessi88} In
179 all of these models, atoms are treated as a positively charged
180 core with a radially-decaying valence electron distribution. To
181 calculate the energy for embedding the core at a particular location,
182 the electron density due to the valence electrons at all of the other
183 atomic sites is computed at atom $i$'s location,
184 \begin{equation*}
185 \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
186 \end{equation*}
187 Here, $\rho_j(r_{ij})$ is the function that describes the distance
188 dependence of the valence electron distribution of atom $j$. The
189 contribution to the potential that comes from placing atom $i$ at that
190 location is then
191 \begin{equation*}
192 V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
193 \end{equation*}
194 where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
195 $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
196 repulsive overlap of the two positively charged cores.
197
198 % The {\it modified} embedded atom method (MEAM) adds angular terms to
199 % the electron density functions and an angular screening factor to the
200 % pairwise interaction between two
201 % atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
202 % MEAM has become widely used to simulate systems in which angular
203 % interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
204 % metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
205 % MEAM presents significant additional computational costs, however.
206
207 The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
208 have all been widely used by the materials simulation community for
209 simulations of bulk and nanoparticle
210 properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
211 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
212 fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
213 propagation,\cite{BECQUART:1993rg} and alloying
214 dynamics.\cite{Shibata:2002hh} One of EAM's strengths
215 is its sensitivity to small changes in structure. This arises
216 because interactions
217 up to the third nearest neighbor were taken into account in the parameterization.\cite{Voter95a}
218 Comparing that to the glue model of Ercolessi {\it et al}.\cite{Ercolessi88}
219 which is only parameterized up to the nearest-neighbor
220 interactions, EAM is a suitable choice for systems where
221 the bulk properties are of secondary importance to low-index
222 surface structures. Additionally, the similarity of EAM's functional
223 treatment of the embedding energy to standard density functional
224 theory (DFT) makes fitting DFT-derived cross potentials with adsorbates somewhat easier.
225 \cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
226
227
228
229
230 \subsection{Carbon Monoxide model}
231 Previous explanations for the surface rearrangements center on
232 the large linear quadrupole moment of carbon monoxide.\cite{Tao:2010}
233 We used a model first proposed by Karplus and Straub to study
234 the photodissociation of CO from myoglobin because it reproduces
235 the quadrupole moment well.\cite{Straub} The Straub and
236 Karplus model treats CO as a rigid three site molecule with a massless M
237 site at the molecular center of mass. The geometry and interaction
238 parameters are reproduced in Table~\ref{tab:CO}. The effective
239 dipole moment, calculated from the assigned charges, is still
240 small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
241 to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
242 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
243 %CO Table
244 \begin{table}[H]
245 \caption{Positions, Lennard-Jones parameters ($\sigma$ and
246 $\epsilon$), and charges for the CO-CO
247 interactions in Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
248 in kcal/mol, and charges are in atomic units.}
249 \centering
250 \begin{tabular}{| c | c | ccc |}
251 \hline
252 & {\it z} & $\sigma$ & $\epsilon$ & q\\
253 \hline
254 \textbf{C} & -0.6457 & 3.83 & 0.0262 & -0.75 \\
255 \textbf{O} & 0.4843 & 3.12 & 0.1591 & -0.85 \\
256 \textbf{M} & 0.0 & - & - & 1.6 \\
257 \hline
258 \end{tabular}
259 \label{tab:CO}
260 \end{table}
261
262 \subsection{Cross-Interactions between the metals and carbon monoxide}
263
264 Since the adsorption of CO onto a Pt surface has been the focus
265 of much experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
266 and theoretical work
267 \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
268 there is a significant amount of data on adsorption energies for CO on
269 clean metal surfaces. An earlier model by Korzeniewski {\it et
270 al.}\cite{Pons:1986} served as a starting point for our fits. The parameters were
271 modified to ensure that the Pt-CO interaction favored the atop binding
272 position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}.
273 The modified parameters yield binding energies that are slightly higher
274 than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
275 {\it et al}.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
276 Lennard-Jones interaction to mimic strong, but short-ranged, partial
277 binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
278 Pt-O interaction was modeled with a Morse potential with a large
279 equilibrium distance, ($r_o$). These choices ensure that the C is preferred
280 over O as the surface-binding atom. In most geometries, the Pt-O parameterization contributes a weak
281 repulsion which favors the atop site. The resulting potential-energy
282 surface suitably recovers the calculated Pt-C separation length
283 (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
284 position.\cite{Deshlahra:2012, Hopster:1978}
285
286 %where did you actually get the functionals for citation?
287 %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
288 %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
289 The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
290 Morse potentials, respectively, to reproduce Au-CO binding energies.
291 The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
292 Adsorption energies were obtained from gas-surface DFT calculations with a
293 periodic supercell plane-wave basis approach, as implemented in the
294 {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores were
295 described with the projector augmented-wave (PAW)
296 method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
297 included to an energy cutoff of 20 Ry. Electronic energies are
298 computed with the PBE implementation of the generalized gradient
299 approximation (GGA) for gold, carbon, and oxygen that was constructed
300 by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
301 In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
302 Au x 2 Au surface planes and separated from vertical images by six
303 layers of vacuum space. The surface atoms were all allowed to relax
304 before CO was added to the system. Electronic relaxations were
305 performed until the energy difference between subsequent steps
306 was less than $10^{-8}$ Ry. Nonspin-polarized supercell calculations
307 were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
308 zone.\cite{Monkhorst:1976} The relaxed gold slab was
309 then used in numerous single point calculations with CO at various
310 heights (and angles relative to the surface) to allow fitting of the
311 empirical force field.
312
313 %Hint at future work
314 The parameters employed for the metal-CO cross-interactions in this work
315 are shown in Table~\ref{tab:co_parameters} and the binding energies on the
316 (111) surfaces are displayed in Table~\ref{tab:co_energies}. Charge transfer
317 and polarization are neglected in this model, although these effects could have
318 an effect on binding energies and binding site preferences.
319
320 %Table of Parameters
321 %Pt Parameter Set 9
322 %Au Parameter Set 35
323 \begin{table}[H]
324 \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
325 interactions are modeled with Lennard-Jones potentials. While the
326 metal-O interactions were fit to Morse
327 potentials. Distances are given in \AA~and energies in kcal/mol. }
328 \centering
329 \begin{tabular}{| c | cc | c | ccc |}
330 \hline
331 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
332 \hline
333 \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
334 \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
335
336 \hline
337 \end{tabular}
338 \label{tab:co_parameters}
339 \end{table}
340
341 %Table of energies
342 \begin{table}[H]
343 \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
344 described in this work. All values are in eV.}
345 \centering
346 \begin{tabular}{| c | cc |}
347 \hline
348 & Calculated & Experimental \\
349 \hline
350 \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
351 (Ref. \protect\cite{Kelemen:1979}) \\
352 & & -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
353 \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{TPDGold}) \\
354 \hline
355 \end{tabular}
356 \label{tab:co_energies}
357 \end{table}
358
359 \subsection{Pt(557) and Au(557) metal interfaces}
360 Our Pt system is an orthorhombic periodic box of dimensions
361 54.482~x~50.046~x~120.88~\AA~while our Au system has
362 dimensions of 57.4~x~51.9285~x~100~\AA. The metal slabs
363 are 9 and 8 atoms deep respectively, corresponding to a slab
364 thickness of $\sim$21~\AA~ for Pt and $\sim$19~\AA~for Au.
365 The systems are arranged in a FCC crystal that have been cut
366 along the (557) plane so that they are periodic in the {\it x} and
367 {\it y} directions, and have been oriented to expose two aligned
368 (557) cuts along the extended {\it z}-axis. Simulations of the
369 bare metal interfaces at temperatures ranging from 300~K to
370 1200~K were performed to confirm the relative
371 stability of the surfaces without a CO overlayer.
372
373 The different bulk melting temperatures predicted by EAM (1345~$\pm$~10~K for Au\cite{Au:melting}
374 and $\sim$~2045~K for Pt\cite{Pt:melting}) suggest that any possible reconstruction should happen at
375 different temperatures for the two metals. The bare Au and Pt surfaces were
376 initially run in the canonical (NVT) ensemble at 800~K and 1000~K
377 respectively for 100 ps. The two surfaces were relatively stable at these
378 temperatures when no CO was present, but experienced increased surface
379 mobility on addition of CO. Each surface was then dosed with different concentrations of CO
380 that was initially placed in the vacuum region. Upon full adsorption,
381 these concentrations correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
382 coverage. Higher coverages resulted in the formation of a double layer of CO,
383 which introduces artifacts that are not relevant to (557) reconstruction.
384 Because of the difference in binding energies, nearly all of the CO was bound to the Pt surface, while
385 the Au surfaces often had a significant CO population in the gas
386 phase. These systems were allowed to reach thermal equilibrium (over
387 5~ns) before being run in the microcanonical (NVE) ensemble for
388 data collection. All of the systems examined had at least 40~ns in the
389 data collection stage, although simulation times for some Pt of the
390 systems exceeded 200~ns. Simulations were carried out using the open
391 source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
392
393
394
395
396 % RESULTS
397 %
398 \section{Results}
399 \subsection{Structural remodeling}
400 The bare metal surfaces experienced minor roughening of the
401 step-edge because of the elevated temperatures, but the (557)
402 face was stable throughout the simulations. The surface of both
403 systems, upon dosage of CO, began to undergo extensive remodeling
404 that was not observed in the bare systems. Reconstructions of
405 the Au systems were limited to breakup of the step-edges and
406 some step wandering. The lower coverage Pt systems experienced
407 similar restructuring but to a greater extent. The 50\% coverage
408 Pt system was unique among our simulations in that it formed
409 well-defined and stable double layers through step coalescence,
410 similar to results reported by Tao {\it et al}.\cite{Tao:2010}
411
412
413 \subsubsection{Step wandering}
414 The 0\% coverage surfaces for both metals showed minimal
415 step-wandering at their respective temperatures. As the CO
416 coverage increased however, the mobility of the surface atoms,
417 described through adatom diffusion and step-edge wandering,
418 also increased. Except for the 50\% Pt system where step
419 coalescence occurred, the step-edges in the other simulations
420 preferred to keep nearly the same distance between steps as in
421 the original (557) lattice, $\sim$13\AA~for Pt and $\sim$14\AA~for Au.
422 Previous work by Williams {\it et al}.\cite{Williams:1991, Williams:1994}
423 highlights the repulsion that exists between step-edges even
424 when no direct interactions are present in the system. This
425 repulsion is caused by an entropic barrier that arises from
426 the fact that steps cannot cross over one another. This entropic
427 repulsion does not completely define the interactions between
428 steps, however, so it is possible to observe step coalescence
429 on some surfaces.\cite{Williams:1991} The presence and
430 concentration of adsorbates, as shown in this work, can
431 affect step-step interactions, potentially leading to a new
432 surface structure as the thermodynamic equilibrium.
433
434 \subsubsection{Double layers}
435 Tao {\it et al}.\cite{Tao:2010} have shown experimentally that the Pt(557) surface
436 undergoes two separate reconstructions upon CO adsorption.
437 The first involves a doubling of the step height and plateau length.
438 Similar behavior has been seen on a number of surfaces
439 at varying conditions, including Ni(977) and Si(111).\cite{Williams:1994,Williams:1991,Pearl}
440 Of the two systems we examined, the Pt system showed a greater
441 propensity for reconstruction
442 because of the larger surface mobility and the greater extent of step wandering.
443 The amount of reconstruction was strongly correlated to the amount of CO
444 adsorbed upon the surface. This appears to be related to the
445 effect that adsorbate coverage has on edge breakup and on the
446 surface diffusion of metal adatoms. Only the 50\% Pt surface underwent the
447 doubling seen by Tao {\it et al}.\cite{Tao:2010} within the time scales studied here.
448 Over a longer time scale (150~ns) two more double layers formed
449 on this surface. Although double layer formation did not occur
450 in the other Pt systems, they exhibited more step-wandering and
451 roughening compared to their Au counterparts. The
452 50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
453 various times along the simulation showing the evolution of a double layer step-edge.
454
455 The second reconstruction observed by
456 Tao {\it et al}.\cite{Tao:2010} involved the formation of triangular clusters that stretched
457 across the plateau between two step-edges. Neither metal, within
458 the 40~ns time scale or the extended simulation time of 150~ns for
459 the 50\% Pt system, experienced this reconstruction.
460
461 %Evolution of surface
462 \begin{figure}[H]
463 \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
464 \caption{The Pt(557) / 50\% CO system at a sequence of times after
465 initial exposure to the CO: (a) 258~ps, (b) 19~ns, (c) 31.2~ns, and
466 (d) 86.1~ns. Disruption of the (557) step-edges occurs quickly. The
467 doubling of the layers appears only after two adjacent step-edges
468 touch. The circled spot in (b) nucleated the growth of the double
469 step observed in the later configurations.}
470 \label{fig:reconstruct}
471 \end{figure}
472
473 \subsection{Dynamics}
474 Previous experimental work by Pearl and Sibener\cite{Pearl},
475 using STM, has been able to capture the coalescence of steps
476 on Ni(977). The time scale of the image acquisition, $\sim$70~s/image,
477 provides an upper bound for the time required for the doubling
478 to occur. By utilizing Molecular Dynamics we are able to probe
479 the dynamics of these reconstructions at elevated temperatures
480 and in this section we provide data on the timescales for transport
481 properties, e.g. diffusion and layer formation time.
482
483
484 \subsubsection{Transport of surface metal atoms}
485 %forcedSystems/stepSeparation
486 The wandering of a step-edge is a cooperative effect
487 arising from the individual movements of the atoms making up the steps. An ideal metal surface
488 displaying a low index facet, (111) or (100), is unlikely to experience
489 much surface diffusion because of the large energetic barrier that must
490 be overcome to lift an atom out of the surface. The presence of step-edges and other surface features
491 on higher-index facets provides a lower energy source for mobile metal atoms.
492 Single-atom break-away from a step-edge on a clean surface still imposes an
493 energetic penalty around $\sim$~45 kcal/mol, but this is easier than lifting
494 the same metal atom vertically out of the surface, \textgreater~60 kcal/mol.
495 The penalty lowers significantly when CO is present in sufficient quantities
496 on the surface. For certain distributions of CO, see Discussion, the penalty can fall to as low as
497 $\sim$~20 kcal/mol. Once an adatom exists on the surface, the barrier for
498 diffusion is negligible (\textless~4 kcal/mol for a Pt adatom). These adatoms are then
499 able to explore the terrace before rejoining either their original step-edge or
500 becoming a part of a different edge. It is an energetically unfavorable process with a high barrier for an atom
501 to traverse to a separate terrace although the presence of CO can lower the
502 energy barrier required to lift or lower an adatom. By tracking the mobility of individual
503 metal atoms on the Pt and Au surfaces we were able to determine the relative
504 diffusion constants, as well as how varying coverages of CO affect the diffusion. Close
505 observation of the mobile metal atoms showed that they were typically in
506 equilibrium with the step-edges.
507 At times, their motion was concerted and two or more adatoms would be
508 observed moving together across the surfaces.
509
510 A particle was considered ``mobile'' once it had traveled more than 2~\AA~
511 between saved configurations of the system (typically 10-100 ps). A mobile atom
512 would typically travel much greater distances than this, but the 2~\AA~cutoff
513 was used to prevent swamping the diffusion data with the in-place vibrational
514 movement of buried atoms. Diffusion on a surface is strongly affected by
515 local structures and in this work, the presence of single and double layer
516 step-edges causes the diffusion parallel to the step-edges to be larger than
517 the diffusion perpendicular to these edges. Parallel and perpendicular
518 diffusion constants are shown in Figure \ref{fig:diff}.
519
520 %Diffusion graph
521 \begin{figure}[H]
522 \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade_20ns.pdf}
523 \caption{Diffusion constants for mobile surface atoms along directions
524 parallel ($\mathbf{D}_{\parallel}$) and perpendicular
525 ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
526 surface coverage. Diffusion parallel to the step-edge is higher
527 than that perpendicular to the edge because of the lower energy
528 barrier associated with traversing along the edge as compared to
529 completely breaking away. The two reported diffusion constants for
530 the 50\% Pt system arise from different sample sets. The lower values
531 correspond to the same 40~ns amount that all of the other systems were
532 examined at, while the larger values correspond to a 20~ns period }
533 \label{fig:diff}
534 \end{figure}
535
536 The weaker Au-CO interaction is evident in the weak CO-coverage
537 dependance of Au diffusion. This weak interaction leads to lower
538 observed coverages when compared to dosage amounts. This further
539 limits the effect the CO can have on surface diffusion. The correlation
540 between coverage and Pt diffusion rates shows a near linear relationship
541 at the earliest times in the simulations. Following double layer formation,
542 however, there is a precipitous drop in adatom diffusion. As the double
543 layer forms, many atoms that had been tracked for mobility data have
544 now been buried resulting in a smaller reported diffusion constant. A
545 secondary effect of higher coverages is CO-CO cross interactions that
546 lower the effective mobility of the Pt adatoms that are bound to each CO.
547 This effect would become evident only at higher coverages. A detailed
548 account of Pt adatom energetics follows in the Discussion.
549
550
551 \subsubsection{Dynamics of double layer formation}
552 The increased diffusion on Pt at the higher CO coverages is the primary
553 contributor to double layer formation. However, this is not a complete
554 explanation -- the 33\%~Pt system has higher diffusion constants, but
555 did not show any signs of edge doubling in 40~ns. On the 50\%~Pt
556 system, one double layer formed within the first 40~ns of simulation time,
557 while two more were formed as the system was allowed to run for an
558 additional 110~ns (150~ns total). This suggests that this reconstruction
559 is a rapid process and that the previously mentioned upper bound is a
560 very large overestimate.\cite{Williams:1991,Pearl} In this system the first
561 appearance of a double layer appears at 19~ns into the simulation.
562 Within 12~ns of this nucleation event, nearly half of the step has formed
563 the double layer and by 86~ns the complete layer has flattened out.
564 From the appearance of the first nucleation event to the first observed
565 double layer, the process took $\sim$20~ns. Another $\sim$40~ns was
566 necessary for the layer to completely straighten. The other two layers in
567 this simulation formed over periods of 22~ns and 42~ns respectively.
568 A possible explanation for this rapid reconstruction is the elevated
569 temperatures under which our systems were simulated. The process
570 would almost certainly take longer at lower temperatures. Additionally,
571 our measured times for completion of the doubling after the appearance
572 of a nucleation site are likely affected by our periodic boxes. A longer
573 step-edge will likely take longer to ``zipper''.
574
575
576 %Discussion
577 \section{Discussion}
578 We have shown that a classical potential model is able to model the
579 initial reconstruction of the Pt(557) surface upon CO adsorption as
580 shown by Tao {\it et al}.\cite{Tao:2010}. More importantly, we were
581 able to observe features of the dynamic processes necessary for
582 this reconstruction. Here we discuss the features of the model that
583 give rise to the observed dynamical properties of the (557) reconstruction.
584
585 \subsection{Diffusion}
586 The perpendicular diffusion constant
587 appears to be the most important indicator of double layer
588 formation. As highlighted in Figure \ref{fig:reconstruct}, the
589 formation of the double layer did not begin until a nucleation
590 site appeared. And as mentioned by Williams {\it et al}.\cite{Williams:1991, Williams:1994},
591 the inability for edges to cross leads to an effective edge-edge repulsion that
592 must be overcome to allow step coalescence.
593 A greater $\textbf{D}_\perp$ implies more step-wandering
594 and a larger chance for the stochastic meeting of two edges
595 to create a nucleation point. Parallel diffusion along the step-edge can help ``zipper'' up a nascent double
596 layer. This helps explain why the time scale for formation after
597 the appearance of a nucleation site was rapid, while the initial
598 appearance of the nucleation site was unpredictable.
599
600 \subsection{Mechanism for restructuring}
601 Since the Au surface showed no large scale restructuring in any of
602 our simulations, our discussion will focus on the 50\% Pt-CO system
603 which did exhibit doubling featured in Figure \ref{fig:reconstruct}. A
604 number of possible mechanisms exist to explain the role of adsorbed
605 CO in restructuring the Pt surface. Quadrupolar repulsion between
606 adjacent CO molecules adsorbed on the surface is one possibility.
607 However, the quadrupole-quadrupole interaction is short-ranged and
608 is attractive for some orientations. If the CO molecules are ``locked'' in
609 a specific orientation relative to each other, through atop adsorption for
610 example, this explanation would gain credence. The energetic repulsion
611 between two CO molecules located a distance of 2.77~\AA~apart
612 (nearest-neighbor distance of Pt) and both in a vertical orientation,
613 is 8.62 kcal/mol. Moving the CO to the second nearest-neighbor distance
614 of 4.8~\AA~drops the repulsion to nearly 0. Allowing the CO to rotate away
615 from a purely vertical orientation also lowers the repulsion. When the
616 carbons are locked at a distance of 2.77~\AA, a minimum of 6.2 kcal/mol is
617 reached when the angle between the 2 CO is $\sim$24\textsuperscript{o}.
618 The barrier for surface diffusion of a Pt adatom is only 4 kcal/mol, so
619 repulsion between adjacent CO molecules bound to Pt could increase the surface
620 diffusion. However, the residence time of CO on Pt suggests that these
621 molecules are extremely mobile, with diffusion constants 40 to 2500 times
622 larger than surface Pt atoms. This mobility suggests that the CO are more
623 likely to shift their positions without dragging the Pt along with them.
624
625 A different interpretation of the above mechanism, taking into account the large
626 mobility of the CO, looks at how instantaneous and short-lived configurations of
627 CO on the surface can destabilize Pt-Pt interactions leading to increased step-edge
628 breakup and diffusion. On the bare Pt(557) surface the barrier to completely detach
629 an edge atom is $\sim$~43~kcal/mol, as is shown in configuration (a) in Figures
630 \ref{fig:SketchGraphic} \& \ref{fig:SketchEnergies}. For certain configurations, cases
631 (e), (g), and (h), the barrier can be lowered to $\sim$~23~kcal/mole. In these instances,
632 it becomes quite energetically favorable to roughen the edge by introducing a small
633 separation of 0.5 to 1.0~\AA. This roughening becomes immediately obvious in
634 simulations with significant CO populations, although it is present to a lesser extent
635 on lower coverage surfaces and even on the bare surfaces. In these cases it is likely
636 due to stochastic vibrational processes that squeeze out step-edge atoms. The mechanism
637 of step-edge breakup suggested by these energy curves is one the most difficult
638 processes, a complete break-away from the step-edge in one unbroken movement.
639 Easier multistep mechanisms likely exist where an adatom moves laterally on the surface
640 after being ejected so it is sitting on the edge. This provides the atom with 5 nearest
641 neighbors, which while lower than the 7 if it had stayed a part of the step-edge, is higher
642 than the 3 it could maintain located on the terrace. In this proposed mechanism, the CO
643 quadrupolar repulsion is still playing a primary role, but for its importance in roughening
644 the step-edge, rather than maintaining long-term bonds with a single Pt atom which is not
645 born out by their mobility data. The requirement for a large density of CO on the surface
646 for some of the more favorable suggested mechanisms in Figure \ref{fig:SketchGraphic}
647 correspond well with the increased mobility seen on higher coverage surfaces.
648
649 %Sketch graphic of different configurations
650 \begin{figure}[H]
651 \includegraphics[width=0.8\linewidth, height=0.8\textheight]{COpathsSketch.pdf}
652 \caption{The dark grey atoms refer to the upper ledge, while the white atoms are
653 the lower terrace. The blue highlighted atoms had a CO in a vertical atop position
654 upon them. These are a sampling of the configurations examined to gain a more
655 complete understanding of the effects CO has on surface diffusion and edge breakup.
656 Energies associated with each configuration are displayed in Figure \ref{fig:SketchEnergies}.}
657 \label{fig:SketchGraphic}
658 \end{figure}
659
660 %energy graph corresponding to sketch graphic
661 \begin{figure}[H]
662 \includegraphics[width=\linewidth]{stepSeparationComparison.pdf}
663 \caption{The energy curves directly correspond to the labeled model
664 surface in Figure \ref{fig:SketchGraphic}. All energy curves are relative
665 to their initial configuration so the energy of a and h do not have the
666 same zero value. As is seen, certain arrangements of CO can lower
667 the energetic barrier that must be overcome to create an adatom.
668 However, it is the highest coverages where these higher-energy
669 configurations of CO will be more likely. }
670 \label{fig:SketchEnergies}
671 \end{figure}
672
673 While configurations of CO on the surface are able to increase diffusion,
674 this does not immediately provide an explanation for the formation of double
675 layers. If adatoms were constrained to their terrace then doubling would be
676 much less likely to occur. Nucleation sites could still potentially form, but there
677 would not be enough atoms to finish the doubling. Real materials, where the
678 step lengths can be taken as infinite, local doubling would be possible, but in
679 our simulations with our periodic treatment of the system, this is not possible.
680 Thus, there must be a mechanism that explains how adatoms are able to move
681 amongst terraces. Figure \ref{fig:lambda} shows points along a reaction coordinate
682 where an adatom along the step-edge with an adsorbed CO ``burrows'' into the
683 edge displacing an atom onto the higher terrace. This mechanism was chosen
684 because of similar events that were observed during the simulations. The barrier
685 heights we obtained are only approximations because we constrained the movement
686 of the highlighted atoms along a specific concerted path. The calculated $\Delta E$'s
687 are the more interesting results from this investigation. When CO is not present and
688 this reaction coordinate is followed, the $\Delta E > 3$~kcal/mol. The example shown
689 in the figure, where CO is present in the atop position, has a $\Delta E < -15$~kcal/mol.
690 While the barrier height is comparable to the non-CO case, that is a nearly a 20~kcal/mol
691 difference in energies and moves the process from slightly unfavorable to energetically favorable.
692
693 %lambda progression of Pt -> shoving its way into the step
694 \begin{figure}[H]
695 \includegraphics[width=\linewidth]{lambdaProgression_atopCO_withLambda.png}
696 \caption{A model system of the Pt(557) surface was used as the framework
697 for exploring energy barriers along a reaction coordinate. Various numbers,
698 placements, and rotations of CO were examined as they affect Pt movement.
699 The coordinate displayed in this Figure was a representative run. relative to the energy of the system at 0\%, there
700 is a slight decrease upon insertion of the Pt atom into the step-edge along
701 with the resultant lifting of the other Pt atom when CO is present at certain positions.}
702 \label{fig:lambda}
703 \end{figure}
704
705 The mechanism for doubling on this surface appears to be a convolution of at least
706 these two described processes. For complete doubling of a layer to occur there must
707 be the equivalent removal of a separate terrace. For those atoms to ``disappear'' from
708 that terrace they must either rise up on the ledge above them or drop to the ledge below
709 them. The presence of CO helps with both of these situations. There must be sufficient
710 breakage of the step-edge to increase the concentration of adatoms on the surface.
711 These adatoms must then undergo the burrowing highlighted above or some comparable
712 mechanism to traverse the step-edge. Over time, these mechanisms working in concert
713 led to the formation of a double layer.
714
715 \subsection{CO Removal and double layer stability}
716 Once a double layer had formed on the 50\%~Pt system it
717 remained for the rest of the simulation time with minimal
718 movement. There were configurations that showed small
719 wells or peaks forming, but typically within a few nanoseconds
720 the feature would smooth away. Within our simulation time,
721 the formation of the double layer was irreversible and a double
722 layer was never observed to split back into two single layer
723 step-edges while CO was present. To further gauge the effect
724 CO had on this system, additional simulations were run starting
725 from a late configuration of the 50\%~Pt system that had formed
726 double layers. These simulations then had their CO removed.
727 The double layer breaks rapidly in these simulations, already
728 showing a well-defined splitting after 100~ps. Configurations of
729 this system are shown in Figure \ref{fig:breaking}. The coloring
730 of the top and bottom layers helps to exhibit how much mixing
731 the edges experience as they split. These systems were only
732 examined briefly, 10~ns, and within that time despite the initial
733 rapid splitting, the edges only moved another few \AA~apart.
734 It is possible with longer simulation times that the
735 (557) lattice could be recovered as seen by Tao {\it et al}.\cite{Tao:2010}
736
737
738
739 %breaking of the double layer upon removal of CO
740 \begin{figure}[H]
741 \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
742 \caption{(A) 0~ps, (B) 100~ps, (C) 1~ns, after the removal of CO. The presence of the CO
743 helped maintain the stability of the double layer and upon removal the two layers break
744 and begin separating. The separation is not a simple pulling apart however, rather
745 there is a mixing of the lower and upper atoms at the edge.}
746 \label{fig:breaking}
747 \end{figure}
748
749
750
751
752 %Peaks!
753 %\begin{figure}[H]
754 %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
755 %\caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
756 %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
757 %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
758 %\label{fig:peaks}
759 %\end{figure}
760
761
762 %Don't think I need this
763 %clean surface...
764 %\begin{figure}[H]
765 %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
766 %\caption{}
767
768 %\end{figure}
769 %\label{fig:clean}
770
771
772 \section{Conclusion}
773 In this work we have shown the reconstruction of the Pt(557) crystalline surface upon adsorption of CO in less than a $\mu s$. Only the highest coverage Pt system showed this initial reconstruction similar to that seen previously. The strong interaction between Pt and CO and the limited interaction between Au and CO helps explain the differences between the two systems.
774
775 %Things I am not ready to remove yet
776
777 %Table of Diffusion Constants
778 %Add gold?M
779 % \begin{table}[H]
780 % \caption{}
781 % \centering
782 % \begin{tabular}{| c | cc | cc | }
783 % \hline
784 % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
785 % \hline
786 % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
787 % \hline
788 % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
789 % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
790 % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
791 % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
792 % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
793 % \hline
794 % \end{tabular}
795 % \end{table}
796
797 \begin{acknowledgement}
798 Support for this project was provided by the National Science
799 Foundation under grant CHE-0848243 and by the Center for Sustainable
800 Energy at Notre Dame (cSEND). Computational time was provided by the
801 Center for Research Computing (CRC) at the University of Notre Dame.
802 \end{acknowledgement}
803 \newpage
804 \bibliography{firstTryBibliography}
805 %\end{doublespace}
806
807 \begin{tocentry}
808 %\includegraphics[height=3.5cm]{timelapse}
809 \end{tocentry}
810
811 \end{document}