ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3872
Committed: Mon Mar 11 22:37:32 2013 UTC (11 years, 3 months ago) by jmichalk
Content type: application/x-tex
Original Path: trunk/COonPt/firstTry.tex
File size: 36233 byte(s)
Log Message:
Starting to restructure results section

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{times}
5 \usepackage{mathptm}
6 \usepackage{setspace}
7 \usepackage{endfloat}
8 \usepackage{caption}
9 %\usepackage{tabularx}
10 \usepackage{graphicx}
11 \usepackage{multirow}
12 %\usepackage{booktabs}
13 %\usepackage{bibentry}
14 %\usepackage{mathrsfs}
15 \usepackage[square, comma, sort&compress]{natbib}
16 \usepackage{url}
17 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
18 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
19 9.0in \textwidth 6.5in \brokenpenalty=10000
20
21 % double space list of tables and figures
22 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
23 \setlength{\abovecaptionskip}{20 pt}
24 \setlength{\belowcaptionskip}{30 pt}
25
26 \bibpunct{}{}{,}{s}{}{;}
27 \bibliographystyle{achemso}
28
29 \begin{document}
30
31
32 %%
33 %Introduction
34 % Experimental observations
35 % Previous work on Pt, CO, etc.
36 %
37 %Simulation Methodology
38 % FF (fits and parameters)
39 % MD (setup, equilibration, collection)
40 %
41 % Analysis of trajectories!!!
42 %Discussion
43 % CO preferences for specific locales
44 % CO-CO interactions
45 % Differences between Au & Pt
46 % Causes of 2_layer reordering in Pt
47 %Summary
48 %%
49
50 %Title
51 \title{Molecular Dynamics simulations of the surface reconstructions
52 of Pt(557) and Au(557) under exposure to CO}
53
54 \author{Joseph R. Michalka, Patrick W. McIntyre and J. Daniel
55 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
56 Department of Chemistry and Biochemistry,\\
57 University of Notre Dame\\
58 Notre Dame, Indiana 46556}
59
60 %Date
61 \date{Mar 5, 2013}
62
63 %authors
64
65 % make the title
66 \maketitle
67
68 \begin{doublespace}
69
70 \begin{abstract}
71 We examine surface reconstructions of Pt and Au(557) under
72 various CO coverages using molecular dynamics in order to
73 explore possible mechanisms for any observed reconstructions
74 and their dynamics. The metal-CO interactions were parameterized
75 as part of this work so that an efficient large-scale treatment of
76 this system could be undertaken. The large difference in binding
77 strengths of the metal-CO interactions was found to play a significant
78 role with regards to step-edge stability and adatom diffusion. A
79 small correlation between coverage and the diffusion constant
80 was also determined. The energetics of CO adsorbed to the surface
81 is sufficient to explain the reconstructions observed on the Pt
82 systems and the lack of reconstruction of the Au systems.
83
84 \end{abstract}
85
86 \newpage
87
88
89 \section{Introduction}
90 % Importance: catalytically active metals are important
91 % Sub: Knowledge of how their surface structure affects their ability to catalytically facilitate certain reactions is growing, but is more reactionary than predictive
92 % 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)
93 % Theory can explore temperatures and pressures which are difficult to work with in experiments
94 % Sub: Also, easier to observe what is going on and provide reasons and explanations
95 %
96
97 Industrial catalysts usually consist of small particles that exhibit a
98 high concentration of steps, kink sites, and vacancies at the edges of
99 the facets. These sites are thought to be the locations of catalytic
100 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
101 significant evidence that solid surfaces are often structurally,
102 compositionally, and chemically modified by reactants under operating
103 conditions.\cite{Tao2008,Tao:2010,Tao2011} The coupling between
104 surface oxidation states and catalytic activity for CO oxidation on
105 Pt, for instance, is widely documented.\cite{Ertl08,Hendriksen:2002}
106 Despite the well-documented role of these effects on reactivity, the
107 ability to capture or predict them in atomistic models is somewhat
108 limited. While these effects are perhaps unsurprising on the highly
109 disperse, multi-faceted nanoscale particles that characterize
110 industrial catalysts, they are manifest even on ordered, well-defined
111 surfaces. The Pt(557) surface, for example, exhibits substantial and
112 reversible restructuring under exposure to moderate pressures of
113 carbon monoxide.\cite{Tao:2010}
114
115 This work is an investigation into the mechanism and timescale for
116 surface restructuring using molecular simulations. Since the dynamics
117 of the process are of particular interest, we employ classical force
118 fields that represent a compromise between chemical accuracy and the
119 computational efficiency necessary to simulate the process of interest.
120 Since restructuring typically occurs as a result of specific interactions of the
121 catalyst with adsorbates, in this work, two metal systems exposed
122 to carbon monoxide were examined. The Pt(557) surface has already been shown
123 to undergo a large scale reconstruction under certain conditions.\cite{Tao:2010}
124 The Au(557) surface, because of a weaker interaction with CO, is seen as less
125 likely to undergo this kind of reconstruction. However, Peters et al.\cite{Peters:2000}
126 and Piccolo et al.\cite{Piccolo:2004} have both observed CO-induced
127 reconstruction of a Au(111) surface. Peters et al. saw a relaxation to the
128 22 x $\sqrt{3}$ cell. They argued that only a few Au atoms
129 become adatoms, limiting the stress of this reconstruction while
130 allowing the rest to relax and approach the ideal (111)
131 configuration. They did not see the usual herringbone pattern being greatly
132 affected by this relaxation. Piccolo et al. on the other hand, did see a
133 disruption of the herringbone pattern as CO was adsorbed to the
134 surface. Both groups suggested that the preference CO shows for
135 low-coordinated Au atoms was the primary driving force for the reconstruction.
136
137
138
139 %Platinum molecular dynamics
140 %gold molecular dynamics
141
142 \section{Simulation Methods}
143 The challenge in modeling any solid/gas interface is the
144 development of a sufficiently general yet computationally tractable
145 model of the chemical interactions between the surface atoms and
146 adsorbates. Since the interfaces involved are quite large (10$^3$ -
147 10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
148 molecular dynamics
149 (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
150 methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
151 mechanical potential energy surfaces remain out of reach.
152 Additionally, the ``bonds'' between metal atoms at a surface are
153 typically not well represented in terms of classical pairwise
154 interactions in the same way that bonds in a molecular material are,
155 nor are they captured by simple non-directional interactions like the
156 Coulomb potential. For this work, we have used classical molecular
157 dynamics with potential energy surfaces that are specifically tuned
158 for transition metals. In particular, we used the EAM potential for
159 Au-Au and Pt-Pt interactions\cite{EAM}. The CO was modeled using a rigid
160 three-site model developed by Straub and Karplus for studying
161 photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
162 Pt-CO cross interactions were parameterized as part of this work.
163
164 \subsection{Metal-metal interactions}
165 Many of the potentials used for modeling transition metals are based
166 on a non-pairwise additive functional of the local electron
167 density. The embedded atom method (EAM) is perhaps the best known of
168 these
169 methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
170 but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
171 the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
172 parameter sets. The glue model of Ercolessi et al. is among the
173 fastest of these density functional approaches.\cite{Ercolessi88} In
174 all of these models, atoms are conceptualized as a positively charged
175 core with a radially-decaying valence electron distribution. To
176 calculate the energy for embedding the core at a particular location,
177 the electron density due to the valence electrons at all of the other
178 atomic sites is computed at atom $i$'s location,
179 \begin{equation*}
180 \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
181 \end{equation*}
182 Here, $\rho_j(r_{ij})$ is the function that describes the distance
183 dependence of the valence electron distribution of atom $j$. The
184 contribution to the potential that comes from placing atom $i$ at that
185 location is then
186 \begin{equation*}
187 V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
188 \end{equation*}
189 where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
190 $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
191 repulsive overlap of the two positively charged cores.
192
193 % The {\it modified} embedded atom method (MEAM) adds angular terms to
194 % the electron density functions and an angular screening factor to the
195 % pairwise interaction between two
196 % atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
197 % MEAM has become widely used to simulate systems in which angular
198 % interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
199 % metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
200 % MEAM presents significant additional computational costs, however.
201
202 The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
203 have all been widely used by the materials simulation community for
204 simulations of bulk and nanoparticle
205 properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
206 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
207 fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
208 propagation,\cite{BECQUART:1993rg} and alloying
209 dynamics.\cite{Shibata:2002hh} One of EAM's strengths
210 is its sensitivity to small changes in structure. This arises
211 from the original parameterization, where the interactions
212 up to the third nearest neighbor were taken into account.\cite{Voter95a}
213 Comparing that to the glue model of Ercolessi et al.\cite{Ercolessi88}
214 which is only parameterized up to the nearest-neighbor
215 interactions, EAM is a suitable choice for systems where
216 the bulk properties are of secondary importance to low-index
217 surface structures. Additionally, the similarity of EAMs functional
218 treatment of the embedding energy to standard density functional
219 theory (DFT) makes fitting DFT-derived cross potentials with adsorbates somewhat easier.
220 \cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
221
222
223
224
225 \subsection{Carbon Monoxide model}
226 Previous explanations for the surface rearrangements center on
227 the large linear quadrupole moment of carbon monoxide.\cite{Tao:2010}
228 We used a model first proposed by Karplus and Straub to study
229 the photodissociation of CO from myoglobin because it reproduces
230 the quadrupole moment well.\cite{Straub} The Straub and
231 Karplus model treats CO as a rigid three site molecule with a massless M
232 site at the molecular center of mass. The geometry and interaction
233 parameters are reproduced in Table~\ref{tab:CO}. The effective
234 dipole moment, calculated from the assigned charges, is still
235 small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
236 to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
237 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
238 %CO Table
239 \begin{table}[H]
240 \caption{Positions, Lennard-Jones parameters ($\sigma$ and
241 $\epsilon$), and charges for the CO-CO
242 interactions in Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
243 in kcal/mol, and charges are in atomic units.}
244 \centering
245 \begin{tabular}{| c | c | ccc |}
246 \hline
247 & {\it z} & $\sigma$ & $\epsilon$ & q\\
248 \hline
249 \textbf{C} & -0.6457 & 3.83 & 0.0262 & -0.75 \\
250 \textbf{O} & 0.4843 & 3.12 & 0.1591 & -0.85 \\
251 \textbf{M} & 0.0 & - & - & 1.6 \\
252 \hline
253 \end{tabular}
254 \label{tab:CO}
255 \end{table}
256
257 \subsection{Cross-Interactions between the metals and carbon monoxide}
258
259 Since the adsorption of CO onto a Pt surface has been the focus
260 of much experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
261 and theoretical work
262 \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
263 there is a significant amount of data on adsorption energies for CO on
264 clean metal surfaces. An earlier model by Korzeniewski {\it et
265 al.}\cite{Pons:1986} served as a starting point for our fits. The parameters were
266 modified to ensure that the Pt-CO interaction favored the atop binding
267 position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}.
268 The modified parameters yield binding energies that are slightly higher
269 than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
270 et al.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
271 Lennard-Jones interaction to mimic strong, but short-ranged partial
272 binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
273 Pt-O interaction was modeled with a Morse potential with a large
274 equilibrium distance, ($r_o$). These choices ensure that the C is preferred
275 over O as the surface-binding atom. In most cases, the Pt-O parameterization contributes a weak
276 repulsion which favors the atop site. The resulting potential-energy
277 surface suitably recovers the calculated Pt-C separation length
278 (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
279 position.\cite{Deshlahra:2012, Hopster:1978}
280
281 %where did you actually get the functionals for citation?
282 %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
283 %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
284 The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
285 Morse potentials, respectively, to reproduce Au-CO binding energies.
286 The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
287 Adsorption energies were obtained from gas-surface DFT calculations with a
288 periodic supercell plane-wave basis approach, as implemented in the
289 {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores were
290 described with the projector augmented-wave (PAW)
291 method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
292 included to an energy cutoff of 20 Ry. Electronic energies are
293 computed with the PBE implementation of the generalized gradient
294 approximation (GGA) for gold, carbon, and oxygen that was constructed
295 by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
296 In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
297 Au x 2 Au surface planes and separated from vertical images by six
298 layers of vacuum space. The surface atoms were all allowed to relax
299 before CO was added to the system. Electronic relaxations were
300 performed until the energy difference between subsequent steps
301 was less than $10^{-8}$ Ry. Nonspin-polarized supercell calculations
302 were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
303 zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxed gold slab was
304 then used in numerous single point calculations with CO at various
305 heights (and angles relative to the surface) to allow fitting of the
306 empirical force field.
307
308 %Hint at future work
309 The parameters employed for the metal-CO cross-interactions in this work
310 are shown in Table~\ref{tab:co_parameters} and the binding energies on the
311 (111) surfaces are displayed in Table~\ref{tab:co_energies}. Charge transfer
312 and polarization are neglected in this model, although these effects are likely to
313 affect binding energies and binding site preferences, and will be addressed in
314 future work.
315
316 %Table of Parameters
317 %Pt Parameter Set 9
318 %Au Parameter Set 35
319 \begin{table}[H]
320 \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
321 interactions are modeled with Lennard-Jones potentials. While the
322 metal-O interactions were fit to Morse
323 potentials. Distances are given in \AA~and energies in kcal/mol. }
324 \centering
325 \begin{tabular}{| c | cc | c | ccc |}
326 \hline
327 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
328 \hline
329 \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
330 \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
331
332 \hline
333 \end{tabular}
334 \label{tab:co_parameters}
335 \end{table}
336
337 %Table of energies
338 \begin{table}[H]
339 \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
340 described in this work. All values are in eV.}
341 \centering
342 \begin{tabular}{| c | cc |}
343 \hline
344 & Calculated & Experimental \\
345 \hline
346 \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
347 (Ref. \protect\cite{Kelemen:1979}) \\
348 & & -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
349 \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{TPD_Gold}) \\
350 \hline
351 \end{tabular}
352 \label{tab:co_energies}
353 \end{table}
354
355 \subsection{Pt(557) and Au(557) metal interfaces}
356 Our Pt system is an orthorhombic periodic box of dimensions
357 54.482~x~50.046~x~120.88~\AA~while our Au system has
358 dimensions of 57.4~x~51.9285~x~100~\AA.
359 The systems are arranged in a FCC crystal that have been cut
360 along the (557) plane so that they are periodic in the {\it x} and
361 {\it y} directions, and have been oriented to expose two aligned
362 (557) cuts along the extended {\it z}-axis. Simulations of the
363 bare metal interfaces at temperatures ranging from 300~K to
364 1200~K were performed to confirm the relative
365 stability of the surfaces without a CO overlayer.
366
367 The different bulk melting temperatures (1337~K for Au
368 and 2045~K for Pt) suggest that any possible reconstruction should happen at
369 different temperatures for the two metals. The bare Au and Pt surfaces were
370 initially run in the canonical (NVT) ensemble at 800~K and 1000~K
371 respectively for 100 ps. The two surfaces were relatively stable at these
372 temperatures when no CO was present, but experienced increased surface
373 mobility on addition of CO. Each surface was then dosed with different concentrations of CO
374 that was initially placed in the vacuum region. Upon full adsorption,
375 these concentrations correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
376 coverage. Higher coverages resulted in the formation of a double layer of CO,
377 which introduces artifacts that are not relevant to (557) reconstruction.
378 Because of the difference in binding energies, nearly all of the CO was bound to the Pt surface, while
379 the Au surfaces often had a significant CO population in the gas
380 phase. These systems were allowed to reach thermal equilibrium (over
381 5 ns) before being run in the microcanonical (NVE) ensemble for
382 data collection. All of the systems examined had at least 40 ns in the
383 data collection stage, although simulation times for some Pt of the
384 systems exceeded 200~ns. Simulations were carried out using the open
385 source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
386
387
388
389
390 % RESULTS
391 %
392 \section{Results}
393 \subsection{Structural remodeling}
394 The surfaces of both systems, upon dosage of CO, began
395 to undergo remodeling that was not observed in the bare
396 metal system. The surfaces to which no CO was exposed
397 did experience minor roughening of the step-edge, but the
398 (557) lattice was well-maintained throughout the simulation
399 time. The Au systems were limited to greater amounts of
400 roughening, i.e. breakup of the step-edge, and some step
401 wandering. The lower coverage Pt systems experienced
402 similar restructuring but to a greater extent when
403 compared to the Au systems. The 50\% coverage
404 Pt system formed double layers at numerous spots upon its surface.
405
406
407 \subsubsection{Step wandering}
408 The 0\% coverage surfaces for both metals showed
409 minimal movement at their respective run temperatures.
410 As the coverage increased, the mobility of the surface
411 also increased. Additionally, at the higher coverages
412 on both metals, there was a large increase in the amount
413 of observed step-wandering. Previous work by
414 Williams\cite{Williams:1993} highlighted the entropic
415 contribution to the repulsion felt between step-edges,
416 and situations were that repulsion could be negated, or
417 overcome, to allow for step coalescence or facet formation.
418
419 \subsubsection{Double layers}
420 Tao et al. have shown experimentally that the Pt(557) surface
421 undergoes two separate reconstructions upon CO
422 adsorption.\cite{Tao:2010} The first involves a doubling of
423 the step height and plateau length. Similar behavior has been
424 seen to occur on numerous surfaces at varying conditions: Ni(977), Si(111).
425 \cite{Williams:1994,Williams:1991,Pearl} Of the two systems
426 we examined, the Pt system showed a greater propensity for
427 reconstruction when compared to the Au system. The amount
428 of reconstruction is correlated to the amount of CO
429 adsorbed upon the surface. This appears to be related to the
430 effect that adsorbate coverage has on edge breakup and on the surface
431 diffusion of metal adatoms. While both systems displayed step-edge
432 wandering, only the Pt surface underwent the doubling seen by
433 Tao et al. within the time scales studied here.
434 Only the 50\% coverage Pt system exhibited
435 a complete doubling in the time scales we
436 were able to monitor. Over longer periods (150~ns) two more double layers formed on this interface.
437 Although double layer formation did not occur in the other Pt systems, they show
438 more lateral movement of the step-edges
439 compared to their Au counterparts. The 50\% Pt system is highlighted
440 in Figure \ref{fig:reconstruct} at various times along the simulation
441 showing the evolution of a step-edge.
442
443 The second reconstruction on the Pt(557) surface observed by
444 Tao involved the formation of triangular clusters that stretched
445 across the plateau between two step-edges. Neither system, within
446 the 40~ns time scale, experienced this reconstruction.
447
448 \subsection{Dynamics}
449 Previous atomistic simulations of stepped surfaces dealt largely
450 with the energetics and structures at different conditions
451 \cite{Williams:1991,Williams:1994}. Consequently, the most common
452 technique utilized to date has been Monte Carlo sampling. Monte Carlo gives an efficient
453 sampling of the equilibrium thermodynamic landscape at the expense
454 of ignoring the dynamics of the system. Previous work by Pearl and
455 Sibener\cite{Pearl}, using STM, has been able to show the coalescing
456 of steps on Ni(977). The time scale of the image acquisition,
457 $\sim$70 s/image provides an upper bound for the time required for
458 the doubling to occur. In this section we give data on dynamic and
459 transport properties, e.g. diffusion, layer formation time, etc.
460
461
462 \subsubsection{Transport of surface metal atoms}
463 %forcedSystems/stepSeparation
464 The movement or wandering of a step-edge is a cooperative effect
465 arising from the individual movements, primarily through surface
466 diffusion, of the atoms making up the steps. An ideal metal surface
467 displaying a low index facet, (111) or (100), is unlikely to experience
468 much surface diffusion because of the large energetic barrier that must
469 be overcome to lift an atom out of the surface. The presence of step-edges
470 on higher-index surfaces provide a source for mobile metal atoms.
471 Breaking away from the step-edge on a clean surface still imposes an
472 energetic penalty around $\sim$~40 kcal/mol, but this is significantly easier than lifting
473 the same metal atom vertically out of the surface, \textgreater~60 kcal/mol.
474 The penalty lowers significantly when CO is present in sufficient quantities
475 on the surface. For certain distributions of CO, the penalty can fall as low as
476 $\sim$~20 kcal/mol. Once an adatom exists on the surface, the barrier for
477 diffusion is negligible ( \textless~4 kcal/mol) and these adatoms are
478 able to explore the terrace before rejoining either the original step-edge or
479 becoming a part of a different edge. It is a more difficult process for an atom
480 to traverse to a separate terrace although the presence of CO can lower the
481 energy barrier required to lift or lower the adatom. By tracking the mobility of individual
482 metal atoms on the Pt and Au surfaces we were able to determine the relative
483 diffusion constants, as well as how varying coverages of CO affect the diffusion. Close
484 observation of the mobile metal atoms showed that they were typically in
485 equilibrium with the step-edges, dynamically breaking apart and rejoining the edges.
486 At times, their motion was concerted and two or more adatoms would be
487 observed moving together across the surfaces.
488
489 A particle was considered ``mobile'' once it had traveled more than 2~\AA~
490 between saved configurations of the system (typically 10-100 ps). An atom that was
491 truly mobile would typically travel much greater distances than this, but the 2~\AA~cutoff
492 was used to prevent swamping the diffusion data with the in-place vibrational
493 movement of buried atoms. Diffusion on a surface is strongly affected by
494 local structures and in this work, the presence of single and double layer
495 step-edges causes the diffusion parallel to the step-edges to be different
496 from the diffusion perpendicular to these edges. Parallel and perpendicular
497 diffusion constants are shown in Figure \ref{fig:diff}.
498
499 \subsubsection{Dynamics of double layer formation}
500 The increased diffusion on Pt at the higher
501 CO coverages plays a primary role in double layer formation. However, this is not
502 a complete explanation -- the 33\%~Pt system
503 has higher diffusion constants but did not show
504 any signs of edge doubling. On the
505 50\%~Pt system, three separate layers were formed over
506 150~ns of simulation time. Previous experimental
507 work gives insight into the upper bounds of the
508 time required for step coalescence.\cite{Williams:1991,Pearl}
509 In this system, as seen in Figure \ref{fig:reconstruct}, the first
510 appearance of a double layer, appears at 19~ns
511 into the simulation. Within 12~ns of this nucleation event, nearly half of the step has
512 formed the double layer and by 86 ns, the complete layer
513 has been flattened out. The double layer could be considered
514 ``complete" by 37~ns but remains a bit rough. From the
515 appearance of the first nucleation event to the first observed double layer, the process took $\sim$20~ns. Another
516 $\sim$40~ns was necessary for the layer to completely straighten.
517 The other two layers in this simulation formed over periods of
518 22~ns and 42~ns respectively. Comparing this to the upper
519 bounds of the image scan, it is likely that most aspects of this
520 reconstruction occur very rapidly. A possible explanation
521 for this rapid reconstruction is the elevated temperatures
522 under which our systems were simulated. It is probable that the process would
523 take longer at lower temperatures.
524
525 %Evolution of surface
526 \begin{figure}[H]
527 \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
528 \caption{The Pt(557) / 50\% CO system at a sequence of times after
529 initial exposure to the CO: (a) 258 ps, (b) 19 ns, (c) 31.2 ns, and
530 (d) 86.1 ns. Disruption of the (557) step-edges occurs quickly. The
531 doubling of the layers appears only after two adjacent step-edges
532 touch. The circled spot in (b) nucleated the growth of the double
533 step observed in the later configurations.}
534 \label{fig:reconstruct}
535 \end{figure}
536
537 \begin{figure}[H]
538 \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade.pdf}
539 \caption{Diffusion constants for mobile surface atoms along directions
540 parallel ($\mathbf{D}_{\parallel}$) and perpendicular
541 ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
542 surface coverage. Diffusion parallel to the step-edge is higher
543 than that perpendicular to the edge because of the lower energy
544 barrier associated with traversing along the edge as compared to
545 completely breaking away. Additionally, the observed
546 maximum and subsequent decrease for the Pt system suggests that the
547 CO self-interactions are playing a significant role with regards to
548 movement of the Pt atoms around and across the surface. }
549 \label{fig:diff}
550 \end{figure}
551
552
553
554
555 %Discussion
556 \section{Discussion}
557 We have shown that the classical potential models are able to model the initial reconstruction of the
558 Pt(557) surface upon CO adsorption as shown by Tao et al. \cite{Tao:2010}. More importantly, we
559 were able to observe features of the dynamic processes necessary for this reconstruction.
560
561 \subsection{Mechanism for restructuring}
562 Since the Au surface showed no large scale restructuring throughout
563 our simulation time our discussion will focus on the 50\% Pt-CO system
564 which did undergo the doubling featured in Figure \ref{fig:reconstruct}.
565 Similarities of our results to those reported previously by
566 Tao et al.\cite{Tao:2010} are quite
567 strong. The simulated Pt
568 system exposed to a large dosage of CO readily restructures by doubling the terrace
569 widths and step heights. The restructuring occurs in a piecemeal fashion, one to two Pt atoms at a time, but is rapid on experimental timescales.
570 The adatoms either
571 break away from the step-edge and stay on the lower terrace or they lift
572 up onto a higher terrace. Once ``free'', they diffuse on the terrace
573 until reaching another step-edge or rejoining their original edge.
574 This combination of growth and decay of the step-edges is in a state of
575 dynamic equilibrium. However, once two previously separated edges
576 meet as shown in Figure 1.B, this nucleates the rest of the edge to meet up, forming a double layer.
577 From simulations which exhibit a double layer, the time delay from the initial appearance of a nucleation point to a fully formed double layer is $\sim$35 ns.
578
579 A number of possible mechanisms exist to explain the role of adsorbed
580 CO in restructuring the Pt surface. Quadrupolar repulsion between adjacent
581 CO molecules adsorbed on the surface is one possibility. However,
582 the quadrupole-quadrupole interaction is short-ranged and is attractive for
583 some orientations. If the CO molecules are ``locked'' in a specific orientation
584 relative to each other, through atop adsorption for example, this explanation
585 gains some credence. The energetic repulsion between two CO located a
586 distance of 2.77~\AA~apart (nearest-neighbor distance of Pt) and both in
587 a vertical orientation, is 8.62 kcal/mol. Moving the CO apart to the second
588 nearest-neighbor distance of 4.8~\AA~or 5.54~\AA~drops the repulsion to
589 nearly 0 kcal/mol. Allowing the CO's to leave a purely vertical orientation
590 also quickly drops the repulsion, a minimum of 6.2 kcal/mol is reached at $\sim$24 degrees between the 2 CO when the carbons are locked at a distance of 2.77 \AA apart.
591 As mentioned above, the energy barrier for surface diffusion
592 of a Pt adatom is only 4 kcal/mol. So this repulsion between neighboring CO molecules can
593 increase the surface diffusion. However, the residence time of CO on Pt was
594 examined and while the majority of the CO is on or near the surface throughout
595 the run, most molecules are mobile. This mobility suggests that the CO are more
596 likely to shift their positions without necessarily the Pt along with them.
597
598 Another possible and more likely mechanism for the restructuring is in the
599 destabilization of strong Pt-Pt interactions by CO adsorbed on surface
600 Pt atoms. This would then have the effect of increasing surface mobility
601 of these atoms. To test this hypothesis, numerous configurations of
602 CO in varying quantities were arranged on the higher and lower plateaus
603 around a step on a otherwise clean Pt(557) surface. One representative
604 configuration is displayed in Figure \ref{fig:lambda}. Single or concerted movement
605 of Pt atoms was then examined to determine possible barriers. Because
606 the movement was forced along a pre-defined reaction coordinate that may differ
607 from the true minimum of this path, only the beginning and ending energies
608 are displayed in Table \ref{tab:energies}. These values suggest that the presence of CO at suitable
609 locations can lead to lowered barriers for Pt breaking apart from the step-edge.
610 Additionally, as highlighted in Figure \ref{fig:lambda}, the presence of CO makes the
611 burrowing and lifting of adatoms favorable, whereas without CO, the process is neutral
612 in terms of energetics.
613
614 %lambda progression of Pt -> shoving its way into the step
615 \begin{figure}[H]
616 \includegraphics[width=\linewidth]{lambdaProgression_atopCO.png}
617 \caption{A model system of the Pt(557) surface was used as the framework
618 for exploring energy barriers along a reaction coordinate. Various numbers,
619 placements, and rotations of CO were examined as they affect Pt movement.
620 The coordinate displayed in this Figure was a representative run. As shown
621 in Table \ref{tab:rxcoord}, relative to the energy of the system at 0\%, there
622 is a slight decrease upon insertion of the Pt atom into the step-edge along
623 with the resultant lifting of the other Pt atom when CO is present at certain positions.}
624 \label{fig:lambda}
625 \end{figure}
626
627
628
629 \subsection{Diffusion}
630 The diffusion parallel to the step-edge tends to be
631 much larger than that perpendicular to the step-edge. The dynamic
632 equilibrium that is established between the step-edge and adatom interface. The coverage
633 of CO also appears to play a slight role in relative rates of diffusion, as shown in Figure \ref{fig:diff}.
634 The
635 Thus, the bottleneck of the double layer formation appears to be the initial formation
636 of this growth point, which seems to be somewhat of a stochastic event. Once it
637 appears, parallel diffusion, along the now slightly angled step-edge, will allow for
638 a faster formation of the double layer than if the entire process were dependent on
639 only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the
640 more likely a growth point is to be formed.
641 \\
642
643
644 %breaking of the double layer upon removal of CO
645 \begin{figure}[H]
646 \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
647 %:
648 \caption{(A) 0 ps, (B) 100 ps, (C) 1 ns, after the removal of CO. The presence of the CO
649 helped maintain the stability of the double layer and upon removal the two layers break
650 and begin separating. The separation is not a simple pulling apart however, rather
651 there is a mixing of the lower and upper atoms at the edge.}
652 \label{fig:breaking}
653 \end{figure}
654
655
656
657
658 %Peaks!
659 %\begin{figure}[H]
660 %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
661 %\caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
662 %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
663 %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
664 %\label{fig:peaks}
665 %\end{figure}
666
667
668 %Don't think I need this
669 %clean surface...
670 %\begin{figure}[H]
671 %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
672 %\caption{}
673
674 %\end{figure}
675 %\label{fig:clean}
676
677
678 \section{Conclusion}
679 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.
680
681 %Things I am not ready to remove yet
682
683 %Table of Diffusion Constants
684 %Add gold?M
685 % \begin{table}[H]
686 % \caption{}
687 % \centering
688 % \begin{tabular}{| c | cc | cc | }
689 % \hline
690 % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
691 % \hline
692 % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
693 % \hline
694 % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
695 % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
696 % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
697 % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
698 % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
699 % \hline
700 % \end{tabular}
701 % \end{table}
702
703 \section{Acknowledgments}
704 Support for this project was provided by the National Science
705 Foundation under grant CHE-0848243 and by the Center for Sustainable
706 Energy at Notre Dame (cSEND). Computational time was provided by the
707 Center for Research Computing (CRC) at the University of Notre Dame.
708
709 \newpage
710 \bibliography{firstTryBibliography}
711 \end{doublespace}
712 \end{document}