ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3877
Committed: Fri Mar 15 13:18:17 2013 UTC (11 years, 3 months ago) by jmichalk
Content type: application/x-tex
Original Path: trunk/COonPt/firstTry.tex
File size: 42284 byte(s)
Log Message:
Most recent pdf->bib->pdf->pdf

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
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 seen as less
130 likely to undergo this kind of reconstruction. However, Peters et al.\cite{Peters:2000}
131 and Piccolo et al.\cite{Piccolo:2004} have both observed CO-induced
132 reconstruction of a Au(111) surface. Peters 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 being greatly
137 affected by this relaxation. Piccolo 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$^6$ 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 et al. is among the
178 fastest of these density functional approaches.\cite{Ercolessi88} In
179 all of these models, atoms are conceptualized 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 from the original parameterization, where the interactions
217 up to the third nearest neighbor were taken into account.\cite{Voter95a}
218 Comparing that to the glue model of Ercolessi 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 EAMs 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 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 cases, 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 are likely to
318 affect binding energies and binding site preferences, and will be addressed in
319 future work.
320
321 %Table of Parameters
322 %Pt Parameter Set 9
323 %Au Parameter Set 35
324 \begin{table}[H]
325 \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
326 interactions are modeled with Lennard-Jones potentials. While the
327 metal-O interactions were fit to Morse
328 potentials. Distances are given in \AA~and energies in kcal/mol. }
329 \centering
330 \begin{tabular}{| c | cc | c | ccc |}
331 \hline
332 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
333 \hline
334 \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
335 \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
336
337 \hline
338 \end{tabular}
339 \label{tab:co_parameters}
340 \end{table}
341
342 %Table of energies
343 \begin{table}[H]
344 \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
345 described in this work. All values are in eV.}
346 \centering
347 \begin{tabular}{| c | cc |}
348 \hline
349 & Calculated & Experimental \\
350 \hline
351 \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
352 (Ref. \protect\cite{Kelemen:1979}) \\
353 & & -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
354 \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{TPDGold}) \\
355 \hline
356 \end{tabular}
357 \label{tab:co_energies}
358 \end{table}
359
360 \subsection{Pt(557) and Au(557) metal interfaces}
361 Our Pt system is an orthorhombic periodic box of dimensions
362 54.482~x~50.046~x~120.88~\AA~while our Au system has
363 dimensions of 57.4~x~51.9285~x~100~\AA.
364 The systems are arranged in a FCC crystal that have been cut
365 along the (557) plane so that they are periodic in the {\it x} and
366 {\it y} directions, and have been oriented to expose two aligned
367 (557) cuts along the extended {\it z}-axis. Simulations of the
368 bare metal interfaces at temperatures ranging from 300~K to
369 1200~K were performed to confirm the relative
370 stability of the surfaces without a CO overlayer.
371
372 The different bulk melting temperatures (1345~$\pm$~10~K for Au\cite{Au:melting}
373 and $\sim$~2045~K for Pt\cite{Pt:melting}) suggest that any possible reconstruction should happen at
374 different temperatures for the two metals. The bare Au and Pt surfaces were
375 initially run in the canonical (NVT) ensemble at 800~K and 1000~K
376 respectively for 100 ps. The two surfaces were relatively stable at these
377 temperatures when no CO was present, but experienced increased surface
378 mobility on addition of CO. Each surface was then dosed with different concentrations of CO
379 that was initially placed in the vacuum region. Upon full adsorption,
380 these concentrations correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
381 coverage. Higher coverages resulted in the formation of a double layer of CO,
382 which introduces artifacts that are not relevant to (557) reconstruction.
383 Because of the difference in binding energies, nearly all of the CO was bound to the Pt surface, while
384 the Au surfaces often had a significant CO population in the gas
385 phase. These systems were allowed to reach thermal equilibrium (over
386 5~ns) before being run in the microcanonical (NVE) ensemble for
387 data collection. All of the systems examined had at least 40~ns in the
388 data collection stage, although simulation times for some Pt of the
389 systems exceeded 200~ns. Simulations were carried out using the open
390 source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
391
392
393
394
395 % RESULTS
396 %
397 \section{Results}
398 \subsection{Structural remodeling}
399 The surfaces of both systems, upon dosage of CO, began
400 to undergo extensive remodeling that was not observed in the bare
401 systems. The bare metal surfaces
402 experienced minor roughening of the step-edge because
403 of the elevated temperatures, but the
404 (557) lattice was well-maintained throughout the simulation
405 time. The Au systems were limited to greater amounts of
406 roughening, i.e. breakup of the step-edge, and some step
407 wandering. The lower coverage Pt systems experienced
408 similar restructuring but to a greater extent when
409 compared to the Au systems. The 50\% coverage
410 Pt system was unique among our simulations in that it
411 formed numerous double layers through step coalescence,
412 similar to results reported by Tao et al.\cite{Tao:2010}
413
414
415 \subsubsection{Step wandering}
416 The 0\% coverage surfaces for both metals showed minimal
417 movement at their respective run temperatures. As the CO
418 coverage increased however, the mobility of the surface,
419 described through adatom diffusion and step-edge wandering,
420 also increased. Except for the 50\% Pt system, the step-edges
421 did not coalesce in any of the other simulations, instead
422 preferring to keep nearly the same distance between steps
423 as in the original (557) lattice, $\sim$13\AA for Pt and $\sim$14\AA for Au.
424 Previous work by Williams et al.\cite{Williams:1991, Williams:1994}
425 highlights the repulsion that exists between step-edges even
426 when no direct interactions are present in the system. This
427 repulsion arises because step-edge crossing is not allowed
428 which constrains the entropy. This entropic repulsion does
429 not completely define the interactions between steps, which
430 is why some surfaces will undergo step coalescence, where
431 additional attractive interactions can overcome the repulsion.\cite{Williams:1991}
432 The presence and concentration of adsorbates, as shown in
433 this work, can affect these step interactions, potentially leading
434 to a new surface structure as the thermodynamic minimum.
435
436 \subsubsection{Double layers}
437 Tao et al.\cite{Tao:2010} have shown experimentally that the Pt(557) surface
438 undergoes two separate reconstructions upon CO adsorption.\cite{Tao:2010}
439 The first involves a doubling of the step height and plateau length.
440 Similar behavior has been seen on numerous surfaces
441 at varying conditions: Ni(977), Si(111).\cite{Williams:1994,Williams:1991,Pearl}
442 Of the two systems we examined, the Pt system showed a greater
443 propensity for reconstruction when compared to the Au system
444 because of the larger surface mobility and extent of step wandering.
445 The amount of reconstruction is strongly correlated to the amount of CO
446 adsorbed upon the surface. This appears to be related to the
447 effect that adsorbate coverage has on edge breakup and on the
448 surface diffusion of metal adatoms. While both systems displayed
449 step-edge wandering, only the 50\% Pt surface underwent the
450 doubling seen by Tao et al.\cite{Tao:2010} within the time scales studied here.
451 Over longer periods, (150~ns) two more double layers formed
452 on this interface. Although double layer formation did not occur
453 in the other Pt systems, they show more step-wandering and
454 general roughening compared to their Au counterparts. The
455 50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
456 various times along the simulation showing the evolution of a double layer step-edge.
457
458 The second reconstruction on the Pt(557) surface observed by
459 Tao involved the formation of triangular clusters that stretched
460 across the plateau between two step-edges. Neither system, within
461 the 40~ns time scale or the extended simulation time of 150~ns for
462 the 50\% Pt system, experienced this reconstruction.
463
464 %Evolution of surface
465 \begin{figure}[H]
466 \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
467 \caption{The Pt(557) / 50\% CO system at a sequence of times after
468 initial exposure to the CO: (a) 258~ps, (b) 19~ns, (c) 31.2~ns, and
469 (d) 86.1~ns. Disruption of the (557) step-edges occurs quickly. The
470 doubling of the layers appears only after two adjacent step-edges
471 touch. The circled spot in (b) nucleated the growth of the double
472 step observed in the later configurations.}
473 \label{fig:reconstruct}
474 \end{figure}
475
476 \subsection{Dynamics}
477 Previous atomistic simulations of stepped surfaces dealt largely
478 with the energetics and structures at different conditions.
479 \cite{Williams:1991,Williams:1994} Consequently, the most common
480 technique utilized to date has been Monte Carlo sampling. Monte Carlo approaches give an efficient
481 sampling of the equilibrium thermodynamic landscape at the expense
482 of ignoring the dynamics of the system. Previous experimental work by Pearl and
483 Sibener\cite{Pearl}, using STM, has been able to capture the coalescing
484 of steps on Ni(977). The time scale of the image acquisition,
485 $\sim$70~s/image provides an upper bound for the time required for
486 the doubling to occur. By utilizing Molecular Dynamics we were able to probe the dynamics of these reconstructions and in this section we give data on dynamic and
487 transport properties, e.g. diffusion, layer formation time, etc.
488
489
490 \subsubsection{Transport of surface metal atoms}
491 %forcedSystems/stepSeparation
492 The movement or wandering of a step-edge is a cooperative effect
493 arising from the individual movements of the atoms making up the steps. An ideal metal surface
494 displaying a low index facet, (111) or (100), is unlikely to experience
495 much surface diffusion because of the large energetic barrier that must
496 be overcome to lift an atom out of the surface. The presence of step-edges and other surface features
497 on higher-index facets provides a lower energy source for mobile metal atoms.
498 Breaking away from the step-edge on a clean surface still imposes an
499 energetic penalty around $\sim$~45 kcal/mol, but this is easier than lifting
500 the same metal atom vertically out of the surface, \textgreater~60 kcal/mol.
501 The penalty lowers significantly when CO is present in sufficient quantities
502 on the surface. For certain distributions of CO, see Figures \ref{fig:SketchGraphic} and \ref{fig:SketchEnergies}, the penalty can fall to as low as
503 $\sim$~20 kcal/mol. Once an adatom exists on the surface, the barrier for
504 diffusion is negligible ( \textless~4 kcal/mol for a Pt adatom). These adatoms are then
505 able to explore the terrace before rejoining either their original step-edge or
506 becoming a part of a different edge. It is a difficult process for an atom
507 to traverse to a separate terrace although the presence of CO can lower the
508 energy barrier required to lift or lower an adatom. By tracking the mobility of individual
509 metal atoms on the Pt and Au surfaces we were able to determine the relative
510 diffusion constants, as well as how varying coverages of CO affect the diffusion. Close
511 observation of the mobile metal atoms showed that they were typically in
512 equilibrium with the step-edges, dynamically breaking apart and rejoining the edges.
513 At times, their motion was concerted and two or more adatoms would be
514 observed moving together across the surfaces.
515
516 A particle was considered ``mobile'' once it had traveled more than 2~\AA~
517 between saved configurations of the system (typically 10-100 ps). An atom that was
518 truly mobile would typically travel much greater distances than this, but the 2~\AA~cutoff
519 was used to prevent swamping the diffusion data with the in-place vibrational
520 movement of buried atoms. Diffusion on a surface is strongly affected by
521 local structures and in this work, the presence of single and double layer
522 step-edges causes the diffusion parallel to the step-edges to be larger than
523 the diffusion perpendicular to these edges. Parallel and perpendicular
524 diffusion constants are shown in Figure \ref{fig:diff}.
525
526 %Diffusion graph
527 \begin{figure}[H]
528 \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade_20ns.pdf}
529 \caption{Diffusion constants for mobile surface atoms along directions
530 parallel ($\mathbf{D}_{\parallel}$) and perpendicular
531 ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
532 surface coverage. Diffusion parallel to the step-edge is higher
533 than that perpendicular to the edge because of the lower energy
534 barrier associated with traversing along the edge as compared to
535 completely breaking away. The two reported diffusion constants for
536 the 50\% Pt system arise from different sample sets. The lower values
537 correspond to the same 40~ns amount that all of the other systems were
538 examined at, while the larger values correspond to a 20~ns period }
539 \label{fig:diff}
540 \end{figure}
541
542 The lack of a definite trend in the Au diffusion data in Figure \ref{fig:diff} is likely due
543 to the weaker bonding between Au and CO. This leads to a lower observed
544 coverage ({\it x}-axis) when compared to dosage amount, which
545 then further limits the effect the CO can have on surface diffusion. The correlation
546 between coverage and Pt diffusion rates conversely shows a
547 definite trend marred by the highest coverage surface. Two
548 explanations arise for this drop. First, upon a visual inspection of
549 the system, after a double layer has been formed, it maintains its
550 stability strongly and many atoms that had been tracked for mobility
551 data have now been buried. By performing the same diffusion
552 calculation but on a shorter run time (20~ns), only including data
553 before the formation of the first double layer, we obtain the larger
554 values for both $\mathbf{D}_{\parallel}$ and $\mathbf{D}_{\perp}$
555 at the 50\% coverage as seen in Figure \ref{fig:diff}.
556 This places the parallel diffusion constant more closely in line with the
557 expected trend, while the perpendicular diffusion constant does not
558 drop as far. A secondary explanation arising from our analysis of the
559 mechanism of double layer formation focuses on the effect that CO on the
560 surface has with respect to overcoming surface diffusion of Pt. If the
561 coverage is too sparse, the Pt engages in minimal interactions and
562 thus minimal diffusion. As coverage increases, there are more favorable
563 arrangements of CO on the surface allowing for the formation of a path,
564 a minimum energy trajectory, for the adatom to explore the surface.
565 As the CO is constantly moving on the surface, this path is constantly
566 changing. If the coverage becomes too great, the paths could
567 potentially be clogged leading to a decrease in diffusion despite
568 their being more adatoms and step-wandering.
569
570
571
572 \subsubsection{Dynamics of double layer formation}
573 The increased diffusion on Pt at the higher CO coverages
574 plays a primary role in double layer formation. However,
575 this is not a complete explanation -- the 33\%~Pt system
576 has higher diffusion constants but did not show any signs
577 of edge doubling in the observed run time. On the
578 50\%~Pt system, one layer formed within the first 40~ns
579 of simulation time, while two more were formed as the
580 system was allowed to run for an additional
581 110~ns (150~ns total). This suggests that this reconstruction is
582 a rapid process and that the previously mentioned upper bound
583 will be lowered as experimental techniques continue to improve.\cite{Williams:1991,Pearl}
584 In this system, as seen in Figure \ref{fig:reconstruct}, the first
585 appearance of a double layer, appears at 19~ns
586 into the simulation. Within 12~ns of this nucleation event, nearly half of the step has
587 formed the double layer and by 86~ns, the complete layer
588 has been flattened out. The double layer could be considered
589 ``complete" by 37~ns but remains a bit rough. From the
590 appearance of the first nucleation event to the first observed double layer, the process took $\sim$20~ns. Another
591 $\sim$40~ns was necessary for the layer to completely straighten.
592 The other two layers in this simulation formed over periods of
593 22~ns and 42~ns respectively. A possible explanation
594 for this rapid reconstruction is the elevated temperatures
595 under which our systems were simulated. It is probable that the process would
596 take longer at lower temperatures. Additionally, our measured times for completion
597 of the doubling after the appearance of a nucleation site are likely affected by our
598 constrained axes. A longer step-edge will likely take longer to ``zipper''. However,
599 the first appearance of a nucleation site will likely occur more quickly due to its stochastic nature.
600
601
602
603
604
605
606 %Sketch graphic of different configurations
607 \begin{figure}[H]
608 \includegraphics[width=0.8\linewidth, height=0.8\textheight]{COpathsSketch.pdf}
609 \caption{The dark grey atoms refer to the upper ledge, while the white atoms are
610 the lower terrace. The blue highlighted atoms had a CO in a vertical atop position
611 upon them. These are a sampling of the configurations examined to gain a more
612 complete understanding of the effects CO has on surface diffusion and edge breakup.
613 Energies associated with each configuration are displayed in Figure \ref{fig:SketchEnergies}.}
614 \label{fig:SketchGraphic}
615 \end{figure}
616
617 %energy graph corresponding to sketch graphic
618 \begin{figure}[H]
619 \includegraphics[width=\linewidth]{stepSeparationComparison.pdf}
620 \caption{The energy curves directly correspond to the labeled model
621 surface in Figure \ref{fig:SketchGraphic}. All energy curves are relative
622 to their initial configuration so the energy of a and h do not have the
623 same zero value. As is seen, certain arrangements of CO can lower
624 the energetic barrier that must be overcome to create an adatom.
625 However, it is the highest coverages where these higher-energy
626 configurations of CO will be more likely. }
627 \label{fig:SketchEnergies}
628 \end{figure}
629
630 %Discussion
631 \section{Discussion}
632 We have shown that the classical potential models are able to model the initial reconstruction of the
633 Pt(557) surface upon CO adsorption as shown by Tao et al. \cite{Tao:2010}. More importantly, we
634 were able to observe features of the dynamic processes necessary for this reconstruction.
635
636 \subsection{Diffusion}
637 As shown in Figure \ref{fig:diff}, for the Pt systems, there
638 is a strong trend toward higher diffusion constants as
639 surface coverage of CO increases. The drop for the 50\%
640 case being explained as double layer formation already
641 beginning to occur in the analyzed 40~ns, which lowered
642 the calculated diffusion rates. Between the parallel and
643 perpendicular rates, the perpendicular diffusion constant
644 appears to be the most important indicator of double layer
645 formation. As highlighted in Figure \ref{fig:reconstruct}, the
646 formation of the double layer did not begin until a nucleation
647 site appeared. And as mentioned by Williams et al.\cite{Williams:1991, Williams:1994},
648 the inability for edges to cross leads to an effective repulsion.
649 This repulsion must be overcome to allow step coalescence.
650 A greater $\textbf{D}_\perp$ implies more step-wandering
651 and a larger chance for the stochastic meeting of two edges
652 to form the nucleation point. Upon that appearance, parallel
653 diffusion along the step-edge can help ``zipper'' up the double
654 layer. This helps explain why the time scale for formation after
655 the appearance of a nucleation site was rapid, while the initial
656 appearance of said site was unpredictable.
657
658 \subsection{Mechanism for restructuring}
659 Since the Au surface showed no large scale restructuring throughout
660 our simulation time our discussion will focus on the 50\% Pt-CO system
661 which did undergo the doubling featured in Figure \ref{fig:reconstruct}.
662 Similarities of our results to those reported previously by Tao et al.\cite{Tao:2010}
663 are quite strong. The simulated Pt system exposed to a large dosage
664 of CO readily restructures by doubling the terrace widths and step heights.
665 The restructuring occurs in a piecemeal fashion, one to two Pt atoms at a
666 time, but is rapid on experimental timescales. The adatoms either break
667 away from the step-edge and stay on the lower terrace or they lift up onto
668 a higher terrace. Once ``free'', they diffuse on the terrace until reaching
669 another step-edge or rejoining their original edge. This combination of
670 growth and decay of the step-edges is in a state of dynamic equilibrium.
671 However, once two previously separated edges meet as shown in Figure 1.B,
672 this nucleates the rest of the edge to meet up, forming a double layer.
673 From simulations which exhibit a double layer, the time delay from the
674 initial appearance of a nucleation point to a fully formed double layer is $\sim$35~ns.
675
676 A number of possible mechanisms exist to explain the role of adsorbed
677 CO in restructuring the Pt surface. Quadrupolar repulsion between adjacent
678 CO molecules adsorbed on the surface is one possibility. However,
679 the quadrupole-quadrupole interaction is short-ranged and is attractive for
680 some orientations. If the CO molecules are ``locked'' in a specific orientation
681 relative to each other, through atop adsorption for example, this explanation
682 gains some credence. The energetic repulsion between two CO located a
683 distance of 2.77~\AA~apart (nearest-neighbor distance of Pt) and both in
684 a vertical orientation, is 8.62 kcal/mol. Moving the CO apart to the second
685 nearest-neighbor distance of 4.8~\AA~or 5.54~\AA~drops the repulsion to
686 nearly 0 kcal/mol. Allowing the CO to rotate away from a purely vertical orientation
687 also lowers the repulsion. A minimum of 6.2 kcal/mol is reached at when the
688 angle between the 2 CO is $\sim$24\textsuperscript{o}, when the carbons are
689 locked at a distance of 2.77 \AA apart. As mentioned above, the energy barrier
690 for surface diffusion of a Pt adatom is only 4 kcal/mol. So this repulsion between
691 neighboring CO molecules can increase the surface diffusion. However, the
692 residence time of CO on Pt was examined and while the majority of the CO is
693 on or near the surface throughout the run, the molecules are extremely mobile,
694 with diffusion constants 40 to 2500 times larger, depending on coverage. This
695 mobility suggests that the CO are more likely to shift their positions without
696 necessarily the Pt along with them.
697
698 Another possible and more likely mechanism for the restructuring is in the
699 destabilization of strong Pt-Pt interactions by CO adsorbed on surface
700 Pt atoms. To test this hypothesis, numerous configurations of
701 CO in varying quantities were arranged on the higher and lower plateaus
702 around a step on a otherwise clean Pt(557) surface. A few sample
703 configurations are displayed in Figure \ref{fig:SketchGraphic}, with
704 energies at various positions along the path displayed in Table
705 NO TABLE. Certain configurations of CO, cases B and D for
706 example, can have quite strong energetic reasons for breaking
707 away from the step-edge. Although the packing of these configurations
708 is unlikely until CO coverage has reached a high enough value.
709 These examples are showing the most difficult cases, immediate
710 adatom formation through breakage away from the step-edge, which
711 is why their energies at large distances are relatively high. There are
712 mechanistic paths where an edge atom could get shifted to onto the
713 step-edge to form a small peak before fully breaking away. And again,
714 once the adatom is formed, the barrier for diffusion on the surface is
715 negligible. These sample configurations help explain CO's effect on
716 general surface mobility and step wandering, but they are lacking in
717 providing a mechanism for the formation of double layers. One possible
718 mechanism is elucidated in Figure \ref{fig:lambda}, where a burrowing
719 and lifting process of an adatom and step-edge atom respectively is
720 examined. The system, without CO present, is nearly energetically
721 neutral, whereas with CO present there is a $\sim$ 15 kcal/mol drop
722 in the energy of the system.
723
724 %lambda progression of Pt -> shoving its way into the step
725 \begin{figure}[H]
726 \includegraphics[width=\linewidth]{lambdaProgression_atopCO_withLambda.png}
727 \caption{A model system of the Pt(557) surface was used as the framework
728 for exploring energy barriers along a reaction coordinate. Various numbers,
729 placements, and rotations of CO were examined as they affect Pt movement.
730 The coordinate displayed in this Figure was a representative run. relative to the energy of the system at 0\%, there
731 is a slight decrease upon insertion of the Pt atom into the step-edge along
732 with the resultant lifting of the other Pt atom when CO is present at certain positions.}
733 \label{fig:lambda}
734 \end{figure}
735
736
737
738
739
740 %breaking of the double layer upon removal of CO
741 \begin{figure}[H]
742 \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
743 \caption{(A) 0~ps, (B) 100~ps, (C) 1~ns, after the removal of CO. The presence of the CO
744 helped maintain the stability of the double layer and upon removal the two layers break
745 and begin separating. The separation is not a simple pulling apart however, rather
746 there is a mixing of the lower and upper atoms at the edge.}
747 \label{fig:breaking}
748 \end{figure}
749
750
751
752
753 %Peaks!
754 %\begin{figure}[H]
755 %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
756 %\caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
757 %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
758 %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
759 %\label{fig:peaks}
760 %\end{figure}
761
762
763 %Don't think I need this
764 %clean surface...
765 %\begin{figure}[H]
766 %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
767 %\caption{}
768
769 %\end{figure}
770 %\label{fig:clean}
771
772
773 \section{Conclusion}
774 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.
775
776 %Things I am not ready to remove yet
777
778 %Table of Diffusion Constants
779 %Add gold?M
780 % \begin{table}[H]
781 % \caption{}
782 % \centering
783 % \begin{tabular}{| c | cc | cc | }
784 % \hline
785 % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
786 % \hline
787 % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
788 % \hline
789 % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
790 % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
791 % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
792 % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
793 % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
794 % \hline
795 % \end{tabular}
796 % \end{table}
797
798 \begin{acknowledgement}
799 Support for this project was provided by the National Science
800 Foundation under grant CHE-0848243 and by the Center for Sustainable
801 Energy at Notre Dame (cSEND). Computational time was provided by the
802 Center for Research Computing (CRC) at the University of Notre Dame.
803 \end{acknowledgement}
804 \newpage
805 \bibliography{firstTryBibliography}
806 %\end{doublespace}
807
808 \begin{tocentry}
809 %\includegraphics[height=3.5cm]{timelapse}
810 \end{tocentry}
811
812 \end{document}