ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3867
Committed: Tue Mar 5 00:18:33 2013 UTC (11 years, 4 months ago) by jmichalk
Content type: application/x-tex
Original Path: trunk/COonPt/firstTry.tex
File size: 34104 byte(s)
Log Message:
Ever marching onwards to completion

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 4, 2013}
62
63 %authors
64
65 % make the title
66 \maketitle
67
68 \begin{doublespace}
69
70 \begin{abstract}
71 We examine potential surface reconstructions of Pt and Au(557)
72 under various CO coverages using molecular dynamics in order
73 to explore possible mechanisms for any observed reconstructions and their dynamics.
74 The metal-CO interactions were parameterized as part of this
75 work so that an efficient large-scale treatment of this system could be
76 undertaken. The relative binding strengths of the metal-CO
77 interactions were found to play a large role with regards to
78 step-edge stability and adatom diffusion. A small correlation
79 between coverage and the size of the diffusion constant was
80 also determined. An in-depth examination of the energetics of CO
81 adsorbed to the surface provides results that appear sufficient to explain the
82 reconstructions observed on the Pt systems and the corresponding lack
83 on the Au systems.
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 attempt to understand 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 Restructuring can occur 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 reconstruct under certain conditions. The Au(557) surface, because
124 of a weaker interaction with CO, is less likely to undergo this kind
125 of reconstruction. MORE HERE ON PT AND AU PREVIOUS WORK.
126
127 %Platinum molecular dynamics
128 %gold molecular dynamics
129
130 \section{Simulation Methods}
131 The challenge in modeling any solid/gas interface problem is the
132 development of a sufficiently general yet computationally tractable
133 model of the chemical interactions between the surface atoms and
134 adsorbates. Since the interfaces involved are quite large (10$^3$ -
135 10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
136 molecular dynamics
137 (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
138 methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
139 mechanical potential energy surfaces remain out of reach.
140 Additionally, the ``bonds'' between metal atoms at a surface are
141 typically not well represented in terms of classical pairwise
142 interactions in the same way that bonds in a molecular material are,
143 nor are they captured by simple non-directional interactions like the
144 Coulomb potential. For this work, we have used classical molecular
145 dynamics with potential energy surfaces that are specifically tuned
146 for transition metals. In particular, we used the EAM potential for
147 Au-Au and Pt-Pt interactions\cite{EAM}, while modeling the CO using a rigid
148 three-site model developed by Straub and Karplus for studying
149 photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
150 Pt-CO cross interactions were parameterized as part of this work.
151
152 \subsection{Metal-metal interactions}
153 Many of the potentials used for modeling transition metals are based
154 on a non-pairwise additive functional of the local electron
155 density. The embedded atom method (EAM) is perhaps the best known of
156 these
157 methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
158 but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
159 the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
160 parameter sets. The glue model of Ercolessi et al. is among the
161 fastest of these density functional approaches.\cite{Ercolessi88} In
162 all of these models, atoms are conceptualized as a positively charged
163 core with a radially-decaying valence electron distribution. To
164 calculate the energy for embedding the core at a particular location,
165 the electron density due to the valence electrons at all of the other
166 atomic sites is computed at atom $i$'s location,
167 \begin{equation*}
168 \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
169 \end{equation*}
170 Here, $\rho_j(r_{ij})$ is the function that describes the distance
171 dependence of the valence electron distribution of atom $j$. The
172 contribution to the potential that comes from placing atom $i$ at that
173 location is then
174 \begin{equation*}
175 V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
176 \end{equation*}
177 where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
178 $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
179 repulsive overlap of the two positively charged cores.
180
181 % The {\it modified} embedded atom method (MEAM) adds angular terms to
182 % the electron density functions and an angular screening factor to the
183 % pairwise interaction between two
184 % atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
185 % MEAM has become widely used to simulate systems in which angular
186 % interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
187 % metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
188 % MEAM presents significant additional computational costs, however.
189
190 The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
191 have all been widely used by the materials simulation community for
192 simulations of bulk and nanoparticle
193 properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
194 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
195 fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
196 propagation,\cite{BECQUART:1993rg} and alloying
197 dynamics.\cite{Shibata:2002hh} All of these potentials have their
198 strengths and weaknesses. One of the strengths common to all of the
199 methods is the relatively large library of metals for which these
200 potentials have been
201 parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
202
203 \subsection{Carbon Monoxide model}
204 Previous explanations for the surface rearrangements center on
205 the large linear quadrupole moment of carbon monoxide.
206 We used a model first proposed by Karplus and Straub to study
207 the photodissociation of CO from myoglobin because it reproduces
208 the quadrupole moment well.\cite{Straub} The Straub and
209 Karplus model, treats CO as a rigid three site molecule which places a massless M
210 site at the center of mass position along the CO bond. The geometry used along
211 with the interaction parameters are reproduced in Table~\ref{tab:CO}. The effective
212 dipole moment, calculated from the assigned charges, is still
213 small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
214 to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
215 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
216 %CO Table
217 \begin{table}[H]
218 \caption{Positions, Lennard-Jones parameters ($\sigma$ and
219 $\epsilon$), and charges for the CO-CO
220 interactions borrowed from Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
221 in kcal/mol, and charges are in atomic units.}
222 \centering
223 \begin{tabular}{| c | c | ccc |}
224 \hline
225 & {\it z} & $\sigma$ & $\epsilon$ & q\\
226 \hline
227 \textbf{C} & -0.6457 & 0.0262 & 3.83 & -0.75 \\
228 \textbf{O} & 0.4843 & 0.1591 & 3.12 & -0.85 \\
229 \textbf{M} & 0.0 & - & - & 1.6 \\
230 \hline
231 \end{tabular}
232 \label{tab:CO}
233 \end{table}
234
235 \subsection{Cross-Interactions between the metals and carbon monoxide}
236
237 Since the adsorption of CO onto a Pt surface has been the focus
238 of much experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
239 and theoretical work
240 \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
241 there is a significant amount of data on adsorption energies for CO on
242 clean metal surfaces. Parameters reported by Korzeniewski {\it et
243 al.}\cite{Pons:1986} were a starting point for our fits, which were
244 modified to ensure that the Pt-CO interaction favored the atop binding
245 position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}
246 This resulted in binding energies that are slightly higher
247 than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
248 et al.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
249 Lennard-Jones interaction to mimic strong, but short-ranged partial
250 binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
251 Pt-O interaction was parameterized to a Morse potential at a larger
252 minimum distance, ($r_o$). This was chosen so that the C would be preferred
253 over O as the binder to the surface. In most cases, this parameterization contributes a weak
254 repulsion which favors the atop site. The resulting potential-energy
255 surface suitably recovers the calculated Pt-C separation length
256 (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
257 position.\cite{Deshlahra:2012, Hopster:1978}
258
259 %where did you actually get the functionals for citation?
260 %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
261 %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
262 The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
263 Morse potentials, respectively, to reproduce Au-CO binding energies.
264 The limited experimental data for CO adsorption on Au lead us to refine our fits against DFT.
265 Adsorption energies were obtained from gas-surface DFT calculations with a
266 periodic supercell plane-wave basis approach, as implemented in the
267 {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores are
268 described with the projector augmented-wave (PAW)
269 method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
270 included to an energy cutoff of 20 Ry. Electronic energies are
271 computed with the PBE implementation of the generalized gradient
272 approximation (GGA) for gold, carbon, and oxygen that was constructed
273 by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
274 In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
275 Au x 2 Au surface planes and separated from vertical images by six
276 layers of vacuum space. The surface atoms were all allowed to relax
277 before CO was added to the system. Electronic relaxations were
278 performed until the energy difference between subsequent steps
279 was less than $10^{-8}$ Ry. Nonspin-polarized supercell calculations
280 were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
281 zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxed gold slab was
282 then used in numerous single point calculations with CO at various
283 heights (and angles relative to the surface) to allow fitting of the
284 empirical force field.
285
286 %Hint at future work
287 The parameters employed for the metal-CO cross-interactions in this work
288 are shown in Table~\ref{co_parameters} and the binding energies on the
289 (111) surfaces are displayed in Table~\ref{co_energies}. Charge transfer
290 and polarization are neglected in this model, although these effects are likely to
291 affect binding energies and binding site preferences, and will be added in
292 a future work.\cite{Deshlahra:2012,StreitzMintmire:1994}
293
294 %Table of Parameters
295 %Pt Parameter Set 9
296 %Au Parameter Set 35
297 \begin{table}[H]
298 \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
299 interactions are modeled with Lennard-Jones potential, while the
300 metal-O interactions were fit to Morse
301 potentials. Distances are given in \AA~and energies in kcal/mol. }
302 \centering
303 \begin{tabular}{| c | cc | c | ccc |}
304 \hline
305 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
306 \hline
307 \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
308 \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
309
310 \hline
311 \end{tabular}
312 \label{tab:co_parameters}
313 \end{table}
314
315 %Table of energies
316 \begin{table}[H]
317 \caption{Adsorption energies for CO on M(111) at the atop site using the potentials
318 described in this work. All values are in eV.}
319 \centering
320 \begin{tabular}{| c | cc |}
321 \hline
322 & Calculated & Experimental \\
323 \hline
324 \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
325 (Ref. \protect\cite{Kelemen:1979}) \\
326 & & -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
327 \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{TPD_Gold}) \\
328 \hline
329 \end{tabular}
330 \label{tab:co_energies}
331 \end{table}
332
333 \subsection{Pt(557) and Au(557) metal interfaces}
334
335 Our model systems are composed of 3888 Pt atoms and 3384 Au atoms in a
336 FCC crystal that have been cut along the (557) plane so that they are
337 periodic in the {\it x} and {\it y} directions, and have been oriented
338 to expose two aligned (557) cuts along the extended {\it
339 z}-axis. Simulations of the bare metal interfaces at temperatures
340 ranging from 300~K to 1200~K were performed to observe the relative
341 stability of the surfaces without a CO overlayer.
342
343 The different bulk (and surface) melting temperatures (1337~K for Au
344 and 2045~K for Pt) suggest that any possible reconstruction may happen at
345 different temperatures for the two metals. The bare Au and Pt surfaces were
346 initially run in the canonical (NVT) ensemble at 800~K and 1000~K
347 respectively for 100 ps. These temperatures were chosen because the
348 surfaces were relatively stable at these temperatures when no CO was
349 present, but experienced additional instability upon addition of CO in the time
350 frames we were examining. Each surface was exposed to a range of CO
351 that was initially placed in the vacuum region. Upon full adsorption,
352 these amounts correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
353 coverage. Higher coverages were tried, but the CO-CO repulsion was preventing
354 a higher amount of adsorption. Because of the difference in binding energies, the Pt
355 systems very rarely had CO that was not bound to the surface, while
356 the Au surfaces often had a significant CO population in the gas
357 phase. These systems were allowed to reach thermal equilibrium (over
358 5 ns) before being run in the microcanonical (NVE) ensemble for
359 data collection. All of the systems examined had at least 40 ns in the
360 data collection stage, although simulation times for some of the
361 systems exceeded 200ns. All simulations were run using the open
362 source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
363
364 % Just results, leave discussion for discussion section
365 % structure
366 % Pt: step wandering, double layers, no triangular motifs
367 % Au: step wandering, no double layers
368 % dynamics
369 % diffusion
370 % time scale, formation, breakage
371 \section{Results}
372 \subsection{Structural remodeling}
373 Tao et al. showed experimentally that the Pt(557) surface
374 undergoes two separate reconstructions upon CO
375 adsorption.\cite{Tao:2010} The first involves a doubling of
376 the step height and plateau length. Similar behavior has been
377 seen to occur on numerous surfaces at varying conditions (Ni 977, Si 111, etc).
378 \cite{Williams:1994,Williams:1991,Pearl} Of the two systems
379 we examined, the Pt system showed a larger amount of
380 reconstruction when compared to the Au system. The amount
381 of reconstruction appears to be correlated to the amount of CO
382 adsorbed upon the surface. We believe this is related to the
383 effect that adsorbate coverage has on edge breakup and surface
384 diffusion of adatoms. While both systems displayed step-edge
385 wandering, only the Pt surface underwent the doubling seen by
386 Tao et al., within the time scales we were modeling. Specifically,
387 only the 50~\% coverage Pt system was observed to have a
388 step-edge undergo a complete doubling in the time scales we
389 were able to monitor. This event encouraged us to allow that
390 specific system to run for much longer periods during which two
391 more double layers were created. The other systems, not displaying
392 any large scale changes of interest, were all stopped after running
393 for 40 ns in the microcanonical ensemble. Despite no observation
394 of double layer formation, the other Pt systems tended to show
395 more cumulative lateral movement of the step-edges when
396 compared to the Au systems. The 50\% Pt system is highlighted
397 in Figure \ref{fig:reconstruct} at various times along the simulation
398 showing the evolution of the system.
399
400 The second reconstruction on the Pt(557) surface observed by
401 Tao involved the formation of triangular clusters that stretched
402 across the plateau between two step-edges. Neither system, within
403 our simulated time scales, experiences this reconstruction. A constructed
404 system in which the triangular motifs were constructed on the surface
405 will be explored in future work and is shown in the supporting information.
406
407 \subsection{Dynamics}
408 While atomistic-like simulations of stepped surfaces have been
409 performed before, they tend to be performed using Monte Carlo
410 techniques\cite{Williams:1991,Williams:1994}. This allows them
411 to efficiently sample the equilibrium thermodynamic landscape
412 but at the expense of ignoring the dynamics of the system. Previous
413 work by Pearl and Sibener\cite{Pearl}, using STM, has been able to
414 visualize the coalescing of steps of Ni(977). The time scale of the image
415 acquisition, $\sim$70 s/image provides an upper bounds for the time
416 required for the doubling to actually occur. Statistical treatments of step-edges
417 are adept at analyzing such systems. However, in a system where
418 the number of steps is limited, examining the individual atoms that make
419 up the steps can provide useful information as well.
420
421
422 \subsubsection{Transport of surface metal atoms}
423 %forcedSystems/stepSeparation
424 The movement or wandering of a step-edge is a cooperative effect
425 arising from the individual movements, primarily through surface
426 diffusion, of the atoms making up the step. An ideal metal surface
427 displaying a low index facet, (111) or (100) is unlikely to experience
428 much surface diffusion because of the large energetic barrier that must
429 be overcome to lift an atom out of the surface. The presence of step-edges
430 on higher-index surfaces provide a source for mobile metal atoms.
431 Breaking away from the step-edge on a clean surface still imposes an
432 energetic penalty around $\sim$~40 kcal/mole, but is much less than lifting
433 the same metal atom out from the surface, \textgreater~60 kcal/mole, and
434 the penalty lowers even further when CO is present in sufficient quantities
435 on the surface. For certain tested distributions of CO, the penalty was lowered
436 to $\sim$~20 kcal/mole. Once an adatom exists on the surface, its barrier for
437 diffusion is negligible ( \textless~4 kcal/mole) and is well able to explore the
438 terrace before potentially rejoining its original step-edge or becoming a part
439 of a different edge. Atoms traversing separate terraces is a more difficult
440 process, but can be overcome through a joining and lifting stage which is
441 examined in the discussion section. By tracking the mobility of individual
442 metal atoms on the Pt and Au surfaces we were able to determine the relative
443 diffusion rates and how varying coverages of CO affected the rates. Close
444 observation of the mobile metal atoms showed that they were typically in
445 equilibrium with the step-edges, constantly breaking apart and rejoining.
446 At times their motion was concerted and two or more adatoms would be
447 observed moving together across the surfaces. The primary challenge in
448 quantifying the overall surface mobility was in defining ``mobile" vs. ``static" atoms.
449
450 A particle was considered mobile once it had traveled more than 2~\AA~
451 between saved configurations of the system (10-100 ps). An atom that was
452 truly mobile would typically travel much greater than this, but the 2~\AA~ cutoff
453 was to prevent the in-place vibrational movement of non-surface atoms from
454 being included in the analysis. Diffusion on a surface is strongly affected by
455 local structures and in this work the presence of single and double layer
456 step-edges causes the diffusion parallel to the step-edges to be different
457 from the diffusion perpendicular to these edges. This led us to compute
458 those diffusions separately as seen in Figure \ref{fig:diff}.
459
460 \subsubsection{Double layer formation}
461 The increased amounts of diffusion on Pt at the higher CO coverages appears
462 to play a primary role in the formation of double layers, although this conclusion
463 does not explain the 33\% coverage Pt system. On the 50\% system, three
464 separate layers were formed over the extended run time of this system. As
465 mentioned earlier, previous experimental work has given some insight into the
466 upper bounds of the time required for enough atoms to move around to allow two
467 steps to coalesce\cite{Williams:1991,Pearl}. As seen in Figure \ref{fig:reconstruct},
468 the first appearance of a double layer, a nodal site, appears at 19 ns into the
469 simulation. Within 12 ns, nearly half of the step has formed the double layer and
470 by 86 ns, a smooth complete layer has formed. The double layer is ``complete" by
471 37 ns but is a bit rough. From the appearance of the first node to the initial doubling
472 of the layers ignoring their roughness took $\sim$~20 ns. Another ~40 ns was
473 necessary for the layer to completely straighten. The other two layers in this
474 simulation form over a period of 22 ns and 42 ns respectively. Comparing this to
475 the upper bounds of the image scan, it is likely that aspects of this reconstruction
476 occur very quickly.
477
478 %Evolution of surface
479 \begin{figure}[H]
480 \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
481 \caption{The Pt(557) / 50\% CO system at a sequence of times after
482 initial exposure to the CO: (a) 258 ps, (b) 19 ns, (c) 31.2 ns, and
483 (d) 86.1 ns. Disruption of the (557) step-edges occurs quickly. The
484 doubling of the layers appears only after two adjacent step-edges
485 touch. The circled spot in (b) nucleated the growth of the double
486 step observed in the later configurations.}
487 \label{fig:reconstruct}
488 \end{figure}
489
490 \begin{figure}[H]
491 \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade.pdf}
492 \caption{Diffusion constants for mobile surface atoms along directions
493 parallel ($\mathbf{D}_{\parallel}$) and perpendicular
494 ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
495 surface coverage. Diffusion parallel to the step-edge is higher
496 than that perpendicular to the edge because of the lower energy
497 barrier associated with traversing along the edge as compared to
498 completely breaking away. Additionally, the observed
499 maximum and subsequent decrease for the Pt system suggests that the
500 CO self-interactions are playing a significant role with regards to
501 movement of the Pt atoms around and across the surface. }
502 \label{fig:diff}
503 \end{figure}
504
505
506
507
508 %Discussion
509 \section{Discussion}
510 In this paper we have shown that we were able to accurately model the initial reconstruction of the
511 Pt(557) surface upon CO adsorption as shown by Tao et al. \cite{Tao:2010}. More importantly, we
512 were able to observe the dynamic processes necessary for this reconstruction.
513
514 \subsection{Mechanism for restructuring}
515 Comparing the results from simulation to those reported previously by
516 Tao et al.\cite{Tao:2010} the similarities in the Pt-CO system are quite
517 strong. As shown in Figure \ref{fig:reconstruct}, the simulated Pt
518 system under a CO atmosphere will restructure by doubling the terrace
519 heights. The restructuring occurs slowly, one to two Pt atoms at a time.
520 Looking at individual configurations of the system, the adatoms either
521 break away from the step-edge and stay on the lower terrace or they lift
522 up onto the higher terrace. Once ``free'' they will diffuse on the terrace
523 until reaching another step-edge or coming back to their original edge.
524 This combination of growth and decay of the step-edges is in a state of
525 dynamic equilibrium. However, once two previously separated edges
526 meet as shown in Figure 1.B, this meeting point tends to act as a focus
527 or growth point for the rest of the edge to meet up, akin to that of a zipper.
528 From the handful of cases where a double layer was formed during the
529 simulation, measuring from the initial appearance of a growth point, the
530 double layer tends to be fully formed within $\sim$~35 ns.
531
532 A number of possible mechanisms exist to explain the role of adsorbed
533 CO in restructuring the Pt surface. Quadrupolar repulsion between adjacent
534 CO molecules adsorbed on the surface is one likely possibility. However,
535 the quadrupole-quadrupole interaction is short-ranged and is attractive for
536 some orientations. If the CO molecules are ``locked'' in a specific orientation
537 relative to each other, through atop adsorption perhaps, this explanation
538 gains some weight. The energetic repulsion between two CO located a
539 distance of 2.77~\AA~apart (nearest-neighbor distance of Pt) with both in
540 a vertical orientation is 8.62 kcal/mole. Moving the CO apart to the second
541 nearest-neighbor distance of 4.8~\AA~or 5.54~\AA~drops the repulsion to
542 nearly 0 kcal/mole. Allowing the CO's to leave a purely vertical orientation
543 also quickly drops the repulsion, a minimum is reached at $\sim$24 degrees
544 of 6.2 kcal/mole. As mentioned above, the energy barrier for surface diffusion
545 of a Pt adatom is only 4 kcal/mole. So this repulsion between CO can help
546 increase the surface diffusion. However, the residence time of CO was
547 examined and while the majority of the CO is on or near the surface throughout
548 the run, it is extremely mobile. This mobility suggests that the CO are more
549 likely to shift their positions without necessarily dragging the Pt along with them.
550
551 Another possible and more likely mechanism for the restructuring is in the
552 destabilization of strong Pt-Pt interactions by CO adsorbed on surface
553 Pt atoms. This would then have the effect of increasing surface mobility
554 of these atoms. To test this hypothesis, numerous configurations of
555 CO in varying quantities were arranged on the higher and lower plateaus
556 around a step on a otherwise clean Pt(557) surface. One representative
557 configuration is displayed in Figure \ref{fig:lambda}. Single or concerted movement
558 of Pt atoms was then examined to determine possible barriers. Because
559 the movement was forced along a pre-defined reaction coordinate that may differ
560 from the true minimum of this path, only the beginning and ending energies
561 are displayed in Table \ref{tab:energies}. These values suggest that the presence of CO at suitable
562 locations can lead to lowered barriers for Pt breaking apart from the step-edge.
563 Additionally, as highlighted in Figure \ref{fig:lambda}, the presence of CO makes the
564 burrowing and lifting of adatoms favorable, whereas without CO, the process is neutral
565 in terms of energetics.
566
567 %lambda progression of Pt -> shoving its way into the step
568 \begin{figure}[H]
569 \includegraphics[width=\linewidth]{lambdaProgression_atopCO.png}
570 \caption{A model system of the Pt(557) surface was used as the framework
571 for exploring energy barriers along a reaction coordinate. Various numbers,
572 placements, and rotations of CO were examined as they affect Pt movement.
573 The coordinate displayed in this Figure was a representative run. As shown
574 in Table \ref{tab:rxcoord}, relative to the energy of the system at 0\%, there
575 is a slight decrease upon insertion of the Pt atom into the step-edge along
576 with the resultant lifting of the other Pt atom when CO is present at certain positions.}
577 \label{fig:lambda}
578 \end{figure}
579
580
581
582 \subsection{Diffusion}
583 As shown in the results section, the diffusion parallel to the step-edge tends to be
584 much larger than that perpendicular to the step-edge, likely because of the dynamic
585 equilibrium that is established between the step-edge and adatom interface. The coverage
586 of CO also appears to play a slight role in relative rates of diffusion, as shown in Figure \ref{fig:diff}.
587 The
588 Thus, the bottleneck of the double layer formation appears to be the initial formation
589 of this growth point, which seems to be somewhat of a stochastic event. Once it
590 appears, parallel diffusion, along the now slightly angled step-edge, will allow for
591 a faster formation of the double layer than if the entire process were dependent on
592 only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the
593 more likely a growth point is to be formed.
594 \\
595
596
597 %breaking of the double layer upon removal of CO
598 \begin{figure}[H]
599 \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
600 %:
601 \caption{(A) 0 ps, (B) 100 ps, (C) 1 ns, after the removal of CO. The presence of the CO
602 helped maintain the stability of the double layer and upon removal the two layers break
603 and begin separating. The separation is not a simple pulling apart however, rather
604 there is a mixing of the lower and upper atoms at the edge.}
605 \label{fig:breaking}
606 \end{figure}
607
608
609
610
611 %Peaks!
612 \begin{figure}[H]
613 \includegraphics[width=\linewidth]{doublePeaks_noCO.png}
614 \caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
615 of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
616 aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
617 \label{fig:peaks}
618 \end{figure}
619
620
621 %Don't think I need this
622 %clean surface...
623 %\begin{figure}[H]
624 %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
625 %\caption{}
626
627 %\end{figure}
628 %\label{fig:clean}
629
630
631 \section{Conclusion}
632 In this work we have shown the reconstruction of the Pt(557) crystalline surface upon adsorption of CO in < $\mu s$. 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.
633
634 %Things I am not ready to remove yet
635
636 %Table of Diffusion Constants
637 %Add gold?M
638 % \begin{table}[H]
639 % \caption{}
640 % \centering
641 % \begin{tabular}{| c | cc | cc | }
642 % \hline
643 % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
644 % \hline
645 % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
646 % \hline
647 % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
648 % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
649 % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
650 % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
651 % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
652 % \hline
653 % \end{tabular}
654 % \end{table}
655
656 \section{Acknowledgments}
657 Support for this project was provided by the National Science
658 Foundation under grant CHE-0848243 and by the Center for Sustainable
659 Energy at Notre Dame (cSEND). Computational time was provided by the
660 Center for Research Computing (CRC) at the University of Notre Dame.
661
662 \newpage
663 \bibliography{firstTryBibliography}
664 \end{doublespace}
665 \end{document}