ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/firstTry.tex
Revision: 3873
Committed: Tue Mar 12 21:33:15 2013 UTC (11 years, 3 months ago) by jmichalk
Content type: application/x-tex
File size: 39194 byte(s)
Log Message:
New figure, restructured results...

File Contents

# User Rev Content
1 gezelter 3808 \documentclass[11pt]{article}
2     \usepackage{amsmath}
3     \usepackage{amssymb}
4 gezelter 3818 \usepackage{times}
5     \usepackage{mathptm}
6 jmichalk 3802 \usepackage{setspace}
7 gezelter 3826 \usepackage{endfloat}
8 gezelter 3808 \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 jmichalk 3802
21 gezelter 3808 % double space list of tables and figures
22 gezelter 3820 %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
23 gezelter 3808 \setlength{\abovecaptionskip}{20 pt}
24     \setlength{\belowcaptionskip}{30 pt}
25    
26 gezelter 3820 \bibpunct{}{}{,}{s}{}{;}
27 gezelter 3808 \bibliographystyle{achemso}
28    
29     \begin{document}
30    
31    
32 jmichalk 3802 %%
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 gezelter 3818 \title{Molecular Dynamics simulations of the surface reconstructions
52     of Pt(557) and Au(557) under exposure to CO}
53    
54 jmichalk 3816 \author{Joseph R. Michalka, Patrick W. McIntyre and J. Daniel
55 gezelter 3808 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 gezelter 3818
60 jmichalk 3802 %Date
61 jmichalk 3868 \date{Mar 5, 2013}
62 gezelter 3818
63 jmichalk 3802 %authors
64 gezelter 3808
65 jmichalk 3802 % make the title
66 jmichalk 3817 \maketitle
67 jmichalk 3802
68 gezelter 3808 \begin{doublespace}
69 jmichalk 3802
70 gezelter 3808 \begin{abstract}
71 jmichalk 3869 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 gezelter 3808 \end{abstract}
85 jmichalk 3802
86 gezelter 3808 \newpage
87    
88    
89 jmichalk 3802 \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 gezelter 3826 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 gezelter 3808 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
101 gezelter 3826 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 jmichalk 3802
115 jmichalk 3872 This work is an investigation into the mechanism and timescale for
116     surface restructuring using molecular simulations. Since the dynamics
117 jmichalk 3866 of the process are of particular interest, we employ classical force
118 gezelter 3826 fields that represent a compromise between chemical accuracy and the
119 jmichalk 3866 computational efficiency necessary to simulate the process of interest.
120 jmichalk 3868 Since restructuring typically occurs as a result of specific interactions of the
121     catalyst with adsorbates, in this work, two metal systems exposed
122 jmichalk 3866 to carbon monoxide were examined. The Pt(557) surface has already been shown
123 jmichalk 3870 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 jmichalk 3872 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 jmichalk 3870 affected by this relaxation. Piccolo et al. on the other hand, did see a
133 jmichalk 3872 disruption of the herringbone pattern as CO was adsorbed to the
134 jmichalk 3870 surface. Both groups suggested that the preference CO shows for
135 jmichalk 3872 low-coordinated Au atoms was the primary driving force for the reconstruction.
136 gezelter 3826
137 jmichalk 3868
138    
139 jmichalk 3811 %Platinum molecular dynamics
140     %gold molecular dynamics
141 jmichalk 3802
142     \section{Simulation Methods}
143 jmichalk 3869 The challenge in modeling any solid/gas interface is the
144 gezelter 3808 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 gezelter 3826 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 jmichalk 3869 Au-Au and Pt-Pt interactions\cite{EAM}. The CO was modeled using a rigid
160 gezelter 3826 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 gezelter 3808
164     \subsection{Metal-metal interactions}
165 gezelter 3826 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 gezelter 3808 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 jmichalk 3867 parameter sets. The glue model of Ercolessi et al. is among the
173 gezelter 3808 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 gezelter 3826 atomic sites is computed at atom $i$'s location,
179 gezelter 3808 \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 jmichalk 3866 $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
191     repulsive overlap of the two positively charged cores.
192 jmichalk 3807
193 gezelter 3826 % 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 jmichalk 3807
202 jmichalk 3866 The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
203 gezelter 3808 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 jmichalk 3870 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 jmichalk 3872 up to the third nearest neighbor were taken into account.\cite{Voter95a}
213 jmichalk 3870 Comparing that to the glue model of Ercolessi et al.\cite{Ercolessi88}
214 jmichalk 3872 which is only parameterized up to the nearest-neighbor
215 jmichalk 3870 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 jmichalk 3872 theory (DFT) makes fitting DFT-derived cross potentials with adsorbates somewhat easier.
220 jmichalk 3870 \cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
221 gezelter 3808
222 jmichalk 3870
223    
224    
225 gezelter 3826 \subsection{Carbon Monoxide model}
226 jmichalk 3866 Previous explanations for the surface rearrangements center on
227 jmichalk 3869 the large linear quadrupole moment of carbon monoxide.\cite{Tao:2010}
228 jmichalk 3866 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 jmichalk 3872 Karplus model treats CO as a rigid three site molecule with a massless M
232 jmichalk 3869 site at the molecular center of mass. The geometry and interaction
233     parameters are reproduced in Table~\ref{tab:CO}. The effective
234 jmichalk 3827 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 jmichalk 3812 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
238 jmichalk 3802 %CO Table
239     \begin{table}[H]
240 gezelter 3826 \caption{Positions, Lennard-Jones parameters ($\sigma$ and
241     $\epsilon$), and charges for the CO-CO
242 jmichalk 3869 interactions in Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
243 gezelter 3826 in kcal/mol, and charges are in atomic units.}
244 jmichalk 3802 \centering
245 jmichalk 3810 \begin{tabular}{| c | c | ccc |}
246 jmichalk 3802 \hline
247 jmichalk 3814 & {\it z} & $\sigma$ & $\epsilon$ & q\\
248 jmichalk 3802 \hline
249 jmichalk 3869 \textbf{C} & -0.6457 & 3.83 & 0.0262 & -0.75 \\
250     \textbf{O} & 0.4843 & 3.12 & 0.1591 & -0.85 \\
251 jmichalk 3814 \textbf{M} & 0.0 & - & - & 1.6 \\
252 jmichalk 3802 \hline
253     \end{tabular}
254 jmichalk 3866 \label{tab:CO}
255 jmichalk 3802 \end{table}
256 gezelter 3808
257 gezelter 3826 \subsection{Cross-Interactions between the metals and carbon monoxide}
258 jmichalk 3802
259 jmichalk 3867 Since the adsorption of CO onto a Pt surface has been the focus
260 gezelter 3826 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 jmichalk 3869 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 gezelter 3826 modified to ensure that the Pt-CO interaction favored the atop binding
267 jmichalk 3869 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 jmichalk 3866 than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
270 jmichalk 3867 et al.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
271 gezelter 3826 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 jmichalk 3869 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 gezelter 3826 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 jmichalk 3811
281 jmichalk 3812 %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 jmichalk 3866 The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
285 gezelter 3818 Morse potentials, respectively, to reproduce Au-CO binding energies.
286 jmichalk 3869 The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
287 jmichalk 3866 Adsorption energies were obtained from gas-surface DFT calculations with a
288 gezelter 3826 periodic supercell plane-wave basis approach, as implemented in the
289 jmichalk 3869 {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores were
290 gezelter 3818 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 jmichalk 3866 In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
297 gezelter 3818 Au x 2 Au surface planes and separated from vertical images by six
298 jmichalk 3866 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 gezelter 3826 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 gezelter 3818
308 jmichalk 3812 %Hint at future work
309 jmichalk 3866 The parameters employed for the metal-CO cross-interactions in this work
310 jmichalk 3869 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 jmichalk 3866 and polarization are neglected in this model, although these effects are likely to
313 jmichalk 3869 affect binding energies and binding site preferences, and will be addressed in
314 jmichalk 3872 future work.
315 jmichalk 3811
316 jmichalk 3802 %Table of Parameters
317     %Pt Parameter Set 9
318     %Au Parameter Set 35
319     \begin{table}[H]
320 jmichalk 3867 \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
321 jmichalk 3869 interactions are modeled with Lennard-Jones potentials. While the
322 jmichalk 3867 metal-O interactions were fit to Morse
323 gezelter 3826 potentials. Distances are given in \AA~and energies in kcal/mol. }
324 jmichalk 3802 \centering
325     \begin{tabular}{| c | cc | c | ccc |}
326     \hline
327 gezelter 3826 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
328 jmichalk 3802 \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 jmichalk 3866 \label{tab:co_parameters}
335 jmichalk 3802 \end{table}
336    
337     %Table of energies
338     \begin{table}[H]
339 jmichalk 3869 \caption{Adsorption energies for a single CO at the atop site on M(111) at the atop site using the potentials
340 jmichalk 3867 described in this work. All values are in eV.}
341 jmichalk 3802 \centering
342     \begin{tabular}{| c | cc |}
343 gezelter 3826 \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 jmichalk 3802 \end{tabular}
352 jmichalk 3866 \label{tab:co_energies}
353 jmichalk 3802 \end{table}
354    
355 gezelter 3826 \subsection{Pt(557) and Au(557) metal interfaces}
356 jmichalk 3872 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 jmichalk 3870 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 jmichalk 3872 1200~K were performed to confirm the relative
365 gezelter 3826 stability of the surfaces without a CO overlayer.
366 jmichalk 3802
367 jmichalk 3869 The different bulk melting temperatures (1337~K for Au
368     and 2045~K for Pt) suggest that any possible reconstruction should happen at
369 jmichalk 3867 different temperatures for the two metals. The bare Au and Pt surfaces were
370 gezelter 3826 initially run in the canonical (NVT) ensemble at 800~K and 1000~K
371 jmichalk 3869 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 gezelter 3826 that was initially placed in the vacuum region. Upon full adsorption,
375 jmichalk 3869 these concentrations correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
376 jmichalk 3872 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 jmichalk 3869 Because of the difference in binding energies, nearly all of the CO was bound to the Pt surface, while
379 jmichalk 3867 the Au surfaces often had a significant CO population in the gas
380 gezelter 3826 phase. These systems were allowed to reach thermal equilibrium (over
381 jmichalk 3873 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 jmichalk 3872 data collection stage, although simulation times for some Pt of the
384     systems exceeded 200~ns. Simulations were carried out using the open
385 jmichalk 3867 source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
386 jmichalk 3802
387 jmichalk 3872
388    
389    
390     % RESULTS
391     %
392 jmichalk 3802 \section{Results}
393 jmichalk 3860 \subsection{Structural remodeling}
394 jmichalk 3872 The surfaces of both systems, upon dosage of CO, began
395     to undergo remodeling that was not observed in the bare
396 jmichalk 3873 metal system. The surfaces which were not exposed to CO
397     did experience minor roughening of the step-edge because
398     of the elevated temperatures, but the
399 jmichalk 3872 (557) lattice was well-maintained throughout the simulation
400     time. The Au systems were limited to greater amounts of
401     roughening, i.e. breakup of the step-edge, and some step
402     wandering. The lower coverage Pt systems experienced
403     similar restructuring but to a greater extent when
404     compared to the Au systems. The 50\% coverage
405 jmichalk 3873 Pt system was unique among our simulations in that it
406     formed numerous double layers through step coalescence,
407     similar to results reported by Tao et al.\cite{Tao:2010}
408 jmichalk 3872
409    
410 jmichalk 3871 \subsubsection{Step wandering}
411 jmichalk 3873 The 0\% coverage surfaces for both metals showed minimal
412     movement at their respective run temperatures. As the CO
413     coverage increased however, the mobility of the surface,
414     adatoms and step-edges alike, also increased. Additionally,
415     at the higher coverages on both metals, there was more
416     step-wandering. Except for the 50\% Pt system, the step-edges
417     did not coalesce in any of the other simulations, instead preferring
418     to keep nearly the same distance between steps as in the
419     original (557) lattice. Previous work by Williams et al.\cite{Williams:1991, Williams:1994}
420     highlights the repulsion that exists between step-edges even
421     when no direct interactions are present in the system. This
422     repulsion exists because the entropy of the step-edges is constrained
423     since step-edge crossing is not allowed. This entropic repulsion
424     does not completely define the interactions between steps,
425     which is why some surfaces will undergo step coalescence,
426     where additional attractive interactions can overcome the
427     repulsion\cite{Williams:1991} and others will not. The presence
428     of adsorbates can affect these step interactions, potentially
429     leading to a new surface structure as the thermodynamic minimum.
430 jmichalk 3872
431 jmichalk 3871 \subsubsection{Double layers}
432 jmichalk 3869 Tao et al. have shown experimentally that the Pt(557) surface
433 jmichalk 3873 undergoes two separate reconstructions upon CO adsorption.\cite{Tao:2010}
434     The first involves a doubling of the step height and plateau length.
435     Similar behavior has been seen to occur on numerous surfaces
436     at varying conditions: Ni(977), Si(111).\cite{Williams:1994,Williams:1991,Pearl}
437     Of the two systems we examined, the Pt system showed a greater
438     propensity for reconstruction when compared to the Au system
439     because of the larger surface mobility and extent of step wandering.
440     The amount of reconstruction is correlated to the amount of CO
441 jmichalk 3869 adsorbed upon the surface. This appears to be related to the
442 jmichalk 3873 effect that adsorbate coverage has on edge breakup and on the
443     surface diffusion of metal adatoms. While both systems displayed
444     step-edge wandering, only the 50\% Pt surface underwent the
445     doubling seen by Tao et al. within the time scales studied here.
446     Over longer periods (150~ns) two more double layers formed
447     on this interface. Although double layer formation did not occur
448     in the other Pt systems, they show more step-wandering and
449     general roughening compared to their Au counterparts. The
450     50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
451     various times along the simulation showing the evolution of a step-edge.
452 jmichalk 3802
453 jmichalk 3867 The second reconstruction on the Pt(557) surface observed by
454     Tao involved the formation of triangular clusters that stretched
455     across the plateau between two step-edges. Neither system, within
456 jmichalk 3873 the 40~ns time scale or the extended simulation time of 150~ns for
457     the 50\% Pt system, experienced this reconstruction.
458 jmichalk 3817
459 jmichalk 3860 \subsection{Dynamics}
460 jmichalk 3872 Previous atomistic simulations of stepped surfaces dealt largely
461     with the energetics and structures at different conditions
462 jmichalk 3870 \cite{Williams:1991,Williams:1994}. Consequently, the most common
463 jmichalk 3872 technique utilized to date has been Monte Carlo sampling. Monte Carlo gives an efficient
464 jmichalk 3870 sampling of the equilibrium thermodynamic landscape at the expense
465 jmichalk 3873 of ignoring the dynamics of the system. Previous experimental work by Pearl and
466     Sibener\cite{Pearl}, using STM, has been able to capture the coalescing
467 jmichalk 3870 of steps on Ni(977). The time scale of the image acquisition,
468     $\sim$70 s/image provides an upper bound for the time required for
469     the doubling to occur. In this section we give data on dynamic and
470     transport properties, e.g. diffusion, layer formation time, etc.
471 gezelter 3826
472 jmichalk 3867
473 jmichalk 3860 \subsubsection{Transport of surface metal atoms}
474 jmichalk 3862 %forcedSystems/stepSeparation
475 jmichalk 3867 The movement or wandering of a step-edge is a cooperative effect
476 jmichalk 3873 arising from the individual movements of the atoms making up the steps. An ideal metal surface
477 jmichalk 3872 displaying a low index facet, (111) or (100), is unlikely to experience
478 jmichalk 3867 much surface diffusion because of the large energetic barrier that must
479 jmichalk 3873 be overcome to lift an atom out of the surface. The presence of step-edges and other surface features
480     on higher-index facets provide a lower energy source for mobile metal atoms.
481 jmichalk 3867 Breaking away from the step-edge on a clean surface still imposes an
482 jmichalk 3872 energetic penalty around $\sim$~40 kcal/mol, but this is significantly easier than lifting
483 jmichalk 3870 the same metal atom vertically out of the surface, \textgreater~60 kcal/mol.
484     The penalty lowers significantly when CO is present in sufficient quantities
485 jmichalk 3872 on the surface. For certain distributions of CO, the penalty can fall as low as
486 jmichalk 3870 $\sim$~20 kcal/mol. Once an adatom exists on the surface, the barrier for
487 jmichalk 3873 diffusion is negligible ( \textless~4 kcal/mol for a Pt adatom). These adatoms are
488 jmichalk 3872 able to explore the terrace before rejoining either the original step-edge or
489     becoming a part of a different edge. It is a more difficult process for an atom
490     to traverse to a separate terrace although the presence of CO can lower the
491     energy barrier required to lift or lower the adatom. By tracking the mobility of individual
492 jmichalk 3867 metal atoms on the Pt and Au surfaces we were able to determine the relative
493 jmichalk 3870 diffusion constants, as well as how varying coverages of CO affect the diffusion. Close
494 jmichalk 3867 observation of the mobile metal atoms showed that they were typically in
495 jmichalk 3870 equilibrium with the step-edges, dynamically breaking apart and rejoining the edges.
496     At times, their motion was concerted and two or more adatoms would be
497 jmichalk 3872 observed moving together across the surfaces.
498 gezelter 3826
499 jmichalk 3872 A particle was considered ``mobile'' once it had traveled more than 2~\AA~
500 jmichalk 3870 between saved configurations of the system (typically 10-100 ps). An atom that was
501 jmichalk 3872 truly mobile would typically travel much greater distances than this, but the 2~\AA~cutoff
502     was used to prevent swamping the diffusion data with the in-place vibrational
503 jmichalk 3873 movement of buried atoms. Diffusion on a surface is strongly affected by
504 jmichalk 3870 local structures and in this work, the presence of single and double layer
505 jmichalk 3867 step-edges causes the diffusion parallel to the step-edges to be different
506 jmichalk 3870 from the diffusion perpendicular to these edges. Parallel and perpendicular
507     diffusion constants are shown in Figure \ref{fig:diff}.
508 gezelter 3826
509 jmichalk 3873 The lack of a definite trend in the Au diffusion data is likely due
510     to the weaker bonding between Au and CO. This leads to a lower
511     coverage ({\it x}-axis) when compared to dosage amount, which
512     then further limits the affects of the surface diffusion. The correlation
513     between coverage and Pt diffusion rates conversely shows a
514     definite trend marred by the highest coverage surface. Two
515     explanations arise for this drop. First, upon a visual inspection of
516     the system, after a double layer has been formed, it maintains its
517     stability strongly and is no longer a good source for adatoms. By
518     performing the same diffusion calculation but on a shorter run time
519     (20~ns), only including data before the formation of the double layer,
520     provides a $\mathbf{D}_{\perp}$ diffusion constant of $1.69~\pm~0.08$
521     and a $\mathbf{D}_{\parallel}$ diffusion constant of $6.30~\pm~0.08$.
522     This places the parallel diffusion constant more closely in line with the
523     expected trend, while the perpendicular diffusion constant does not
524     drop as far. A secondary explanation arising from our analysis of the
525     mechanism of double layer formation show the affect that CO on the
526     surface has with respect to overcoming surface diffusion of Pt. If the
527     coverage is too sparse, the Pt engages in minimal interactions and
528     thus minimal diffusion. As coverage increases, there are more favorable
529     arrangements of CO on the surface allowing the formation of a path,
530     a minimum energy trajectory, for the adatom to explore the surface.
531     As the CO is constantly moving on the surface, this path is constantly
532     changing. If the coverage becomes too great, the paths could
533     potentially be clogged leading to a decrease in diffusion despite
534     their being more adatoms and step-wandering.
535    
536 jmichalk 3871 \subsubsection{Dynamics of double layer formation}
537 jmichalk 3872 The increased diffusion on Pt at the higher
538     CO coverages plays a primary role in double layer formation. However, this is not
539     a complete explanation -- the 33\%~Pt system
540     has higher diffusion constants but did not show
541 jmichalk 3873 any signs of edge doubling in the observed run time. On the
542     50\%~Pt system, one layer formed within the first 40~ns of simulation time, while two more were formed as the system was run for an additional
543     110~ns (150~ns total). Previous experimental
544 jmichalk 3872 work gives insight into the upper bounds of the
545     time required for step coalescence.\cite{Williams:1991,Pearl}
546     In this system, as seen in Figure \ref{fig:reconstruct}, the first
547     appearance of a double layer, appears at 19~ns
548     into the simulation. Within 12~ns of this nucleation event, nearly half of the step has
549 jmichalk 3873 formed the double layer and by 86~ns, the complete layer
550 jmichalk 3872 has been flattened out. The double layer could be considered
551     ``complete" by 37~ns but remains a bit rough. From the
552     appearance of the first nucleation event to the first observed double layer, the process took $\sim$20~ns. Another
553     $\sim$40~ns was necessary for the layer to completely straighten.
554     The other two layers in this simulation formed over periods of
555     22~ns and 42~ns respectively. Comparing this to the upper
556     bounds of the image scan, it is likely that most aspects of this
557     reconstruction occur very rapidly. A possible explanation
558     for this rapid reconstruction is the elevated temperatures
559     under which our systems were simulated. It is probable that the process would
560     take longer at lower temperatures.
561 jmichalk 3817
562 jmichalk 3862 %Evolution of surface
563 jmichalk 3816 \begin{figure}[H]
564 jmichalk 3862 \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
565     \caption{The Pt(557) / 50\% CO system at a sequence of times after
566 jmichalk 3873 initial exposure to the CO: (a) 258~ps, (b) 19~ns, (c) 31.2~ns, and
567     (d) 86.1~ns. Disruption of the (557) step-edges occurs quickly. The
568 jmichalk 3867 doubling of the layers appears only after two adjacent step-edges
569 jmichalk 3862 touch. The circled spot in (b) nucleated the growth of the double
570     step observed in the later configurations.}
571     \label{fig:reconstruct}
572     \end{figure}
573    
574     \begin{figure}[H]
575 jmichalk 3839 \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade.pdf}
576 gezelter 3826 \caption{Diffusion constants for mobile surface atoms along directions
577     parallel ($\mathbf{D}_{\parallel}$) and perpendicular
578 jmichalk 3867 ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
579     surface coverage. Diffusion parallel to the step-edge is higher
580 gezelter 3826 than that perpendicular to the edge because of the lower energy
581 jmichalk 3867 barrier associated with traversing along the edge as compared to
582     completely breaking away. Additionally, the observed
583 gezelter 3826 maximum and subsequent decrease for the Pt system suggests that the
584     CO self-interactions are playing a significant role with regards to
585 jmichalk 3867 movement of the Pt atoms around and across the surface. }
586 gezelter 3826 \label{fig:diff}
587 jmichalk 3816 \end{figure}
588    
589 jmichalk 3802
590 jmichalk 3862
591    
592 jmichalk 3802 %Discussion
593     \section{Discussion}
594 jmichalk 3872 We have shown that the classical potential models are able to model the initial reconstruction of the
595 jmichalk 3867 Pt(557) surface upon CO adsorption as shown by Tao et al. \cite{Tao:2010}. More importantly, we
596 jmichalk 3872 were able to observe features of the dynamic processes necessary for this reconstruction.
597 jmichalk 3802
598 jmichalk 3862 \subsection{Mechanism for restructuring}
599 jmichalk 3870 Since the Au surface showed no large scale restructuring throughout
600     our simulation time our discussion will focus on the 50\% Pt-CO system
601     which did undergo the doubling featured in Figure \ref{fig:reconstruct}.
602 jmichalk 3872 Similarities of our results to those reported previously by
603     Tao et al.\cite{Tao:2010} are quite
604     strong. The simulated Pt
605     system exposed to a large dosage of CO readily restructures by doubling the terrace
606     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.
607     The adatoms either
608 jmichalk 3867 break away from the step-edge and stay on the lower terrace or they lift
609 jmichalk 3872 up onto a higher terrace. Once ``free'', they diffuse on the terrace
610 jmichalk 3870 until reaching another step-edge or rejoining their original edge.
611 jmichalk 3867 This combination of growth and decay of the step-edges is in a state of
612     dynamic equilibrium. However, once two previously separated edges
613 jmichalk 3872 meet as shown in Figure 1.B, this nucleates the rest of the edge to meet up, forming a double layer.
614 jmichalk 3873 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.
615 gezelter 3826
616 jmichalk 3867 A number of possible mechanisms exist to explain the role of adsorbed
617     CO in restructuring the Pt surface. Quadrupolar repulsion between adjacent
618 jmichalk 3872 CO molecules adsorbed on the surface is one possibility. However,
619 jmichalk 3867 the quadrupole-quadrupole interaction is short-ranged and is attractive for
620     some orientations. If the CO molecules are ``locked'' in a specific orientation
621 jmichalk 3870 relative to each other, through atop adsorption for example, this explanation
622 jmichalk 3872 gains some credence. The energetic repulsion between two CO located a
623     distance of 2.77~\AA~apart (nearest-neighbor distance of Pt) and both in
624     a vertical orientation, is 8.62 kcal/mol. Moving the CO apart to the second
625 jmichalk 3867 nearest-neighbor distance of 4.8~\AA~or 5.54~\AA~drops the repulsion to
626 jmichalk 3870 nearly 0 kcal/mol. Allowing the CO's to leave a purely vertical orientation
627     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.
628     As mentioned above, the energy barrier for surface diffusion
629 jmichalk 3872 of a Pt adatom is only 4 kcal/mol. So this repulsion between neighboring CO molecules can
630 jmichalk 3870 increase the surface diffusion. However, the residence time of CO on Pt was
631 jmichalk 3867 examined and while the majority of the CO is on or near the surface throughout
632 jmichalk 3872 the run, most molecules are mobile. This mobility suggests that the CO are more
633     likely to shift their positions without necessarily the Pt along with them.
634 gezelter 3826
635 jmichalk 3862 Another possible and more likely mechanism for the restructuring is in the
636     destabilization of strong Pt-Pt interactions by CO adsorbed on surface
637 jmichalk 3867 Pt atoms. This would then have the effect of increasing surface mobility
638 jmichalk 3862 of these atoms. To test this hypothesis, numerous configurations of
639     CO in varying quantities were arranged on the higher and lower plateaus
640 jmichalk 3867 around a step on a otherwise clean Pt(557) surface. One representative
641     configuration is displayed in Figure \ref{fig:lambda}. Single or concerted movement
642     of Pt atoms was then examined to determine possible barriers. Because
643     the movement was forced along a pre-defined reaction coordinate that may differ
644 jmichalk 3862 from the true minimum of this path, only the beginning and ending energies
645 jmichalk 3873 are displayed in Table \ref{tab:energies} with the corresponding beginning and ending reaction coordinates in Figure \ref{fig:lambdaTable}. These values suggest that the presence of CO at suitable
646 jmichalk 3867 locations can lead to lowered barriers for Pt breaking apart from the step-edge.
647     Additionally, as highlighted in Figure \ref{fig:lambda}, the presence of CO makes the
648     burrowing and lifting of adatoms favorable, whereas without CO, the process is neutral
649 jmichalk 3862 in terms of energetics.
650    
651     %lambda progression of Pt -> shoving its way into the step
652     \begin{figure}[H]
653 jmichalk 3873 \includegraphics[width=\linewidth]{lambdaProgression_atopCO_withLambda.png}
654 jmichalk 3867 \caption{A model system of the Pt(557) surface was used as the framework
655     for exploring energy barriers along a reaction coordinate. Various numbers,
656     placements, and rotations of CO were examined as they affect Pt movement.
657     The coordinate displayed in this Figure was a representative run. As shown
658     in Table \ref{tab:rxcoord}, relative to the energy of the system at 0\%, there
659     is a slight decrease upon insertion of the Pt atom into the step-edge along
660     with the resultant lifting of the other Pt atom when CO is present at certain positions.}
661 jmichalk 3862 \label{fig:lambda}
662     \end{figure}
663    
664 jmichalk 3873 \begin{figure}[H]
665     \includegraphics[totalheight=0.9\textheight]{lambdaTable.png}
666     \caption{}
667     \label{fig:lambdaTable}
668     \end{figure}
669 jmichalk 3862
670    
671 jmichalk 3802 \subsection{Diffusion}
672 jmichalk 3872 The diffusion parallel to the step-edge tends to be
673     much larger than that perpendicular to the step-edge. The dynamic
674 jmichalk 3867 equilibrium that is established between the step-edge and adatom interface. The coverage
675     of CO also appears to play a slight role in relative rates of diffusion, as shown in Figure \ref{fig:diff}.
676     The
677 jmichalk 3862 Thus, the bottleneck of the double layer formation appears to be the initial formation
678     of this growth point, which seems to be somewhat of a stochastic event. Once it
679 jmichalk 3867 appears, parallel diffusion, along the now slightly angled step-edge, will allow for
680 jmichalk 3862 a faster formation of the double layer than if the entire process were dependent on
681     only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the
682     more likely a growth point is to be formed.
683 jmichalk 3802 \\
684 jmichalk 3862
685    
686     %breaking of the double layer upon removal of CO
687 jmichalk 3802 \begin{figure}[H]
688 jmichalk 3862 \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
689 jmichalk 3873 \caption{(A) 0~ps, (B) 100~ps, (C) 1~ns, after the removal of CO. The presence of the CO
690 jmichalk 3867 helped maintain the stability of the double layer and upon removal the two layers break
691     and begin separating. The separation is not a simple pulling apart however, rather
692     there is a mixing of the lower and upper atoms at the edge.}
693 jmichalk 3862 \label{fig:breaking}
694 jmichalk 3802 \end{figure}
695    
696    
697 jmichalk 3862
698    
699 jmichalk 3802 %Peaks!
700 jmichalk 3872 %\begin{figure}[H]
701     %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
702     %\caption{At the initial formation of this double layer ( $\sim$ 37 ns) there is a degree
703     %of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
704     %aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
705     %\label{fig:peaks}
706     %\end{figure}
707 jmichalk 3862
708 jmichalk 3867
709     %Don't think I need this
710 jmichalk 3862 %clean surface...
711 jmichalk 3867 %\begin{figure}[H]
712     %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
713     %\caption{}
714 jmichalk 3862
715 jmichalk 3867 %\end{figure}
716     %\label{fig:clean}
717    
718    
719 jmichalk 3802 \section{Conclusion}
720 jmichalk 3870 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.
721 jmichalk 3802
722 jmichalk 3862 %Things I am not ready to remove yet
723    
724     %Table of Diffusion Constants
725     %Add gold?M
726     % \begin{table}[H]
727     % \caption{}
728     % \centering
729     % \begin{tabular}{| c | cc | cc | }
730     % \hline
731     % &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
732     % \hline
733     % \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ \\
734     % \hline
735     % 50\% & 4.32(2) & 1.185(8) & 1.72(2) & 0.455(6) \\
736     % 33\% & 5.18(3) & 1.999(5) & 1.95(2) & 0.337(4) \\
737     % 25\% & 5.01(2) & 1.574(4) & 1.26(3) & 0.377(6) \\
738     % 5\% & 3.61(2) & 0.355(2) & 1.84(3) & 0.169(4) \\
739     % 0\% & 3.27(2) & 0.147(4) & 1.50(2) & 0.194(2) \\
740     % \hline
741     % \end{tabular}
742     % \end{table}
743    
744 gezelter 3808 \section{Acknowledgments}
745     Support for this project was provided by the National Science
746     Foundation under grant CHE-0848243 and by the Center for Sustainable
747     Energy at Notre Dame (cSEND). Computational time was provided by the
748     Center for Research Computing (CRC) at the University of Notre Dame.
749 jmichalk 3802
750 gezelter 3808 \newpage
751     \bibliography{firstTryBibliography}
752     \end{doublespace}
753     \end{document}