ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
(Generate patch)

Comparing:
trunk/COonPt/firstTry.tex (file contents), Revision 3873 by jmichalk, Tue Mar 12 21:33:15 2013 UTC vs.
trunk/COonPt/COonPtAu.tex (file contents), Revision 3892 by gezelter, Wed Jun 5 21:22:46 2013 UTC

# Line 1 | Line 1
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}
1 > \documentclass[journal = jpccck, manuscript = article]{achemso}
2 > \setkeys{acs}{usetitle = true}
3 > \usepackage{achemso}
4 > \usepackage{natbib}
5   \usepackage{multirow}
6 < %\usepackage{booktabs}
7 < %\usepackage{bibentry}
8 < %\usepackage{mathrsfs}
9 < \usepackage[square, comma, sort&compress]{natbib}
6 > \usepackage{wrapfig}
7 > \usepackage{fixltx2e}
8 > %\mciteErrorOnUnknownfalse
9 >
10 > \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
11   \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
12  
13 < % double space list of tables and figures
14 < %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
23 < \setlength{\abovecaptionskip}{20 pt}
24 < \setlength{\belowcaptionskip}{30 pt}
13 > \title{Molecular Dynamics simulations of the surface reconstructions
14 >  of Pt(557) and Au(557) under exposure to CO}
15  
16 < \bibpunct{}{}{,}{s}{}{;}
17 < \bibliographystyle{achemso}
16 > \author{Joseph R. Michalka}
17 > \author{Patrick W. McIntyre}
18 > \author{J. Daniel Gezelter}
19 > \email{gezelter@nd.edu}
20 > \affiliation[University of Notre Dame]{251 Nieuwland Science Hall\\
21 >  Department of Chemistry and Biochemistry\\ University of Notre
22 >  Dame\\ Notre Dame, Indiana 46556}
23  
24 + \keywords{}
25 +
26   \begin{document}
27  
28 <
28 >
29   %%
30   %Introduction
31   %       Experimental observations
# Line 47 | Line 44
44   %Summary
45   %%
46  
50 %Title
51 \title{Molecular Dynamics simulations of the surface reconstructions
52  of Pt(557) and Au(557) under exposure to CO}
47  
54 \author{Joseph R. Michalka, Patrick W. McIntyre and J. Daniel
55 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
56 Department of Chemistry and Biochemistry,\\
57 University of Notre Dame\\
58 Notre Dame, Indiana 46556}
59
60 %Date
61 \date{Mar 5, 2013}
62
63 %authors
64
65 % make the title
66 \maketitle
67
68 \begin{doublespace}
69
48   \begin{abstract}
49 < We examine surface reconstructions of Pt and Au(557) under
50 < various CO coverages using molecular dynamics in order to
51 < explore possible mechanisms for any observed reconstructions
52 < and their dynamics. The metal-CO interactions were parameterized
53 < as part of this work so that an efficient large-scale treatment of
54 < this system could be undertaken. The large difference in binding
55 < strengths of the metal-CO interactions was found to play a significant
56 < role with regards to step-edge stability and adatom diffusion. A
57 < small correlation between coverage and the diffusion constant
58 < was also determined. The energetics of CO adsorbed to the surface
59 < is sufficient to explain the reconstructions observed on the Pt
60 < systems and the lack  of reconstruction of the Au systems.
61 <
49 >  The mechanism and dynamics of surface reconstructions of Pt(557) and
50 >  Au(557) exposed to various coverages of carbon monoxide (CO) were
51 >  investigated using molecular dynamics simulations.  Metal-CO
52 >  interactions were parameterized from experimental data and
53 >  plane-wave Density Functional Theory (DFT) calculations.  The large
54 >  difference in binding strengths of the Pt-CO and Au-CO interactions
55 >  was found to play a significant role in step-edge stability and
56 >  adatom diffusion constants.  Various mechanisms for CO-mediated step
57 >  wandering and step doubling were investigated on the Pt(557)
58 >  surface.  We find that the energetics of CO adsorbed to the surface
59 >  can explain the step-doubling reconstruction observed on Pt(557) and
60 >  the lack of such a reconstruction on the Au(557) surface.  However,
61 >  more complicated reconstructions into triangular clusters that have
62 >  been seen in recent experiments were not observed in these
63 >  simulations.
64   \end{abstract}
65  
66   \newpage
# Line 112 | Line 92 | This work is an investigation into the mechanism and t
92   reversible restructuring under exposure to moderate pressures of
93   carbon monoxide.\cite{Tao:2010}
94  
95 < This work is an investigation into the mechanism and timescale for
96 < surface restructuring using molecular simulations.  Since the dynamics
97 < of the process are of particular interest, we employ classical force
98 < fields that represent a compromise between chemical accuracy and the
99 < computational efficiency necessary to simulate the process of interest.
100 < Since restructuring typically occurs as a result of specific interactions of the
101 < catalyst with adsorbates, in this work, two metal systems exposed
102 < to carbon monoxide were examined. The Pt(557) surface has already been shown
103 < to undergo a large scale reconstruction under certain conditions.\cite{Tao:2010}
104 < The Au(557) surface, because of a weaker interaction with CO, is seen as less
105 < likely to undergo this kind of reconstruction. However, Peters et al.\cite{Peters:2000}
106 < and Piccolo et al.\cite{Piccolo:2004} have both observed CO-induced
107 < reconstruction of a Au(111) surface. Peters et al. saw a relaxation to the
108 < 22 x $\sqrt{3}$ cell. They argued that only a few Au atoms
109 < become adatoms, limiting the stress of this reconstruction while
110 < allowing the rest to relax and approach the ideal (111)
111 < configuration. They did not see the usual herringbone pattern being greatly
112 < affected by this relaxation. Piccolo et al. on the other hand, did see a
113 < disruption of the herringbone pattern as CO was adsorbed to the
114 < surface. Both groups suggested that the preference CO shows for
115 < low-coordinated Au atoms was the primary driving force for the reconstruction.
116 <
117 <
95 > This work is an investigation into the mechanism and timescale for the
96 > Pt(557) \& Au(557) surface restructuring using molecular simulation.
97 > Since the dynamics of the process are of particular interest, we
98 > employ classical force fields that represent a compromise between
99 > chemical accuracy and the computational efficiency necessary to
100 > simulate the process of interest.  Since restructuring typically
101 > occurs as a result of specific interactions of the catalyst with
102 > adsorbates, in this work, two metal systems exposed to carbon monoxide
103 > were examined. The Pt(557) surface has already been shown to undergo a
104 > large scale reconstruction under certain conditions.\cite{Tao:2010}
105 > The Au(557) surface, because of weaker interactions with CO, is less
106 > likely to undergo this kind of reconstruction. However, Peters {\it et
107 >  al}.\cite{Peters:2000} and Piccolo {\it et al}.\cite{Piccolo:2004}
108 > have both observed CO-induced modification of reconstructions to the
109 > Au(111) surface. Peters {\it et al}. observed the Au(111)-($22 \times
110 > \sqrt{3}$) ``herringbone'' reconstruction relaxing slightly under CO
111 > adsorption. They argued that only a few Au atoms become adatoms,
112 > limiting the stress of this reconstruction, while allowing the rest to
113 > relax and approach the ideal (111) configuration.  Piccolo {\it et
114 >  al}. on the other hand, saw a more significant disruption of the
115 > Au(111)-($22 \times \sqrt{3}$) herringbone pattern as CO adsorbed on
116 > the surface. Both groups suggested that the preference CO shows for
117 > low-coordinated Au atoms was the primary driving force for the
118 > relaxation.  Although the Au(111) reconstruction was not the primary
119 > goal of our work, the classical models we have fit may be of future
120 > use in simulating this reconstruction.
121  
122   %Platinum molecular dynamics
123   %gold molecular dynamics
124  
125   \section{Simulation Methods}
126 < The challenge in modeling any solid/gas interface is the
127 < development of a sufficiently general yet computationally tractable
128 < model of the chemical interactions between the surface atoms and
129 < adsorbates.  Since the interfaces involved are quite large (10$^3$ -
130 < 10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
126 > The challenge in modeling any solid/gas interface is the development
127 > of a sufficiently general yet computationally tractable model of the
128 > chemical interactions between the surface atoms and adsorbates.  Since
129 > the interfaces involved are quite large (10$^3$ - 10$^4$ atoms), have
130 > many electrons, and respond slowly to perturbations, {\it ab initio}
131   molecular dynamics
132   (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
133   methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
# Line 156 | Line 139 | Au-Au and Pt-Pt interactions\cite{EAM}. The CO was mod
139   Coulomb potential.  For this work, we have used classical molecular
140   dynamics with potential energy surfaces that are specifically tuned
141   for transition metals.  In particular, we used the EAM potential for
142 < Au-Au and Pt-Pt interactions\cite{EAM}. The CO was modeled using a rigid
143 < three-site model developed by Straub and Karplus for studying
142 > Au-Au and Pt-Pt interactions.\cite{Foiles86} The CO was modeled using
143 > a rigid three-site model developed by Straub and Karplus for studying
144   photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
145   Pt-CO cross interactions were parameterized as part of this work.
146    
# Line 169 | Line 152 | parameter sets. The glue model of Ercolessi et al. is
152   methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
153   but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
154   the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
155 < parameter sets. The glue model of Ercolessi et al. is among the
156 < fastest of these density functional approaches.\cite{Ercolessi88} In
157 < all of these models, atoms are conceptualized as a positively charged
158 < core with a radially-decaying valence electron distribution. To
159 < calculate the energy for embedding the core at a particular location,
160 < the electron density due to the valence electrons at all of the other
161 < atomic sites is computed at atom $i$'s location,
155 > parameter sets. The glue model of Ercolessi {\it et
156 >  al}.\cite{Ercolessi88} is among the fastest of these density
157 > functional approaches. In all of these models, atoms are treated as a
158 > positively charged core with a radially-decaying valence electron
159 > distribution. To calculate the energy for embedding the core at a
160 > particular location, the electron density due to the valence electrons
161 > at all of the other atomic sites is computed at atom $i$'s location,
162   \begin{equation*}
163   \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
164   \end{equation*}
# Line 202 | Line 185 | properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007
185   The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
186   have all been widely used by the materials simulation community for
187   simulations of bulk and nanoparticle
188 < properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
188 > properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq,mishin99:_inter}
189   melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
190 < fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
191 < propagation,\cite{BECQUART:1993rg} and alloying
192 < dynamics.\cite{Shibata:2002hh} One of EAM's strengths
193 < is its sensitivity to small changes in structure. This arises
194 < from the original parameterization, where the interactions
195 < up to the third nearest neighbor were taken into account.\cite{Voter95a}
196 < Comparing that to the glue model of Ercolessi et al.\cite{Ercolessi88}
197 < which is only parameterized up to the nearest-neighbor
198 < interactions, EAM is a suitable choice for systems where
199 < the bulk properties are of secondary importance to low-index
200 < surface structures. Additionally, the similarity of EAMs functional
201 < treatment of the embedding energy to standard density functional
202 < theory (DFT) makes fitting DFT-derived cross potentials with adsorbates somewhat easier.
203 < \cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}  
190 > fracture,\cite{Shastry:1996qg,Shastry:1998dx,mishin01:cu} crack
191 > propagation,\cite{BECQUART:1993rg,Rifkin1992} and alloying
192 > dynamics.\cite{Shibata:2002hh,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
193 > One of EAM's strengths is its sensitivity to small changes in
194 > structure. This is due to the inclusion of up to the third nearest
195 > neighbor interactions during fitting of the parameters.\cite{Voter95a}
196 > In comparison, the glue model of Ercolessi {\it et
197 >  al}.\cite{Ercolessi88} was only parameterized to include
198 > nearest-neighbor interactions, EAM is a suitable choice for systems
199 > where the bulk properties are of secondary importance to low-index
200 > surface structures. Additionally, the similarity of EAM's functional
201 > treatment of the embedding energy to standard density functional
202 > theory (DFT) makes fitting DFT-derived cross potentials with
203 > adsorbates somewhat easier.
204  
222
223
224
205   \subsection{Carbon Monoxide model}
206 < Previous explanations for the surface rearrangements center on
207 < the large linear quadrupole moment of carbon monoxide.\cite{Tao:2010}  
208 < We used a model first proposed by Karplus and Straub to study
209 < the photodissociation of CO from myoglobin because it reproduces
210 < the quadrupole moment well.\cite{Straub} The Straub and
211 < Karplus model treats CO as a rigid three site molecule with a massless M
212 < site at the molecular center of mass. The geometry and interaction
213 < parameters are reproduced in Table~\ref{tab:CO}. The effective
214 < dipole moment, calculated from the assigned charges, is still
215 < small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
216 < to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
206 > Previous explanations for the surface rearrangements center on the
207 > large linear quadrupole moment of carbon monoxide.\cite{Tao:2010} We
208 > used a model first proposed by Karplus and Straub to study the
209 > photodissociation of CO from myoglobin because it reproduces the
210 > quadrupole moment well.\cite{Straub} The Straub and Karplus model
211 > treats CO as a rigid three site molecule with a massless
212 > charge-carrying ``M'' site at the center of mass. The geometry and
213 > interaction parameters are reproduced in Table~\ref{tab:CO}. The
214 > effective dipole moment, calculated from the assigned charges, is
215 > still small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is
216 > close to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
217   mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
218   %CO Table
219   \begin{table}[H]
220    \caption{Positions, Lennard-Jones parameters ($\sigma$ and
221 <    $\epsilon$), and charges for the CO-CO
222 <    interactions in Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
223 <    in kcal/mol, and charges are in atomic units.}
221 >    $\epsilon$), and charges for CO-CO
222 >    interactions. Distances are in \AA, energies are
223 >    in kcal/mol, and charges are in atomic units.  The CO model
224 >    from Ref.\bibpunct{}{}{,}{n}{}{,}
225 >    \protect\cite{Straub} was used without modification.}
226   \centering
227   \begin{tabular}{| c | c | ccc |}
228   \hline
# Line 267 | Line 249 | et al.,\cite{Pons:1986} the Pt-C interaction was fit t
249   position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}.
250   The modified parameters yield binding energies that are slightly higher
251   than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
252 < et al.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
253 < Lennard-Jones interaction to mimic strong, but short-ranged partial
252 > {\it et al}.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
253 > Lennard-Jones interaction to mimic strong, but short-ranged, partial
254   binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
255   Pt-O interaction was modeled with a Morse potential with a large
256   equilibrium distance, ($r_o$).  These choices ensure that the C is preferred
257 < over O as the surface-binding atom. In most cases, the Pt-O parameterization contributes a weak
257 > over O as the surface-binding atom. In most geometries, the Pt-O parameterization contributes a weak
258   repulsion which favors the atop site.  The resulting potential-energy
259   surface suitably recovers the calculated Pt-C separation length
260   (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
# Line 286 | Line 268 | periodic supercell plane-wave basis approach, as imple
268   The limited experimental data for CO adsorption on Au required refining the fits against plane-wave DFT calculations.
269   Adsorption energies were obtained from gas-surface DFT calculations with a
270   periodic supercell plane-wave basis approach, as implemented in the
271 < {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores were
271 > Quantum ESPRESSO package.\cite{QE-2009} Electron cores were
272   described with the projector augmented-wave (PAW)
273   method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
274   included to an energy cutoff of 20 Ry. Electronic energies are
# Line 300 | Line 282 | zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxe
282   performed until the energy difference between subsequent steps
283   was less than $10^{-8}$ Ry.   Nonspin-polarized supercell calculations
284   were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
285 < zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxed gold slab was
285 > zone.\cite{Monkhorst:1976} The relaxed gold slab was
286   then used in numerous single point calculations with CO at various
287   heights (and angles relative to the surface) to allow fitting of the
288   empirical force field.
# Line 309 | Line 291 | and polarization are neglected in this model, although
291   The parameters employed for the metal-CO cross-interactions in this work
292   are shown in Table~\ref{tab:co_parameters} and the binding energies on the
293   (111) surfaces are displayed in Table~\ref{tab:co_energies}.  Charge transfer
294 < and polarization are neglected in this model, although these effects are likely to
295 < affect binding energies and binding site preferences, and will be addressed in
314 < future work.
294 > and polarization are neglected in this model, although these effects could have
295 > an effect on binding energies and binding site preferences.
296  
297   %Table  of Parameters
298   %Pt Parameter Set 9
299   %Au Parameter Set 35
300   \begin{table}[H]
301 <  \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
302 <    interactions are modeled with Lennard-Jones potentials. While the
303 <    metal-O interactions were fit to Morse
301 >  \caption{Parameters for the metal-CO cross-interactions. Metal-C
302 >    interactions are modeled with Lennard-Jones potentials, while the
303 >    metal-O interactions were fit to broad Morse
304      potentials.  Distances are given in \AA~and energies in kcal/mol. }
305   \centering
306   \begin{tabular}{| c | cc | c | ccc |}
# Line 343 | Line 324 | future work.
324    \hline
325    & Calculated & Experimental \\
326    \hline
327 <  \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
327 >  \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.84} & -1.4 \bibpunct{}{}{,}{n}{}{,}
328    (Ref. \protect\cite{Kelemen:1979}) \\
329   & &  -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
330 <  \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,}  (Ref. \protect\cite{TPD_Gold}) \\
330 >  \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,}  (Ref. \protect\cite{TPDGold}) \\
331    \hline
332   \end{tabular}
333   \label{tab:co_energies}
334   \end{table}
335  
336 +
337 + \subsection{Forcefield validation}
338 + The CO-metal cross interactions were compared directly to DFT results
339 + found in the supporting information of Tao {\it et al.}
340 + \cite{Tao:2010} These calculations are estimates of the stabilization
341 + energy provided to double-layer reconstructions of the perfect 557
342 + surface by an overlayer of CO molecules in a $c (2 \times 4)$ pattern.
343 + To make the comparison, metal slabs that were five atoms thick and
344 + which displayed a 557 facet were constructed.  Double-layer
345 + (reconstructed) systems were created using six atomic layers where
346 + enough of a layer was removed from both exposed 557 facets to create
347 + the double step.  In all cases, the metal slabs contained 480 atoms
348 + and were minimized using steepest descent under the EAM force
349 + field. Both the bare metal slabs and slabs with 50\% carbon monoxide
350 + coverage (arranged in the $c (2 \times 4)$ pattern) were used.  The
351 + systems are periodic along and perpendicular to the step-edge axes
352 + with a large vacuum above the displayed 557 facet.
353 +
354 + Energies using our force field for the various systems are displayed
355 + in Table ~\ref{tab:steps}.  The relative energies are calculated as
356 + $E_{relative} = E_{system} - E_{M-557-S} - N_{CO} E_{CO-M}$,
357 + where $E_{CO-M}$ is -1.84 eV for CO-Pt and -0.39 eV for CO-Au. For
358 + platinum, the bare double layer is slightly less stable then the
359 + original single (557) step. However, addition of carbon monoxide
360 + stabilizes the reconstructed double layer relative to the perfect 557.
361 + This result is in qualitative agreement with DFT calculations in Tao
362 + {\it et al.}\cite{Tao:2010}, who also showed that the addition of CO
363 + leads to a reversal in stability.
364 +
365 + The DFT calculations suggest an increased stability of 0.08 kcal/mol
366 + (0.7128 eV) per Pt atom for going from the single to double step
367 + structure in the presence of carbon monoxide.
368 +
369 + The gold systems show much smaller energy differences between the
370 + single and double layers. The weaker binding of CO to Au is evidenced
371 + by the much smaller change in relative energy between the structures
372 + when carbon monoxide is present.  Additionally, as CO-Au binding is
373 + much weaker than CO-Pt, it would be unlikely that CO would approach
374 + the 50\% coverage levels operating temperatures for the gold surfaces.
375 +
376 + %Table of single step double step calculations
377 + \begin{table}[H]
378 +  \caption{Minimized single point energies of (S)ingle and (D)ouble
379 +    steps.  The addition of CO in a 50\% $c(2 \times 4)$ coverage acts as a
380 +    stabilizing presence and suggests a driving force for the observed
381 +    reconstruction on the highest coverage Pt system. All energies are
382 +    in kcal/mol.}
383 + \centering
384 + \begin{tabular}{| c | c | c | c | c | c |}
385 + \hline
386 + \textbf{Step} & \textbf{N}\textsubscript{M} & \textbf{N\textsubscript{CO}} & \textbf{Relative Energy} & \textbf{$\Delta$E/M} & \textbf{$\Delta$E/CO} \\
387 + \hline
388 + Pt(557)-S & 480 & 0 & 0 & 0 & - \\
389 + Pt(557)-D & 480 & 0 & 114.783 & 0.239 & -\\
390 + Pt(557)-S & 480 & 40 & -124.546 & -0.259 & -3.114\\
391 + Pt(557)-D & 480 & 44 & -34.953 & -0.073 & -0.794\\
392 + \hline
393 + \hline
394 + Au(557)-S & 480 & 0 & 0 & 0 & - \\
395 + Au(557)-D & 480 & 0 & 79.572 & 0.166 & - \\
396 + Au(557)-S & 480 & 40 & -157.199 & -0.327 & -3.930\\
397 + Au(557)-D & 480 & 44 & -123.297 & -0.257 & -2.802 \\
398 + \hline
399 + \end{tabular}
400 + \label{tab:steps}
401 + \end{table}
402 +
403 +
404   \subsection{Pt(557) and Au(557) metal interfaces}
405   Our Pt system is an orthorhombic periodic box of dimensions
406   54.482~x~50.046~x~120.88~\AA~while our Au system has
407 < dimensions of 57.4~x~51.9285~x~100~\AA.
407 > dimensions of 57.4~x~51.9285~x~100~\AA. The metal slabs
408 > are 9 and 8 atoms deep respectively, corresponding to a slab
409 > thickness of $\sim$21~\AA~ for Pt and $\sim$19~\AA~for Au.
410   The systems are arranged in a FCC crystal that have been cut
411   along the (557) plane so that they are periodic in the {\it x} and
412   {\it y} directions, and have been oriented to expose two aligned
# Line 364 | Line 415 | The different bulk melting temperatures (1337~K for Au
415   1200~K were performed to confirm the relative
416   stability of the surfaces without a CO overlayer.  
417  
418 < The different bulk melting temperatures (1337~K for Au
419 < and 2045~K for Pt) suggest that any possible reconstruction should happen at
420 < different temperatures for the two metals.  The bare Au and Pt surfaces were
421 < initially run in the canonical (NVT) ensemble at 800~K and 1000~K
422 < respectively for 100 ps. The two surfaces were relatively stable at these
423 < temperatures when no CO was present, but experienced increased surface
424 < mobility on addition of CO. Each surface was then dosed with different concentrations of CO
425 < that was initially placed in the vacuum region.  Upon full adsorption,
426 < these concentrations correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
427 < coverage. Higher coverages resulted in the formation of a double layer of CO,
428 < which introduces artifacts that are not relevant to (557) reconstruction.
429 < Because of the difference in binding energies, nearly all of the CO was bound to the Pt surface, while
430 < the Au surfaces often had a significant CO population in the gas
431 < phase.  These systems were allowed to reach thermal equilibrium (over
432 < 5~ns) before being run in the microcanonical (NVE) ensemble for
433 < data collection. All of the systems examined had at least 40~ns in the
434 < data collection stage, although simulation times for some Pt of the
435 < systems exceeded 200~ns.  Simulations were carried out using the open
436 < source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
418 > The different bulk melting temperatures predicted by EAM
419 > (1345~$\pm$~10~K for Au\cite{Au:melting} and $\sim$~2045~K for
420 > Pt\cite{Pt:melting}) suggest that any reconstructions should happen at
421 > different temperatures for the two metals.  The bare Au and Pt
422 > surfaces were initially run in the canonical (NVT) ensemble at 800~K
423 > and 1000~K respectively for 100 ps. The two surfaces were relatively
424 > stable at these temperatures when no CO was present, but experienced
425 > increased surface mobility on addition of CO. Each surface was then
426 > dosed with different concentrations of CO that was initially placed in
427 > the vacuum region.  Upon full adsorption, these concentrations
428 > correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface coverage. Higher
429 > coverages resulted in the formation of a double layer of CO, which
430 > introduces artifacts that are not relevant to (557) reconstruction.
431 > Because of the difference in binding energies, nearly all of the CO
432 > was bound to the Pt surface, while the Au surfaces often had a
433 > significant CO population in the gas phase.  These systems were
434 > allowed to reach thermal equilibrium (over 5~ns) before being run in
435 > the microcanonical (NVE) ensemble for data collection. All of the
436 > systems examined had at least 40~ns in the data collection stage,
437 > although simulation times for some Pt of the systems exceeded 200~ns.
438 > Simulations were carried out using the open source molecular dynamics
439 > package, OpenMD.\cite{Ewald,OOPSE,openmd}
440  
441  
388
389
442   % RESULTS
443   %
444   \section{Results}
445   \subsection{Structural remodeling}
446 < The surfaces of both systems, upon dosage of CO, began
447 < to undergo remodeling that was not observed in the bare
448 < metal system. The surfaces which were not exposed to CO
449 < did experience minor roughening of the step-edge because
450 < of the elevated temperatures, but the
451 < (557) lattice was well-maintained throughout the simulation
452 < time. The Au systems were limited to greater amounts of
453 < roughening, i.e. breakup of the step-edge, and some step
454 < wandering. The lower coverage Pt systems experienced
455 < similar restructuring but to a greater extent when
456 < compared to the Au systems. The 50\% coverage
405 < 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}
446 > The bare metal surfaces experienced minor roughening of the step-edge
447 > because of the elevated temperatures, but the (557) face was stable
448 > throughout the simulations. The surfaces of both systems, upon dosage
449 > of CO, began to undergo extensive remodeling that was not observed in
450 > the bare systems. Reconstructions of the Au systems were limited to
451 > breakup of the step-edges and some step wandering. The lower coverage
452 > Pt systems experienced similar step edge wandering but to a greater
453 > extent. The 50\% coverage Pt system was unique among our simulations
454 > in that it formed well-defined and stable double layers through step
455 > coalescence, similar to results reported by Tao {\it et
456 >  al}.\cite{Tao:2010}
457  
409
458   \subsubsection{Step wandering}
459 < The 0\% coverage surfaces for both metals showed minimal
460 < movement at their respective run temperatures. As the CO
461 < coverage increased however, the mobility of the surface,
462 < adatoms and step-edges alike, also increased. Additionally,
463 < at the higher coverages on both metals, there was more
464 < step-wandering. Except for the 50\% Pt system, the step-edges
465 < did not coalesce in any of the other simulations, instead preferring
466 < to keep nearly the same distance between steps as in the
467 < original (557) lattice. Previous work by Williams et al.\cite{Williams:1991, Williams:1994}
468 < highlights the repulsion that exists between step-edges even
469 < when no direct interactions are present in the system. This
470 < repulsion exists because the entropy of the step-edges is constrained
471 < since step-edge crossing is not allowed. This entropic repulsion
472 < does not completely define the interactions between steps,
473 < which is why some surfaces will undergo step coalescence,
474 < where additional attractive interactions can overcome the
475 < repulsion\cite{Williams:1991} and others will not. The presence
476 < of adsorbates can affect these step interactions, potentially
429 < leading to a new surface structure as the thermodynamic minimum.
459 > The bare surfaces for both metals showed minimal step-wandering at
460 > their respective temperatures. As the CO coverage increased however,
461 > the mobility of the surface atoms, described through adatom diffusion
462 > and step-edge wandering, also increased.  Except for the 50\% Pt
463 > system where step coalescence occurred, the step-edges in the other
464 > simulations preferred to keep nearly the same distance between steps
465 > as in the original (557) lattice, $\sim$13\AA~for Pt and
466 > $\sim$14\AA~for Au.  Previous work by Williams {\it et
467 >  al}.\cite{Williams:1991, Williams:1994} highlights the repulsion
468 > that exists between step-edges even when no direct interactions are
469 > present in the system. This repulsion is caused by an entropic barrier
470 > that arises from the fact that steps cannot cross over one
471 > another. This entropic repulsion does not completely define the
472 > interactions between steps, however, so it is possible to observe step
473 > coalescence on some surfaces.\cite{Williams:1991} The presence and
474 > concentration of adsorbates, as shown in this work, can affect
475 > step-step interactions, potentially leading to a new surface structure
476 > as the thermodynamic equilibrium.
477  
478   \subsubsection{Double layers}
479 < Tao et al. have shown experimentally that the Pt(557) surface
480 < undergoes two separate reconstructions upon CO adsorption.\cite{Tao:2010}
481 < The first involves a doubling of the step height and plateau length.
482 < Similar behavior has been seen to occur on numerous surfaces
483 < at varying conditions: Ni(977), Si(111).\cite{Williams:1994,Williams:1991,Pearl}
484 < Of the two systems we examined, the Pt system showed a greater
485 < propensity for reconstruction when compared to the Au system
486 < because of the larger surface mobility and extent of step wandering.
487 < The amount of reconstruction is correlated to the amount of CO
488 < adsorbed upon the surface.  This appears to be related to the
489 < effect that adsorbate coverage has on edge breakup and on the
490 < surface diffusion of metal adatoms. While both systems displayed
491 < step-edge wandering, only the 50\% Pt surface underwent the
492 < doubling seen by Tao et al. within the time scales studied here.
493 < Over longer periods (150~ns) two more double layers formed
494 < on this interface. Although double layer formation did not occur
495 < in the other Pt systems, they show more step-wandering and
496 < general roughening compared to their Au counterparts. The
497 < 50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
498 < various times along the simulation showing the evolution of a step-edge.
452 <
453 < 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 < the 40~ns time scale or the extended simulation time of 150~ns for
457 < the 50\% Pt system, experienced this reconstruction.
458 <
459 < \subsection{Dynamics}
460 < Previous atomistic simulations of stepped surfaces dealt largely
461 < with the energetics and structures at different conditions
462 < \cite{Williams:1991,Williams:1994}. Consequently, the most common
463 < technique utilized to date has been Monte Carlo sampling. Monte Carlo gives an efficient
464 < sampling of the equilibrium thermodynamic landscape at the expense
465 < 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 < 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 <
472 <
473 < \subsubsection{Transport of surface metal atoms}
474 < %forcedSystems/stepSeparation
475 < The movement or wandering of a step-edge is a cooperative effect
476 < arising from the individual movements of the atoms making up the steps. An ideal metal surface
477 < displaying a low index facet, (111) or (100), is unlikely to experience
478 < much surface diffusion because of the large energetic barrier that must
479 < 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 < Breaking away from the step-edge on a clean surface still imposes an
482 < energetic penalty around $\sim$~40 kcal/mol, but this is significantly easier than lifting
483 < 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 < on the surface. For certain distributions of CO, the penalty can fall as low as
486 < $\sim$~20 kcal/mol. Once an adatom exists on the surface, the barrier for
487 < diffusion is negligible ( \textless~4 kcal/mol for a Pt adatom). These adatoms are
488 < 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 < metal atoms on the Pt and Au surfaces we were able to determine the relative
493 < diffusion constants, as well as how varying coverages of CO affect the diffusion. Close
494 < observation of the mobile metal atoms showed that they were typically in
495 < 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 < observed moving together across the surfaces.
498 <
499 < A particle was considered ``mobile'' once it had traveled more than 2~\AA~
500 < between saved configurations of the system (typically 10-100 ps). An atom that was
501 < 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 < movement of buried atoms. Diffusion on a surface is strongly affected by
504 < local structures and in this work, the presence of single and double layer
505 < step-edges causes the diffusion parallel to the step-edges to be different
506 < from the diffusion perpendicular to these edges. Parallel and perpendicular
507 < diffusion constants are shown in Figure \ref{fig:diff}.
479 > Tao {\it et al}.\cite{Tao:2010} have shown experimentally that the
480 > Pt(557) surface undergoes two separate reconstructions upon CO
481 > adsorption.  The first involves a doubling of the step height and
482 > plateau length.  Similar behavior has been seen on a number of
483 > surfaces at varying conditions, including Ni(977) and
484 > Si(111).\cite{Williams:1994,Williams:1991,Pearl} Of the two systems we
485 > examined, the Pt system showed a greater propensity for reconstruction
486 > because of the larger surface mobility and the greater extent of step
487 > wandering.  The amount of reconstruction was strongly correlated to
488 > the amount of CO adsorbed upon the surface.  This appears to be
489 > related to the effect that adsorbate coverage has on edge breakup and
490 > on the surface diffusion of metal adatoms. Only the 50\% Pt surface
491 > underwent the doubling seen by Tao {\it et al}.\cite{Tao:2010} within
492 > the time scales studied here.  Over a longer time scale (150~ns) two
493 > more double layers formed on this surface. Although double layer
494 > formation did not occur in the other Pt systems, they exhibited more
495 > step-wandering and roughening compared to their Au counterparts. The
496 > 50\% Pt system is highlighted in Figure \ref{fig:reconstruct} at
497 > various times along the simulation showing the evolution of a double
498 > layer step-edge.
499  
500 < The lack of a definite trend in the Au diffusion data is likely due
501 < to the weaker bonding between Au and CO. This leads to a lower
502 < coverage ({\it x}-axis) when compared to dosage amount, which
503 < then further limits the affects of the surface diffusion. The correlation
504 < 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.
500 > The second reconstruction observed by Tao {\it et al}.\cite{Tao:2010}
501 > involved the formation of triangular clusters that stretched across
502 > the plateau between two step-edges. Neither of the simulated metal
503 > interfaces, within the 40~ns time scale or the extended time of 150~ns
504 > for the 50\% Pt system, experienced this reconstruction.
505  
536 \subsubsection{Dynamics of double layer formation}
537 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 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 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 formed the double layer and by 86~ns, the complete layer
550 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
506   %Evolution of surface
507   \begin{figure}[H]
508 < \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
509 < \caption{The Pt(557) / 50\% CO system at a sequence of times after
510 <  initial exposure to the CO: (a) 258~ps, (b) 19~ns, (c) 31.2~ns, and
511 <  (d) 86.1~ns. Disruption of the (557) step-edges occurs quickly.  The
508 > \includegraphics[width=\linewidth]{EPS_ProgressionOfDoubleLayerFormation}
509 > \caption{The Pt(557) / 50\% CO interface upon exposure to the CO: (a)
510 >  258~ps, (b) 19~ns, (c) 31.2~ns, and (d) 86.1~ns after
511 >  exposure. Disruption of the (557) step-edges occurs quickly.  The
512    doubling of the layers appears only after two adjacent step-edges
513    touch.  The circled spot in (b) nucleated the growth of the double
514    step observed in the later configurations.}
515    \label{fig:reconstruct}
516   \end{figure}
517  
518 + \subsection{Dynamics}
519 + Previous experimental work by Pearl and Sibener\cite{Pearl}, using
520 + STM, has been able to capture the coalescence of steps on Ni(977). The
521 + time scale of the image acquisition, $\sim$70~s/image, provides an
522 + upper bound for the time required for the doubling to occur. By
523 + utilizing Molecular Dynamics we are able to probe the dynamics of
524 + these reconstructions at elevated temperatures and in this section we
525 + provide data on the timescales for transport properties,
526 + e.g. diffusion and layer formation time.
527 +
528 +
529 + \subsubsection{Transport of surface metal atoms}
530 + %forcedSystems/stepSeparation
531 +
532 + The wandering of a step-edge is a cooperative effect arising from the
533 + individual movements of the atoms making up the steps. An ideal metal
534 + surface displaying a low index facet, (111) or (100), is unlikely to
535 + experience much surface diffusion because of the large energetic
536 + barrier that must be overcome to lift an atom out of the surface. The
537 + presence of step-edges and other surface features on higher-index
538 + facets provides a lower energy source for mobile metal atoms.  Using
539 + our potential model, single-atom break-away from a step-edge on a
540 + clean surface still imposes an energetic penalty around
541 + $\sim$~45~kcal/mol, but this is certainly easier than lifting the same
542 + metal atom vertically out of the surface, \textgreater~60~kcal/mol.
543 + The penalty lowers significantly when CO is present in sufficient
544 + quantities on the surface. For certain distributions of CO, the
545 + energetic penalty can fall to as low as $\sim$~20~kcal/mol. The
546 + configurations that create these lower barriers are detailed in the
547 + discussion section below.
548 +
549 + Once an adatom exists on the surface, the barrier for diffusion is
550 + negligible (\textless~4~kcal/mol for a Pt adatom). These adatoms are
551 + then able to explore the terrace before rejoining either their
552 + original step-edge or becoming a part of a different edge. It is an
553 + energetically unfavorable process with a high barrier for an atom to
554 + traverse to a separate terrace although the presence of CO can lower
555 + the energy barrier required to lift or lower an adatom. By tracking
556 + the mobility of individual metal atoms on the Pt and Au surfaces we
557 + were able to determine the relative diffusion constants, as well as
558 + how varying coverages of CO affect the diffusion. Close observation of
559 + the mobile metal atoms showed that they were typically in equilibrium
560 + with the step-edges.  At times, their motion was concerted, and two or
561 + more adatoms would be observed moving together across the surfaces.
562 +
563 + A particle was considered ``mobile'' once it had traveled more than
564 + 2~\AA~ between saved configurations of the system (typically 10-100
565 + ps). A mobile atom would typically travel much greater distances than
566 + this, but the 2~\AA~cutoff was used to prevent swamping the diffusion
567 + data with the in-place vibrational movement of buried atoms. Diffusion
568 + on a surface is strongly affected by local structures and the presence
569 + of single and double layer step-edges causes the diffusion parallel to
570 + the step-edges to be larger than the diffusion perpendicular to these
571 + edges. Parallel and perpendicular diffusion constants are shown in
572 + Figure \ref{fig:diff}.  Diffusion parallel to the step-edge is higher
573 + than diffusion perpendicular to the edge because of the lower energy
574 + barrier associated with sliding along an edge compared to breaking
575 + away to form an isolated adatom.
576 +
577 + %Diffusion graph
578   \begin{figure}[H]
579 < \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade.pdf}
579 > \includegraphics[width=\linewidth]{Portrait_DiffusionComparison_1}
580   \caption{Diffusion constants for mobile surface atoms along directions
581    parallel ($\mathbf{D}_{\parallel}$) and perpendicular
582    ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
583 <  surface coverage.  Diffusion parallel to the step-edge is higher
584 <  than that perpendicular to the edge because of the lower energy
585 <  barrier associated with traversing along the edge as compared to
586 <  completely breaking away. Additionally, the observed
583 <  maximum and subsequent decrease for the Pt system suggests that the
584 <  CO self-interactions are playing a significant role with regards to
585 <  movement of the Pt atoms around and across the surface. }
583 >  surface coverage.  The two reported diffusion constants for the 50\%
584 >  Pt system correspond to a 20~ns period before the formation of the
585 >  double layer (upper points), and to the full 40~ns sampling period
586 >  (lower points).}
587   \label{fig:diff}
588   \end{figure}
589  
590 + The weaker Au-CO interaction is evident in the weak CO-coverage
591 + dependance of Au diffusion. This weak interaction leads to lower
592 + observed coverages when compared to dosage amounts. This further
593 + limits the effect the CO can have on surface diffusion. The correlation
594 + between coverage and Pt diffusion rates shows a near linear relationship
595 + at the earliest times in the simulations. Following double layer formation,
596 + however, there is a precipitous drop in adatom diffusion. As the double
597 + layer forms, many atoms that had been tracked for mobility data have
598 + now been buried, resulting in a smaller reported diffusion constant. A
599 + secondary effect of higher coverages is CO-CO cross interactions that
600 + lower the effective mobility of the Pt adatoms that are bound to each CO.
601 + This effect would become evident only at higher coverages. A detailed
602 + account of Pt adatom energetics follows in the Discussion.
603 +
604 + \subsubsection{Dynamics of double layer formation}
605 + The increased diffusion on Pt at the higher CO coverages is the primary
606 + contributor to double layer formation. However, this is not a complete
607 + explanation -- the 33\%~Pt system has higher diffusion constants, but
608 + did not show any signs of edge doubling in 40~ns. On the 50\%~Pt
609 + system, one double layer formed within the first 40~ns of simulation time,
610 + while two more were formed as the system was allowed to run for an
611 + additional 110~ns (150~ns total). This suggests that this reconstruction
612 + is a rapid process and that the previously mentioned upper bound is a
613 + very large overestimate.\cite{Williams:1991,Pearl} In this system the first
614 + appearance of a double layer appears at 19~ns into the simulation.
615 + Within 12~ns of this nucleation event, nearly half of the step has formed
616 + the double layer and by 86~ns the complete layer has flattened out.
617 + From the appearance of the first nucleation event to the first observed
618 + double layer, the process took $\sim$20~ns. Another $\sim$40~ns was
619 + necessary for the layer to completely straighten. The other two layers in
620 + this simulation formed over periods of 22~ns and 42~ns respectively.
621 + A possible explanation for this rapid reconstruction is the elevated
622 + temperatures under which our systems were simulated. The process
623 + would almost certainly take longer at lower temperatures. Additionally,
624 + our measured times for completion of the doubling after the appearance
625 + of a nucleation site are likely affected by our periodic boxes. A longer
626 + step-edge will likely take longer to ``zipper''.
627  
628  
591
629   %Discussion
630   \section{Discussion}
631 < We have shown that the classical potential models are able to model the initial reconstruction of the
632 < Pt(557) surface upon CO adsorption as shown by Tao et al. \cite{Tao:2010}. More importantly, we
633 < were able to observe features of the dynamic processes necessary for this reconstruction.
631 > We have shown that a classical potential is able to model the initial
632 > reconstruction of the Pt(557) surface upon CO adsorption, and have
633 > reproduced the double layer structure observed by Tao {\it et
634 >  al}.\cite{Tao:2010}. Additionally, this reconstruction appears to be
635 > rapid -- occurring within 100 ns of the initial exposure to CO.  Here
636 > we discuss the features of the classical potential that are
637 > contributing to the stability and speed of the Pt(557) reconstruction.
638  
639 < \subsection{Mechanism for restructuring}
640 < Since the Au surface showed no large scale restructuring throughout
641 < our simulation time our discussion will focus on the 50\% Pt-CO system
642 < which did undergo the doubling featured in Figure \ref{fig:reconstruct}.
643 < Similarities of our results to those reported previously by
644 < Tao et al.\cite{Tao:2010} are quite
645 < strong. The simulated Pt
646 < system exposed to a large dosage of CO readily restructures by doubling the terrace
647 < 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.
648 < The adatoms either
649 < break away from the step-edge and stay on the lower terrace or they lift
650 < up onto a higher terrace. Once ``free'', they diffuse on the terrace
651 < until reaching another step-edge or rejoining their original edge.  
652 < 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 < meet as shown in Figure 1.B, this nucleates the rest of the edge to meet up, forming a double layer.
614 < 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.
639 > \subsection{Diffusion}
640 > The perpendicular diffusion constant appears to be the most important
641 > indicator of double layer formation. As highlighted in Figure
642 > \ref{fig:reconstruct}, the formation of the double layer did not begin
643 > until a nucleation site appeared.  Williams {\it et
644 >  al}.\cite{Williams:1991,Williams:1994} cite an effective edge-edge
645 > repulsion arising from the inability of edge crossing.  This repulsion
646 > must be overcome to allow step coalescence.  A larger
647 > $\textbf{D}_\perp$ value implies more step-wandering and a larger
648 > chance for the stochastic meeting of two edges to create a nucleation
649 > point.  Diffusion parallel to the step-edge can help ``zipper'' up a
650 > nascent double layer. This helps explain the rapid time scale for
651 > double layer completion after the appearance of a nucleation site, while
652 > the initial appearance of the nucleation site was unpredictable.
653  
654 < A number of possible mechanisms exist to explain the role of adsorbed
655 < CO in restructuring the Pt surface. Quadrupolar repulsion between adjacent
656 < CO molecules adsorbed on the surface is one possibility.  However,
657 < the quadrupole-quadrupole interaction is short-ranged and is attractive for
658 < some orientations.  If the CO molecules are ``locked'' in a specific orientation
659 < relative to each other, through atop adsorption for example, this explanation
660 < gains some credence.  The energetic repulsion between two CO located a
661 < distance of 2.77~\AA~apart (nearest-neighbor distance of Pt) and both in
662 < a  vertical orientation, is 8.62 kcal/mol. Moving the CO apart to the second
663 < nearest-neighbor distance of 4.8~\AA~or 5.54~\AA~drops the repulsion to
664 < nearly 0 kcal/mol. Allowing the CO's to leave a purely vertical orientation
665 < 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.
666 < As mentioned above, the energy barrier for surface diffusion
667 < of a Pt adatom is only 4 kcal/mol. So this repulsion between neighboring CO molecules can
668 < increase the surface diffusion. However, the residence time of CO on Pt was
669 < examined and while the majority of the CO is on or near the surface throughout
670 < the run, most molecules are mobile. This mobility suggests that the CO are more
671 < likely to shift their positions without necessarily the Pt along with them.
654 > \subsection{Mechanism for restructuring}
655 > Since the Au surface showed no large scale restructuring in any of our
656 > simulations, our discussion will focus on the 50\% Pt-CO system which
657 > did exhibit doubling. A number of possible mechanisms exist to explain
658 > the role of adsorbed CO in restructuring the Pt surface. Quadrupolar
659 > repulsion between adjacent CO molecules adsorbed on the surface is one
660 > possibility.  However, the quadrupole-quadrupole interaction is
661 > short-ranged and is attractive for some orientations.  If the CO
662 > molecules are ``locked'' in a vertical orientation, through atop
663 > adsorption for example, this explanation would gain credence. Within
664 > the framework of our classical potential, the calculated energetic
665 > repulsion between two CO molecules located a distance of
666 > 2.77~\AA~apart (nearest-neighbor distance of Pt) and both in a
667 > vertical orientation, is 8.62 kcal/mol. Moving the CO to the second
668 > nearest-neighbor distance of 4.8~\AA~drops the repulsion to nearly
669 > 0. Allowing the CO to rotate away from a purely vertical orientation
670 > also lowers the repulsion. When the carbons are locked at a distance
671 > of 2.77~\AA, a minimum of 6.2 kcal/mol is reached when the angle
672 > between the 2 CO is $\sim$24\textsuperscript{o}.  The calculated
673 > barrier for surface diffusion of a Pt adatom is only 4 kcal/mol, so
674 > repulsion between adjacent CO molecules bound to Pt could indeed
675 > increase the surface diffusion. However, the residence time of CO on
676 > Pt suggests that the CO molecules are extremely mobile, with diffusion
677 > constants 40 to 2500 times larger than surface Pt atoms. This mobility
678 > suggests that the CO molecules jump between different Pt atoms
679 > throughout the simulation.  However, they do stay bound to individual
680 > Pt atoms for long enough to modify the local energy landscape for the
681 > mobile adatoms.
682  
683 < Another possible and more likely mechanism for the restructuring is in the
684 < destabilization of strong Pt-Pt interactions by CO adsorbed on surface
685 < Pt atoms.  This would then have the effect of increasing surface mobility
686 < of these atoms.  To test this hypothesis, numerous configurations of
687 < CO in varying quantities were arranged on the higher and lower plateaus
688 < around a step on a otherwise clean Pt(557) surface. One representative
689 < configuration is displayed in Figure \ref{fig:lambda}. Single or concerted movement
690 < of Pt atoms was then examined to determine possible barriers. Because
691 < the movement was forced along a pre-defined reaction coordinate that may differ
692 < from the true minimum of this path, only the beginning and ending energies
693 < 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
694 < locations can lead to lowered barriers for Pt breaking apart from the step-edge.
695 < Additionally, as highlighted in Figure \ref{fig:lambda}, the presence of CO makes the
696 < burrowing and lifting of adatoms favorable, whereas without CO, the process is neutral
697 < in terms of energetics.
683 > A different interpretation of the above mechanism which takes the
684 > large mobility of the CO into account, would be in the destabilization
685 > of Pt-Pt interactions due to bound CO.  Destabilizing Pt-Pt bonds at
686 > the edges could lead to increased step-edge breakup and diffusion. On
687 > the bare Pt(557) surface the barrier to completely detach an edge atom
688 > is $\sim$43~kcal/mol, as is shown in configuration (a) in Figures
689 > \ref{fig:SketchGraphic} \& \ref{fig:SketchEnergies}. For certain
690 > configurations, cases (e), (g), and (h), the barrier can be lowered to
691 > $\sim$23~kcal/mol by the presence of bound CO molecules. In these
692 > instances, it becomes energetically favorable to roughen the edge by
693 > introducing a small separation of 0.5 to 1.0~\AA. This roughening
694 > becomes immediately obvious in simulations with significant CO
695 > populations. The roughening is present to a lesser extent on surfaces
696 > with lower CO coverage (and even on the bare surfaces), although in
697 > these cases it is likely due to random fluctuations that squeeze out
698 > step-edge atoms. Step-edge breakup by direct single-atom translations
699 > (as suggested by these energy curves) is probably a worst-case
700 > scenario.  Multistep mechanisms in which an adatom moves laterally on
701 > the surface after being ejected would be more energetically favorable.
702 > This would leave the adatom alongside the ledge, providing it with
703 > five nearest neighbors.  While fewer than the seven neighbors it had
704 > as part of the step-edge, it keeps more Pt neighbors than the three
705 > neighbors an isolated adatom has on the terrace. In this proposed
706 > mechanism, the CO quadrupolar repulsion still plays a role in the
707 > initial roughening of the step-edge, but not in any long-term bonds
708 > with individual Pt atoms.  Higher CO coverages create more
709 > opportunities for the crowded CO configurations shown in Figure
710 > \ref{fig:SketchGraphic}, and this is likely to cause an increased
711 > propensity for step-edge breakup.
712  
713 < %lambda progression of Pt -> shoving its way into the step
713 > %Sketch graphic of different configurations
714   \begin{figure}[H]
715 < \includegraphics[width=\linewidth]{lambdaProgression_atopCO_withLambda.png}
716 < \caption{A model system of the Pt(557) surface was used as the framework
717 < for exploring energy barriers along a reaction coordinate. Various numbers,
718 < placements, and rotations of CO were examined as they affect Pt movement.
719 < The coordinate displayed in this Figure was a representative run. As shown
720 < in Table \ref{tab:rxcoord}, relative to the energy of the system at 0\%, there
721 < is a slight decrease upon insertion of the Pt atom into the step-edge along
722 < with the resultant lifting of the other Pt atom when CO is present at certain positions.}
723 < \label{fig:lambda}
715 > \includegraphics[width=\linewidth]{COpaths}
716 > \caption{Configurations used to investigate the mechanism of step-edge
717 >  breakup on Pt(557). In each case, the central (starred) atom was
718 >  pulled directly across the surface away from the step edge.  The Pt
719 >  atoms on the upper terrace are colored dark grey, while those on the
720 >  lower terrace are in white.  In each of these configurations, some
721 >  of the atoms (highlighted in blue) had CO molecules bound in the
722 >  vertical atop position.  The energies of these configurations as a
723 >  function of central atom displacement are displayed in Figure
724 >  \ref{fig:SketchEnergies}.}
725 > \label{fig:SketchGraphic}
726   \end{figure}
727  
728 + %energy graph corresponding to sketch graphic
729   \begin{figure}[H]
730 < \includegraphics[totalheight=0.9\textheight]{lambdaTable.png}
731 < \caption{}
732 < \label{fig:lambdaTable}
730 > \includegraphics[width=\linewidth]{Portrait_SeparationComparison}
731 > \caption{Energies for displacing a single edge atom perpendicular to
732 >  the step edge as a function of atomic displacement. Each of the
733 >  energy curves corresponds to one of the labeled configurations in
734 >  Figure \ref{fig:SketchGraphic}, and the energies are referenced to
735 >  the unperturbed step-edge.  Certain arrangements of bound CO
736 >  (notably configurations g and h) can lower the energetic barrier for
737 >  creating an adatom relative to the bare surface (configuration a).}
738 > \label{fig:SketchEnergies}
739   \end{figure}
740  
741 + While configurations of CO on the surface are able to increase
742 + diffusion and the likelihood of edge wandering, this does not provide
743 + a complete explanation for the formation of double layers. If adatoms
744 + were constrained to their original terraces then doubling could not
745 + occur.  A mechanism for vertical displacement of adatoms at the
746 + step-edge is required to explain the doubling.
747  
748 < \subsection{Diffusion}
749 < The diffusion parallel to the step-edge tends to be
750 < much larger than that perpendicular to the step-edge. The dynamic
751 < equilibrium that is established between the step-edge and adatom interface. The coverage
752 < of CO also appears to play a slight role in relative rates of diffusion, as shown in Figure \ref{fig:diff}.
753 < The
754 < Thus, the bottleneck of the double layer formation appears to be the initial formation
755 < of this growth point, which seems to be somewhat of a stochastic event. Once it
756 < appears, parallel diffusion, along the now slightly angled step-edge, will allow for
757 < a faster formation of the double layer than if the entire process were dependent on
758 < only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the
759 < more likely a growth point is to be formed.
760 < \\
748 > We have discovered one possible mechanism for a CO-mediated vertical
749 > displacement of Pt atoms at the step edge. Figure \ref{fig:lambda}
750 > shows four points along a reaction coordinate in which a CO-bound
751 > adatom along the step-edge ``burrows'' into the edge and displaces the
752 > original edge atom onto the higher terrace.  A number of events
753 > similar to this mechanism were observed during the simulations.  We
754 > predict an energetic barrier of 20~kcal/mol for this process (in which
755 > the displaced edge atom follows a curvilinear path into an adjacent
756 > 3-fold hollow site).  The barrier heights we obtain for this reaction
757 > coordinate are approximate because the exact path is unknown, but the
758 > calculated energy barriers would be easily accessible at operating
759 > conditions.  Additionally, this mechanism is exothermic, with a final
760 > energy 15~kcal/mol below the original $\lambda = 0$ configuration.
761 > When CO is not present and this reaction coordinate is followed, the
762 > process is endothermic by 3~kcal/mol.  The difference in the relative
763 > energies for the $\lambda=0$ and $\lambda=1$ case when CO is present
764 > provides strong support for CO-mediated Pt-Pt interactions giving rise
765 > to the doubling reconstruction.
766  
767 + %lambda progression of Pt -> shoving its way into the step
768 + \begin{figure}[H]
769 + \includegraphics[width=\linewidth]{EPS_rxnCoord}
770 + \caption{Points along a possible reaction coordinate for CO-mediated
771 +  edge doubling. Here, a CO-bound adatom burrows into an established
772 +  step edge and displaces an edge atom onto the upper terrace along a
773 +  curvilinear path.  The approximate barrier for the process is
774 +  20~kcal/mol, and the complete process is exothermic by 15~kcal/mol
775 +  in the presence of CO, but is endothermic by 3~kcal/mol without CO.}
776 + \label{fig:lambda}
777 + \end{figure}
778  
779 + The mechanism for doubling on the Pt(557) surface appears to require
780 + the cooperation of at least two distinct processes. For complete
781 + doubling of a layer to occur there must be a breakup of one
782 + terrace. These atoms must then ``disappear'' from that terrace, either
783 + by travelling to the terraces above or below their original levels.
784 + The presence of CO helps explain mechanisms for both of these
785 + situations. There must be sufficient breakage of the step-edge to
786 + increase the concentration of adatoms on the surface and these adatoms
787 + must then undergo the burrowing highlighted above (or a comparable
788 + mechanism) to create the double layer.  With sufficient time, these
789 + mechanisms working in concert lead to the formation of a double layer.
790 +
791 + \subsection{CO Removal and double layer stability}
792 + Once the double layers had formed on the 50\%~Pt system, they remained
793 + stable for the rest of the simulation time with minimal movement.
794 + Random fluctuations that involved small clusters or divots were
795 + observed, but these features typically healed within a few
796 + nanoseconds.  Within our simulations, the formation of the double
797 + layer appeared to be irreversible and a double layer was never
798 + observed to split back into two single layer step-edges while CO was
799 + present.
800 +
801 + To further gauge the effect CO has on this surface, additional
802 + simulations were run starting from a late configuration of the 50\%~Pt
803 + system that had already formed double layers. These simulations then
804 + had their CO molecules suddenly removed.  The double layer broke apart
805 + rapidly in these simulations, showing a well-defined edge-splitting
806 + after 100~ps. Configurations of this system are shown in Figure
807 + \ref{fig:breaking}. The coloring of the top and bottom layers helps to
808 + show how much mixing the edges experience as they split. These systems
809 + were only examined for 10~ns, and within that time despite the initial
810 + rapid splitting, the edges only moved another few \AA~apart. It is
811 + possible that with longer simulation times, the (557) surface recovery
812 + observed by Tao {\it et al}.\cite{Tao:2010} could also be recovered.
813 +
814   %breaking of the double layer upon removal of CO
815   \begin{figure}[H]
816 < \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
817 < \caption{(A)  0~ps, (B) 100~ps, (C) 1~ns, after the removal of CO. The presence of the CO
818 < helped maintain the stability of the double layer and upon removal the two layers break
819 < and begin separating. The separation is not a simple pulling apart however, rather
820 < there is a mixing of the lower and upper atoms at the edge.}
816 > \includegraphics[width=\linewidth]{EPS_doubleLayerBreaking}
817 > \caption{Behavior of an established (111) double step after removal of
818 >  the adsorbed CO: (A) 0~ps, (B) 100~ps, and (C) 1~ns after the
819 >  removal of CO.  Nearly immediately after the CO is removed, the
820 >  step edge reforms in a (100) configuration, which is also the step
821 >  type seen on clean (557) surfaces. The step separation involves
822 >  significant mixing of the lower and upper atoms at the edge.}
823   \label{fig:breaking}
824   \end{figure}
825  
826  
697
698
827   %Peaks!
828   %\begin{figure}[H]
829   %\includegraphics[width=\linewidth]{doublePeaks_noCO.png}
# Line 709 | Line 837 | more likely a growth point is to be formed.
837   %Don't think I need this
838   %clean surface...
839   %\begin{figure}[H]
840 < %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
840 > %\includegraphics[width=\linewidth]{557_300K_cleanPDF}
841   %\caption{}
842  
843   %\end{figure}
# Line 717 | Line 845 | In this work we have shown the reconstruction of the P
845  
846  
847   \section{Conclusion}
848 < 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.
848 > The strength and directionality of the Pt-CO binding interaction, as
849 > well as the large quadrupolar repulsion between atop-bound CO
850 > molecules, help to explain the observed increase in surface mobility
851 > of Pt(557) and the resultant reconstruction into a double-layer
852 > configuration at the highest simulated CO-coverages.  The weaker Au-CO
853 > interaction results in significantly lower adataom diffusion
854 > constants, less step-wandering, and a lack of the double layer
855 > reconstruction on the Au(557) surface.
856  
857 + An in-depth examination of the energetics shows the important role CO
858 + plays in increasing step-breakup and in facilitating edge traversal
859 + which are both necessary for double layer formation.
860 +
861   %Things I am not ready to remove yet
862  
863   %Table of Diffusion Constants
# Line 741 | Line 880 | In this work we have shown the reconstruction of the P
880   % \end{tabular}
881   % \end{table}
882  
883 < \section{Acknowledgments}
884 < Support for this project was provided by the National Science
885 < Foundation under grant CHE-0848243 and by the Center for Sustainable
886 < Energy at Notre Dame (cSEND). Computational time was provided by the
887 < Center for Research Computing (CRC) at the University of Notre Dame.
888 <
883 > \begin{acknowledgement}
884 >  We gratefully acknowledge conversations with Dr. William
885 >  F. Schneider and Dr. Feng Tao.  Support for this project was
886 >  provided by the National Science Foundation under grant CHE-0848243
887 >  and by the Center for Sustainable Energy at Notre Dame
888 >  (cSEND). Computational time was provided by the Center for Research
889 >  Computing (CRC) at the University of Notre Dame.
890 > \end{acknowledgement}
891   \newpage
892 < \bibliography{firstTryBibliography}
893 < \end{doublespace}
892 > \bibstyle{achemso}
893 > \bibliography{COonPtAu}
894 > %\end{doublespace}
895 >
896 > \begin{tocentry}
897 > \begin{wrapfigure}{l}{0.5\textwidth}
898 > \begin{center}
899 > \includegraphics[width=\linewidth]{TOC_doubleLayer}
900 > \end{center}
901 > \end{wrapfigure}
902 > A reconstructed Pt(557) surface after 86~ns exposure to a half a
903 > monolayer of CO.  The double layer that forms is a result of
904 > CO-mediated step-edge wandering as well as a burrowing mechanism that
905 > helps lift edge atoms onto an upper terrace.
906 > \end{tocentry}
907 >
908   \end{document}

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines