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 3812 by jmichalk, Fri Dec 14 04:02:29 2012 UTC vs.
Revision 3868 by jmichalk, Tue Mar 5 15:43:47 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}
# Line 10 | Line 12
12   %\usepackage{booktabs}
13   %\usepackage{bibentry}
14   %\usepackage{mathrsfs}
13 %\usepackage[ref]{overcite}
15   \usepackage[square, comma, sort&compress]{natbib}
16   \usepackage{url}
17   \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
# Line 18 | Line 19
19   9.0in \textwidth 6.5in \brokenpenalty=10000
20  
21   % double space list of tables and figures
22 < \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22 > %\AtBeginDelayedFloats{\renewcomand{\baselinestretch}{1.66}}
23   \setlength{\abovecaptionskip}{20 pt}
24   \setlength{\belowcaptionskip}{30 pt}
25  
26 < %\renewcommand\citemid{\ } % no comma in optional reference note
26 < \bibpunct{[}{]}{,}{n}{}{;}
26 > \bibpunct{}{}{,}{s}{}{;}
27   \bibliographystyle{achemso}
28  
29   \begin{document}
# Line 48 | Line 48
48   %%
49  
50   %Title
51 < \title{Investigation of the Pt and Au 557 Surface Reconstructions
52 <  under a CO Atmosphere}
53 < \author{Joseph R. Michalka, Patrick W. MacIntyre and J. Daniel
51 > \title{Molecular Dynamics simulations of the surface reconstructions
52 >  of Pt(557) and Au(557) under exposure to CO}
53 >
54 > \author{Joseph R. Michalka, Patrick W. McIntyre and J. Daniel
55   Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
56   Department of Chemistry and Biochemistry,\\
57   University of Notre Dame\\
58   Notre Dame, Indiana 46556}
59 +
60   %Date
61 < \date{Dec 15,  2012}
61 > \date{Mar 5, 2013}
62 >
63   %authors
64  
65   % make the title
# Line 65 | Line 68 | Notre Dame, Indiana 46556}
68   \begin{doublespace}
69  
70   \begin{abstract}
71 <
71 > We examine potential surface reconstructions of Pt and Au(557)
72 > under various CO coverages using molecular dynamics in order
73 > to explore possible mechanisms for any observed reconstructions and their dynamics.
74 > The metal-CO interactions were parameterized as part of this
75 > work so that an efficient large-scale treatment of this system could be
76 > undertaken. The large difference in binding strengths of the metal-CO
77 > interactions was found to play a significant role with regards to
78 > step-edge stability and adatom diffusion. A small correlation
79 > between coverage and the magnitude of the diffusion constant was
80 > also determined. An in-depth examination of the energetics of CO
81 > adsorbed to the surface provides results that appear sufficient to explain the
82 > reconstructions observed on the Pt systems and the corresponding lack  
83 > on the Au systems.
84   \end{abstract}
85  
86   \newpage
# Line 79 | Line 94 | Industrial catalysts usually consist of small particle
94   %       Sub: Also, easier to observe what is going on and provide reasons and explanations
95   %
96  
97 < Industrial catalysts usually consist of small particles exposing
98 < different atomic terminations that exhibit a high concentration of
99 < step, kink sites, and vacancies at the edges of the facets.  These
85 < sites are thought to be the locations of catalytic
97 > Industrial catalysts usually consist of small particles that exhibit a
98 > high concentration of steps, kink sites, and vacancies at the edges of
99 > the facets.  These sites are thought to be the locations of catalytic
100   activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
101 < significant evidence to demonstrate that solid surfaces are often
102 < structurally, compositionally, and chemically {\it modified} by
103 < reactants under operating conditions.\cite{Tao2008,Tao:2010,Tao2011}
104 < The coupling between surface oxidation state and catalytic activity
105 < for CO oxidation on Pt, for instance, is widely
106 < documented.\cite{Ertl08,Hendriksen:2002} Despite the well-documented
107 < role of these effects on reactivity, the ability to capture or predict
108 < them in atomistic models is currently somewhat limited.  While these
109 < effects are perhaps unsurprising on the highly disperse, multi-faceted
110 < nanoscale particles that characterize industrial catalysts, they are
111 < manifest even on ordered, well-defined surfaces. The Pt(557) surface,
112 < for example, exhibits substantial and reversible restructuring under
113 < exposure to moderate pressures of carbon monoxide.\cite{Tao:2010}
101 > significant evidence that solid surfaces are often structurally,
102 > compositionally, and chemically modified by reactants under operating
103 > conditions.\cite{Tao2008,Tao:2010,Tao2011} The coupling between
104 > surface oxidation states and catalytic activity for CO oxidation on
105 > Pt, for instance, is widely documented.\cite{Ertl08,Hendriksen:2002}
106 > Despite the well-documented role of these effects on reactivity, the
107 > ability to capture or predict them in atomistic models is somewhat
108 > limited.  While these effects are perhaps unsurprising on the highly
109 > disperse, multi-faceted nanoscale particles that characterize
110 > industrial catalysts, they are manifest even on ordered, well-defined
111 > surfaces. The Pt(557) surface, for example, exhibits substantial and
112 > reversible restructuring under exposure to moderate pressures of
113 > carbon monoxide.\cite{Tao:2010}
114  
115 < This work is part of an ongoing effort to understand the causes,
116 < mechanisms and timescales for surface restructuring using molecular
117 < simulation methods.  Since the dynamics of the process is of
118 < particular interest, we utilize classical molecular dynamic methods
119 < with force fields that represent a compromise between chemical
120 < accuracy and the computational efficiency necessary to observe the
121 < process of interest.
115 > This work is an attempt to understand the mechanism and timescale for
116 > surface restructuring by using molecular simulations.  Since the dynamics
117 > of the process are of particular interest, we employ classical force
118 > fields that represent a compromise between chemical accuracy and the
119 > computational efficiency necessary to simulate the process of interest.
120 > Since restructuring typically occurs as a result of specific interactions of the
121 > catalyst with adsorbates, in this work, two metal systems exposed
122 > to carbon monoxide were examined. The Pt(557) surface has already been shown
123 > to reconstruct under certain conditions. The Au(557) surface, because
124 > of a weaker interaction with CO, is less likely to undergo this kind
125 > of reconstruction.  
126  
127 < Since restructuring occurs as a result of specific interactions of the catalyst
128 < with adsorbates, two metals systems exposed to the same adsorbate, CO,
111 < were examined in this work. The Pt(557) surface has already been shown to
112 < reconstruct under certain conditions. The Au(557) surface, because of gold's
113 < weaker interaction with CO, is less likely to undergo such a large reconstruction.
127 >
128 >
129   %Platinum molecular dynamics
130   %gold molecular dynamics
131  
117
118
119
120
121
132   \section{Simulation Methods}
133   The challenge in modeling any solid/gas interface problem is the
134   development of a sufficiently general yet computationally tractable
# Line 133 | Line 143 | Coulomb potential.  For this work, we have been using
143   typically not well represented in terms of classical pairwise
144   interactions in the same way that bonds in a molecular material are,
145   nor are they captured by simple non-directional interactions like the
146 < Coulomb potential.  For this work, we have been using classical
147 < molecular dynamics with potential energy surfaces that are
148 < specifically tuned for transition metals.  In particular, we use the
149 < EAM potential for Au-Au and Pt-Pt interactions, while modeling the CO
150 < using a model developed by Straub and Karplus for studying
151 < photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and Pt-CO
152 < cross interactions were parameterized as part of this work.
146 > Coulomb potential.  For this work, we have used classical molecular
147 > dynamics with potential energy surfaces that are specifically tuned
148 > for transition metals.  In particular, we used the EAM potential for
149 > Au-Au and Pt-Pt interactions\cite{EAM}, while modeling the CO using a rigid
150 > three-site model developed by Straub and Karplus for studying
151 > photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and
152 > Pt-CO cross interactions were parameterized as part of this work.
153    
154   \subsection{Metal-metal interactions}
155 < Many of the potentials used for classical simulation of transition
156 < metals are based on a non-pairwise additive functional of the local
157 < electron density. The embedded atom method (EAM) is perhaps the best
158 < known of these
155 > Many of the potentials used for modeling transition metals are based
156 > on a non-pairwise additive functional of the local electron
157 > density. The embedded atom method (EAM) is perhaps the best known of
158 > these
159   methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
160   but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
161   the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
162 < parameter sets. The glue model of Ercolessi {\it et al.} is among the
162 > parameter sets. The glue model of Ercolessi et al. is among the
163   fastest of these density functional approaches.\cite{Ercolessi88} In
164   all of these models, atoms are conceptualized as a positively charged
165   core with a radially-decaying valence electron distribution. To
166   calculate the energy for embedding the core at a particular location,
167   the electron density due to the valence electrons at all of the other
168 < atomic sites is computed at atom $i$'s location,
168 > atomic sites is computed at atom $i$'s location,
169   \begin{equation*}
170   \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
171   \end{equation*}
# Line 167 | Line 177 | $\phi_{ij}(r_{ij})$ is an pairwise term that is meant
177   V_i =  F[ \bar{\rho}_i ]  + \sum_{j \neq i} \phi_{ij}(r_{ij})
178   \end{equation*}
179   where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
180 < $\phi_{ij}(r_{ij})$ is an pairwise term that is meant to represent the
181 < overlap of the two positively charged cores.  
180 > $\phi_{ij}(r_{ij})$ is a pairwise term that is meant to represent the
181 > repulsive overlap of the two positively charged cores.  
182  
183 < The {\it modified} embedded atom method (MEAM) adds angular terms to
184 < the electron density functions and an angular screening factor to the
185 < pairwise interaction between two
186 < atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
187 < MEAM has become widely used to simulate systems in which angular
188 < interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
189 < metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
190 < MEAM presents significant additional computational costs, however.
183 > % The {\it modified} embedded atom method (MEAM) adds angular terms to
184 > % the electron density functions and an angular screening factor to the
185 > % pairwise interaction between two
186 > % atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
187 > % MEAM has become widely used to simulate systems in which angular
188 > % interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
189 > % metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
190 > % MEAM presents significant additional computational costs, however.
191  
192 < The EAM, Finnis-Sinclair, MEAM, and the Quantum Sutton-Chen potentials
192 > The EAM, Finnis-Sinclair, and the Quantum Sutton-Chen (QSC) potentials
193   have all been widely used by the materials simulation community for
194   simulations of bulk and nanoparticle
195   properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
# Line 190 | Line 200 | parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin19
200   strengths and weaknesses.  One of the strengths common to all of the
201   methods is the relatively large library of metals for which these
202   potentials have been
203 < parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
203 > parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}  
204  
205 < \subsection{CO}
206 < Since one explanation for the strong surface CO repulsion on metals is
207 < the large linear quadrupole moment of carbon monoxide, the model
208 < chosen for this molecule exhibits this property in an efficient
209 < manner.  We used a model first proposed by Karplus and Straub to study
210 < the photodissociation of CO from myoglobin.\cite{Straub} The Straub and
211 < Karplus model is a rigid three site model which places a massless M
212 < site at the center of mass along the CO bond.  The geometry used along
213 < with the interaction parameters are reproduced in Table 1. The effective
205 > \subsection{Carbon Monoxide model}
206 > Previous explanations for the surface rearrangements center on
207 > the large linear quadrupole moment of carbon monoxide.  
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 which places a massless M
212 > site at the center of mass position along the CO bond.  The geometry used along
213 > with the interaction 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
217   mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
218   %CO Table
219   \begin{table}[H]
220 < \caption{Positions, $\sigma$, $\epsilon$ and charges for CO geometry and self-interactions\cite{Straub}. Distances are in \AA~, energies are in kcal/mol, and charges are in $e$.}
220 >  \caption{Positions, Lennard-Jones parameters ($\sigma$ and
221 >    $\epsilon$), and charges for the CO-CO
222 >    interactions borrowed from Ref.\bibpunct{}{}{,}{n}{}{,} \protect\cite{Straub}. Distances are in \AA, energies are
223 >    in kcal/mol, and charges are in atomic units.}
224   \centering
225   \begin{tabular}{| c | c | ccc |}
226   \hline
227 < \multicolumn{5}{|c|}{\textbf{Self-Interactions}}\\
227 > &  {\it z} & $\sigma$ & $\epsilon$ & q\\
228   \hline
229 < &  r & $\sigma$ & $\epsilon$ & q\\
229 > \textbf{C} & -0.6457 &  0.0262  & 3.83   &   -0.75 \\
230 > \textbf{O} &  0.4843 &   0.1591 &   3.12 &   -0.85 \\
231 > \textbf{M} & 0.0 & -  &  -  &    1.6 \\
232   \hline
218 \textbf{C} & 0.0 &  0.0262  & 3.83   &   -0.75 \\
219 \textbf{O} &  1.13 &   0.1591 &   3.12 &   -0.85 \\
220 \textbf{M} & 0.6457 & -  &  -  &    1.6 \\
221 \hline
233   \end{tabular}
234 + \label{tab:CO}
235   \end{table}
236  
237 < \subsection{Cross-Interactions}
237 > \subsection{Cross-Interactions between the metals and carbon monoxide}
238  
239 < One hurdle that must be overcome in classical molecular simulations
240 < is the proper parameterization of the potential interactions present
241 < in the system. Since the adsorption of CO onto a platinum surface has been
242 < the focus of many experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
243 < and theoretical studies \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
244 < there is a large amount of data in the literature to fit too. We started with parameters
245 < reported by Korzeniewski et al. \cite{Pons:1986} and then
246 < modified them to ensure that the Pt-CO interaction favored
247 < an atop binding position for the CO upon the Pt surface. This
248 < constraint led to the binding energies being on the higher side
249 < of reported values. Following the method laid out by Korzeniewski,
250 < the Pt-C interaction was fit to a strong Lennard-Jones 12-6
251 < interaction to mimic binding, while the Pt-O interaction
252 < was parameterized to a Morse potential with a large $r_o$
253 < to contribute a weak repulsion. The resultant potential-energy
254 < surface suitably recovers the calculated Pt-C bond length ( 1.6\AA)\cite{Beurden:2002ys} and affinity
255 < for the atop binding position.\cite{Deshlahra:2012, Hopster:1978}
239 > Since the adsorption of CO onto a Pt surface has been the focus
240 > of much experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
241 > and theoretical work
242 > \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
243 > there is a significant amount of data on adsorption energies for CO on
244 > clean metal surfaces. Parameters reported by Korzeniewski {\it et
245 >  al.}\cite{Pons:1986} were a starting point for our fits, which were
246 > modified to ensure that the Pt-CO interaction favored the atop binding
247 > position on Pt(111). These parameters are reproduced in Table~\ref{tab:co_parameters}
248 > This resulted in binding energies that are slightly higher
249 > than the experimentally-reported values as shown in Table~\ref{tab:co_energies}. Following Korzeniewski
250 > et al.,\cite{Pons:1986} the Pt-C interaction was fit to a deep
251 > Lennard-Jones interaction to mimic strong, but short-ranged partial
252 > binding between the Pt $d$ orbitals and the $\pi^*$ orbital on CO. The
253 > Pt-O interaction was parameterized to a Morse potential at a larger
254 > minimum distance, ($r_o$).  This was chosen so that the C would be preferred
255 > over O as the binder to the surface. In most cases, this parameterization contributes a weak
256 > repulsion which favors the atop site.  The resulting potential-energy
257 > surface suitably recovers the calculated Pt-C separation length
258 > (1.6~\AA)\cite{Beurden:2002ys} and affinity for the atop binding
259 > position.\cite{Deshlahra:2012, Hopster:1978}
260  
261   %where did you actually get the functionals for citation?
262   %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
263   %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
264 < The Au-C and Au-O interaction parameters were also fit to a Lennard-Jones
265 < and Morse potential respectively, to reproduce Au-CO binding energies.
266 < These energies were obtained from quantum calculations carried out using
267 < the PBE GGA exchange-correlation functionals\cite{Perdew_GGA} for gold, carbon, and oxygen
268 < constructed by Rappe, Rabe, Kaxiras, and Joannopoulos. \cite{RRKJ_PP}.
269 < All calculations were run using the {\sc Quantum ESPRESSO} package. \cite{QE-2009}  
270 < First, a four layer slab of gold comprised of 32 atoms displaying a (111) surface was
271 < converged using a 4X4X4 grid of Monkhorst-Pack \emph{k}-points.\cite{Monkhorst:1976}
272 < The kinetic energy of the wavefunctions were truncated at 20 Ry while the
273 < cutoff for the charge density and potential was set at 80 Ry. This relaxed
274 < gold slab was then used in numerous single point calculations  with CO at various heights
275 < to create a potential energy surface for the Au-CO interaction.
264 > The Au-C and Au-O cross-interactions were also fit using Lennard-Jones and
265 > Morse potentials, respectively, to reproduce Au-CO binding energies.
266 > The limited experimental data for CO adsorption on Au lead us to refine our fits against DFT.
267 > Adsorption energies were obtained from gas-surface DFT calculations with a
268 > periodic supercell plane-wave basis approach, as implemented in the
269 > {\sc Quantum ESPRESSO} package.\cite{QE-2009} Electron cores are
270 > described with the projector augmented-wave (PAW)
271 > method,\cite{PhysRevB.50.17953,PhysRevB.59.1758} with plane waves
272 > included to an energy cutoff of 20 Ry. Electronic energies are
273 > computed with the PBE implementation of the generalized gradient
274 > approximation (GGA) for gold, carbon, and oxygen that was constructed
275 > by Rappe, Rabe, Kaxiras, and Joannopoulos.\cite{Perdew_GGA,RRKJ_PP}
276 > In testing the Au-CO interaction, Au(111) supercells were constructed of four layers of 4
277 > Au x 2 Au surface planes and separated from vertical images by six
278 > layers of vacuum space. The surface atoms were all allowed to relax
279 > before CO was added to the system. Electronic relaxations were
280 > performed until the energy difference between subsequent steps
281 > was less than $10^{-8}$ Ry.   Nonspin-polarized supercell calculations
282 > were performed with a 4~x~4~x~4 Monkhorst-Pack {\bf k}-point sampling of the first Brillouin
283 > zone.\cite{Monkhorst:1976,PhysRevB.13.5188} The relaxed gold slab was
284 > then used in numerous single point calculations with CO at various
285 > heights (and angles relative to the surface) to allow fitting of the
286 > empirical force field.
287  
288   %Hint at future work
289 < The fit parameter sets employed in this work are shown in Table 2 and their
290 < reproduction of the binding energies are displayed in Table 3. Currently,
291 < charge transfer is not being treated in this system, however, that is a goal
292 < for future work as the effect has been seen to affect binding energies and
293 < binding site preferences. \cite{Deshlahra:2012}
289 > The parameters employed for the metal-CO cross-interactions in this work
290 > are shown in Table~\ref{co_parameters} and the binding energies on the
291 > (111) surfaces are displayed in Table~\ref{co_energies}.  Charge transfer
292 > and polarization are neglected in this model, although these effects are likely to
293 > affect binding energies and binding site preferences, and will be added in
294 > a future work.\cite{Deshlahra:2012,StreitzMintmire:1994}
295  
268
269
270
271 \subsection{Construction and Equilibration of 557 Metal interfaces}
272
273 Our model systems are composed of approximately 4000 metal atoms cut along the 557 plane so that they are periodic in the \it{x} and \it{y} directions exposing the 557 plane in the \it{z} direction. Runs at various temperatures ranging from 300~K to 1200~K were started with the intent of viewing relative stability of the surface when CO was not present in the system.  Owing to the different melting points (1337~K for Au and 2045~K for Pt), the bare crystal systems were initially run in the Canonical ensemble for at 800~K and 1000~K respectively for 100 ps. Various amounts of CO were placed in the vacuum portion which upon full adsorption to the surface corresponded to 5\%, 25\%, 33\%, and 50\% coverages. These systems were again allowed to reach thermal equilibrium before being run in the micro canonical ensemble. All of the systems examined were run for at least 40 ns. A subset that were undergoing interesting effects have been allowed to continue running with one system approaching 200 ns.em
274
275
276
277
278
279
280 %\subsection{System}
281 %Once equilibration was reached, the systems were exposed to various sub-monolayer coverage of CO: $0, \frac{1}{10}, \frac{1}{4}, \frac{1}{3},\frac{1}{2}$. The CO was started many \AA~above the surface with random velocity and rotational velocity vectors sampling from a Gaussian distribution centered on the temperature of the equilibrated metal block.  Full adsorption occurred over the period of approximately 10 ps for Pt, while the binding energy between Au and CO is smaller and led to an incomplete adsorption. The metal-metal interactions were treated using the Embedded Atom Method while the Pt-CO and Au-CO interactions were fit to experimental data and quantum calculations. The raised temperature helped shorten the length of the simulations by allowing the activation barrier of reconstruction to be more easily overcome. A few runs at lower temperatures showed the very beginnings of reconstructions, but their simulation lengths limited their usefulness.
282
283
296   %Table  of Parameters
297   %Pt Parameter Set 9
298   %Au Parameter Set 35
299   \begin{table}[H]
300 < \caption{Best fit parameters for metal-adsorbate interactions. Distances are in \AA~and energies are in kcal/mol}
300 >  \caption{Best fit parameters for metal-CO cross-interactions. Metal-C
301 >    interactions are modeled with Lennard-Jones potential, while the
302 >    metal-O interactions were fit to Morse
303 >    potentials.  Distances are given in \AA~and energies in kcal/mol. }
304   \centering
305   \begin{tabular}{| c | cc | c | ccc |}
306   \hline
307 < \multicolumn{7}{| c |}{\textbf{Cross-Interactions} }\\
307 > &  $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
308   \hline
294 &  $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ \\
295 \hline
309   \textbf{Pt-C} & 1.3 & 15  & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
310   \textbf{Au-C} & 1.9 & 6.5  & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
311  
312   \hline
313   \end{tabular}
314 + \label{tab:co_parameters}
315   \end{table}
316  
317   %Table of energies
318   \begin{table}[H]
319 < \caption{Adsorption energies in eV}
319 >  \caption{Adsorption energies for CO on M(111) at the atop site using the potentials
320 >    described in this work.  All values are in eV.}
321   \centering
322   \begin{tabular}{| c | cc |}
323 < \hline
324 < & Calc. & Exp. \\
325 < \hline
326 < \textbf{Pt-CO} & -1.9 & -1.4~\cite{Kelemen:1979}-- -1.9~\cite{Yeo} \\
327 < \textbf{Au-CO} & -0.39 & -0.44~\cite{TPD_Gold_CO} \\
328 < \hline
323 >  \hline
324 >  & Calculated & Experimental \\
325 >  \hline
326 >  \multirow{2}{*}{\textbf{Pt-CO}} & \multirow{2}{*}{-1.9} & -1.4 \bibpunct{}{}{,}{n}{}{,}
327 >  (Ref. \protect\cite{Kelemen:1979}) \\
328 > & &  -1.9 \bibpunct{}{}{,}{n}{}{,} (Ref. \protect\cite{Yeo}) \\ \hline
329 >  \textbf{Au-CO} & -0.39 & -0.40 \bibpunct{}{}{,}{n}{}{,}  (Ref. \protect\cite{TPD_Gold}) \\
330 >  \hline
331   \end{tabular}
332 + \label{tab:co_energies}
333   \end{table}
334  
335 + \subsection{Pt(557) and Au(557) metal interfaces}
336  
337 + Our model systems are composed of 3888 Pt atoms and 3384 Au atoms in a
338 + FCC crystal that have been cut along the (557) plane so that they are
339 + periodic in the {\it x} and {\it y} directions, and have been oriented
340 + to expose two aligned (557) cuts along the extended {\it
341 +  z}-axis.  Simulations of the bare metal interfaces at temperatures
342 + ranging from 300~K to 1200~K were performed to observe the relative
343 + stability of the surfaces without a CO overlayer.  
344  
345 <
346 <
345 > The different bulk (and surface) melting temperatures (1337~K for Au
346 > and 2045~K for Pt) suggest that any possible reconstruction may happen at
347 > different temperatures for the two metals.  The bare Au and Pt surfaces were
348 > initially run in the canonical (NVT) ensemble at 800~K and 1000~K
349 > respectively for 100 ps. These temperatures were chosen because the
350 > surfaces were relatively stable at these temperatures when no CO was
351 > present, but experienced additional instability upon addition of CO in the time
352 > frames we were examining. Each surface was exposed to a range of CO
353 > that was initially placed in the vacuum region.  Upon full adsorption,
354 > these amounts correspond to 0\%, 5\%, 25\%, 33\%, and 50\% surface
355 > coverage. Higher coverages were tried, but the CO-CO repulsion was preventing
356 > a higher amount of adsorption.  Because of the difference in binding energies, the Pt
357 > systems very rarely had CO that was not bound to the surface, while
358 > the Au surfaces often had a significant CO population in the gas
359 > phase.  These systems were allowed to reach thermal equilibrium (over
360 > 5 ns) before being run in the microcanonical (NVE) ensemble for
361 > data collection. All of the systems examined had at least 40 ns in the
362 > data collection stage, although simulation times for some of the
363 > systems exceeded 200ns.  All simulations were run using the open
364 > source molecular dynamics package, OpenMD.\cite{Ewald,OOPSE}
365  
366   % Just results, leave discussion for discussion section
367 + % structure
368 + %       Pt: step wandering, double layers, no triangular motifs
369 + %       Au: step wandering, no double layers
370 + % dynamics
371 + %       diffusion
372 + %       time scale, formation, breakage
373   \section{Results}
374 < \subsection{Diffusion}
375 < While an ideal metallic surface is unlikely to experience much surface diffusion, high-index surfaces have large numbers of low-coordinated atoms which have a much easier time overcoming the energetic barriers limiting diffusion, leading to easier surface reconstructions. Surface movement was divided between the parallel ($\parallel$) and perpendicular ($\perp$) directions relative to the step edge. We were then able to calculate diffusion constants as a function of CO coverage. As can be seen in Table 4, the presence and amount of CO directly affects the diffusion constants of surface platinum atoms. The presence of two 50\% coverage systems is to show how the diffusion process is affected by time. The majority of the systems were run for approximately 50 ns while the half monolayer system been running continuously. The lowered diffusion constant at longer run times will be examined in-depth in the discussion section.
374 > \subsection{Structural remodeling}
375 > Tao et al. showed experimentally that the Pt(557) surface
376 > undergoes two separate reconstructions upon CO
377 > adsorption.\cite{Tao:2010} The first involves a doubling of
378 > the step height and plateau length. Similar behavior has been
379 > seen to occur on numerous surfaces at varying conditions (Ni 977, Si 111, etc).
380 > \cite{Williams:1994,Williams:1991,Pearl} Of the two systems
381 > we examined, the Pt system showed a larger amount of
382 > reconstruction when compared to the Au system. The amount
383 > of reconstruction appears to be correlated to the amount of CO
384 > adsorbed upon the surface.  We believe this is related to the
385 > effect that adsorbate coverage has on edge breakup and surface
386 > diffusion of adatoms. While both systems displayed step-edge
387 > wandering, only the Pt surface underwent the doubling seen by
388 > Tao et al., within the time scales we were modeling. Specifically,
389 > only the 50~\% coverage Pt system was observed to have a
390 > step-edge undergo a complete doubling in the time scales we
391 > were able to monitor. This event encouraged us to allow that
392 > specific system to run for much longer periods during which two
393 > more double layers were created. The other systems, not displaying
394 > any large scale changes of interest, were all stopped after running
395 > for 40 ns in the microcanonical ensemble. Despite no observation
396 > of double layer formation, the other Pt systems tended to show
397 > more cumulative lateral movement of the step-edges when
398 > compared to the Au systems. The 50\% Pt system is highlighted
399 > in Figure \ref{fig:reconstruct} at various times along the simulation
400 > showing the evolution of the system.
401  
402 < %Table of Diffusion Constants
403 < %Add gold?M
404 < \begin{table}[H]
405 < \caption{Platinum diffusion constants parallel and perpendicular to the 557 step edge. As the coverage increases, the diffusion constants parallel and  perpendicular to the step edge both initially increase and then decrease slightly. There were two approaches of analysis. One looking at the surface atoms that had moved more than a prescribed amount over the run time and the other looking at all surface atoms. Units are \AA\textsuperscript{2}/ns}
406 < \centering
407 < \begin{tabular}{| c | ccc | ccc | c |}
333 < \hline
334 < \textbf{System Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & \textbf{Atoms} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & \textbf{Atoms} & \textbf{Time (ns)}\\
335 < \hline
336 < &\multicolumn{3}{c|}{\textbf{Mobile}}&\multicolumn{3}{c|}{\textbf{Surface Atoms}} & \\
337 < \hline
338 < 50\% & 3.74 & 0.89 & 497 & 2.05 & 0.49 & 912 & 116 \\
339 < 50\% & 5.81 & 1.59 & 365 & 2.41 & 0.68 & 912 & 46   \\
340 < 33\% & 6.73 & 2.47 & 332 & 2.51 & 0.93 & 912 & 46   \\
341 < 25\% & 5.38 & 2.04 & 361 & 2.18 & 0.84 & 912 & 46  \\
342 < 5\% & 5.54 & 0.63 & 230 & 1.44 & 0.19 & 912 & 46  \\
343 < 0\% & 3.53 & 0.61 & 282 & 1.11 & 0.22 & 912 & 56  \\
344 < \hline
345 < 50\%-r & 6.91 & 2.00 & 198 & 2.23 & 0.68  & 925 & 25\\
346 < 0\%-r  & 4.73 & 0.27 & 128 & 0.72 & 0.05 & 925 & 43\\
347 < \hline
348 < \end{tabular}
349 < \end{table}
402 > The second reconstruction on the Pt(557) surface observed by
403 > Tao involved the formation of triangular clusters that stretched
404 > across the plateau between two step-edges. Neither system, within
405 > our simulated time scales, experiences this reconstruction. A constructed
406 > system in which the triangular motifs were constructed on the surface
407 > will be explored in future work and is shown in the supporting information.
408  
409 + \subsection{Dynamics}
410 + While atomistic-like simulations of stepped surfaces have been
411 + performed before, they tend to be performed using Monte Carlo
412 + techniques\cite{Williams:1991,Williams:1994}. This allows them
413 + to efficiently sample the equilibrium thermodynamic landscape
414 + but at the expense of ignoring the dynamics of the system. Previous
415 + work by Pearl and Sibener\cite{Pearl}, using STM, has been able to
416 + visualize the coalescing of steps of Ni(977). The time scale of the image
417 + acquisition, $\sim$70 s/image provides an upper bounds for the time
418 + required for the doubling to actually occur. Statistical treatments of step-edges
419 + are adept at analyzing such systems. However, in a system where
420 + the number of steps is limited, examining the individual atoms that make
421 + up the steps can provide useful information as well.
422  
423  
424 + \subsubsection{Transport of surface metal atoms}
425 + %forcedSystems/stepSeparation
426 + The movement or wandering of a step-edge is a cooperative effect
427 + arising from the individual movements, primarily through surface
428 + diffusion, of the atoms making up the step. An ideal metal surface
429 + displaying a low index facet, (111) or (100) is unlikely to experience
430 + much surface diffusion because of the large energetic barrier that must
431 + be overcome to lift an atom out of the surface. The presence of step-edges
432 + on higher-index surfaces provide a source for mobile metal atoms.
433 + Breaking away from the step-edge on a clean surface still imposes an
434 + energetic penalty around $\sim$~40 kcal/mole, but is much less than lifting
435 + the same metal atom out from the surface,  \textgreater~60 kcal/mole, and
436 + the penalty lowers even further when CO is present in sufficient quantities
437 + on the surface. For certain tested distributions of CO, the penalty was lowered
438 + to $\sim$~20 kcal/mole. Once an adatom exists on the surface, its barrier for
439 + diffusion is negligible ( \textless~4 kcal/mole) and is well able to explore the
440 + terrace before potentially rejoining its original step-edge or becoming a part
441 + of a different edge. Atoms traversing separate terraces is a more difficult
442 + process, but can be overcome through a joining and lifting stage which is
443 + examined in the discussion section. By tracking the mobility of individual
444 + metal atoms on the Pt and Au surfaces we were able to determine the relative
445 + diffusion rates and how varying coverages of CO affected the rates. Close
446 + observation of the mobile metal atoms showed that they were typically in
447 + equilibrium with the step-edges, constantly breaking apart and rejoining.
448 + At times their motion was concerted and two or more adatoms would be
449 + observed moving together across the surfaces. The primary challenge in
450 + quantifying the overall surface mobility was in defining ``mobile" vs. ``static" atoms.
451 +
452 + A particle was considered mobile once it had traveled more than 2~\AA~
453 + between saved configurations of the system (10-100 ps). An atom that was
454 + truly mobile would typically travel much greater than this, but the 2~\AA~ cutoff
455 + was to prevent the in-place vibrational movement of non-surface atoms from
456 + being included in the analysis. Diffusion on  a surface is strongly affected by
457 + local structures and in this work the presence of single and double layer
458 + step-edges causes the diffusion parallel to the step-edges to be different
459 + from the diffusion perpendicular to these edges. This led us to compute
460 + those diffusions separately as seen in Figure \ref{fig:diff}.
461 +
462 + \subsubsection{Double layer formation}
463 + The increased amounts of diffusion on Pt at the higher CO coverages appears
464 + to play a primary role in the formation of double layers, although this conclusion
465 + does not explain the 33\% coverage Pt system. On the 50\% system, three
466 + separate layers were formed over the extended run time of this system. As
467 + mentioned earlier, previous experimental work has given some insight into the
468 + upper bounds of the time required for enough atoms to move around to allow two
469 + steps to coalesce\cite{Williams:1991,Pearl}. As seen in Figure \ref{fig:reconstruct},
470 + the first appearance of a double layer, a nodal site, appears at 19 ns into the
471 + simulation. Within 12 ns, nearly half of the step has formed the double layer and
472 + by 86 ns, a smooth complete layer has formed. The double layer is ``complete" by
473 + 37 ns but is a bit rough. From the appearance of the first node to the initial doubling
474 + of the layers ignoring their roughness took $\sim$~20 ns. Another ~40 ns was
475 + necessary for the layer to completely straighten. The other two layers in this
476 + simulation form over a period of 22 ns and 42 ns respectively. Comparing this to
477 + the upper bounds of the image scan, it is likely that aspects of this reconstruction
478 + occur very quickly.
479 +
480 + %Evolution of surface
481 + \begin{figure}[H]
482 + \includegraphics[width=\linewidth]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
483 + \caption{The Pt(557) / 50\% CO system at a sequence of times after
484 +  initial exposure to the CO: (a) 258 ps, (b) 19 ns, (c) 31.2 ns, and
485 +  (d) 86.1 ns. Disruption of the (557) step-edges occurs quickly.  The
486 +  doubling of the layers appears only after two adjacent step-edges
487 +  touch.  The circled spot in (b) nucleated the growth of the double
488 +  step observed in the later configurations.}
489 +  \label{fig:reconstruct}
490 + \end{figure}
491 +
492 + \begin{figure}[H]
493 + \includegraphics[width=\linewidth]{DiffusionComparison_errorXY_remade.pdf}
494 + \caption{Diffusion constants for mobile surface atoms along directions
495 +  parallel ($\mathbf{D}_{\parallel}$) and perpendicular
496 +  ($\mathbf{D}_{\perp}$) to the (557) step-edges as a function of CO
497 +  surface coverage.  Diffusion parallel to the step-edge is higher
498 +  than that perpendicular to the edge because of the lower energy
499 +  barrier associated with traversing along the edge as compared to
500 +  completely breaking away. Additionally, the observed
501 +  maximum and subsequent decrease for the Pt system suggests that the
502 +  CO self-interactions are playing a significant role with regards to
503 +  movement of the Pt atoms around and across the surface. }
504 + \label{fig:diff}
505 + \end{figure}
506 +
507 +
508 +
509 +
510   %Discussion
511   \section{Discussion}
512 < Comparing the results from simulation to those reported previously by Tao et al. the similarities in the platinum and CO system are quite strong. As shown in figure 1, the simulated platinum system under a CO atmosphere will restructure slightly by doubling the terrace heights. The restructuring appears to occur slowly, one to two platinum atoms at a time. Looking at individual snapshots, these adatoms tend to either rise on top of the plateau or break away from the step edge and then diffuse perpendicularly to the step direction until reaching another step edge. This combination of growth and decay of the step edges appears to be in somewhat of a state of dynamic equilibrium. However, once two previously separated edges meet as shown in figure 1.B, this point tends to act as a focus or growth point for the rest of the edge to meet up, akin to that of a zipper. From the handful of cases where a double layer was formed during the simulation. Measuring from the initial appearance of a growth point, the double layer tends to be fully formed within $\sim$~35 ns.
512 > In this paper we have shown that we were able to accurately model the initial reconstruction of the
513 > Pt(557) surface upon CO adsorption as shown by Tao et al. \cite{Tao:2010}. More importantly, we
514 > were able to observe the dynamic processes necessary for this reconstruction.
515  
516 + \subsection{Mechanism for restructuring}
517 + Comparing the results from simulation to those reported previously by
518 + Tao et al.\cite{Tao:2010} the similarities in the Pt-CO system are quite
519 + strong. As shown in Figure \ref{fig:reconstruct}, the simulated Pt
520 + system under a CO atmosphere will restructure by doubling the terrace
521 + heights. The restructuring occurs slowly, one to two Pt atoms at a time.
522 + Looking at individual configurations of the system, the adatoms either
523 + break away from the step-edge and stay on the lower terrace or they lift
524 + up onto the higher terrace. Once ``free'' they will diffuse on the terrace
525 + until reaching another step-edge or coming back to their original edge.  
526 + This combination of growth and decay of the step-edges is in a state of
527 + dynamic equilibrium. However, once two previously separated edges
528 + meet as shown in Figure 1.B, this meeting point tends to act as a focus
529 + or growth point for the rest of the edge to meet up, akin to that of a zipper.
530 + From the handful of cases where a double layer was formed during the
531 + simulation, measuring from the initial appearance of a growth point, the
532 + double layer tends to be fully formed within $\sim$~35 ns.
533 +
534 + A number of possible mechanisms exist to explain the role of adsorbed
535 + CO in restructuring the Pt surface. Quadrupolar repulsion between adjacent
536 + CO molecules adsorbed on the surface is one likely possibility.  However,
537 + the quadrupole-quadrupole interaction is short-ranged and is attractive for
538 + some orientations.  If the CO molecules are ``locked'' in a specific orientation
539 + relative to each other, through atop adsorption perhaps, this explanation
540 + gains some weight.  The energetic repulsion between two CO located a
541 + distance of 2.77~\AA~apart (nearest-neighbor distance of Pt) with both in
542 + a  vertical orientation is 8.62 kcal/mole. Moving the CO apart to the second
543 + nearest-neighbor distance of 4.8~\AA~or 5.54~\AA~drops the repulsion to
544 + nearly 0 kcal/mole. Allowing the CO's to leave a purely vertical orientation
545 + also quickly drops the repulsion, a minimum is reached at $\sim$24 degrees
546 + of 6.2 kcal/mole. As mentioned above, the energy barrier for surface diffusion
547 + of a Pt adatom is only 4 kcal/mole. So this repulsion between CO can help
548 + increase the surface diffusion. However, the residence time of CO was
549 + examined and while the majority of the CO is on or near the surface throughout
550 + the run, it is extremely mobile. This mobility suggests that the CO are more
551 + likely to shift their positions without necessarily dragging the Pt along with them.
552 +
553 + Another possible and more likely mechanism for the restructuring is in the
554 + destabilization of strong Pt-Pt interactions by CO adsorbed on surface
555 + Pt atoms.  This would then have the effect of increasing surface mobility
556 + of these atoms.  To test this hypothesis, numerous configurations of
557 + CO in varying quantities were arranged on the higher and lower plateaus
558 + around a step on a otherwise clean Pt(557) surface. One representative
559 + configuration is displayed in Figure \ref{fig:lambda}. Single or concerted movement
560 + of Pt atoms was then examined to determine possible barriers. Because
561 + the movement was forced along a pre-defined reaction coordinate that may differ
562 + from the true minimum of this path, only the beginning and ending energies
563 + are displayed in Table \ref{tab:energies}. These values suggest that the presence of CO at suitable
564 + locations can lead to lowered barriers for Pt breaking apart from the step-edge.
565 + Additionally, as highlighted in Figure \ref{fig:lambda}, the presence of CO makes the
566 + burrowing and lifting of adatoms favorable, whereas without CO, the process is neutral
567 + in terms of energetics.
568 +
569 + %lambda progression of Pt -> shoving its way into the step
570 + \begin{figure}[H]
571 + \includegraphics[width=\linewidth]{lambdaProgression_atopCO.png}
572 + \caption{A model system of the Pt(557) surface was used as the framework
573 + for exploring energy barriers along a reaction coordinate. Various numbers,
574 + placements, and rotations of CO were examined as they affect Pt movement.
575 + The coordinate displayed in this Figure was a representative run. As shown
576 + in Table \ref{tab:rxcoord}, relative to the energy of the system at 0\%, there
577 + is a slight decrease upon insertion of the Pt atom into the step-edge along
578 + with the resultant lifting of the other Pt atom when CO is present at certain positions.}
579 + \label{fig:lambda}
580 + \end{figure}
581 +
582 +
583 +
584   \subsection{Diffusion}
585 < As shown in the results section, the diffusion parallel to the step edge tends to be much faster than that perpendicular to the step edge. Additionally, the coverage of CO appears to play a slight role in relative rates of diffusion, as shown in Table 4. Thus, the bottleneck of the double layer formation appears to be the initial formation of this growth point, which seems to be somewhat of a stochastic event. Once it appears, parallel diffusion, along the now slightly angled step edge, will allow for a faster formation of the double layer than if the entire process were dependent on only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the more likely a growth point is to be formed. One driving force behind this reconstruction appears to be the lowering of surface energy that occurs by doubling the terrace widths. (I'm not really proving this... I have the surface flatness to show it, but surface energy?)
585 > As shown in the results section, the diffusion parallel to the step-edge tends to be
586 > much larger than that perpendicular to the step-edge, likely because of the dynamic
587 > equilibrium that is established between the step-edge and adatom interface. The coverage
588 > of CO also appears to play a slight role in relative rates of diffusion, as shown in Figure \ref{fig:diff}.
589 > The
590 > Thus, the bottleneck of the double layer formation appears to be the initial formation
591 > of this growth point, which seems to be somewhat of a stochastic event. Once it
592 > appears, parallel diffusion, along the now slightly angled step-edge, will allow for
593 > a faster formation of the double layer than if the entire process were dependent on
594 > only perpendicular diffusion across the plateaus. Thus, the larger $D_{\perp}$, the
595 > more likely a growth point is to be formed.
596   \\
597 < \\
598 < %Evolution of surface
597 >
598 >
599 > %breaking of the double layer upon removal of CO
600   \begin{figure}[H]
601 < \includegraphics[scale=0.5]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
602 < \caption{Four snapshots at various times a) 258 ps b) 19 ns c) 31.2 ns d) 86.1 ns. Slight disruption of the surface occurs fairly quickly. However, the doubling of the layers seems to be very dependent on the initial linking of two separate step edges. The focal point in b, appears to be a growth spot for the rest of the double layer.}
601 > \includegraphics[width=\linewidth]{doubleLayerBreaking_greenBlue_whiteLetters.png}
602 > %:
603 > \caption{(A)  0 ps, (B) 100 ps, (C) 1 ns, after the removal of CO. The presence of the CO
604 > helped maintain the stability of the double layer and upon removal the two layers break
605 > and begin separating. The separation is not a simple pulling apart however, rather
606 > there is a mixing of the lower and upper atoms at the edge.}
607 > \label{fig:breaking}
608   \end{figure}
609  
610  
611  
612  
613   %Peaks!
614 < \includegraphics[scale=0.25]{doublePeaks_noCO.png}
614 > \begin{figure}[H]
615 > \includegraphics[width=\linewidth]{doublePeaks_noCO.png}
616 > \caption{At the initial formation of this double layer  ( $\sim$ 37 ns) there is a degree
617 > of roughness inherent to the edge. The next $\sim$ 40 ns show the edge with
618 > aspects of waviness and by 80 ns the double layer is completely formed and smooth. }
619 > \label{fig:peaks}
620 > \end{figure}
621 >
622 >
623 > %Don't think I need this
624 > %clean surface...
625 > %\begin{figure}[H]
626 > %\includegraphics[width=\linewidth]{557_300K_cleanPDF.pdf}
627 > %\caption{}
628 >
629 > %\end{figure}
630 > %\label{fig:clean}
631 >
632 >
633   \section{Conclusion}
634 + In this work we have shown the reconstruction of the Pt(557) crystalline surface upon adsorption of CO in < $\mu s$. Only the highest coverage Pt system showed this initial reconstruction similar to that seen previously. The strong interaction between Pt and CO and the limited interaction between Au and CO helps explain the differences between the two systems.
635  
636 + %Things I am not ready to remove yet
637  
638 + %Table of Diffusion Constants
639 + %Add gold?M
640 + % \begin{table}[H]
641 + %   \caption{}
642 + %   \centering
643 + % \begin{tabular}{| c | cc | cc | }
644 + %   \hline
645 + %   &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} \\
646 + %   \hline
647 + %   \textbf{Surface Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$  \\
648 + %   \hline
649 + %   50\% & 4.32(2) & 1.185(8)  & 1.72(2) & 0.455(6) \\
650 + %   33\% & 5.18(3)  & 1.999(5)  & 1.95(2) & 0.337(4)   \\
651 + %   25\% & 5.01(2)  & 1.574(4)  & 1.26(3) & 0.377(6) \\
652 + %   5\%   & 3.61(2)  & 0.355(2)  & 1.84(3)  & 0.169(4)  \\
653 + %   0\%   & 3.27(2)  & 0.147(4)  & 1.50(2)  & 0.194(2)   \\
654 + %   \hline
655 + % \end{tabular}
656 + % \end{table}
657 +
658   \section{Acknowledgments}
659   Support for this project was provided by the National Science
660   Foundation under grant CHE-0848243 and by the Center for Sustainable

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines