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

Comparing trunk/PtPd/PtPd/tex/draft.tex (file contents):
Revision 4295 by gezelter, Tue Feb 10 20:46:59 2015 UTC vs.
Revision 4296 by jmichalk, Thu Feb 12 21:22:42 2015 UTC

# Line 14 | Line 14
14   \usepackage[version=3]{mhchem}  % this is a great package for formatting chemical reactions
15   \usepackage{url}
16  
17 < \title{CO-induced island formation on Pt@Pd(557) subsurface alloys: A
17 > \title{\ce{CO}-induced island formation on \ce{Pt}/\ce{Pd}(557) subsurface alloys: A
18    molecular dynamics study}
19    
20  
# Line 51 | Line 51 | currently under intense investigation\cite{Kim:2013jt}
51  
52   \section{Introduction}
53   Bimetallic alloys, subsurface alloys, and core-shell nanostructures are
54 < currently under intense investigation\cite{Kim:2013jt} because of their large accesible
54 > currently under intense investigation\cite{Kim:2013jt, Gao:2009oj, Gao:2009wo} because of their large accesible
55   design space for various catalytic processes.  The presence of two (or more)
56   components in these structures allows for a high degree of tuning of the
57 < specific characteristics, whether that be catalytic activity\cite{a}, thermal
57 > specific characteristics, whether that be catalytic activity\cite{Kim:2013jt}, thermal
58   stability\cite{a}, or resistance to deactivation\cite{a}. As seen in many
59   experimental\cite{Ertl:1989} and theoretical studies\cite{a}, the potential
60   energy landscape of the surface is often modified by the presence of adsorbates
61   leading to large-scale reconstructions of the surface. This reconstruction
62   could be a simple refaceting or a more complicated process that leads to the
63   formation of significant nano-features. Both situations will provide additional
64 < or different active sites, chaning the activity and selectivity of the
64 > or different active sites, changing the activity and selectivity of the
65   catalyst.
66  
67   Tuning catalyst work...
68  
69   As Tao et al.\cite{Tao:2010} showed and we modeled\cite{Michalka:2013}, the
70 < steps of the Pt(557) system when exposed to a CO atmosphere undergo doubling.
71 < To the extent of our knowledge there has been no similar work done with CO on
72 < Pd(557) and this work is an attempt to explore that system as well as what
73 < happens to a bimetallic system containing both Pt and Pd.
70 > steps of the Pt(557) system when exposed to a \ce{CO} atmosphere undergo doubling.
71 > To the extent of our knowledge there has been no similar work done with \ce{CO} on
72 > \ce{Pd}(557) and this work is an attempt to explore that system as well as what
73 > happens to a bimetallic system containing both \ce{Pt} and \ce{Pd}.
74  
75 < This work is an investigation into the effect of CO adsorption on surface
76 < restructuring of a Pd(557) and Pt@Pd(557) shell surface using molecular
75 > This work is an investigation into the effect of \ce{CO} adsorption on surface
76 > restructuring of a \ce{Pd}(557) and \ce{Pt}/\ce{Pd} (557) subsurface alloy using molecular
77   simulation. Since the mechanism and dynamics of the restructuring are of
78   particular interest, classical force fields which balance computational
79   efficiency against chemical accuracy were employed. A more complete
# Line 137 | Line 137 | describe the Pt and Pd electron densities, embedding f
137   neighbor interactions during parameterization.\cite{a}
138  
139   In this work, we have employed the embedded atom method (EAM) to
140 < describe the Pt and Pd electron densities, embedding functionals, and
140 > describe the \ce{Pt} and \ce{Pd} electron densities, embedding functionals, and
141   pair potentials,\cite{EAM} utilizing the Johnson mixing rules for the
142 < Pt-Pd cross-interactions.\cite{johnson89}
142 > \ce{Pt\bond{-}Pd} cross-interactions.\cite{johnson89}
143  
144   The carbon monoxide (\ce{CO}) self-interactions were modeled using a
145   rigid three-site model developed by Straub and Karplus for studying
146 < photodissociation of CO from myoglobin.\cite{Straub} This model
146 > photodissociation of \ce{CO} from myoglobin.\cite{Straub} This model
147   accurately captures the large linear quadrupole (and weak dipole) of
148 < the CO molecule.
148 > the \ce{CO} molecule.
149  
150 < The Pt-CO interactions have been modified from previous fits to
150 > The \ce{Pt\bond{-}CO} interactions have been modified from previous fits to
151   account for recently-published DFT
152   data.\cite{Michalka:2013,Deshlahra:2012} This modification yields a
153 < slightly weaker Pt-CO binding energy.
153 > slightly weaker \ce{Pt\bond{-}CO} binding energy.
154  
155 < The Pd-CO interaction potential was parameterized as part of this
156 < work, and uses similar functional forms to the Pt-CO
155 > The \ce{Pd\bond{-}CO} interaction potential was parameterized as part of this
156 > work, and uses similar functional forms to the \ce{Pt\bond{-}CO}
157   model.\cite{Michalka:2013} Our starting point is a model introduced by
158   Korzeniewski \textit{et al.}\cite{Pons:1986} The parameters were
159   modified to reflect binding energies and binding site preferences on
160 < the M(111) surfaces.  One key difference from the potential in
161 < Ref. \citenum{Michalka:2013} is that the M-O bond is modeled using a
160 > the \ce{M} (111) surfaces.  One key difference from the potential in
161 > Ref. \citenum{Michalka:2013} is that the \ce{M\bond{-}O} bond is modeled using a
162   purely repulsive Morse potential, $D e^{-2\gamma(r-r_e)}$.  The
163 < functional forms and the broad repulsive M-O contribution are flexible
164 < enough to reproduce the atop preference for Pt-CO as well as the
165 < bridge/hollow - preference for Pd-CO.  Parameters for the potentials
163 > functional forms and the broad repulsive \ce{M\bond{-}O} contribution are flexible
164 > enough to reproduce the atop preference for \ce{Pt\bond{-}CO} as well as the
165 > bridge/hollow - preference for \ce{Pd\bond{-}CO}.  Parameters for the potentials
166   are given in Table~\ref{tab:CO_parameters} and the calculated binding
167   energies at various binding sites are shown in
168   Table~\ref{tab:CO_energies}.
169   \begin{table}
170 < \caption{Parameters for the metal-CO cross-interactions. Metal-Carbon
170 > \caption{Parameters for the metal-\ce{CO} cross-interactions. Metal-Carbon
171    interactions are modeled with Lennard-Jones potentials, while the
172    metal-Oxygen interactions are fit using repulsive Morse potentials.
173    Distances are given in \AA~and energies in
# Line 177 | Line 177 | Table~\ref{tab:CO_energies}.
177   \hline
178   &  $\sigma$ & $\epsilon$ & & $r_e$ & $D$ & $\gamma$ (\AA$^{-1}$) \\
179   \hline
180 < \textbf{Pt-C} & 1.41 & 45  & \textbf{Pt-O} & 4.4  & 0.05 & 1.8 \\
181 < \textbf{Pd-C} & 1.6 &  40  & \textbf{Pd-O} & 4.95 & 0.05 & 1.45\\
180 > \textbf{\ce{Pt\bond{-}C}} & 1.41 & 45  & \textbf{\ce{Pt\bond{-}O}} & 4.4  & 0.05 & 1.8 \\
181 > \textbf{\ce{Pd\bond{-}C}} & 1.6 &  40  & \textbf{\ce{Pd\bond{-}O}} & 4.95 & 0.05 & 1.45\\
182   \hline
183   \end{tabular}
184   \end{table}
185  
186   %Table of energies
187   \begin{table}
188 <  \caption{Adsorption energies for a CO molecule at the three special sites
189 <    on M(111) using the potentials described in table
188 >  \caption{Adsorption energies for a \ce{CO} molecule at the three special sites
189 >    on \ce{M} (111) using the potentials described in table
190      \ref{tab:CO_parameters}.   These values are compared with DFT
191 <    calculations of XXX along with relevant experimental desorption
192 <    data. Reference \citenum{Deshlahra:2012} values are reported at $\frac{1}{4}$ ML. All values are in eV.}
191 >    calculations of XXX along with experimental desorption
192 >    data when available. Reference \citenum{Deshlahra:2012} values are reported at $\frac{1}{4}$ ML. All values are in eV.}
193   \centering
194   \begin{tabular}{| cc | ccc |}
195    \hline
196    & Site & This Model & DFT & Experimental \\
197    \hline
198 <  \textbf{Pt-CO} & atop   & -1.47 & -1.48\cite{Deshlahra:2012} & -1.39\cite{Kelemen:1979}, -1.43\cite{Ertl:1977}, -1.90\cite{Yeo:1997} \\
198 >  \textbf{\ce{Pt\bond{-}CO}} & atop   & -1.47 & -1.48\cite{Deshlahra:2012} & -1.39\cite{Kelemen:1979}, -1.43\cite{Ertl:1977}, -1.90\cite{Yeo:1997} \\
199                   & bridge & -1.13 & -1.47\cite{Deshlahra:2012} &  \\
200                   & hollow & -1.02 & -1.45\cite{Deshlahra:2012} &  \\
201 <  \textbf{Pd-CO} & atop   & -1.54 & -1.44\cite{Honkala:2001sf} &  \\
201 >  \textbf{\ce{Pd\bond{-}CO}} & atop   & -1.54 & -1.44\cite{Honkala:2001sf} &  \\
202                   & bridge & -1.65 & -1.83\cite{Honkala:2001sf} &  \\
203                   & hollow & -1.60 & -1.99\cite{Honkala:2001sf} & -1.47\cite{Ertl:1970}, -1.54\cite{Guo:1989} \\
204    \hline
# Line 206 | Line 206 | This Pd-CO model does not have a strong preference for
206   \label{tab:CO_energies}
207   \end{table}
208  
209 < This Pd-CO model does not have a strong preference for either the
209 > This \ce{Pd\bond{-}CO} model does not have a strong preference for either the
210   bridge or hollow binding sites, so it may overestimate the bridge-site
211   binding at low coverages, but at higher coverages, the situation is
212   somewhat less clear.\cite{Wong:1991ta} Studies using low-energy
213 < elecron diffraction (LEED) and C-O stretching frequencies of \ce{CO}
213 > elecron diffraction (LEED) and \ce{C\bond{-}O} stretching frequencies of \ce{CO}
214   bound to \ce{Pd}(111) suggest that the 3-fold hollow sites are
215   preferred at low
216   coverages,\cite{Bradshaw:1978uf,Conrad:1978fx,Ohtani:1987zh} where it
# Line 221 | Line 221 | of CO on Pd(111) appears to be a $c(4\times2)$ ordered
221   reported to lie between 1.3 and 1.54 eV.
222  
223   At higher \ce{CO} coverages (e.g. $> 0.5$ ML), the preferred binding
224 < of CO on Pd(111) appears to be a $c(4\times2)$ ordered structure with
225 < the CO bound to the bridge sites.\cite{Bradshaw:1978uf}
224 > of \ce{CO} on \ce{Pd}(111) appears to be a $c(4\times2)$ ordered structure with
225 > the \ce{CO} bound to the bridge sites.\cite{Bradshaw:1978uf}
226  
227   Theoretical work by Honkala \textit{et al.}\cite{Honkala:2001sf} using
228   DFT with the generalized gradient approximation (GGA) to describe
229 < electron exchange correlation and pseudopotentials for the Pd atoms
229 > electron exchange correlation and pseudopotentials for the \ce{Pd} atoms
230   also reported the fcc site as the most favorable binding position with
231   a binding energy of 2.00 eV compared to the bridge site binding
232   energy of 1.83 eV at $1/3$ monolayer.
# Line 242 | Line 242 | captured by the Pd-CO fit, it does represent a signifi
242   energy difference is 0.06 eV (-1.85 hollow, -1.79 bridge).
243  
244   Although the weak preference for hollow vs. bridge sites is not
245 < captured by the Pd-CO fit, it does represent a significant change from
246 < the atop preference of the Pt-CO model.  The dynamics of the metal
247 < bound to the CO is significantly altered as a result of this
245 > captured by the \ce{Pd\bond{-}CO} fit, it does represent a significant change from
246 > the atop preference of the \ce{Pt\bond{-}CO} model.  The dynamics of the metal
247 > bound to the \ce{CO} is significantly altered as a result of this
248   difference.
249  
250   \subsection{557 interfaces and subsurface alloys}
251 < The Pd(557) model is an orthorhombic periodic box with dimensions of
251 > The \ce{Pd}(557) model is an orthorhombic periodic box with dimensions of
252   $55.09 \times 49.48 \times 120$~\AA~ while the subsurface alloys
253   (Pt(557) surface layers, with Pd bulk) have dimensions of $54.875
254 < \times 49.235 \times 120$~\AA.  The Pd system consists of 9 layers of
255 < Pd while our subsurface alloys consist of 7 layers of Pd sandwiched
256 < between 2 layers of Pt.  Both the pure Pd slab and the subsurface
257 < alloy systems are $\sim$22~\AA~ thick. The lattice constants for Pd
258 < and Pt, 3.89 and 3.92~\AA, respectively, provide minimal strain energy
254 > \times 49.235 \times 120$~\AA.  The \ce{Pd} system consists of 9 layers of
255 > \ce{Pd} while our subsurface alloys consist of 7 layers of \ce{Pd} sandwiched
256 > between 2 layers of \ce{Pt}.  Both the pure \ce{Pd} slab and the subsurface
257 > alloy systems are $\sim$22~\AA~ thick. The lattice constants for \ce{Pd}
258 > and \ce{Pt}, 3.89 and 3.92~\AA, respectively, provide minimal strain energy
259   in the alloy, and the relaxed geometries of the two interfaces are
260   therefore quite similar.
261  
# Line 266 | Line 266 | of the 557 surface without a CO overlayer.  The bare s
266  
267   Simulations of the metal without any adsorbate present were performed
268   at temperatures ranging from 300 to 900~K to establish the stability
269 < of the 557 surface without a CO overlayer.  The bare systems were run
269 > of the 557 surface without a \ce{CO} overlayer.  The bare systems were run
270   in the canonical (NVT) ensemble at 850~K for 200 ps and the
271   microcanonical (NVE) ensemble for 1 ns, and displayed no changes in
272   the 557 structure during this period.
273  
274 < Ten systems were constructed, corresponding to five CO-coverage levels
275 < for each metallic system.  The number of CO molecules (0, 48, 240,
274 > Ten systems were constructed, corresponding to five \ce{CO}-coverage levels
275 > for each metallic system.  The number of \ce{CO} molecules (0, 48, 240,
276   320, and 480) yield surface coverages of 0, 0.05, 0.25, 0.33, and 0.5
277 < monolayers (ML) assuming that every CO adsorbs on the surface.
277 > monolayers (ML) assuming that every \ce{CO} adsorbs on the surface.
278  
279   Simulation boxes of the same sizes as the metallic systems were
280 < constructed with appropriate densities of CO and equilibrated to
281 < 850~K. The gas-phase CO and surface simulation boxes were then
282 < combined, using a 5~\AA~ cutoff between metallic atoms and CO to
283 < prevent overlap. The remaining CO population was further reduced to
280 > constructed with appropriate densities of \ce{CO} and equilibrated to
281 > 850~K. The gas-phase \ce{CO} and surface simulation boxes were then
282 > combined, using a 5~\AA~ cutoff between metallic atoms and \ce{CO} to
283 > prevent overlap. The remaining \ce{CO} population was further reduced to
284   match the required number for the correct surface coverage.
285   Velocities were resampled from a Boltzmann distribution, and any net
286   linear momentum was subtracted from the entire system.  The combined
287   systems were run for 1 ns in the NVT ensemble, before being run in the
288   NVE ensemble for data collection.
289  
290 < All of the Pd systems were run in the microcanonical ensemble for a
291 < minimum of 40 ns to collect statistics. The Pt/Pd subsurface alloy
290 > All of the \ce{Pd} systems were run in the microcanonical ensemble for a
291 > minimum of 40 ns to collect statistics. The \ce{Pt/Pd} subsurface alloy
292   systems, which were observed to undergo significant restructuring,
293   were each run for a total simulation time of 110 ns.  All simulations
294   were carried out with the open source molecular dynamics package,
295   OpenMD.\cite{openmd,OOPSE}
296  
297   \section{Results}
298 < In our earlier work on Pt(557) we observed CO-induced restructuring
299 < into relatively clean double-layer structures. For the pure Pd(557)
298 > In our earlier work on \ce{Pt}(557) we observed \ce{CO}-induced restructuring
299 > into relatively clean double-layer structures. For the pure \ce{Pd}(557)
300   studied here, the 557 facet retains the plateaus and steps with only
301   minimal adatom movement, and with almost no surface reconstruction.
302 < Higher CO coverages appear to have minimal effect on the pure Pd(557)
302 > Higher \ce{CO} coverages appear to have minimal effect on the pure \ce{Pd}(557)
303   systems.
304  
305 < However, the Pt/Pd subsurface alloy exhibits a CO-induced speedup of
305 > However, the \ce{Pt/Pd} subsurface alloy exhibits a \ce{CO}-induced speedup of
306   the diffusion of surface metal atoms, as well as a large-scale
307 < restructuring of the well-ordered surface into Pt-rich islands, and
307 > restructuring of the well-ordered surface into \ce{Pt}-rich islands, and
308   will therefore be the focus of most of our analysis.
309  
310   \begin{figure}
311    \includegraphics[width=\linewidth]{../figures/SystemFigures/systems_ochre2.png}
312    \caption{Snapshots of the some of the simulated systems. Panel A is the
313 <    pure Pd(557) $\sim$40 ns after being dosed with  $\frac{1}{3}$
314 <    monolayer of CO.  Panels B-D are the subsurface alloy 80 ns after
313 >    pure \ce{Pd}(557) $\sim$40 ns after being dosed with  $\frac{1}{3}$
314 >    monolayer of {CO}.  Panels B-D are the subsurface alloy 80 ns after
315      being dosed with 0, $\frac{1}{3}$, and
316 <    $\frac{1}{2}$ ML of CO, respectively.  Pt atoms are shown in gray,
317 <    Pd in orange, while the CO molecules are shown in black / red. }
316 >    $\frac{1}{2}$ ML of \ce{CO}, respectively.  \ce{Pt} atoms are shown in gray,
317 >    \ce{Pd} in orange, while the \ce{CO} molecules are shown in black / red. }
318   \label{fig:systems}
319   \end{figure}
320  
321   Figure \ref{fig:systems} shows representative configurations of the
322 < various systems after significant exposure to the CO. We see that the
322 > various systems after significant exposure to the \ce{CO}. We see that the
323   Pd system highlighted in panel A has undergone no surface
324   restructuring. The other three panels highlight the effect of varying
325 < CO concentrations on the subsurface alloys, which do exhibit
325 > \ce{CO} concentrations on the subsurface alloys, which do exhibit
326   structural reorganization.
327  
328   \subsection{Diffusion of Surface Metal Atoms in the Subsurface Alloy}
# Line 337 | Line 337 | increasing CO coverage. To estimate the surface diffus
337  
338   However, there is significant movement of surface Pt in the subsurface
339   alloys, and the mobility of the surface Pt layer increases with
340 < increasing CO coverage. To estimate the surface diffusion, we define a
340 > increasing \ce{CO} coverage. To estimate the surface diffusion, we define a
341   ``mobile'' atom as one which moves at least 2~\AA~ in any 10 ps window
342   during the simulation. Once an atom has been labeled as mobile, we
343   analyze the entire simulation to find the planar ($xy$) diffusion
344   constant for the mobile atoms of a particular type. The calculated
345   diffusion constants of mobile Pt atoms from the subsurface alloys are
346   shown in Table \ref{tab:diffusion}. The absolute number of mobile Pt
347 < atoms ($\sim 600$) was similar between all systems, independent of CO
348 < coverage. There is a correlation between increasing CO coverage and Pt
347 > atoms ($\sim 600$) was similar between all systems, independent of \ce{CO}
348 > coverage. There is a correlation between increasing \ce{CO} coverage and Pt
349   diffusion rates of $\sim 1.6$ \AA\textsuperscript{2}/ns/ML.
350  
351   \begin{table} \centering \begin{tabular}{| c | c |} \hline
352 <    CO Coverage & Diffusion Constant\footnotemark[1] (\AA\textsuperscript{2}/ns) \\
352 >    \ce{CO} Coverage & Diffusion Constant\footnotemark[1] (\AA\textsuperscript{2}/ns) \\
353      \hline
354      0    & 2.779(2) \\
355      0.05 & 3.992(6) \\
# Line 357 | Line 357 | diffusion rates of $\sim 1.6$ \AA\textsuperscript{2}/n
357      0.33 & 4.180(7) \\
358      0.50 & 3.935(5) \\
359      \hline \end{tabular}
360 < \caption{Diffusion constants of mobile Pt
360 > \caption{Diffusion constants of mobile \ce{Pt}
361      atoms for the subsurface alloys.\label{tab:diffusion}}
362  
363    \footnotemark[1]{Uncertainties in the last digit
# Line 366 | Line 366 | In a similar manner to the Pt(557) surfaces, the struc
366  
367   \subsection{Island Formation and Clustering in the Subsurface Alloy}
368  
369 < In a similar manner to the Pt(557) surfaces, the structural
369 > In a similar manner to the \ce{Pt} (557) surfaces, the structural
370   reconstructions that occur for the subsurface alloy are influenced by
371 < the presence of the CO adsorbate. In Figure \ref{fig:domainAreasPd},
372 < the area of exposed Pd increases both over time, and as a function of
373 < CO coverage. The appearance of the subsurface Pd requires a
374 < simultaneous reduction in the surface area of the outer Pt skin. Two
375 < scenarios could explain the reduction of exposed Pt: either the Pt
376 < atoms are being buried under the Pd bulk, or islands of Pt are forming
377 < on top of the Pd surface.
371 > the presence of the \ce{CO} adsorbate. In Figure \ref{fig:domainAreasPd},
372 > the area of exposed \ce{Pd} increases both over time, and as a function of
373 > \ce{CO} coverage. The appearance of the subsurface \ce{Pd} requires a
374 > simultaneous reduction in the surface area of the outer \ce{Pt} skin. Two
375 > scenarios could explain the reduction of exposed \ce{Pt}: either the \ce{Pt}
376 > atoms are being buried under the \ce{Pd} bulk, or islands of \ce{Pt} are forming
377 > on top of the \ce{Pd} surface.
378  
379 < Both mechanisms would explain the decreased Pt surface area (see Fig.
379 > Both mechanisms would explain the decreased \ce{Pt} surface area (see Fig.
380   \ref{fig:domainAreasPt}). To discern which of these mechanisms is
381   taking place, the identity of nearest metal atom neighbors can be
382 < tabulated. Single-layer Pt skins have atoms with 6 Pt nearest
383 < neighbors. Islands of Pt require the presence of Pt atoms with 7-9 Pt
382 > tabulated. Single-layer \ce{Pt} skins have atoms with 6 \ce{Pt} nearest
383 > neighbors. Islands of \ce{Pt} require the presence of \ce{Pt} atoms with 7-9 \ce{Pt}
384   nearest neighbors. In figure \ref{fig:nearestNeighbors}, we see an
385 < increase in Pt population with 9 Pt nearest neighbors along with the
386 < simultaneous decrease in Pt atoms with only 6 Pt nearest neighbors.
387 < This is evidence for the formation of multi-layer Pt features since
388 < single layers of Pt are restricted to having 6 Pt nearest neighbors.
385 > increase in \ce{Pt} population with 9 \ce{Pt} nearest neighbors along with the
386 > simultaneous decrease in \ce{Pt} atoms with only 6 \ce{Pt} nearest neighbors.
387 > This is evidence for the formation of multi-layer \ce{Pt} features since
388 > single layers of \ce{Pt} are restricted to having 6 \ce{Pt} nearest neighbors.
389   We note that nearest-neighbor population analysis provides information
390   similar to the information one might obtain from an XAFS experiment,
391   which could make this phenomenon experimentally observable.
# Line 393 | Line 393 | which could make this phenomenon experimentally observ
393   \begin{figure}
394   \includegraphics[width=\linewidth]{../figures/domainAreas/domainSize_Pd_110ns_deCluttered_color.pdf}
395   %\includegraphics[width=\linewidth]{../figures/domainAreas/final_domain_Pd.pdf}
396 < \caption{Distributions of Pd domain size as a function of time and CO coverage.
396 > \caption{Distributions of \ce{Pd} domain size as a function of time and \ce{CO} coverage.
397   Data is averaged over $\sim$20~ns segments to help show progression,
398   additionally, the data is shown as a percentage of the total surface area of
399 < the Pt@Pd system with the integration of the curves equaling the percentage
400 < surface area of Pd, shown in Table \ref{tab:integratedArea}. The presence of CO
401 < leads to more exposure of the underlying Pd, which is quantified here by an
402 < increasing number and increasing size of Pd domains. The bare Pt@Pd surface,
399 > the \ce{Pt/Pd} system with the integration of the curves equaling the percentage
400 > surface area of \ce{Pd}, shown in Table \ref{tab:integratedArea}. The presence of \ce{CO}
401 > leads to more exposure of the underlying \ce{Pd}, which is quantified here by an
402 > increasing number and increasing size of \ce{Pd} domains. The bare \ce{Pt/Pd} surface,
403   as seen in Figure \ref{fig:systems}.B, undergoes some restructuring, however, the
404   extent is much less when compared to the 25\% and 50\% monolayer (ML) systems.}
405   \label{fig:domainAreasPd}
# Line 408 | Line 408 | extent is much less when compared to the 25\% and 50\%
408   \begin{figure}
409   \includegraphics[width=\linewidth]{../figures/domainAreas/domainSize_Pt_110ns_deCluttered_color.pdf}
410   %\includegraphics[width=\linewidth]{../figures/domainAreas/final_domain_Pt.pdf}
411 < \caption{Distributions of Pt domain size as a function of time and CO coverage.
412 < Here the presence of CO facilitates the clustering of Pt into smaller domains
413 < by forming multilayer features which leads to a reduction of Pt surface coverage and concomitant increased exposure of the Pd.}
411 > \caption{Distributions of \ce{Pt} domain size as a function of time and \ce{CO} coverage.
412 > Here the presence of \ce{CO} facilitates the clustering of \ce{Pt} into smaller domains
413 > by forming multilayer features which leads to a reduction of \ce{Pt} surface coverage and concomitant increased exposure of the \ce{Pd}.}
414   \label{fig:domainAreasPt}
415   \end{figure}
416  
417  
418   \begin{figure}
419    \includegraphics[width=\linewidth]{../figures/nearestNeighbor/NearestNeighbor_110ns_color.pdf}
420 <  \caption{Population of Pt atoms with either 6 (solid) or 9 (hollow)
421 <    Pt nearest neighbors averaged over similar blocks of time as in
420 >  \caption{Population of \ce{Pt} atoms with either 6 (solid) or 9 (hollow)
421 >    \ce{Pt} nearest neighbors averaged over similar blocks of time as in
422      Figure \ref{fig:domainAreasPd} and \ref{fig:domainAreasPt}. At
423 <    time 0, the majority ($\frac{2}{3}$) of Pt is located in the (111)
424 <    plateaus where the number of Pt nearest neighbors is 6. A sizeable
423 >    time 0, the majority ($\frac{2}{3}$) of \ce{Pt} is located in the (111)
424 >    plateaus where the number of \ce{Pt} nearest neighbors is 6. A sizeable
425      minority ($\frac{1}{3}$) is located at the step edge, or beneath a
426 <    step edge with a nearest neighbor number of 5. The decrease in Pt
427 <    with 6 nearest neighbors, while Pt with 9 nearest neighbors rises
428 <    implies that Pt atoms are being incorporated into multilayer
426 >    step edge with a nearest neighbor number of 5. The decrease in \ce{Pt}
427 >    with 6 nearest neighbors, while \ce{Pt} with 9 nearest neighbors rises
428 >    implies that \ce{Pt} atoms are being incorporated into multilayer
429      features. } \label{fig:nearestNeighbors}
430   \end{figure}
431  
432   The small amount of restructuring observed in the zero coverage system suggests
433 < that there are two driving forces for restructuring, with the CO playing one
433 > that there are two driving forces for restructuring, with the \ce{CO} playing one
434   role.
435  
436  
437  
438   \begin{table}
439 <  \caption{Percent Pd surface coverage as a function of time. The following values were obtained by integrating the data in Figure \ref{fig:domainAreasPd}.}
439 >  \caption{Percent \ce{Pd} surface coverage as a function of time. The following values were obtained by integrating the data in Figure \ref{fig:domainAreasPd}.}
440    \begin{tabular}{| c || c | c | c | c | c | c |}
441    \hline
442    & 0-18 ns & 19-37 ns & 38-56 ns & 57-75 ns & 76-94 ns & 95-113 ns \\
# Line 455 | Line 455 | Pt maximizes Pt-Pt bonds. (C) and (D) have undergone g
455   \section{Discussion}
456  
457   Explaining figure 1:  The minor restructuring in B is due to the energy benefit gained when
458 < Pt maximizes Pt-Pt bonds. (C) and (D) have undergone greater
459 < remodeling because the presence of CO helps speed up adatom mobility
460 < and enables the vertical displacement of Pt adatoms leading to more
458 > \ce{Pt} maximizes \ce{Pt\bond{-}Pt} bonds. (C) and (D) have undergone greater
459 > remodeling because the presence of \ce{CO} helps speed up adatom mobility
460 > and enables the vertical displacement of \ce{Pt} adatoms leading to more
461   clustering.
462  
463 < The stronger Pd-CO binding energy when compared to Pt-CO is hypothesized to
463 > The stronger \ce{Pd\bond{-}CO} binding energy when compared to \ce{Pt\bond{-}CO} is hypothesized to
464   play a role in disrupting the surface and in the case of the shell system in
465 < revealing the underlying Pd by causing clustering and island formation of the
466 < Pt shell.
465 > revealing the underlying \ce{Pd} by causing clustering and island formation of the
466 > \ce{Pt} shell.
467  
468   \subsection{Diffusion}
469 < As noted above, their is limited movement of Pd in any of the systems we
470 < examined. In a few instances, inversion is observed where a Pd and a Pt atom
471 < are swapped in the shell systems. But on the whole the Pd is overwhelmingly
469 > As noted above, their is limited movement of \ce{Pd} in any of the systems we
470 > examined. In a few instances, inversion is observed where a \ce{Pd} and a \ce{Pt} atom
471 > are swapped in the shell systems. But on the whole the \ce{Pd} is overwhelmingly
472   stationary. Time scales and kinetic barriers are possible explanations for the
473   lack of movement, but for the shell systems what seems to be the most likely is
474 < that the Pt is acting as a protective layer. Even with significant
475 < restructuring of the Pt overlayer, the underlying Pd is unlikely to be located
474 > that the \ce{Pt} is acting as a protective layer. Even with significant
475 > restructuring of the \ce{Pt} overlayer, the underlying \ce{Pd} is unlikely to be located
476   in a position where an energetically easier break from a step-edge will be
477   possible. However, this explanation does not explain the stability of the pure
478 < Pd systems and is an area for further exploration.
478 > \ce{Pd} systems and is an area for further exploration.
479  
480 < An analysis of Pt's perpendicular (across the plateaus) and parallel (along the
480 > An analysis of \ce{Pt}'s perpendicular (across the plateaus) and parallel (along the
481   steps) diffusion constants on the various shell systems is shown in the
482   supporting information. Unlike in our previous work\cite{Michalka:2013}, where
483   the step-edges were maintained throughout the restructuring of the surface from
# Line 489 | Line 489 | increasing CO coverage. This correlation likely stems
489   previous work which is easily explained by the lower temperature these systems
490   were run at (850~K compared to 1000~K). While the 5\% data is abnormally high,
491   the other coverages show a strong correlation of increasing diffusion with
492 < increasing CO coverage. This correlation likely stems from the same mechanism
493 < we reported previously, where the presence of CO, coupled with its large
492 > increasing \ce{CO} coverage. This correlation likely stems from the same mechanism
493 > we reported previously, where the presence of \ce{CO}, coupled with its large
494   quadrupolar moment assists in the initial break-up of the step-edges allowing
495 < for consistent adatom formation.  Once the Pt adatoms are formed, the barrier
495 > for consistent adatom formation.  Once the \ce{Pt} adatoms are formed, the barrier
496   for diffusion is negligible ($<$4 kcal/mol using the EAM forcefield) and the
497   adatom will continue to diffuse until it is reincorporated, with most diffusion
498 < occuring along the front of the step edges. Thus, the more CO present on the
498 > occuring along the front of the step edges. Thus, the more \ce{CO} present on the
499   surface, the more likely adatoms will form and explore the surface before
500   reaching a more stable state.
501  
502   \subsection{Relative Metallic Binding Energies}
503 < The presence and amount of CO is one of the driving forces for the observed
503 > The presence and amount of \ce{CO} is one of the driving forces for the observed
504   reconstruction, however, this doesn't explain the minor restructuring observed
505 < for the shell system that had no CO present. Rather, there appears to be two
505 > for the shell system that had no \ce{CO} present. Rather, there appears to be two
506   factors that are both responsible for aspects of the restructuring. This other
507 < driving force is that Pt-Pt interactions are stronger and thus more favored
508 < when compared to Pt-Pd interactions, as established by the EAM forcefield.
509 < Removing a Pt surface atom from a (111) plateau on a pure Pt (557) surface,
510 < shows that the Pt was contributing (-$\infty$ kcal/mol) to the energy of the
511 < system, while a Pt taken from a similar spot in our shell system was only
512 < contributing (-$\infty$ kcal/mol).  In the first instance, the Pt had 9
513 < nearest neighbors, all Pt, while in the second the three atoms underneath the
514 < surface are now Pd, which contribute a smaller electron density, leading to a
515 < weaker binding between the Pt and Pd. As Figure \ref{fig:nearestNeighbors}
516 < shows, over the 110 ns of the simulation, the number of Pt with increasing
517 < number of Pt-Pt nearest neighbors grows. Thus, the restructuring of the surface
518 < for the 0\%~coverage system can be explained by the stronger Pt-Pt binding
519 < interaction, while the presence of CO is what appears to allow or speed up the
507 > driving force is that \ce{Pt\bond{-}Pt} interactions are stronger and thus more favored
508 > when compared to \ce{Pt\bond{-}Pd} interactions, as established by the EAM forcefield.
509 > Removing a \ce{Pt} surface atom from a (111) plateau on a pure \ce{Pt} (557) surface,
510 > shows that the \ce{Pt} was contributing (-$\infty$ kcal/mol) to the energy of the
511 > system, while a \ce{Pt} taken from a similar spot in our shell system was only
512 > contributing (-$\infty$ kcal/mol).  In the first instance, the \ce{Pt} had 9
513 > nearest neighbors, all \ce{Pt}, while in the second the three atoms underneath the
514 > surface are now \ce{Pd}, which contribute a smaller electron density, leading to a
515 > weaker binding between the \ce{Pt} and \ce{Pd}. As Figure \ref{fig:nearestNeighbors}
516 > shows, over the 110 ns of the simulation, the number of \ce{Pt} with increasing
517 > number of \ce{Pt\bond{-}Pt} nearest neighbors grows. Thus, the restructuring of the surface
518 > for the 0\%~coverage system can be explained by the stronger \ce{Pt\bond{-}Pt} binding
519 > interaction, while the presence of \ce{CO} is what appears to allow or speed up the
520   mechanism of step traversal, leading to larger scale reconstructions and for
521   the shell systems, clustering and island formation.
522  
# Line 527 | Line 527 | site, Pt or Pd. The resulting Ising-like grids were th
527   exposed surfaces were first simplified by projecting the 3-dimensional surface
528   onto a 2-dimensional grid (with two grids per system to capture the surfaces on
529   both sides of the system). The grids could only have one of two values at each
530 < site, Pt or Pd. The resulting Ising-like grids were then deconvoluted into
530 > site, \ce{Pt} or \ce{Pd}. The resulting Ising-like grids were then deconvoluted into
531   separate domains based on nearest-neighbor connectivity (up, down, left, right;
532   corners were not included). The resulting data was aggregated and normalized
533   and is presented in Figures \ref{fig:domainAreasPd} and
# Line 536 | Line 536 | beginning of the simulations, the surface layer of Pt
536  
537   This analysis allows us to focus on collective motion of the surface atoms as
538   measured by the domain sizes, rather than individual adatom movement.  At the
539 < beginning of the simulations, the surface layer of Pt makes up one domain of
539 > beginning of the simulations, the surface layer of \ce{Pt} makes up one domain of
540   size $\sim$2625~\AA\textsuperscript{2}. This domain begins to shrink relatively
541 < quickly and is matched by a growth in the number and size of Pd domains.  The
542 < presence of CO in the system allows further clustering
543 < of the Pt domains, which requires a larger amount of exposed
544 < Pd of various domain sizes. For clarity purposes, there is a small peak in the
545 < Pt graphs around 0-100~\AA~that is not shown in Figure \ref{fig:domainAreasPt}
541 > quickly and is matched by a growth in the number and size of \ce{Pd} domains.  The
542 > presence of \ce{CO} in the system allows further clustering
543 > of the \ce{Pt} domains, which requires a larger amount of exposed
544 > \ce{Pd} of various domain sizes. For clarity purposes, there is a small peak in the
545 > \ce{Pt} graphs around 0-100~\AA~that is not shown in Figure \ref{fig:domainAreasPt}
546   but can be seen in the supporting information. These data poins arise from 1 to
547 < 2 atom clusters of Pt embedded in the Pd.
547 > 2 atom clusters of \ce{Pt} embedded in the \ce{Pd}.
548  
549   The quantification of the surface composition that these figures display is
550   helpful, but is more easily seen when the curves are integrated, which is shown
# Line 553 | Line 553 | small but significant amount of restructuring, despite
553  
554   \subsection{Equilibrium state}
555   As shown in Figure \ref{fig:systems}.B, the 0\% coverage system has undergone a
556 < small but significant amount of restructuring, despite no CO being present.
557 < This is due to the stronger Pt-Pt compared to Pt-Pd binding energy.  Movement
558 < of Pt from one layer onto the top of another layer without vertical
559 < displacement benefits both layers of Pt, and the small energy barrier
556 > small but significant amount of restructuring, despite no \ce{CO} being present.
557 > This is due to the stronger \ce{Pt\bond{-}Pt} compared to \ce{Pt\bond{-}Pd} binding energy.  Movement
558 > of \ce{Pt} from one layer onto the top of another layer without vertical
559 > displacement benefits both layers of \ce{Pt}, and the small energy barrier
560   preventing it is overcome by the increased thermal motion at elevated
561 < temperatures. The now underlying Pt has approximately 9 nearest neighbors of Pt
562 < and 3 of Pd and is essentially in bulk.  The upper layer of Pt also benefits
561 > temperatures. The now underlying \ce{Pt} has approximately 9 nearest neighbors of \ce{Pt}
562 > and 3 of \ce{Pd} and is essentially in bulk.  The upper layer of \ce{Pt} also benefits
563   because it is now experiencing 9 nearest neighbor interactions, all with other
564 < Pt. The ideal case would involve the majority of Pt maximizing their Pt-Pt
565 < interactions which could lead to massive disruption without any need for CO,
564 > \ce{Pt}. The ideal case would involve the majority of \ce{Pt} maximizing their \ce{Pt\bond{-}Pt}
565 > interactions which could lead to massive disruption without any need for \ce{CO},
566   but as seen in Figure \ref{fig:systems}.B, the (557) crystal facet is still
567 < present, just with Pt plateaus moved slightly forward and backward. Without the
568 < presence of CO, very little vertical displacement is observed, which is what is
567 > present, just with \ce{Pt} plateaus moved slightly forward and backward. Without the
568 > presence of \ce{CO}, very little vertical displacement is observed, which is what is
569   hypothesized to facilite the multiple layer features observed in the higher
570   coverage systems. The systems were run for approximately 110 nanoseconds and
571   then stopped, primarily because, large scale changes had drastically slowed.
# Line 575 | Line 575 | while possible, was not judged to be useful at this ti
575   state, at least for the time scales we were able to explore. Increased run time
576   while possible, was not judged to be useful at this time.
577  
578 < \subsection{Role of CO: Presence and Absence}
579 < As shown in the previous sections, the presence of CO plays a large role in the
580 < restructuring of the Pt@Pd shell systems. The small amount of restructuring due
581 < to favorable Pt-Pt interactions is greatly enhanced when CO is added to the
582 < system. As concluded in our previous paper\cite{Michalka:2013}, CO helps enable
578 > \subsection{Role of \ce{CO}: Presence and Absence}
579 > As shown in the previous sections, the presence of \ce{CO} plays a large role in the
580 > restructuring of the \ce{Pt/Pd} shell systems. The small amount of restructuring due
581 > to favorable \ce{Pt\bond{-}Pt} interactions is greatly enhanced when \ce{CO} is added to the
582 > system. As concluded in our previous paper\cite{Michalka:2013}, \ce{CO} helps enable
583   vertical displacement of adatoms between layers, which is also seen here by
584 < examining the degree of clustering that occurred for various CO coverages.
584 > examining the degree of clustering that occurred for various \ce{CO} coverages.
585  
586   %One
587   %final test we performed, already mentioned in Figure \ref{fig:domainAreasNoCO},
# Line 599 | Line 599 | The favorable Pt-Pt interactions, coupled with the str
599  
600  
601   \section{Conclusion}
602 < The favorable Pt-Pt interactions, coupled with the stronger Pd-CO binding
603 < energy help to explain the clustering seen on the Pt@Pd (557) systems. The lack
604 < of any surface disruption on the Pd (557) surfaces at all coverages, suggests
605 < that the presence of CO is not enough of a perturbation to overcome the
602 > The favorable \ce{Pt\bond{-}Pt} interactions, coupled with the stronger \ce{Pd\bond{-}CO} binding
603 > energy help to explain the clustering seen on the \ce{Pt/Pd} (557) systems. The lack
604 > of any surface disruption on the \ce{Pd} (557) surfaces at all coverages, suggests
605 > that the presence of \ce{CO} is not enough of a perturbation to overcome the
606   thermodynamic barriers hindering reconstruction.
607  
608   This work suggests that bimetallic and subsurface alloys could be tailored to

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines