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 4297 by jmichalk, Tue Feb 17 15:29:25 2015 UTC vs.
Revision 4298 by gezelter, Tue Feb 17 17:10:04 2015 UTC

# Line 50 | Line 50 | Bimetallic alloys, subsurface alloys, and core-shell n
50  
51  
52   \section{Introduction}
53 +
54 + Pt-based and Pd-based bimetallic materials are crucial catalysts in
55 + oxygen reduction reactions (ORR),\cite{Lim:2009fk} and oxygen
56 + evolution reactions (OER),\cite{Reier:2012uq} that are important in
57 + charging and discharging of Li-air batteries\cite{Lu:2011vn} and in
58 + fuel cell processes.\cite{Bliznakov:2012kx} Oxide-supported noble
59 + metal particles are also important in the water-gas shift
60 + reaction.\cite{Bunluesin:1998ys}
61 +
62 +
63   Bimetallic alloys, subsurface alloys, and core-shell nanostructures are
64   currently under intense investigation\cite{Kim:2013jt, Gao:2009oj, Gao:2009wo}
65   because of their large accesible design space for various catalytic processes.
# Line 137 | Line 147 | describe the \ce{Pt} and \ce{Pd} electron densities, e
147   neighbor interactions during parameterization.\cite{a}
148  
149   In this work, we have employed the embedded atom method (EAM) to
150 < describe the \ce{Pt} and \ce{Pd} electron densities, embedding functionals, and
151 < pair potentials,\cite{EAM} utilizing the Johnson mixing rules for the
152 < \ce{Pt\bond{-}Pd} cross-interactions.\cite{johnson89}
150 > describe the \ce{Pt} and \ce{Pd} electron densities, embedding
151 > functionals, and pair potentials,\cite{EAM} utilizing the Johnson
152 > mixing rules for the \ce{Pt\bond{-}Pd}
153 > cross-interactions.\cite{johnson89}
154  
155   The carbon monoxide (\ce{CO}) self-interactions were modeled using a
156   rigid three-site model developed by Straub and Karplus for studying
# Line 152 | Line 163 | The \ce{Pd\bond{-}CO} interaction potential was parame
163   data.\cite{Michalka:2013,Deshlahra:2012} This modification yields a
164   slightly weaker \ce{Pt\bond{-}CO} binding energy.
165  
166 < The \ce{Pd\bond{-}CO} interaction potential was parameterized as part of this
167 < work, and uses similar functional forms to the \ce{Pt\bond{-}CO}
168 < model.\cite{Michalka:2013} Our starting point is a model introduced by
169 < Korzeniewski \textit{et al.}\cite{Pons:1986} The parameters were modified to
170 < reflect binding energies and binding site preferences on the \ce{M} (111)
171 < surfaces.  One key difference from the potential in Ref.
172 < \citenum{Michalka:2013} is that the \ce{M\bond{-}O} bond is modeled using a
173 < purely repulsive Morse potential, $D e^{-2\gamma(r-r_e)}$.  The functional
174 < forms and the broad repulsive \ce{M\bond{-}O} contribution are flexible enough
175 < to reproduce the atop preference for \ce{Pt\bond{-}CO} as well as the
176 < bridge/hollow - preference for \ce{Pd\bond{-}CO}.  Parameters for the
177 < potentials are given in Table~\ref{tab:CO_parameters} and the calculated
178 < binding energies at various binding sites are shown in
179 < Table~\ref{tab:CO_energies}.  
166 > The \ce{Pd\bond{-}CO} interaction potential was parameterized as part
167 > of this work, and uses similar functional forms to the
168 > \ce{Pt\bond{-}CO} model.\cite{Michalka:2013} Our starting point is a
169 > model introduced by Korzeniewski \textit{et al.}\cite{Pons:1986} The
170 > parameters were modified to reflect binding energies and binding site
171 > preferences on the \ce{M} (111) surfaces.  One key difference from the
172 > potential in Ref.  \citenum{Michalka:2013} is that the \ce{M\bond{-}O}
173 > bond is modeled using a purely repulsive Morse potential, $D
174 > e^{-2\gamma(r-r_e)}$.  The functional forms and the broad repulsive
175 > \ce{M\bond{-}O} contribution are flexible enough to reproduce the atop
176 > preference for \ce{Pt\bond{-}CO} as well as the bridge/hollow -
177 > preference for \ce{Pd\bond{-}CO}.  Parameters for the potentials are
178 > given in Table~\ref{tab:CO_parameters} and the calculated binding
179 > energies at various binding sites are shown in
180 > Table~\ref{tab:CO_energies}.
181  
182   \begin{table}
183   \caption{Parameters for the metal-\ce{CO} cross-interactions. Metal-Carbon
# Line 207 | Line 219 | This \ce{Pd\bond{-}CO} model does not have a strong pr
219   \label{tab:CO_energies}
220   \end{table}
221  
222 < This \ce{Pd\bond{-}CO} model does not have a strong preference for either the
223 < bridge or hollow binding sites, so it may overestimate the bridge-site binding
224 < at low coverages, but at higher coverages, the situation is somewhat less
225 < clear.\cite{Wong:1991ta} Studies using low-energy elecron diffraction (LEED)
226 < and \ce{C\bond{-}O} stretching frequencies of \ce{CO} bound to \ce{Pd}(111)
227 < suggest that the 3-fold hollow sites are preferred at low
228 < coverages,\cite{Bradshaw:1978uf,Conrad:1978fx,Ohtani:1987zh} where it forms a
229 < $(\sqrt{3} \times \sqrt{3}) R~30^{\circ}$ pattern.  These observations are
230 < supported by temperature desorption spectroscopy,\cite{Guo:1989} and infrared
231 < absorption spectroscopy~\cite{Szanyi:1992} where binding energies have been
222 > This \ce{Pd\bond{-}CO} model does not have a strong preference for
223 > either the bridge or hollow binding sites, so it may overestimate the
224 > bridge-site binding at low coverages, but at higher coverages, the
225 > situation is somewhat less clear.\cite{Wong:1991ta} Studies using
226 > low-energy elecron diffraction (LEED) and \ce{C\bond{-}O} stretching
227 > frequencies of \ce{CO} bound to \ce{Pd}(111) suggest that the 3-fold
228 > hollow sites are preferred at low
229 > coverages,\cite{Bradshaw:1978uf,Conrad:1978fx,Ohtani:1987zh} where it
230 > forms a $(\sqrt{3} \times \sqrt{3}) R~30^{\circ}$ pattern.  These
231 > observations are supported by temperature desorption
232 > spectroscopy,\cite{Guo:1989} and infrared absorption
233 > spectroscopy~\cite{Szanyi:1992} where binding energies have been
234   reported to lie between 1.3 and 1.54 eV.
235  
236   At higher \ce{CO} coverages (e.g. $> 0.5$ ML), the preferred binding
237 < of \ce{CO} on \ce{Pd}(111) appears to be a $c(4\times2)$ ordered structure with
238 < the \ce{CO} bound to the bridge sites.\cite{Bradshaw:1978uf}
237 > of \ce{CO} on \ce{Pd}(111) appears to be a $c(4\times2)$ ordered
238 > structure with the \ce{CO} bound to the bridge
239 > sites.\cite{Bradshaw:1978uf}
240  
241   Theoretical work by Honkala \textit{et al.}\cite{Honkala:2001sf} using
242   DFT with the generalized gradient approximation (GGA) to describe
243 < electron exchange correlation and pseudopotentials for the \ce{Pd} atoms
244 < also reported the fcc site as the most favorable binding position with
245 < a binding energy of 2.00 eV compared to the bridge site binding
246 < energy of 1.83 eV at $1/3$ monolayer.
243 > electron exchange correlation and pseudopotentials for the \ce{Pd}
244 > atoms also reported the fcc site as the most favorable binding
245 > position with a binding energy of 2.00 eV compared to the bridge site
246 > binding energy of 1.83 eV at $1/3$ monolayer.
247  
248   High resolution x-ray photoelectron spectroscopy (XPS) results from
249   Surnev \textit{et al.}\cite{Surnev:2000uk} confirm that the preferred
# Line 240 | Line 255 | Although the weak preference for hollow vs. bridge sit
255   binding energy difference shrinks, as at $1/2$ ML the hollow to bridge
256   energy difference is 0.06 eV (-1.85 hollow, -1.79 bridge).
257  
258 < Although the weak preference for hollow vs. bridge sites is not captured by the
259 < \ce{Pd\bond{-}CO} fit, it does represent a significant change from the atop
260 < preference of the \ce{Pt\bond{-}CO} model.  The dynamics of the metal bound to
261 < the \ce{CO} is significantly altered as a result of this difference.
258 > Although the weak preference for hollow vs. bridge sites is not
259 > captured by the \ce{Pd\bond{-}CO} fit, the slight favoring of the
260 > bridge adsorption site in this model does result in an accurate
261 > reproduction of the $c(4\times2)$ adsorption structure at higher
262 > coverages, which would be the most relevant regime for catalytic
263 > behavior.
264  
248
249 Although the weak preference for hollow vs. bridge sites is not captured by the
250 \ce{Pd\bond{-}CO} fit, the slight favoring of the bridge adsorption site in
251 this model does result in an accurate reproduction of the $c(4\times2)$
252 adsorption structure at higher coverages, which is where most of the
253 restructuring was observed.
254
255
265   \subsection{557 interfaces and subsurface alloys}
266 < The \ce{Pd}(557) model is an orthorhombic periodic box with dimensions of
267 < $55.09 \times 49.48 \times 120$~\AA~ while the subsurface alloys
266 > The \ce{Pd}(557) model is an orthorhombic periodic box with dimensions
267 > of $55.09 \times 49.48 \times 120$~\AA~ while the subsurface alloys
268   (Pt(557) surface layers, with Pd bulk) have dimensions of $54.875
269 < \times 49.235 \times 120$~\AA.  The \ce{Pd} system consists of 9 layers of
270 < \ce{Pd} while our subsurface alloys consist of 7 layers of \ce{Pd} sandwiched
271 < between 2 layers of \ce{Pt}.  Both the pure \ce{Pd} slab and the subsurface
272 < alloy systems are $\sim$22~\AA~ thick. The lattice constants for \ce{Pd}
273 < and \ce{Pt}, 3.89 and 3.92~\AA, respectively, provide minimal strain energy
274 < in the alloy, and the relaxed geometries of the two interfaces are
275 < therefore quite similar.
269 > \times 49.235 \times 120$~\AA.  The \ce{Pd} system consists of 9
270 > layers of \ce{Pd} while our subsurface alloys consist of 7 layers of
271 > \ce{Pd} sandwiched between 2 layers of \ce{Pt}.  Both the pure \ce{Pd}
272 > slab and the subsurface alloy systems are $\sim$22~\AA~ thick. The
273 > lattice constants for \ce{Pd} and \ce{Pt}, 3.89 and 3.92~\AA,
274 > respectively, provide minimal strain energy in the alloy, and the
275 > relaxed geometries of the two interfaces are therefore quite similar.
276  
277   The systems are cut from a FCC crystal along the 557 plane, and are
278   rotated so that they are periodic in the $x$ and $y$ directions,
# Line 301 | Line 310 | In our earlier work on \ce{Pt}(557) we observed \ce{CO
310   OpenMD.\cite{openmd,OOPSE}
311  
312   \section{Results}
313 < In our earlier work on \ce{Pt}(557) we observed \ce{CO}-induced restructuring
314 < into relatively clean double-layer structures. For the pure \ce{Pd}(557)
315 < studied here, the 557 facet retains the plateaus and steps with only
316 < minimal adatom movement, and with almost no surface reconstruction.
317 < Higher \ce{CO} coverages appear to have minimal effect on the pure \ce{Pd}(557)
318 < systems.
313 > In our earlier work on \ce{Pt}(557), we observed \ce{CO}-induced
314 > restructuring into relatively clean double-layer structures. For the
315 > pure \ce{Pd}(557) studied here, the 557 facet retains the plateaus and
316 > steps with only minimal adatom movement, and with almost no surface
317 > reconstruction.  Higher \ce{CO} coverages appear to have minimal
318 > effect on the pure \ce{Pd}(557) systems.
319  
320 < However, the \ce{Pt/Pd} subsurface alloy exhibits a \ce{CO}-induced speedup of
321 < the diffusion of surface metal atoms, as well as a large-scale
322 < restructuring of the well-ordered surface into \ce{Pt}-rich islands, and
323 < will therefore be the focus of most of our analysis.
320 > However, the \ce{Pt}-coated \ce{Pd} alloy exhibits a \ce{CO}-induced
321 > speedup of the diffusion of surface metal atoms, as well as a
322 > large-scale restructuring of the well-ordered surface into
323 > \ce{Pt}-rich islands, and will therefore be the focus of most of our
324 > analysis.
325  
326   \begin{figure}
327    \includegraphics[width=\linewidth]{../figures/SystemFigures/systems_ochre2.png}
# Line 325 | Line 335 | various systems after significant exposure to the \ce{
335   \end{figure}
336  
337   Figure \ref{fig:systems} shows representative configurations of the
338 < various systems after significant exposure to the \ce{CO}. We see that the
339 < Pd system highlighted in panel A has undergone no surface
338 > various systems after significant exposure to the \ce{CO}. We see that
339 > the Pd system highlighted in panel A has undergone no surface
340   restructuring. The other three panels highlight the effect of varying
341 < \ce{CO} concentrations on the subsurface alloys, which do exhibit
341 > \ce{CO} concentrations on the surface alloys, which do exhibit
342   structural reorganization.
343  
344 < \subsection{Diffusion of Surface Metal Atoms in the Subsurface Alloy}
344 > \subsection{Diffusion of Surface Metal Atoms in the Surface Alloy}
345  
346   Figure \ref{fig:systems} suggests that there is limited to no mobility
347   on the pure Pd systems. Analysis of the surface atom mobility showed
# Line 372 | Line 382 | In a similar manner to the \ce{Pt} (557) surfaces, the
382  
383   \subsection{Island Formation and Clustering in the Subsurface Alloy}
384  
385 < In a similar manner to the \ce{Pt} (557) surfaces, the structural
385 > In a similar manner to the \ce{Pt}(557) surfaces, the structural
386   reconstructions that occur for the subsurface alloy are influenced by
387 < the presence of the \ce{CO} adsorbate. In Figure \ref{fig:domainAreasPd},
388 < the area of exposed \ce{Pd} increases both over time, and as a function of
389 < \ce{CO} coverage. The appearance of the subsurface \ce{Pd} requires a
390 < simultaneous reduction in the surface area of the outer \ce{Pt} skin. Two
391 < scenarios could explain the reduction of exposed \ce{Pt}: either the \ce{Pt}
392 < atoms are being buried under the \ce{Pd} bulk, or islands of \ce{Pt} are forming
393 < on top of the \ce{Pd} surface.
387 > the presence of the \ce{CO} adsorbate. In Figure
388 > \ref{fig:domainAreasPd}, the area of exposed \ce{Pd} increases both
389 > over time, and as a function of \ce{CO} coverage. The presence of
390 > \ce{CO} leads to more exposure of the underlying \ce{Pd}, measured by
391 > the increasing number and size of \ce{Pd} domains. Without \ce{CO}
392 > exposure, the bare \ce{Pt/Pd} surface does undergo some restructuring,
393 > although both the rate and extent is significantly smaller than in the
394 > 0.25 and 0.50 monolayer (ML) systems.
395  
396 < Both mechanisms would explain the decreased \ce{Pt} surface area (see Fig.
397 < \ref{fig:domainAreasPt}). To discern which of these mechanisms is
398 < taking place, the identity of nearest metal atom neighbors can be
399 < tabulated. Single-layer \ce{Pt} skins have atoms with 6 \ce{Pt} nearest
400 < neighbors. Islands of \ce{Pt} require the presence of \ce{Pt} atoms with 7-9 \ce{Pt}
390 < nearest neighbors. In figure \ref{fig:nearestNeighbors}, we see an
391 < increase in \ce{Pt} population with 9 \ce{Pt} nearest neighbors along with the
392 < simultaneous decrease in \ce{Pt} atoms with only 6 \ce{Pt} nearest neighbors.
393 < This is evidence for the formation of multi-layer \ce{Pt} features since
394 < single layers of \ce{Pt} are restricted to having 6 \ce{Pt} nearest neighbors.
395 < We note that nearest-neighbor population analysis provides information
396 < similar to the information one might obtain from an XAFS experiment,
397 < which could make this phenomenon experimentally observable.
396 > The appearance of \ce{Pd} from the bulk layers on the surface requires
397 > a simultaneous reduction in the surface area of the outer \ce{Pt}
398 > skin. Two scenarios could explain the reduction of exposed \ce{Pt}:
399 > either the \ce{Pt} atoms are being buried under the \ce{Pd} bulk, or
400 > islands of \ce{Pt} are forming on top of the \ce{Pd} surface.
401  
402 + Both mechanisms would explain the decreased \ce{Pt} surface area (see
403 + Fig.  \ref{fig:domainAreasPt}).  To discern which of these mechanisms
404 + is taking place, the identity of nearest metal atom neighbors can be
405 + tabulated as a function of time of exposure to \ce{CO}. Single-layer
406 + \ce{Pt} skins have atoms with 6 \ce{Pt} nearest neighbors. Islands of
407 + \ce{Pt} require the presence of \ce{Pt} atoms with 7-9 \ce{Pt} nearest
408 + neighbors. In figure \ref{fig:nearestNeighbors}, we see an increase in
409 + \ce{Pt} population with 9 \ce{Pt} nearest neighbors along with the
410 + simultaneous decrease in \ce{Pt} atoms with only 6 \ce{Pt} nearest
411 + neighbors.  This is evidence for the formation of multi-layer \ce{Pt}
412 + features since single layers of \ce{Pt} are restricted to having 6
413 + \ce{Pt} nearest neighbors.
414 +
415 + The presence of \ce{CO} therefore appears to facilitate the clustering
416 + of \ce{Pt} into smaller domains by forming multilayer features which
417 + leads to a reduction of \ce{Pt} surface coverage and concomitant
418 + increased exposure of the \ce{Pd}.  We note that nearest-neighbor
419 + population analysis provides information similar to the information
420 + one might obtain from an XAFS experiment, which could make this
421 + phenomenon experimentally observable.
422 +
423   \begin{figure}
424   \includegraphics[width=\linewidth]{../figures/domainAreas/domainSize_Pd_110ns_deCluttered_color.pdf}
425   %\includegraphics[width=\linewidth]{../figures/domainAreas/final_domain_Pd.pdf}
426 < \caption{Distributions of \ce{Pd} domain size as a function of time and \ce{CO} coverage.
427 < Data is averaged over $\sim$20~ns segments to help show progression,
404 < additionally, the data is shown as a percentage of the total surface area of
405 < the \ce{Pt/Pd} system with the integration of the curves equaling the percentage
406 < surface area of \ce{Pd}, shown in Table \ref{tab:integratedArea}. The presence of \ce{CO}
407 < leads to more exposure of the underlying \ce{Pd}, which is quantified here by an
408 < increasing number and increasing size of \ce{Pd} domains. The bare \ce{Pt/Pd} surface,
409 < as seen in Figure \ref{fig:systems}.B, undergoes some restructuring, however, the
410 < extent is much less when compared to the 25\% and 50\% monolayer (ML) systems.}
426 > \caption{Distributions of \ce{Pd} domain sizes at different \ce{CO}
427 >  coverages and at different times after exposure to \ce{CO}.}
428   \label{fig:domainAreasPd}
429   \end{figure}
430  
431   \begin{figure}
432   \includegraphics[width=\linewidth]{../figures/domainAreas/domainSize_Pt_110ns_deCluttered_color.pdf}
433   %\includegraphics[width=\linewidth]{../figures/domainAreas/final_domain_Pt.pdf}
434 < \caption{Distributions of \ce{Pt} domain size as a function of time and \ce{CO} coverage.
435 < Here the presence of \ce{CO} facilitates the clustering of \ce{Pt} into smaller domains
419 < by forming multilayer features which leads to a reduction of \ce{Pt} surface coverage and concomitant increased exposure of the \ce{Pd}.}
434 > \caption{Distributions of \ce{Pt} domain sizes at different \ce{CO}
435 >  coverages and at different times after exposure to \ce{CO}.}
436   \label{fig:domainAreasPt}
437   \end{figure}
438  
423
439   \begin{figure}
440    \includegraphics[width=\linewidth]{../figures/nearestNeighbor/NearestNeighbor_110ns_color.pdf}
441 <  \caption{Population of \ce{Pt} atoms with either 6 (solid) or 9 (hollow)
442 <    \ce{Pt} nearest neighbors averaged over similar blocks of time as in
443 <    Figure \ref{fig:domainAreasPd} and \ref{fig:domainAreasPt}. At
444 <    time 0, the majority ($\frac{2}{3}$) of \ce{Pt} is located in the (111)
445 <    plateaus where the number of \ce{Pt} nearest neighbors is 6. A sizeable
446 <    minority ($\frac{1}{3}$) is located at the step edge, or beneath a
432 <    step edge with a nearest neighbor number of 5. The decrease in \ce{Pt}
433 <    with 6 nearest neighbors, while \ce{Pt} with 9 nearest neighbors rises
434 <    implies that \ce{Pt} atoms are being incorporated into multilayer
435 <    features. } \label{fig:nearestNeighbors}
441 >  \caption{Population of \ce{Pt} atoms with either 6 (solid) or 9
442 >    (hollow) \ce{Pt} nearest neighbors averaged over 18 ns blocks of
443 >    time.  At $t=0$, the majority ($\frac{2}{3}$) of \ce{Pt} is
444 >    located in the (111) plateaus where the number of \ce{Pt} nearest
445 >    neighbors is 6. The remaining \ce{Pt} is located at step edges,
446 >    with a nearest neighbor \ce{Pt} count of 5.} \label{fig:nearestNeighbors}
447   \end{figure}
448  
449 < The small amount of restructuring observed in the zero coverage system suggests
450 < that there are two driving forces for restructuring, with the \ce{CO} playing one
451 < role.
449 > The small amount of restructuring observed in the bare metal system
450 > suggests that the relative surface energies of the two metals provides
451 > some of the driving force for the restructuring, while the \ce{CO}
452 > significantly speeds up the effects (and may help to drive the process
453 > at lower temperatures).
454  
442
443
455   \begin{table}
456 <  \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}.}
456 >  \caption{\ce{Pd} surface coverage (in \% of total surface area)
457 >    averaged over 18 ns blocks of time.}
458    \begin{tabular}{| c || c | c | c | c | c | c |}
459    \hline
460 <  & 0-18 ns & 19-37 ns & 38-56 ns & 57-75 ns & 76-94 ns & 95-113 ns \\
460 > \ce{CO} coverage  & 0-18 ns & 19-37 ns & 38-56 ns & 57-75 ns & 76-94 ns & 95-113 ns \\
461    \hline
462    0.00 &  6.6 & 16.2 & 20.1 & 21.7 & 23.5 & 25.2 \\
463    0.05 &  8.0 & 15.8 & 20.2 & 25.1 & 27.6 & 30.9 \\

Diff Legend

Removed lines
+ Added lines
< Changed lines
> Changed lines