ViewVC Help
View File | Revision Log | Show Annotations | View Changeset | Root Listing
root/group/trunk/COonPt/COonPtAu.tex
Revision: 3814
Committed: Fri Dec 14 19:19:04 2012 UTC (11 years, 6 months ago) by jmichalk
Content type: application/x-tex
Original Path: trunk/COonPt/firstTry.tex
File size: 20835 byte(s)
Log Message:
Table 4, diffusion constants updated

File Contents

# Content
1 \documentclass[11pt]{article}
2 \usepackage{amsmath}
3 \usepackage{amssymb}
4 \usepackage{setspace}
5 \usepackage{endfloat}
6 \usepackage{caption}
7 %\usepackage{tabularx}
8 \usepackage{graphicx}
9 \usepackage{multirow}
10 %\usepackage{booktabs}
11 %\usepackage{bibentry}
12 %\usepackage{mathrsfs}
13 %\usepackage[ref]{overcite}
14 \usepackage[square, comma, sort&compress]{natbib}
15 \usepackage{url}
16 \pagestyle{plain} \pagenumbering{arabic} \oddsidemargin 0.0cm
17 \evensidemargin 0.0cm \topmargin -21pt \headsep 10pt \textheight
18 9.0in \textwidth 6.5in \brokenpenalty=10000
19
20 % double space list of tables and figures
21 \AtBeginDelayedFloats{\renewcommand{\baselinestretch}{1.66}}
22 \setlength{\abovecaptionskip}{20 pt}
23 \setlength{\belowcaptionskip}{30 pt}
24
25 %\renewcommand\citemid{\ } % no comma in optional reference note
26 \bibpunct{[}{]}{,}{n}{}{;}
27 \bibliographystyle{achemso}
28
29 \begin{document}
30
31
32 %%
33 %Introduction
34 % Experimental observations
35 % Previous work on Pt, CO, etc.
36 %
37 %Simulation Methodology
38 % FF (fits and parameters)
39 % MD (setup, equilibration, collection)
40 %
41 % Analysis of trajectories!!!
42 %Discussion
43 % CO preferences for specific locales
44 % CO-CO interactions
45 % Differences between Au & Pt
46 % Causes of 2_layer reordering in Pt
47 %Summary
48 %%
49
50 %Title
51 \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
54 Gezelter\footnote{Corresponding author. \ Electronic mail: gezelter@nd.edu} \\
55 Department of Chemistry and Biochemistry,\\
56 University of Notre Dame\\
57 Notre Dame, Indiana 46556}
58 %Date
59 \date{Dec 15, 2012}
60 %authors
61
62 % make the title
63 \maketitle
64
65 \begin{doublespace}
66
67 \begin{abstract}
68
69 \end{abstract}
70
71 \newpage
72
73
74 \section{Introduction}
75 % Importance: catalytically active metals are important
76 % Sub: Knowledge of how their surface structure affects their ability to catalytically facilitate certain reactions is growing, but is more reactionary than predictive
77 % Sub: Designing catalysis is the future, and will play an important role in numerous processes (ones that are currently seen to be impractical, or at least inefficient)
78 % Theory can explore temperatures and pressures which are difficult to work with in experiments
79 % Sub: Also, easier to observe what is going on and provide reasons and explanations
80 %
81
82 Industrial catalysts usually consist of small particles exposing
83 different atomic terminations that exhibit a high concentration of
84 step, kink sites, and vacancies at the edges of the facets. These
85 sites are thought to be the locations of catalytic
86 activity.\cite{ISI:000083038000001,ISI:000083924800001} There is now
87 significant evidence to demonstrate that solid surfaces are often
88 structurally, compositionally, and chemically {\it modified} by
89 reactants under operating conditions.\cite{Tao2008,Tao:2010,Tao2011}
90 The coupling between surface oxidation state and catalytic activity
91 for CO oxidation on Pt, for instance, is widely
92 documented.\cite{Ertl08,Hendriksen:2002} Despite the well-documented
93 role of these effects on reactivity, the ability to capture or predict
94 them in atomistic models is currently somewhat limited. While these
95 effects are perhaps unsurprising on the highly disperse, multi-faceted
96 nanoscale particles that characterize industrial catalysts, they are
97 manifest even on ordered, well-defined surfaces. The Pt(557) surface,
98 for example, exhibits substantial and reversible restructuring under
99 exposure to moderate pressures of carbon monoxide.\cite{Tao:2010}
100
101 This work is part of an ongoing effort to understand the causes,
102 mechanisms and timescales for surface restructuring using molecular
103 simulation methods. Since the dynamics of the process is of
104 particular interest, we utilize classical molecular dynamic methods
105 with force fields that represent a compromise between chemical
106 accuracy and the computational efficiency necessary to observe the
107 process of interest.
108
109 Since restructuring occurs as a result of specific interactions of the catalyst
110 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.
114 %Platinum molecular dynamics
115 %gold molecular dynamics
116
117
118
119
120
121
122 \section{Simulation Methods}
123 The challenge in modeling any solid/gas interface problem is the
124 development of a sufficiently general yet computationally tractable
125 model of the chemical interactions between the surface atoms and
126 adsorbates. Since the interfaces involved are quite large (10$^3$ -
127 10$^6$ atoms) and respond slowly to perturbations, {\it ab initio}
128 molecular dynamics
129 (AIMD),\cite{KRESSE:1993ve,KRESSE:1993qf,KRESSE:1994ul} Car-Parrinello
130 methods,\cite{CAR:1985bh,Izvekov:2000fv,Guidelli:2000fy} and quantum
131 mechanical potential energy surfaces remain out of reach.
132 Additionally, the ``bonds'' between metal atoms at a surface are
133 typically not well represented in terms of classical pairwise
134 interactions in the same way that bonds in a molecular material are,
135 nor are they captured by simple non-directional interactions like the
136 Coulomb potential. For this work, we have been using classical
137 molecular dynamics with potential energy surfaces that are
138 specifically tuned for transition metals. In particular, we use the
139 EAM potential for Au-Au and Pt-Pt interactions, while modeling the CO
140 using a model developed by Straub and Karplus for studying
141 photodissociation of CO from myoglobin.\cite{Straub} The Au-CO and Pt-CO
142 cross interactions were parameterized as part of this work.
143
144 \subsection{Metal-metal interactions}
145 Many of the potentials used for classical simulation of transition
146 metals are based on a non-pairwise additive functional of the local
147 electron density. The embedded atom method (EAM) is perhaps the best
148 known of these
149 methods,\cite{Daw84,Foiles86,Johnson89,Daw89,Plimpton93,Voter95a,Lu97,Alemany98}
150 but other models like the Finnis-Sinclair\cite{Finnis84,Chen90} and
151 the quantum-corrected Sutton-Chen method\cite{QSC,Qi99} have simpler
152 parameter sets. The glue model of Ercolessi {\it et al.} is among the
153 fastest of these density functional approaches.\cite{Ercolessi88} In
154 all of these models, atoms are conceptualized as a positively charged
155 core with a radially-decaying valence electron distribution. To
156 calculate the energy for embedding the core at a particular location,
157 the electron density due to the valence electrons at all of the other
158 atomic sites is computed at atom $i$'s location,
159 \begin{equation*}
160 \bar{\rho}_i = \sum_{j\neq i} \rho_j(r_{ij})
161 \end{equation*}
162 Here, $\rho_j(r_{ij})$ is the function that describes the distance
163 dependence of the valence electron distribution of atom $j$. The
164 contribution to the potential that comes from placing atom $i$ at that
165 location is then
166 \begin{equation*}
167 V_i = F[ \bar{\rho}_i ] + \sum_{j \neq i} \phi_{ij}(r_{ij})
168 \end{equation*}
169 where $F[ \bar{\rho}_i ]$ is an energy embedding functional, and
170 $\phi_{ij}(r_{ij})$ is an pairwise term that is meant to represent the
171 overlap of the two positively charged cores.
172
173 The {\it modified} embedded atom method (MEAM) adds angular terms to
174 the electron density functions and an angular screening factor to the
175 pairwise interaction between two
176 atoms.\cite{BASKES:1994fk,Lee:2000vn,Thijsse:2002ly,Timonova:2011ve}
177 MEAM has become widely used to simulate systems in which angular
178 interactions are important (e.g. silicon,\cite{Timonova:2011ve} bcc
179 metals,\cite{Lee:2001qf} and also interfaces.\cite{Beurden:2002ys})
180 MEAM presents significant additional computational costs, however.
181
182 The EAM, Finnis-Sinclair, MEAM, and the Quantum Sutton-Chen potentials
183 have all been widely used by the materials simulation community for
184 simulations of bulk and nanoparticle
185 properties,\cite{Chui:2003fk,Wang:2005qy,Medasani:2007uq}
186 melting,\cite{Belonoshko00,sankaranarayanan:155441,Sankaranarayanan:2005lr}
187 fracture,\cite{Shastry:1996qg,Shastry:1998dx} crack
188 propagation,\cite{BECQUART:1993rg} and alloying
189 dynamics.\cite{Shibata:2002hh} All of these potentials have their
190 strengths and weaknesses. One of the strengths common to all of the
191 methods is the relatively large library of metals for which these
192 potentials have been
193 parameterized.\cite{Foiles86,PhysRevB.37.3924,Rifkin1992,mishin99:_inter,mishin01:cu,mishin02:b2nial,zope03:tial_ap,mishin05:phase_fe_ni}
194
195 \subsection{CO}
196 Since one explanation for the strong surface CO repulsion on metals is
197 the large linear quadrupole moment of carbon monoxide, the model
198 chosen for this molecule exhibits this property in an efficient
199 manner. We used a model first proposed by Karplus and Straub to study
200 the photodissociation of CO from myoglobin.\cite{Straub} The Straub and
201 Karplus model is a rigid three site model which places a massless M
202 site at the center of mass along the CO bond. The geometry used along
203 with the interaction parameters are reproduced in Table 1. The effective
204 dipole moment, calculated from the assigned charges, is still
205 small (0.35 D) while the linear quadrupole (-2.40 D~\AA) is close
206 to the experimental (-2.63 D~\AA)\cite{QuadrupoleCO} and quantum
207 mechanical predictions (-2.46 D~\AA)\cite{QuadrupoleCOCalc}.
208 %CO Table
209 \begin{table}[H]
210 \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 $e$.}
211 \centering
212 \begin{tabular}{| c | c | ccc |}
213 \hline
214 \multicolumn{5}{|c|}{\textbf{Self-Interactions}}\\
215 \hline
216 & {\it z} & $\sigma$ & $\epsilon$ & q\\
217 \hline
218 \textbf{C} & -0.6457 & 0.0262 & 3.83 & -0.75 \\
219 \textbf{O} & 0.4843 & 0.1591 & 3.12 & -0.85 \\
220 \textbf{M} & 0.0 & - & - & 1.6 \\
221 \hline
222 \end{tabular}
223 \end{table}
224
225 \subsection{Cross-Interactions}
226
227 One hurdle that must be overcome in classical molecular simulations
228 is the proper parameterization of the potential interactions present
229 in the system. Since the adsorption of CO onto a platinum surface has been
230 the focus of many experimental \cite{Yeo, Hopster:1978, Ertl:1977, Kelemen:1979}
231 and theoretical studies \cite{Beurden:2002ys,Pons:1986,Deshlahra:2009,Feibelman:2001,Mason:2004}
232 there is a large amount of data in the literature to fit too. We started with parameters
233 reported by Korzeniewski et al. \cite{Pons:1986} and then
234 modified them to ensure that the Pt-CO interaction favored
235 an atop binding position for the CO upon the Pt surface. This
236 constraint led to the binding energies being on the higher side
237 of reported values. Following the method laid out by Korzeniewski,
238 the Pt-C interaction was fit to a strong Lennard-Jones 12-6
239 interaction to mimic binding, while the Pt-O interaction
240 was parameterized to a Morse potential with a large $r_o$
241 to contribute a weak repulsion. The resultant potential-energy
242 surface suitably recovers the calculated Pt-C bond length ( 1.6\AA)\cite{Beurden:2002ys} and affinity
243 for the atop binding position.\cite{Deshlahra:2012, Hopster:1978}
244
245 %where did you actually get the functionals for citation?
246 %scf calculations, so initial relaxation was of the four layers, but two layers weren't kept fixed, I don't think
247 %same cutoff for slab and slab + CO ? seems low, although feibelmen had values around there...
248 The Au-C and Au-O interaction parameters were also fit to a Lennard-Jones
249 and Morse potential respectively, to reproduce Au-CO binding energies.
250 These energies were obtained from quantum calculations carried out using
251 the PBE GGA exchange-correlation functionals\cite{Perdew_GGA} for gold, carbon, and oxygen
252 constructed by Rappe, Rabe, Kaxiras, and Joannopoulos. \cite{RRKJ_PP}.
253 All calculations were run using the {\sc Quantum ESPRESSO} package. \cite{QE-2009}
254 First, a four layer slab of gold comprised of 32 atoms displaying a (111) surface was
255 converged using a 4X4X4 grid of Monkhorst-Pack \emph{k}-points.\cite{Monkhorst:1976}
256 The kinetic energy of the wavefunctions were truncated at 20 Ry while the
257 cutoff for the charge density and potential was set at 80 Ry. This relaxed
258 gold slab was then used in numerous single point calculations with CO at various heights
259 to create a potential energy surface for the Au-CO interaction.
260
261 %Hint at future work
262 The fit parameter sets employed in this work are shown in Table 2 and their
263 reproduction of the binding energies are displayed in Table 3. Currently,
264 charge transfer is not being treated in this system, however, that is a goal
265 for future work as the effect has been seen to affect binding energies and
266 binding site preferences. \cite{Deshlahra:2012}
267
268
269
270
271 \subsection{Construction and Equilibration of 557 Metal interfaces}
272
273 Our model systems are composed of approximately 4000 metal atoms
274 cut along the 557 plane so that they are periodic in the {\it x} and {\it y}
275 directions exposing the 557 plane in the {\it z} direction. Runs at various
276 temperatures ranging from 300~K to 1200~K were started with the intent
277 of viewing relative stability of the surface when CO was not present in the
278 system. Owing to the different melting points (1337~K for Au and 2045~K for Pt),
279 the bare crystal systems were initially run in the Canonical ensemble at
280 800~K and 1000~K respectively for 100 ps. Various amounts of CO were
281 placed in the vacuum region, which upon full adsorption to the surface
282 corresponded to 5\%, 25\%, 33\%, and 50\% coverages. These systems
283 were again allowed to reach thermal equilibrium before being run in the
284 microcanonical ensemble. All of the systems examined in this work were
285 run for at least 40 ns. A subset that were undergoing interesting effects
286 have been allowed to continue running with one system approaching 200 ns.
287 All simulations were run using the open source molecular dynamics package, OpenMD. \cite{Ewald, OOPSE}
288
289
290
291
292
293
294 %\subsection{System}
295 %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.
296
297
298 %Table of Parameters
299 %Pt Parameter Set 9
300 %Au Parameter Set 35
301 \begin{table}[H]
302 \caption{Best fit parameters for metal-adsorbate interactions. Distances are in \AA~and energies are in kcal/mol}
303 \centering
304 \begin{tabular}{| c | cc | c | ccc |}
305 \hline
306 \multicolumn{7}{| c |}{\textbf{Cross-Interactions} }\\
307 \hline
308 & $\sigma$ & $\epsilon$ & & $r$ & $D$ & $\gamma$ \\
309 \hline
310 \textbf{Pt-C} & 1.3 & 15 & \textbf{Pt-O} & 3.8 & 3.0 & 1 \\
311 \textbf{Au-C} & 1.9 & 6.5 & \textbf{Au-O} & 3.8 & 0.37 & 0.9\\
312
313 \hline
314 \end{tabular}
315 \end{table}
316
317 %Table of energies
318 \begin{table}[H]
319 \caption{Adsorption energies in eV}
320 \centering
321 \begin{tabular}{| c | cc |}
322 \hline
323 & Calc. & Exp. \\
324 \hline
325 \textbf{Pt-CO} & -1.9 & -1.4~\cite{Kelemen:1979}-- -1.9~\cite{Yeo} \\
326 \textbf{Au-CO} & -0.39 & -0.44~\cite{TPD_Gold_CO} \\
327 \hline
328 \end{tabular}
329 \end{table}
330
331
332
333
334
335
336 % Just results, leave discussion for discussion section
337 \section{Results}
338 \subsection{Diffusion}
339 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.
340
341 %Table of Diffusion Constants
342 %Add gold?M
343 \begin{table}[H]
344 \caption{Platinum and gold 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. Units are \AA\textsuperscript{2}/ns}
345 \centering
346 \begin{tabular}{| c | cc | cc | c |}
347 \hline
348 \textbf{System Coverage} & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & $\mathbf{D}_{\parallel}$ & $\mathbf{D}_{\perp}$ & \textbf{Time (ns)}\\
349 \hline
350 &\multicolumn{2}{c|}{\textbf{Platinum}}&\multicolumn{2}{c|}{\textbf{Gold}} & \\
351 \hline
352 50\% & 4.32 $\pm$ 0.02 & 1.185 $\pm$ 0.008 & 1.72 $\pm$ 0.02 & 0.455 $\pm$ 0.006 & 40 \\
353 33\% & 5.18 $\pm$ 0.03 & 1.999 $\pm$ 0.005 & 1.95 $\pm$ 0.02 & 0.337 $\pm$ 0.004 & 40 \\
354 25\% & 5.01 $\pm$ 0.02 & 1.574 $\pm$ 0.004 & 1.26 $\pm$ 0.03 & 0.377 $\pm$ 0.006 & 40 \\
355 5\% & 3.61 $\pm$ 0.02 & 0.355 $\pm$ 0.002 & 1.84 $\pm$ 0.03 & 0.169 $\pm$ 0.004 & 40 \\
356 0\% & 3.27 $\pm$ 0.02 & 0.147 $\pm$ 0.004 & 1.50 $\pm$ 0.02 & 0.194 $\pm$ 0.002 & 40 \\
357 \hline
358 \end{tabular}
359 \end{table}
360
361
362
363 %Discussion
364 \section{Discussion}
365 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.
366
367 \subsection{Diffusion}
368 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?)
369 \\
370 \\
371 %Evolution of surface
372 \begin{figure}[H]
373 \includegraphics[scale=0.5]{ProgressionOfDoubleLayerFormation_yellowCircle.png}
374 \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.}
375 \end{figure}
376
377
378
379
380 %Peaks!
381 \includegraphics[scale=0.25]{doublePeaks_noCO.png}
382 \section{Conclusion}
383
384
385 \section{Acknowledgments}
386 Support for this project was provided by the National Science
387 Foundation under grant CHE-0848243 and by the Center for Sustainable
388 Energy at Notre Dame (cSEND). Computational time was provided by the
389 Center for Research Computing (CRC) at the University of Notre Dame.
390
391 \newpage
392 \bibliography{firstTryBibliography}
393 \end{doublespace}
394 \end{document}