cond-mat0311541/prl.tex
1: \documentclass[prl,aps,twocolumn,showpacs,home,floats]{revtex4}
2: %\documentclass[prl,superscriptaddress,aps,twocolumn,showpacs,home,floats]{revtex4}
3: %\documentclass[prl,aps,preprint,showpacs,home,floats]{revtex4}
4: 
5: \usepackage{graphics}
6: \bibliographystyle{apsrev}
7: \begin{document}
8: 
9: \title{{\rm\small\hfill (Phys. Rev. Lett., in press)}\\
10: Kinetic hindrance during the initial oxidation of Pd(100) at ambient pressures}
11: 
12: 
13: \author{E. Lundgren}
14: \email{edvin.lundgren@sljus.lu.se}
15: \affiliation{Department of Synchrotron Radiation Research,
16: Institute of Physics, University
17: of Lund, Box 118, S-221 00 Lund, Sweden}
18: \author{A. Stierle}
19: \affiliation{Max-Planck Institut f\"ur Metallforschung,
20: Heisenbergstr.3, D-70569 Stuttgart, Germany}
21: \author{M. Todorova}
22: \affiliation{Fritz-Haber Institut der Max-Planck Gesellschaft,
23: Faradayweg 4-6, D-14195 Berlin, Germany}
24: \author{J. Gustafson}
25: \affiliation{Department of Synchrotron Radiation Research,
26: Institute of Physics, University of Lund, Box 118, S-221 00 Lund,
27: Sweden}
28: \author{A. Mikkelsen}
29: \affiliation{Department of Synchrotron Radiation Research,
30: Institute of Physics, University of Lund, Box 118, S-221 00 Lund,
31: Sweden}
32: \author{J. Rogal}
33: \affiliation{Fritz-Haber Institut der Max-Planck Gesellschaft,
34: Faradayweg 4-6, D-14195 Berlin, Germany}
35: \author{K. Reuter}
36: \affiliation{Fritz-Haber Institut der Max-Planck Gesellschaft,
37: Faradayweg 4-6, D-14195 Berlin, Germany}
38: \author{J.N. Andersen}
39: \affiliation{Department of Synchrotron Radiation Research,
40: Institute of Physics, University of Lund, Box 118, S-221 00 Lund,
41: Sweden}
42: \author{H. Dosch}
43: \affiliation{Max-Planck Institut f\"ur Metallforschung,
44: Heisenbergstr.3, D-70569 Stuttgart, Germany}
45: \author{M. Scheffler}
46: \affiliation{Fritz-Haber Institut der Max-Planck Gesellschaft,
47: Faradayweg 4-6, D-14195 Berlin, Germany}
48: 
49: 
50: \begin{abstract}
51: The oxidation of the Pd(100) surface at oxygen pressures in the
52: 10${}^{-6}$ to 10${}^3$\,mbar range and temperatures up to 1000 K
53: has been studied {\em in-situ} by surface x-ray diffraction (SXRD)
54: The results provide direct structural information on the phases
55: present in the surface region and on the kinetics of the oxide
56: formation. Depending on the $(T,p)$ environmental conditions we
57: either observe a thin $(\sqrt{5} \times \sqrt{5})R27^{\circ}$
58: surface oxide or the growth of a rough, poorly ordered bulk oxide
59: film of PdO predominantly with (001) orientation. By either
60: comparison to the surface phase diagram from first-principles
61: atomistic thermodynamics or by explicit time-resolved measurements
62: we identify a strong kinetic hindrance to the bulk oxide formation
63: even at temperatures as high as 675\,K.
64: \end{abstract}
65: 
66: \date{\today}
67: \pacs{61.10.-i, 81.65.Mq, 68.55.Jk, 68.43.Bc}
68: 
69: %%KR Edvin: Here's two possible PACS suggestions. I don't know,
70: %%KR whether there is a special PACS for high-pressure experiments,
71: %%KR but there is certainly one for SXRD. The 68.43.Bc should come
72: %%KR last - we don't want to get theoretical referees...
73: 
74: %61.10.-i %X-ray diffraction and scattering
75: %81.65.Mq  %Oxidation
76: %68.43.Bc % Ab initio calculations of adsorbate structure and
77: %reactions
78: 
79: \maketitle
80: 
81: Many technologically important materials containing transition
82: metals are intended for use under oxygen pressures much higher
83: than those of the high or ultra high vacuum (UHV) environment
84: typically employed in surface science related investigations of
85: the structural, electronic, and chemical properties of these
86: materials. As the surface properties of such materials may be
87: significantly altered by the oxidation or corrosion
88: \cite{over00,hendriksen02,stampfl02} occurring at oxygen pressures
89: difficult to achieve in conventional UHV experiments, it is
90: disconcerting that most of our present atomic-scale knowledge
91: derives from such experiments or from theoretical treatments which
92: neglect the surrounding gas-phase. Despite the pressure
93: limitations a few such experiments and theoretical simulations
94: have in recent years significantly advanced our atomic-scale
95: understanding of the initial oxidation of metal surfaces,
96: demonstrating e.g. how radically a surface may change its
97: structure and functionality under conditions appropriate for high
98: pressure oxidation catalysis \cite{over00,hendriksen02} and how
99: the growth of bulk oxide films is often preceded by the formation
100: of few-atomic-layer-thin so-called surface oxides of complex
101: geometrical structure and with properties often unknown from the
102: bulk oxides \cite{carlisle00,lundgren02,todorova03}. The
103: unexpected and complex behavior revealed by these experiments
104: emphasizes the need for further {\em in-situ} investigations of
105: the structural, electronic and chemical surface properties at
106: higher oxygen pressures but maintaining the accuracy known from
107: UHV studies - in particular to also address the kinetics of
108: oxidation and corrosion processes.
109: 
110: A main reason for the lack of {\em in-situ} investigations has
111: been the scarcity of appropriate experimental techniques that
112: provide the aspired atomically-resolved information at high
113: pressure {\em in situ}, or theories that explicitly include the
114: effect of the surrounding gas-phase. Recently enforced attempts to
115: overcome this limitation have on the theoretical side led to the
116: development of first-principles atomistic thermodynamics, where
117: electronic structure theory calculations are combined with
118: thermodynamic considerations to address the surface structure and
119: composition of a metal surface in {\em equilibrium} with arbitrary
120: environments (See Refs. \cite{reuter02,li03} and references
121: therein). On the experimental side the state-of-the-art is,
122: however, currently still characterized by either traditional {\em
123: ex-situ} atomic-scale investigations \cite{over00}, or new high
124: pressure techniques like high pressure scanning tunneling
125: microscopy (STM) that still lack atomic resolution
126: \cite{hendriksen02}.
127: 
128: 
129: In this Letter we demonstrate how precise knowledge of the
130: $(T,p)$-conditions under which such various oxides, -bulk and
131: surface-, exist can be obtained by means of {\em in-situ} surface
132: X-ray diffraction (SXRD) measurements. We have chosen to study Pd
133: as the metal is a highly active oxidation catalyst of hydrocarbons
134: under oxygen-rich conditions \cite{ziauddin97}. Still it is
135: unknown whether Pd or PdO is the active phase \cite{veser99}.
136: Monitoring the oxidation of the Pd(100) surface over the pressure
137: range from 10${}^{-6}$ to 10${}^3$\,mbar and up to sample
138: temperatures of 1000\,K, we observe the formation of both the
139: previously characterized $(\sqrt{5} \times \sqrt{5})R27^{\circ}$
140: surface oxide \cite{todorova03} and the transform to
141: three-dimensional bulk oxide films. Framing our experimental data
142: with the equilibrium results from atomistic thermodynamics
143: calculations we can identify kinetically inhibited bulk oxide
144: growth even at temperatures as high as 675\,K. Under suitable
145: conditions this hindered transformation to the bulk oxide can be
146: followed on a time scale currently accessible to the experiment,
147: opening the door to time-resolved atomic-scale studies of the
148: initial oxidation process of metal surfaces at ambient pressures.
149: 
150: The SXRD measurements were performed at the Angstr{\o}m Quelle
151: Karlsruhe (ANKA) beamline in Germany \cite{beamline}. A photon
152: energy of 10.5\,keV was used and the experiments were conducted in
153: a six-circle diffraction mode. The crystal basis used to describe
154: the (H K L) diffraction is a tetragonal basis set \
155: ({\bf{a}$_{1}$},{\bf{a}$_{2}$},{\bf{a}$_{3}$}), with
156: {\bf{a}$_{1}$} and {\bf{a}$_{2}$} lying in the surface plane and
157: of length equal to the nearest neighbor surface distance
158: a$_{\circ}$/$\sqrt{2}$, and a$_{3}$ out-of-plane with length
159: a$_{\circ}$ (a$_{\circ}$(Pd) = 3.89{\AA}). The UHV x-ray
160: diffraction chamber allowed partial oxygen pressures of up to
161: 10${}^3$ mbar, and the temperature was estimated from a
162: thermocouple mounted behind the transferable sample holder,
163: resulting in an uncertainty of the sample temperature of $\pm
164: 25$\,K. The sample and the cleaning procedure are identical to the
165: one described in an earlier UHV study \cite{todorova03}. The oxide
166: films grown were found to be metastable on the time scale of hours
167: under UHV conditions, and could readily be desorbed at 1175\,K.
168: 
169: The atomistic thermodynamics results are based on
170: density-functional theory (DFT) calculations performed within the
171: Full-Potential Linear Augmented Plane Wave (FP-LAPW) scheme
172: \cite{blaha99} using the generalized gradient approximation (GGA)
173: \cite{perdew96} for the exchange-correlation functional. The
174: supercell setup and the highly converged basis-set \cite{basis}
175: have been detailed in our preceding study \cite{todorova03}. To
176: determine the range of $(T,p)$-conditions in which this surface
177: oxide would represent the thermodynamically most stable state, we
178: evaluate the Gibbs free energy of adsorption \cite{reuter02,li03}
179: and compare it to other possible states of the system, like the
180: reported $p(2 \times 2)$ on-surface phase with O in fcc sites
181: \cite{zheng00,todorova03} or the tetragonal PdO bulk oxide
182: \cite{rogers71}.
183: 
184: \begin{figure}[t!]
185: \scalebox{0.28}{\includegraphics{fig1.ps}}
186: \caption{\label{fig1} SXRD in-plane K-scan in reciprocal lattice
187: units (r.l.u) along (0.6 K 0.3) at $T=575$\,K and
188: $p=10^{-3}$\,mbar. The location of this scan line in reciprocal
189: space is indicated in the inset showing the diffraction from the
190: $(\sqrt{5} \times \sqrt{5})R27^{\circ}$ surface oxide as observed
191: in UHV-LEED. The direction of the in-plane (H K) basis vectors and
192: the unit-cell of one of the surface oxide domains are also plotted
193: in the inset.}
194: \end{figure}
195: 
196: Figure \ref{fig1} depicts a SXRD K-scan at a sample temperature of
197: 575\,K and partial oxygen pressure of $10^{-3}$\,mbar. From the
198: inset showing the reciprocal space as observed with UHV low-energy
199: electron diffraction (LEED) it is obvious that the two observed
200: peaks in the SXRD scan arise from the $(\sqrt{5} \times
201: \sqrt{5})R27^{\circ}$ surface oxide. The width of the rocking scan
202: at the (0.6 0.8 0.3) reflection is (0.13${}^{\circ}$) equal to the
203: substrate rocking scan in the minimum of the crystal truncation
204: rod (CTR) before dosing, allowing us to conclude that the surface
205: oxide domains extend over complete substrate terraces -- in
206: agreement with our previous STM results \cite{todorova03}.
207: 
208: \begin{figure}[t!]
209: \scalebox{0.32}{\includegraphics{fig2.ps}}
210: \caption{\label{fig2} Top: SXRD out-of-plane L-scan along (0.6 0.8
211: L). Bottom: corresponding schematic of the observed out-of-plane
212: diffraction (full lines - rods due to both Pd(100) and the surface
213: oxide, dashed lines - rods only due to the surface oxide. Bragg
214: reflections from bulk Pd and bulk PdO are marked as black and grey
215: circles respectively). a) $p= 10^{-3}$\,mbar and T=575 K, only the
216: surface oxide diffraction is apparent, b) p= 10${}^3$\,mbar and
217: T=675 K, diffraction is now due to a bulk-like PdO film.}
218: \end{figure}
219: 
220: 
221: 
222: The  out-of-plane diffraction from the (0.6 0.8 L) reflection
223: under the same temperature and pressure conditions is shown as an
224: L-scan in the top section of Fig \ref{fig2}a. The smooth decrease
225: of the intensity and the absence of sharper peaks with increasing
226: L is a clear fingerprint of a single diffracting layer, in
227: agreement with the recent finding \cite{todorova03} that the
228: $(\sqrt{5} \times \sqrt{5})R27^{\circ}$ surface oxide consists of
229: a single PdO(101) layer adsorbed on the Pd(001) surface. The
230: observed diffraction changes significantly as the oxygen pressure
231: and temperature is increased to 10${}^3$\,mbar and 675 K,
232: respectively, as shown in Fig. \ref{fig2}b. Instead of a smoothly
233: decreasing diffraction intensity with increasing L, a peak is now
234: observed at L=0.74. Since the reciprocal lattice is defined by the
235: lattice constant of Pd, this peak corresponds to a lattice
236: distance of a$_{\circ}$(Pd)/0.74 = 5.26\,{\AA}, which is very
237: close to the $c$-lattice constant of bulk PdO, namely 5.33\,{\AA}.
238: Thus, we observe bulk-like diffraction from PdO, indicating the
239: formation of several layers of PdO on the Pd(001) surface.
240: Interestingly, we also note that no diffracted intensity is
241: observed at L=0 anymore demonstrating that the $(\sqrt{5} \times
242: \sqrt{5})R27^{\circ}$ surface oxide has completely disappeared
243: from the surface, i.e. the initially formed PdO(101) plane does
244: not continue to grow but instead restructures. This is in
245: agreement with our DFT calculations identifying the (101)
246: orientation as a higher-energy facet of PdO \cite{todorova03}, and
247: also with experimental observations on the preferred growth
248: direction of PdO \cite{mcbride91}. Further, since no finite
249: thickness oscillations are observed along the rod, the observed
250: oxide film must be rough, as has also been reported in a recent
251: high-pressure STM study of this surface \cite{hendriksen03}.
252: 
253: A more detailed analysis of our diffraction data allows us to even
254: draw some more quantitative conclusions about the properties of
255: the grown oxide. We observe no change in the overall shape of the
256: (1 1 L) CTR upon oxidation. Without apparent relaxation, we
257: therefore attribute the change in integrated intensity at the
258: minimum (1 1 1) to a change in surface roughness. Estimating the
259: latter within the $\beta$-model \cite{robinson88} yields
260: approximately root-mean-square roughness of 10\,{\AA}, whereas the
261: measured width of $\sim 0.1$ of the L=0.74 peak indicates an
262: average film thickness of around $\sim 3.89$\,{\AA}/0.1 = $\sim
263: 40$\,{\AA}. The poor order of the formed oxide fringe is further
264: reflected in the 1${}^{\circ}$ width of the rocking scan at the
265: L=0.74 peak, indicating either an enhanced mosaic spread, or a
266: small domain size of the PdO at the surface, which would comply
267: with the previous STM observations by Hendriksen {\em et al.}
268: \cite{hendriksen03}. This poor order is probably also reflected by
269: our inability to detect reflections from the O sublattice. In
270: fact, the (0.6 0.8 0.74) peak corresponds to the (1 0 1)
271: reflection in PdO bulk coordinates. In addition to this reflection
272: we also found the (103), (200) and the (202), and equivalents when
273: rotating by 90 degrees. By observing that this indexing is similar
274: to the selection rule for a bcc lattice for which the sum of all
275: indices must be even, we conclude that what we observe is the Pd
276: sublattice in PdO, which forms a distorted bcc lattice. The
277: orientation and geometry of the ordered domains are then
278: predominantly PdO(001)$\|$Pd(100).
279: 
280: 
281: \begin{figure}[t!]
282: \scalebox{0.32}{\includegraphics{fig3.ps}} \caption{\label{fig3}
283: Left: $(T,p)$-diagram showing the measured phases in the whole
284: range of experimentally accessible conditions from UHV to ambient
285: pressure. The ``phase boundaries'' (see text) are rough estimates
286: to guide the eye. Right: Corresponding surface phase diagram, as
287: calculated by first-principles atomistic thermodynamics (see
288: text). The dashed line indicates the thermodynamic stability range
289: of the $p(2 \times 2)$ adphase, if formation of the surface oxide
290: was kinetically inhibited.}
291: \end{figure}
292: 
293: Having thus established the means to distinguish and characterize
294: the formation of either surface or thicker oxide films at higher O
295: partial pressures from the diffraction signals, we may in a
296: straightforward way construct a diagram showing which phase we
297: measure under which $(T,p)$-conditions. The result is shown in
298: Fig. \ref{fig3}, covering the whole range of now experimentally
299: accessible gas pressures from 10${}^{-6}$\,mbar to ambient
300: pressure. Most strikingly, the $(\sqrt{5} \times
301: \sqrt{5})R27^{\circ}$ structure is found under a wide variety of
302: conditions -- even at an oxygen pressure of 10${}^3$\,mbar and a
303: sample temperature of 575\,K we still observe only the formation
304: of this surface oxide and no indications for the growth of a
305: thicker oxide film on the time scale of 1\,hr currently accessible
306: to the experiment.
307: 
308: As a first step to understand this data we proceed by comparing
309: it with the surface phase diagram obtained by DFT and atomistic
310: thermodynamics. Figure \ref{fig3}b gives the corresponding
311: $(T,p)$-diagram, showing which phase would be most stable on the
312: basis of our DFT data, if the surface were in full thermodynamic
313: equilibrium with the surrounding oxygen gas phase. We immediately stress
314: that although state-of-the-art, some approximations like e.g. the
315: present treatment of vibrational contributions to the free energies,
316: as well as the uncertainty introduced by finite basis set and the
317: employed exchange-correlation functional may well allow for errors
318: in the phase boundaries of the order of $\pm 100$\,K and (depending on
319: temperature) of up to several orders in pressure \cite{reuter02,li03}.
320: 
321: Taking this into account we notice a gratifying overall agreement
322: between theory and experiment in this wide range of environmental
323: conditions. At a closer look there is, however, a notable
324: difference that is beyond the uncertainties underlying both the
325: experimental and theoretical approach: the experimental
326: observation of the $(\sqrt{5} \times \sqrt{5})R27^{\circ}$ surface
327: oxide in the top left corner of the drawn diagram, i.e. at lower
328: temperatures and high pressures. Since the central assumption of
329: theory, which predicts the stability of the bulk oxide under such
330: conditions, is the full thermodynamic equilibrium between surface
331: and gas phase, we interpret this difference as reflecting kinetic
332: limitations to the growth of the bulk oxide under such conditions.
333: This is also apparent even within the experimental data set alone.
334: If the surface was fully equilibrated with the environment, the
335: evaluated phase boundaries would have to follow lines of constant
336: oxygen chemical potential, which is the single determining
337: quantity if thermodynamics applies. In the drawn $(T,p)$-plots
338: such lines of constant chemical potential would always be parallel
339: to the phase boundaries as drawn in the theoretical diagram, cf.
340: Fig. \ref{fig3}b, which the bulk/surface oxide boundary drawn in
341: Fig. \ref{fig3}a (even considering all uncertainties) is clearly
342: not. Similarly we also understand the experimental observation of
343: the $p(2 \times 2)$ adsorbate phase, which according to theory is
344: never a thermodynamically stable phase, as a sign of kinetic
345: hindrance to the formation of the $(\sqrt{5} \times
346: \sqrt{5})R27^{\circ}$ phase. Correspondingly, we have also marked
347: in the theoretical plot the area, where the $p(2 \times 2)$ would
348: turn out more stable than the clean surface, if the surface oxide
349: can not form.
350: 
351: \begin{figure}[t!]
352: \scalebox{0.29}{\includegraphics{fig4.ps}}
353: \caption{\label{fig4} Six consecutive L-scans along (0.6 0.8 L) at
354: 675\,K and 50\,mbar. Each scan took 240 sec to record, showing the
355: gradual transformation from the diffraction signal characteristic
356: of the surface oxide to the one of the bulk oxide, cf. Fig.
357: \ref{fig2}. Solid circles: first spectrum; empty circles: last
358: spectrum taken.}
359: \end{figure}
360: 
361: Under suitable conditions we can even follow the kinetically
362: limited transition from surface to bulk oxide within the time
363: resolution of our current experiment. This is illustrated in Fig.
364: \ref{fig4} showing how the diffraction signal at 675\,K and $p =
365: 50$\,mbar slowly transforms at every consecutive scan from the one
366: characteristic for the surface oxide phase, cf. Fig. \ref{fig2}a,
367: to the one of the bulk oxide film, cf. Fig. \ref{fig2}b.
368: Obviously, this transformation will be faster for higher
369: temperatures, so that we immediately observe the bulk oxide, or
370: eventually so slow at lower temperatures that we can no longer
371: measure the transition within the time frame open to our
372: experiment. In passing we finally note that these clear kinetic
373: limitations at time scales of 1\,hr even at technologically
374: relevant temperatures as high as 675\,K, are somewhat at variance
375: with the well-known theoretical notion by King and coworkers
376: suggesting that oxide growth should immediately set in as soon as
377: it is thermodynamically possible \cite{carlisle00b}.
378: 
379: 
380: 
381: In conclusion we have studied the oxidation of the Pd(100) surface
382: from $10^{-6}$\,mbar to ambient pressure by {\em in-situ} SXRD.
383: Depending on the environmental conditions we observe either the
384: formation of the $(\sqrt{5} \times \sqrt{5})R27^{\circ}$ surface
385: oxide (essentially a well-ordered layer of PdO(101)), or the
386: growth of  $\sim 40$\,{\AA} poorly ordered and rough PdO bulk
387: oxide, predominantly with PdO(001) orientation. The range of
388: $(T,p)$-conditions where we measure the surface oxide is
389: surprisingly large, and goes for $T < 600$\,K even up to ambient
390: pressures. Comparing with the theoretical surface phase diagram
391: from first-principles atomistic thermodynamics we interpret this
392: as reflecting a kinetic hindrance to the formation of the bulk
393: oxide, which is clearly an activated process due to the involved
394: massive restructuring at the surface. Such kinetic limitations in
395: the initial oxide formation process have hitherto barely been
396: addressed, but could be crucial for understanding high pressure
397: applications like e.g. oxidation catalysis. Our measurements
398: demonstrate the usefulness of SXRD for the study of the oxidation
399: process of almost any material in such high pressure environments,
400: providing the aspired atomically resolved structural information
401: of any thin film or nano-based structure exposing its surface to
402: an ambient working atmosphere.
403: 
404: This work was financially supported by the
405: Swedish Natural Science Council and the DFG-priority
406: program ``Realkatalyse''.
407: 
408: \begin{references}
409: 
410: \bibitem{over00}
411: H. Over {\em et al.}, Science {\bf 287}, 1474 (2000).
412: 
413: \bibitem{hendriksen02}
414: B.L.M. Hendriksen and J.W.M. Frenken, Phys. Rev. Lett. {\bf 89},
415: 046101 (2002).
416: 
417: \bibitem{stampfl02}
418: C. Stampfl {\em et al.}, Surf. Sci. {\bf 500}, 368 (2002).
419: 
420: \bibitem{carlisle00}
421: C.I. Carlisle {\em et al.}, Phys. Rev. Lett. {\bf 84}, 3899
422: (2000).
423: 
424: \bibitem{lundgren02}
425: E. Lundgren {\em et. al.}, Phys. Rev. Lett. {\bf 88}, 246103
426: (2002).
427: 
428: \bibitem{todorova03}
429: M. Todorova {\em et al.}, Surf. Sci. {\bf 541},101 (2003).
430: 
431: \bibitem{reuter02}
432: K. Reuter and M. Scheffler, Phys. Rev. B {\bf 65}, 035406 (2002).
433: 
434: \bibitem{li03}
435: W.-X. Li, C. Stampfl and M. Scheffler, Phys. Rev. Lett. {\bf 90}, 256102 (2003).
436: 
437: \bibitem{ziauddin97}
438: M. Ziauddin, G. Veser and L. D. Schmidt, Catal. Lett. {\bf 46},
439: 159 (1997).
440: 
441: \bibitem{veser99}
442: G. Veser, A. Wright, and R. Caretta, Catal. Lett {\bf 58}, 199
443: (1999).
444: 
445: \bibitem{beamline}
446: http://hikwww1.fzk.de/iss/beamlinebook.html
447: 
448: \bibitem{blaha99}
449: P. Blaha, K. Schwarz and J. Luitz, {\bf WIEN97}, Techn.
450: Universit\"at Wien, Austria, (1999). ISBN 3-9501031-0-4.
451: 
452: \bibitem{perdew96}
453: J.P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett. {\bf 77},
454: 3865 (1996).
455: 
456: \bibitem{basis}
457: FP-LAPW basis set parameters: $R_{\rm{MT}}^{\rm{Pd}}=$1.8\,bohr,
458: $R_{\rm{MT}}^{\rm{O}}=$1.3\,bohr, $l_{\rm{max}}^{\rm{wf}} = 12$,
459: $l_{\rm{max}}^{\rm{pot}} = 4$, $E^{\rm max}_{\rm wf} = 20$\,Ry,
460: $E^{\rm max}_{\rm pot} = 169$\,Ry, and a $(4 \times 4 \times 1)$
461: Monkhorst-Pack grid with 28 {\bf k}-points in the full Brillouin
462: zone.
463: 
464: \bibitem{zheng00}
465: G. Zheng and E.I. Altman, Surf. Sci. {\bf 462}, 151 (2000).
466: 
467: \bibitem{rogers71}
468: D. Rogers, R. Shannon and J. Gillson, J. Solid State Chem. {\bf 3}, 314 (1971).
469: 
470: \bibitem{mcbride91}
471: J. McBride, K. Hass and W. Weber, Phys. Rev. B {\bf 44}, 5016 (1991).
472: 
473: \bibitem{hendriksen03}
474: B.L.M. Hendriksen {\em et al.}, Surf. Sci. ({\em submitted}).
475: %, {\em Model Catalysts in Action: High-Pressure Scanning
476: %Tunneling Microscopy}, Ph.D. thesis, Universiteit Leiden (2003).
477: 
478: \bibitem{robinson88}
479: I.K. Robinson, Phys. Rev. B {\bf 33}, 3830 (1986).
480: 
481: \bibitem{carlisle00b}
482: C.I. Carlisle {\em et al.}, Surf. Sci. {\bf 470}, 15 (2000), and
483: references therein.
484: 
485: 
486: \end{references}
487: \end{document}
488: