Pt–Zn Clusters on Stoichiometric MgO(100) and TiO2(110

Mar 2, 2015 - The effect of temperature is included via both the Metropolis algorithm and the Boltzmann-weighted populations of the global and thermal...
0 downloads 0 Views 541KB Size
Article pubs.acs.org/JPCC

Pt−Zn Clusters on Stoichiometric MgO(100) and TiO2(110): Dramatically Different Sintering Behavior Jonny Dadras,† Lu Shen,† and Anastassia Alexandrova*,†,‡ †

Department of Chemistry & Biochemistry, University of California, Los Angeles, California 90095, United States California NanoSystems Institute, Los Angeles, California 90095, United States



S Supporting Information *

ABSTRACT: Zn was suggested to be a promising additive to Pt in the catalysis of dehydrogenation reactions. In this work, mixed Pt− Zn clusters deposited on two simple oxides, MgO(100) and TiO2(110), were investigated. The stability of these systems against cluster sintering, one of the major mechanisms of catalyst deactivation, is simulated using a Metropolis Monte Carlo scheme under the assumption of the Ostwald ripening mechanism. Particle migration, association to and dissociation from clusters, and evaporation and redeposition of monomers were all included in the simulations. Simulations are done at several high temperatures relevant to reactions of catalytic dehydrogenation. The effect of temperature is included via both the Metropolis algorithm and the Boltzmann-weighted populations of the global and thermally accessible local minima on the density functional theory potential energy surfaces of clusters of all sizes and compositions up to tetramers. On both surfaces, clusters are shown to sinter quite rapidly. However, the resultant compositions of the clusters most resistant to sintering are quite different on the two supports. On TiO2(110), Pt and Zn appear to phase separate, preferentially forming clusters rich in just one or the other metal. On MgO(100), Pt and Zn remain well-mixed and form a range of bimetallic clusters of various compositions that appear relatively stable. However, Zn is more easily lost from MgO through evaporation. These phenomena were rationalized by several means of chemical bonding analysis.



INTRODUCTION Platinum-based nano and subnanoclusters show great catalytic activity and selectivity for alkane dehydrogenation and cracking.1−4 Catalytic properties of such nanoscale clusters can be superb but strongly depend on cluster size, composition, and the type of support.5−16 Further dehydrogenation to alkynes can occur, as well as methane formation from cracking, and both reactions may result in coke deposits. Coking along with particle sintering are the two main causes for supportedcatalyst deactivation. A popular approach to mitigating coke involves alloying Pt with main group metals and some transition metals.17−22 Also, Gutierrez et al. have shown that coke-induced deactivation of certain Pt-based bimetallic clusters can be reduced by controlling the acidity of the support.23 Sintering of metal nanoparticles is a result of the particles minimizing their surface energy. This leads to populations of small-sized particles “dying off”, directly leading to an increase in the population of large-sized particles. Some of the authors (Dadras and Alexandrova) have previously shown that pure Pd and mixed Pt−Pd clusters deposited on TiO2 readily sinter, by the Ostwald ripening mechanism.24,25 Indeed, Ostwald ripening rather than particle coalescence is expected to be the dominant mode of sintering for Pt-based clusters, at least when supported on TiO2.26 Given that cluster activity and selectivity depend on the number of available edge and vertex atoms, rapid sintering © XXXX American Chemical Society

decreases the catalysts’ useful lifetime. Quantifying supportdependent effects on cluster stability and selectivity becomes a matter of central importance for catalysis. The present work attempts to address the effect of the support on cluster stability, through a fundamental understanding of electronic structure. The supports often play the role of a ligand to tiny (1000 K) this latter effect would lead to the appearance of large nanoparticles that are more Pt-rich than the model predicts. Again, what is more surprising is the predicted “bimodal” distributions on TiO2. For that system, the noted limitations would not have as great an effect based on the aforementioned previous results. By applying reasonable physical-chemical assumptions, that have proven appropriate in prior simulations, clear predictions for the presented systems have been made. These results are in need of experimental validation or falsification. More generally, it is noted that metal−metal covalent bonding between the atoms in the support and the atoms in the deposited clusters may play a significant role in modulating the shapes and properties of deposited clusters. This adds to the already rich and diverse roles that the support can play in impacting the catalytic properties of small surface-mounted clusters. In view of the interest in Pt−Zn clusters for catalytic dehydrogenation, our results suggest the promise to not be very high because of Zn loss via evaporation and, on TiO2, Zn separation from Pt.



The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This material is based in part on work supported by the Air Force Office of Scientific Research (AFOSR) under grant number 10029173-S3 and NSF CAREER Award CHE1351968. Computational resources were provided by the UCLA-IDRE cluster. A portion of the research was performed using EMSL, a national scientific user facility sponsored by the Department of Energy’s Office of Biological and Environmental Research and located at Pacific Northwest National Laboratory.



(1) Bhasin, M.; McCain, J.; Vora, B.; Imai, T.; Pujado, P. Dehydrogenation and oxydehydrogenation of paraffins to olefins. Appl. Catal. A: Gen. 2001, 221, 397−419. (2) Galvita, V.; Siddiqi, G.; Sun, P.; Bell, A. Ethane dehydrogenation on Pt/Mg(Al)O and PtSn/Mg(Al)O catalysts. J. Catal. 2010, 271, 209−219. (3) Burch, R.; Garla, L. Platinum-tin reforming catalysts: II. Activity and selectivity in hydrocarbon reactions. J. Catal. 1981, 71, 360−372. (4) Weckhuysen, B.; Schoonheydt, R. Alkane dehydrogenation over supported chromium oxide catalysts. Catal. Today 1999, 51, 223−232. (5) Guisbiers, G.; Abudukelimu, G.; Hourlier, D. Size-dependent catalytic and melting properties of platinum-palladium nanoparticles. Nano Res. Lett. 2011, 6, 396. (6) Cuenya, B. R. Synthesis and catalytic properties of metal nanoparticles: Size, shape, support, composition, and oxidation state effects. Thin Solid Films 2010, 518, 3127−3150. (7) Yamazoe, S.; Koyasu, K.; Tsukuda, T. Nonscalable Oxidation Catalysis of Gold Clusters. Acc. Chem. Res. 2014, 47, 816−824. (8) Kaden, W. E.; Wu, T. P.; Kunkel, W. A.; Anderson, S. L. Electronic Structure Controls Reactivity of Size-Selected Pd Clusters Adsorbed on TiO2 Surfaces. Science 2009, 326, 826−829. (9) Lee, S. S.; Fan, C. Y.; Wu, T. P.; Anderson, S. L. CO Oxidation on Aun/TiO2 Catalysts Produced by Size-Selected Cluster Deposition. J. Am. Chem. Soc. 2004, 126, 5682−5683. (10) Kaden, W. E.; Kunkel, W. A.; Kane, M. D.; Roberts, F. S.; Anderson, S. L. Size-Dependent Oxygen Activation Efficiency over Pdn/TiO2(110) for the CO Oxidation Reaction. J. Am. Chem. Soc. 2010, 132, 13097−13099. (11) Feng, L.; Hoang, D. T.; Tsung, C. K.; Huang, W. Y.; Lo, S. H. Y.; Wood, J. B.; Wang, H. T.; Tang, J. Y.; Yang, P. D. Catalytic properties of Pt cluster-decorated CeO2 nanostructures. Nano Res. 2011, 4, 61−71. (12) Alexeev, O. S.; Graham, G. W.; Shelef, M.; Gates, B. C. γAl2O3-Supported Pt Catalysts with Extremely High Dispersions Resulting from Pt−W Interactions. J. Catal. 2000, 190, 157−172. (13) Cuenya, B. R.; Baeck, S. H.; Jaramillo, T. F.; McFarland, E. W. Size- and Support-Dependent Electronic and Catalytic Properties of Au0/Au3+ Nanoparticles Synthesized from Block Copolymer Micelles. J. Am. Chem. Soc. 2003, 125, 12928−12934. (14) Kulkarni, A.; Lobo-Lapidus, R. J.; Gates, B. C. Metal clusters on supports: Synthesis, structure, reactivity, and catalytic properties. Chem. Commun. 2010, 46, 5997−6015. (15) Zhang, J.; Alexandrova, A. N. The Golden Crown: A Single Au Atom that Boosts the CO Oxidation Catalyzed by a Palladium Cluster on Titania Surfaces. J. Phys. Chem. Lett. 2013, 4, 2250−2255. (16) Zhang, J.; Alexandrova, A. N. Double σ-Aromaticity in a SurfaceDeposited Cluster: Pd4 on TiO2 (110). J. Phys. Chem. Lett. 2012, 3, 751−754.

ASSOCIATED CONTENT

S Supporting Information *

Global and low-energy local isomers in the gas phase, calculated using a number of ab initio methods, plotted for the empirical scaling of the Zn chemical potential used in GCMC simulations. PDOS of Pt and/or Zn for pure and equimolar Pt−Zn clusters. Intracluster Coulomb potentials for tetramers. This material is available free of charge via the Internet at http://pubs.acs.org.



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. G

DOI: 10.1021/jp512277x J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C (17) Bariås, O.; Holmen, A.; Blekkan, E. Propane Dehydrogenation over Supported Pt and Pt−Sn Catalysts: Catalyst Preparation, Characterization, and Activity Measurements. J. Catal. 1996, 158, 1− 12. (18) Jablonski, E.; Castro, A.; Scelza, O.; de Miguel, S. Effect of Ga addition to Pt/Al2O3 on the activity, selectivity and deactivation in the propane dehydrogenation. Appl. Catal. A: Gen. 1999, 183, 189−198. (19) Sun, P.; Siddiqi, G.; Vining, W.; Chi, M.; Bell, A. Novel Pt/ Mg(In)(Al)O catalysts for ethane and propane dehydrogenation. J. Catal. 2011, 282, 165−174. (20) Siddiqi, G.; Sun, P.; Galvita, V.; Bell, A. Catalyst performance of novel Pt/Mg(Ga)(Al)O catalysts for alkane dehydrogenation. J. Catal. 2010, 274, 200−206. (21) Ren, H.; Humbert, M.; Menning, C.; Chen, J.; Shu, Y.; Singh, U.; Cheng, W. Catalyst performance of novel Pt/Mg(Ga)(Al)O catalysts for alkane dehydrogenation. Appl. Catal. A: Gen. 2010, 375, 303−309. (22) Jovanovic, M.; Putanov, P. Nature and distribution of coke formed on mono-metallic platinum and bimetallic platinum-rhenium catalysts. Appl. Catal. A: Gen. 1997, 159, 1−7. (23) Gutiérrez, A.; Arandes, J.; Castaño, P.; Aguayo, A.; Bilbao, J. Role of Acidity in the Deactivation and Steady Hydroconversion of Light Cycle Oil on Noble Metal Supported Catalysts. Energy Fuels 2011, 25, 3389−3399. (24) Zhang, J.; Alexandrova, A. Structure, stability, and mobility of small Pd clusters on the stoichiometric and defective TiO2 (110) surfaces. J. Chem. Phys. 2011, 135, 174702. (25) Ha, M.-A.; Dadras, J.; Alexandrova, A. N. Rutile-Deposited Pt− Pd clusters: A Hypothesis Regarding the Stability at 50/50 Ratio. ACS Catal. 2014, 4, 3570−3580. (26) Hansen, T.; DeLaRiva, A.; Challa, S.; Datye, A. Sintering of Catalytic Nanoparticles: Particle Migration or Ostwald Ripening? Acc. Chem. Res. 2013, 46, 1720−1730. (27) Henrich, V.; Cox, P. The Surface Science of Metal Oxides; Cambridge University Press: Cambridge, England, 1996. (28) Henry, C. R. Surface studies of supported model catalysts. Surf. Sci. Rep. 1998, 31, 231−325. (29) Diebold, U. The surface science of titanium dioxide. Surf. Sci. Rep. 2003, 48, 53−229. (30) Goellner, J. F.; Neyman, K. M.; Mayer, M.; Nortemann, F.; Gates, B. C.; Rosch, N. Ligand-Free Osmium Clusters Supported on MgO. A Density Functional Study. Langmuir 2000, 16, 2736−2743. (31) Campbell, C. T. The Energetics of Supported Metal Nanoparticles: Relationships to Sintering Rates and Catalytic Activity. Acc. Chem. Res. 2013, 46, 1712−1719. (32) Wang, Y. G.; Yoon, Y.; Glezakou, V. A.; Li, J.; Rousseau, R. The Role of Reducible Oxide−Metal Cluster Charge Transfer in Catalytic Processes: New Insights on the Catalytic Mechanism of CO Oxidation on Au/TiO2 from ab Initio Molecular Dynamics. J. Am. Chem. Soc. 2013, 135, 10673−10683. (33) Yoon, B.; Hakkinen, H.; Landman, U.; Worz, A. S.; Antonietti, J. M.; Abbet, S.; Judai, K.; Heiz, U. Charging Effects on Bonding and Catalyzed Oxidation of CO on Au8 Clusters on MgO. Science 2005, 307, 403−407. (34) De Cola, P. L.; Glaser, R.; Weitkamp, J. Erratum to “Nonoxidative propane dehydrogenation over Pt−Zn-containing zeolites” [Appl. Catal. A: Gen. 306 (2006) 85−97]. Appl. Catal. A: Gen. 2006, 310, 205−206. (35) Shen, L.; Dadras, J.; Alexandrova, A. N. Pure and Zn-doped Pt clusters go flat and upright on MgO(100). Phys. Chem. Chem. Phys. 2014, 16, 264366−26442. (36) Perdew, J.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865−3868. (37) Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials. J. Phys.: Condens. Matter 2009, 21, 395502.

(38) Kohn, W.; Sham, L. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133− A1138. (39) Vanderbilt, D. Soft self-consistent pseudopotentials in a generalized eigenvalue formalism. Phys. Rev. B 1990, 41, 7892−5. (40) Burke, K.; Werschnik, J.; Gross, E. Time-dependent density functional theory: Past, present, and future. J. Chem. Phys. 2005, 123, 062206. (41) Kowalski, P.; Meyer, B.; Marx, D. Composition, structure, and stability of the rutile TiO2(110) surface: Oxygen depletion, hydroxylation, hydrogen migration, and water adsorption. Phys. Rev. B 2009, 79, 115410. (42) Kowalski, P.; Camellone, M.; Nair, N.; Meyer, B.; Marx, D. Charge Localization Dynamics Induced by Oxygen Vacancies on the TiO2(110) Surface. Phys. Rev. Lett. 2010, 105, 146405. (43) Becke, A. D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648−52. (44) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785−89. (45) Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results obtained with the correlation energy density functionals of Becke and Lee, Yang and Parr. Chem. Phys. Lett. 1989, 157, 200−06. (46) Peterson, K.; Puzzarini, C. Systematically convergent basis sets for transition metals. II. Pseudopotential-based correlation consistent basis sets for the group 11 (Cu, Ag, Au) and 12 (Zn, Cd, Hg) elements. Theor. Chem. Acc. 2005, 114, 283−296. (47) Figgen, D.; Rauhut, G.; Dolg, M.; Stoll, H. Energy-consistent pseudopotentials for group 11 and 12 atoms: adjustment to multiconfiguration Dirac−Hartree−Fock data. Chem. Phys. 2005, 311, 227− 244. (48) Dunning, T. H.; Hay, P. J. Gaussian Basis Sets for Molecular Calculations. Modern Theoretical Chemistry; Schaefer, H. F., III, Ed.; Plenum: New York, 1977; Vol. 3, pp 1−28. (49) Hay, P.; Wadt, W. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270−283. (50) Wadt, W.; Hay, P. Ab initio effective core potentials for molecular calculations. Potentials for main group elements Na to Bi. J. Chem. Phys. 1985, 82, 284−298. (51) Hay, P.; Wadt, W. Ab initio effective core potentials for molecular calculations. Potentials for K to Au including the outermost core orbitals. J. Chem. Phys. 1985, 82, 299−310. (52) Hegarty, D.; Robb, M. Application of unitary group methods to configuration interaction calculations. Mol. Phys. 1979, 38, 1795− 1812. (53) Eade, R.; Robb, M. Direct minimization in mc scf theory. the quasi-newton method. Chem. Phys. Lett. 1981, 83, 362−368. (54) Schlegel, H.; Robb, M. MC SCF gradient optimization of the H2CO→H2 + CO transition structure. Chem. Phys. Lett. 1982, 93, 43− 46. (55) Bernardi, F.; Bottoni, A.; Mcdouall, J.; Robb, M.; Schlegel, H. MCSCF gradient calculation of transition structures in organic reactions. Faraday Symp. Chem. Soc. 1984, 19, 137−147. (56) Frisch, M.; Ragazos, I.; Robb, M.; Schlegel, H. An evaluation of three direct MC-SCF procedures. Chem. Phys. Lett. 1992, 189, 524− 528. (57) Yamamoto, N.; Vreven, T.; Robb, M.; Frisch, M.; Schlegel, H. A direct derivative MC-SCF procedure. Chem. Phys. Lett. 1996, 250, 373−378. (58) M?ller, C.; Plesset, M. S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 0618−0622. (59) Headgordon, M.; Pople, J. A.; Frisch, M. J. MP2 energy evaluation by direct methods. Chem. Phys. Lett. 1988, 153, 503−506. (60) Saebo, S.; Almlof, J. Avoiding the integral storage bottleneck in LCAO calculations of electron correlation. Chem. Phys. Lett. 1989, 154, 83−89. (61) Frisch, M. J.; Head-Gordon, M.; Pople, J. A. A direct MP2 gradient method. Chem. Phys. Lett. 1990, 166, 275−280. H

DOI: 10.1021/jp512277x J. Phys. Chem. C XXXX, XXX, XXX−XXX

Article

The Journal of Physical Chemistry C (62) Frisch, M. J.; Head-Gordon, M.; Pople, J. A. Semi-direct algorithms for the MP2 energy and gradient. Chem. Phys. Lett. 1990, 166, 281−289. (63) Head-Gordon, M.; Head-Gordon, T. Analytic MP2 frequencies without fifth-order storage. Theory and application to bifurcated hydrogen bonds in the water hexamer. Chem. Phys. Lett. 1994, 220, 122−128. (64) Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.; et al., Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford CT, 2009. (65) Tang, W.; Sanville, E.; Henkelman, G. A grid-based Bader analysis algorithm without lattice bias. J. Phys.: Condens. Matter 2009, 21, 084204. (66) Sanville, E.; Kenny, S.; Smith, R.; Henkelman, G. Improved gridbased algorithm for Bader charge allocation. J. Comput. Chem. 2007, 28, 899−908. (67) Henkelman, G.; Arnaldsson, A.; Jonsson, H. A fast and robust algorithm for Bader decomposition of charge density. Comput. Mater. Sci. 2006, 36, 354−360. (68) Glendening, E.; Reed, A.; Carpenter, J.; Weinhold, F. NBO Version 3.1, 1998. (69) Metropolis, N.; Rosenbluth, A.; Rosenbluth, M.; Teller, A.; Teller, E. Equation of State Calculations by Fast Computing Machines. J. Chem. Phys. 1953, 21, 1087−1092. (70) Hastings, W. Monte Carlo sampling methods using Markov chains and their applications. Biometrika 1970, 57, 97−109. (71) Wei, S.; Zeng, Z.; You, J.; Yan, X.; Gong, X. A density-functional study of small titanium clusters. J. Chem. Phys. 2000, 113, 11127− 11133. (72) Addou, R.; Senftle, T. P.; O’Connor, N.; Janik, M. J.; van Duin, A. C. T.; Batzill, M. Influence of Hydroxyls on Pd Atom Mobility and Clustering on Rutile TiO2(011)-2 × 1. ACS Nano 2014, 8, 6321− 6333. (73) Adams, D. J. Grand canonical ensemble Monte Carlo for a Lennard-Jones fluid. Mol. Phys. 1975, 29, 307−311. (74) Wolf, R. J.; Lee, M. W.; Davis, R. C.; Fay, P. J.; Ray, J. R. Pressure-composition isotherms for palladium hydride. Phys. Rev. B 1993, 48, 12415. (75) Lachet, V.; Boutin, A.; Tavitian, B.; Fuchs, A. H. Grand canonical Monte Carlo simulations of adsorption of mixtures of xylene molecules in faujasite zeolites. Faraday Discuss. 1997, 106, 307−323. (76) Senftle, T. P.; Meyer, R. J.; Janik, M. J.; van Duin, A. C. T. Development of a ReaxFF potential for Pd/O and application to palladium oxide formation. J. Chem. Phys. 2013, 139, 044109. (77) Chase, M. W. National Institute of Standards and Technology. NIST-JANAF Thermochemical Tables; American Chemical Society: Washington, D.C., 1998. (78) Florez, E.; Fuentealba, P.; Mondragon, F. Chemical reactivity of oxygen vacancies on the MgO surface: Reactions with CO2, NO2 and metals. Catal. Today 2008, 133, 216−222.

I

DOI: 10.1021/jp512277x J. Phys. Chem. C XXXX, XXX, XXX−XXX