First principles study of benzene adsorption on transition metal surfaces

142 511 0
First principles study of benzene adsorption on transition metal surfaces

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

FIRST PRINCIPLES STUDY OF BENZENE ADSORPTION ON TRANSITION METAL SURFACES HONG WON KEON NATIONAL UNIVERSITY OF SINGAPORE 2008 FIRST PRINCIPLES STUDY OF BENZENE ADSORPTION ON TRANSITION METAL SURFACES HONG WON KEON (B. Eng.(Hons.), Sungkyunkwan University, Republic of Korea) A THESIS SUBMITTED FOR THE DEGREE OF MASTER OF ENGINEERING DEPARTMENT OF CHEMICAL & BIOMELEULAR ENGINEERING NATIONAL UNIVERSITY OF SINGAPORE 2008 ACKNOWLEDGEMENTS I would like to express my sincere appreciation to my supervisor, Prof. Mark Saeys, for his encouragement, insight, support and incessant guidance throughout the course of this research project. I am extremely grateful to him for spending so much time on explaining my questions on the research work and sharing his broad and profound knowledge with me. I also feel thankful to his high integrity and dedication in the scientific research, which have greatly inspired me. I am very thankful to Ms Khoh Leng Khim, Mdm. Jamie Siew, Mr. Mao Ning, Mr. Chia Phai Ann, Dr. Yuan Ze Liang, Ms Lee Chai Keng, Mdm Sam Fam Hwee Koong, Ms Tay Choon Yen, and Shang Zhenhua for their technical and kind support. I would sincerely like to thank our group members Xu Jing, Sun Wenjie, Tan Kong Fei, Zhuo Mingkun, Fan Xuexiang, Su Mingjuan and Ravi Kumar Tiwari for many useful discussions and their help in carrying out my research work in the lab. I also thank all my friends both in Singapore and abroad, who have enriched my life personally and professionally. Finally, special thanks must go to my family for their kind understanding, encouragement, and support during my pursuit of M. Eng. degree. i TABLE of CONTENTS ACKNOWLEDGEMENTS ···························································································· i TABLE OF CONTENTS ······························································································· ii SUMMARY ···················································································································· v SYMBOLS AND ABBREVIATIONS········································································· vii LIST OF TABLES ········································································································· ix LIST OF FIGURES········································································································ xi CHAPTER 1 Introduction ·······························································································1 1.1 First Principles-based Modeling ···········································································1 1.2 Principles in Heterogeneous Catalysis···································································3 1.3 Application of First Principles-based Modeling ····················································7 1.4 Electronic Interaction in Heterogeneous Catalysis ··············································10 1.4.1 Surface-adsorbate Interaction·········································································11 1.4.1.1 CO Interaction with the Pt(111) Surface ················································12 1.4.1.2 C2H4 Interaction with the Pd(111) Surface ·············································13 1.4.2 Adsorbate-adsorbate Interaction ····································································15 1.4.3 Electronic Interaction on Benzene Adsorption on Pt(111) ····························16 1.5 Organization of the Thesis ·················································································20 1.6 References···········································································································21 CHAPTER 2 Computational Method ·············································································23 2.1 First Principles Quantum Chemical Methods ······················································23 2.1.1 Time Independent Schrödinger Equation ·······················································23 2.1.2 Hartree-Fock Approximation·········································································25 2.1.3 Electron Correlation Methods········································································27 2.2.3.1 Configurational Interaction·····································································28 2.2.3.2 Møller-Plesset Perturbation Theory ························································29 2.2.3.3 Coupled-Cluster Theory ·········································································30 ii 2.1.4 Density Functional Theory (DFT) ································································31 2.1.5 Exchange-Correlation (XC) Functionals ·······················································33 2.2.5.1 Local Density Approximation (LDA) ····················································33 2.2.5.2 Generalized Gradient Approximation (GGA) ········································34 2.1.6 Basis Sets ······································································································36 2.1.7 Plane Wave Basis Sets ··················································································40 2.1.8 Pseudopotentials ···························································································41 2.2.8.1 Norm-Conserving (NC) Pseudopotentials ··············································43 2.2.8.2 Ultra-Soft (US) Pseudopotentials ···························································43 2.2.8.3 Projector Augmented Wave (PAW) Method ·········································45 2.2 Computational Codes ·························································································47 2.2.1 Vienna Ab-initio Simulation Package (VASP) ·············································47 2.2.2 Gaussian03 (G03) ··························································································47 2.3 References ··········································································································49 CHAPTER 3 Benzene Chemisorption on Pt(111) ·························································51 3.1 Converged DFT Benzene Adsorption Energy on Pt(111) ··································52 3.1.1 Review of the Literature: Benzene Adsorption Studies on Pt(111) ···············54 3.1.2 Convergence Test for DFT Benzene Adsorption Energy on Pt(111) ············59 3.1.2.1 Convergence Test on Vacuum Thickness ··············································63 3.1.2.2 Convergence Test on the number of k-points and the Slab Thickness ···63 3.1.3 Electronic Analysis of DFT Benzene Adsorption Energy on Pt(111) ··········65 3.1.4 Summary ······································································································73 3.2 Accuracy of DFT Adsorption Energy on Pt(111) ··············································74 3.2.1 Review of Computational Studies of CO Adsorption on Pt(111) ················75 3.2.2 Exchange-Correlation (XC) Correction Approach ·······································81 3.2.2.1 DFT Adsorption Energy in a Periodic Slab Calculation ·························82 3.2.2.2 DFT-PBE Adsorption Energy on a Small Cluster in VASP ···················82 3.2.2.3 DFT-PBE Adsorption Energy on a Small Cluster in G03 ······················85 3.2.2.4 DFT-B3LYP Adsorption Energy on a Small Cluster in G03 ·················92 3.2.2.5 Adsorption Energy on a Small Cluster validated with correlated wavefunction based methods ·············································································94 3.2.2.6 Comparison with wave function based methods: binding energy of di-σ and π ethylene on a small Pt2 cluster ········································································97 3.2.3 Electronic Structure based Correction Approach ·········································98 3.2.3.1 HOMO-LUMO gap of Gas Phase Molecules ········································99 iii 3.3 Coverage Effects on the Benzene Adsorption Energy and Site Preference on Pt(111) ···················································································································105 3.3.1 Review of the Experimental and Theoretical Literature on the Coverage Effects on Benzene Adsorption on Pt(111) ····································································105 3.3.1.1 Preferential Adsorption Mode for Benzene at High Coverage ············105 3.3.1.2 Chemisorbed Benzene Structure as a Function of Coverage ···············106 3.3.1.3 Benzene Adsorption Energy as a Function of Coverage ·····················107 3.3.2 DFT Study of the Coverage Effect on the Benzene Adsorption Energy ·····108 3.3.2.1 Preferred Benzene Adsorption Sites at High Coverage ·······················109 3.3.2.2 Coverage Effect on the Benzene Adsorption Energy ··························111 3.3.2.3 Coverage Effect on the Benzene Adsorption Structure ······················113 3.4 References········································································································115 CHAPTER 4 Conclusion ·····························································································120 APPENDIX A Conclusion DFT-PBE Calculation Data ··············································126 iv SUMMARY The chemisorption of aromatic molecules on transition metal catalysts is a key step in catalytic processes for the production of fuels and petrochemicals, as well as in the removal of aromatics from exhaust gases. In this work, state-of-the-art molecular modeling is used to theoretically investigate the adsorption of benzene on a model Pt(111) surface. First, numerically converged, low coverage benzene adsorption energies of -107 kJ/mol for the bridge(30) site and -71 kJ/mol for the hollow(0) site of the Pt(111) surface were determined at the Density Functional Theory-Perdew Burke Enzerhoff (DFT-PBE) level of theory using periodic slab calculations as implemented in the Vienna Ab initio Simulation Package (VASP). The calculations indicate that a 5-layer Pt slab is required to accurately describe the surface electronic structure. The commonly used 3- and 4-layer slabs show 8 ~ 30 % of deficiency in the description of the surface d-band and hence do not accurately describe the adsorption. To avoid interaction between periodically repeated slabs, a 14 Å vacuum layer was found to be sufficient. Second, the accuracy of the DFT-PBE description of the interaction between benzene and Pt was investigated. Though DFT-PBE was found to accurately describe the electronic structure of the Pt(111) surface as well as the ionization potential of benzene, it significantly underestimates the electron affinity of benzene by 0.82 eV. As a result, DFTPBE significantly underestimates the interaction between the benzene LUMO and the Pt d-band, and hence does not accurately describe the effect of electron back-donation. v Correlated wave-function-based methods such as MP2 and CCSD(T) were used to begin to correct this problem, and these methods indeed predict a -126 kJ/mol and -134 kJ/mol stronger adsorption of benzene at the hollow site of a small Pt3 cluster. Unfortunately, calculating numerically converged benzene adsorption energies at the MP2 and CCSD(T) level of theory is beyond current computational capabilities since basis set requirements increase exponentially with the number of electrons for correlated wave-function based methods (Duch and Diercksen, 1994). Combining our best estimates at the MP2 and CCSD(T) level of theory, the adsorption energy is predicted about 60 kJ/mol stronger than the value predicted by DFT-PBE. Similarly the reliability of DFT-PBE for the adsorption energies of methyl, CO, ethene, and 1,3-butadiene on Pt was evaluated. Based on the predicted position of the HOMO and the LUMO, it can be expected that DFT-PBE gives a fairly accurate description of methyl and 1,3-butadiene adsorption on Pt(111), while DFT-B3LYP is expected to be more accurate for CO, in agreement with benchmark studies in the literature. However, for ethene and benzene, both DFT-PBE and DFT-B3LYP significantly overestimate the HOMO-LUMO gap by about 1 eV and are hence expected to underestimate adsorption energies. Finally, to elucidate the experimentally observed change in the preferred adsorption site at higher coverages, the adsorption of benzene was studied for coverages of 1/9, 1/7 and 1/6 monolayer. The latter coverage corresponds to the experimentally observed saturation coverage. DFT-PBE calculations did not predict a change in the preferred adsorption site. vi SYMBOLS AND ABBREVIATIONS Symbols E Total electronic energy EF Fermi energy ML Monolayer a0 Bulk lattice constant in equilibrium eV Electron volt fcc Face-centered cubic hcp Hexagonal closed-packed ΔEads Adsorption energy ΔHads Heat of Adsorption ΔΦ Change in workfunction of metal surface upon adsorption Φ Workfunction of metal surface θ Surface coverage in computational methods θexp Surface coverage in experiment Abbreviations ARPEFS Angle-Resolved Photoemission Extended Fine Structure ARUPS Angle Resolved Ultraviolet Photoelectron Spectroscopy B3LYP Becke, three-parameter, Lee-Yang-Parr XC functional BSSE Basis Set Superposition Error CBS Complete Basis Set CCSD(T) coupled cluster theory with single and double excitations and a quasi-perturbative treatment of triple excitations DFT Density Functional Theory DOS Density of States ECP Effective Core Potential EELS Electron Energy Loss Spectroscopy vii GGA Generalized Gradient Approximation G03 Gaussian 03 HF Hartree-Fock method HOMO Highest Occupied Molecular Orbital HREELS High Resolution Electron Energy Loss Spectroscopy LCAO Linear Combination of Atomic Orbital LDA Local Density Approximation LEED Low Energy Electron Diffraction LUMO Lowest Unoccupied Molecular Orbital MP2 Møller-Plesset perturbation theory of the second order NEXAFS Near Edge X-ray Adsorption Fine Structure PAW Projector Augmented Wave method PBE Perdew Burke Enzerhoff exchange-correlation functional PDOS Projected Density of States PW91 Perdew–Wang 1991 exchange-correlation functional RAIRS Reflection Absorption IR Spectroscopy SERS Surface-Enhanced Raman Spectroscopy SCAC Single Crystal Adsorption Calorimetry SCF Self-Consistent Field method STM Scanning Tunneling Microscopy TDS Temperature Desorption Spectroscopy TPD Temperature Programmed Desorption VASP Vienna Ab initio Simulation Package US Ultra-Soft pseudopotential XC Exchange-Correlation viii LIST of TABLES Table 3.1 Summary of the experimental studies of the benzene adsorption sites on Pt(111) ·······················································································································55 Table 3.2 Adsorption energies at various benzene adsorption sites on Pt(111) in the literature ·········································································································56 Table 3.3 Benzene adsorption energy on Pt(111) at 1/9 ML by DFT periodic slab calculation using VASP ·················································································58 Table 3.4 Surface relaxation upon benzene adsorption for various Pt(111) models ·······61 Table 3.5 Convergence test with various vacuum thickness for benzene adsorption energy on Pt(111) at the Bridge(30) site ····································································63 Table 3.6 Slab thickness convergence test for benzene adsorption energy on Pt(111) at Bridge(30) site along with k-point convergence test ······································64 Table 3.7 Surface electronic properties of various Pt(111) slabs ···································68 Table 3.8 Surface electronic properties of various Pt(111) slabs with benzene ·············68 Table 3.9 Changes of surface electronic properties of Pt(111) upon adsorption ············68 Table 3.10 DFT adsorption energy results for molecules on Pt(111) slab ·····················82 Table 3.11 Adsorption energy results for molecules on Pt3 cluster in VASP··················84 Table 3.12 Details of the basis sets used in the molecular calculations ··························86 Table 3.13 Details of the valence basis sets and effective core potentials for the Pt atoms ······················································································································86 Table 3.14 Comparison of the DFT-PBE binding energies on a Pt3 cluster with different basis sets in Gaussian03 ···············································································89 Table 3.15 Comparison of the DFT-PBE binding energies on a Pt3 cluster with different valence basis sets for Pt in Gaussian03 ························································90 Table 3.16 Number of basis functions used for each calculation ···································91 Table 3.17 Comparison of DFT-B3LYP and DFT-PBE binding energies on a Pt3 cluster in Gaussian03 ··································································································93 ix Table 3.18 Benzene binding energies on Pt3 cluster with correlated wavefunction based methods in Gaussian03 ················································································96 Table 3.19 Ethylene binding energies on Pt2 cluster with DFT-PBE and correlated wavefunction based methods in Gaussian03···················································97 Table 3.20 Calculated ionization potential for molecules in the gas-phase···················100 Table 3.21 Calculated electron affinity for molecules in the gas-phase ·······················100 Table 3.22 Comparison of gap between ionization potential and electron affinity for molecules in the gas-phase············································································100 Table 3.23 The electronic interaction strength parameters related to the HOMO and LUMO of molecules compared to the Fermi energy of Pt(111) ·················104 Table 3.24 Benzene adsorption energy results on Pt(111) at low coverage ················112 Table 3.25 Benzene adsorption geometry at the Bridge(30) of Pt(111) at various coverage ······················································································································114 Table 3.26 Benzene adsorption geometry at the Hollow-hcp(0) of Pt(111) slab at various coverage ····································································································114 Table A.1 DFT-PBE total energy calculation results for convergence test with various vacuum thickness for benzene adsorption energy on Pt(111) at the Bridge(30) site ···············································································································126 Table A.2 DFT-PBE total energy calculation results for slab thickness convergence test for benzene adsorption energy on Pt(111) at Bridge(30) site along with k-point convergence test at low coverage of 1/9 ML ················································126 Table A.3 DFT-PBE total energy calculation results for molecular adsorption energy on Pt3 cluster in VASP ····················································································127 Table A.4 DFT-PBE total energy calculation results for slab thickness convergence test for benzene adsorption energy on Pt(111) at Bridge(30) site along with k-point convergence test at moderate coverage of 1/7 ML ·······································127 Table A.5 DFT-PBE total energy calculation results for slab thickness convergence test for benzene adsorption energy on Pt(111) along with k-point convergence test at moderate coverage of 1/6 ML ······································································127 x LIST of FIGURES Figure 1.1 Schematic illustration of a simple heterogeneous catalytic reaction ··············3 Figure 1.2 Schematic illustration of low index surface of the fcc-structured heterogeneous catalyst: the fcc(100) is in the right, the fcc(11) in the middle, and the fcc(111) in the left ···········································································5 Figure 1.3 The calculated potential energy diagram for NH3 synthesis from N2 and H2 over closed-packed and stepped Ru surface (Honkala et al, 2005) ···············8 Figure 1.4 The local density of states at an adsorbate in two limiting cases: (a) for a broad surface band relevant to the interaction with a metal s band; (b) for a narrow metal band representing the interaction with a transition metal d band (Hammer and Nørskov, 2000) ····································································12 Figure 1.5 The local density of states projected onto an adsorbate state interacting with the d bands at a surface (Hammer and Nørskov, 2000) ·······························12 Figure 1.6 The self-consistent electronic DOS projected onto the 5σ and 2π* orbitals of CO: in vacuum and on Al(111) and Pt(111) surface (Hammer et al,1996)··13 Figure 1.7 Frontier orbital interaction in the di-σ adsorption of ethylene on Pd(111) (Pallassana and Neurock, 2000) ··································································14 Figure 1.8 Schematic orbital mixing diagrams for molecular adsorbate. Case (a) displays the Fermi level is closer to the LUMO than the HOMO; case (b) shows the HOMO is closer to the Fermi level (Yamagishi et al, 2001)························17 Figure 1.9 Orbital energy diagram for benzene in the gas phase and adsorbed at the Bridge(30) and Hollow(0) sites (Saeys et al, 2002) ·····································19 Figure 2.1 Lattice model of one-dimension system (Hoffmann, 1988)·························40 Figure 2.2 A schematic illustration of all-electron (solid lines) and pseudo- (dashed lines) potentials and their corresponding wavefunctions ·······································42 Figure 2.3 Comparison on pseudo-wavefunctions generated using the norm-conserving pseudopotential by Hamann, Schlüter and Chiang (dotted line) and US (dashed line) for the oxygen 2p orbital with regards to the oxygen 2p radial wavefunction (solid line) (Vanderbilt, 1990)···············································44 Figure 3.1 Schematic of high-symmetry benzene adsorption sites on a Pt(111) surface (Saeys et al, 2002) ······················································································52 xi Figure 3.2 Examples of a cluster model (left) and a periodic slab model (right) for benzene adsorption on Pt(111) ····································································55 Figure 3.3 Schematic illustration of the Pt(111) slab (left) and the super cell including slab and vacuum (right)···············································································60 Figure 3.4 Schematic illustration of the Pt(111) slab models with various slab-thickness ····················································································································62 Figure 3.5 Benzene adsorption energy at the Bridge(30) site with various slab thickness of Pt(111) slab models at different k-points mesh conditions at low coverage ····················································································································64 Figure 3.6 Schematic illustration of the electron donation (left) and electron backdonation (right) in the d-band model (Bligaard and Nørskov, 2007)············65 Figure 3.7 Electronic density of states (DOS) projected to d-bands of the Pt(111) surface at various slab models··················································································69 Figure 3.8 Electronic density of states (DOS) projected to d-bands of Pt(111) (solid line) and C 2pz orbital for benzene (dotted line) chemisorbed at the Pt(111) surface ····················································································································70 Figure 3.9 Electronic density of states (DOS) projected to d-bands of Pt(111) (solid line) and C 2pz orbital for benzene (dotted line) chemisorbed at the Pt(111) surface at various slab models.·················································································72 Figure 3.10 First-principles extrapolation procedure based on the plot of adsorption energy for CO on Pt(111) Hollow-hcp site versus the singlet-triplet excitation energy difference. From Mason et al (2004)················································80 Figure 3.11 Adsorption structures of molecules adsorbed on Pt3 cluster optimized in VASP ··········································································································84 Figure 3.12 HOMO-LUMO energy diagram of the molecules and Pt(111). (Thick solid line: experiment, dashed lines: CBS-QB3, dotted lines: B3LYP and thin solid lines: PBE) ································································································102 Figure 3.13 Schematic illustration of the surface interaction with the front orbitals: leftside describes electron back-donation and right-side illustrates electron donation. (Yamagishi et al, 2001)······························································103 Figure 3.14 Various coverage simulation with surface unit cells for benzene on Pt(111) ··················································································································109 xii Figure 3.15 Possible adsorption configurations at high coverage of 1/6 ML for benzene on Pt(111) with their adsorption energy. Top-views are presented in left-side, side-views in right-side.·············································································111 Figure 3.16 Adsorption energy calculation results at high coverage of 1/6 ML using 6layered slab with k-point grid of 5×5×1·····················································112 Figure 3.17 Benzene on the Pt(111) Bridge(30) site (left) and Hollow-hcp(0) site (right). (Saeys et al, 2002) ·····················································································114 xiii Chapter 1 Introduction Since the breakthrough in ammonia synthesis, catalysts have become prevalent to daily lives and essential to chemical industries. The automotive exhaust converter under the car is a typical example of catalyst application in common lives. The varieties of catalysts have played their roles in various industrial areas such as petroleum, chemicals, pharmaceuticals, automobiles, and electronic materials etc. It contributes to the feedstock production for synthetic materials for example fuels and fertilizers. Environmental issues, however, have contemporarily become severely critical for humankind to survive in the future world such as greenhouse effect, ozone layer decomposition and air/water pollution. To tackle those environmental confrontations it is strongly required to develop novel catalytic materials whose activity is enhanced to reduce energy demands, and whose selectivity is revised to prevent harmful byproducts or minimize chemical wastes. 1.1 First Principles-based Modeling Conventionally, industrial catalysts have been fabricated by trial-and-error experimentation, for example, the Fe-based ammonia synthesis catalyst has been identified following 6,500 tests with 2,500 different catalysts. (Mittasch and Frankenburg, 1950) Parallel probing schemes have advanced more efficient catalyst design, shortening catalyst screening time. (Jandeleit et al., 1999) To achieve the advancement in selectivity and activity of catalysts, the “rational catalyst development strategy” has been introduced in the design of Co-Mo bimetallic catalyst for ammonia synthesis analyzing the volcano-shaped correlation among ammonia synthesis activity 1 and nitrogen adsorption energy of potential catalysts with the help of density functional theory (DFT) calculations. (Jacobsen et al., 2001) First principles calculations has been employed in the search for novel catalysts demonstrating that a reaction rate of ammonia synthesis on a Ru(0001) step catalytic surface can be directly predicted applying DFT calculations in agreement with the experiment. (Honkala et al., 2005) However, the breakthrough in the advancement of catalysts is only attainable if catalytic reactions can be mastered at the molecular level. First principles-based modeling is a theoretical and computational method, which models molecular systems of interest, finds out the electronic and atomistic information, simulates the same behavior of molecular systems, and designs novel catalysts with enhanced activity, selectivity or stability followed by thorough validation against experiment. First principles-based modeling is serving as a prominent tool for the rational catalyst design and for kinetic modeling of catalytic process. (Xu and Saeys, 2007) In first principles calculations, DFT calculations become the base for the first principles-based modeling endowing adsorption energy, activation energy, potential energy surface, and so on. To predict heterogeneous catalytic reactions with a chemical accuracy, first principles-based modeling requires more accurate outcomes from DFT calculations. The main concern of this thesis is to obtain as accurate as possible DFT calculation results in heterogeneous catalytic reaction, particularly benzene adsorption on Pt(111), to pave the way for the first principle-based modeling for benzene hydrogenation on catalytic surfaces. 2 1.2 Principles in Heterogeneous Catalysis Heterogeneous catalysis reaction is simply comprised of adsorption, surface reaction, and desorption in terms of elementary steps. Suppose a simple chemical reaction ( A + B → P ) happens in the presence of catalyst as illustrated in Fig 1.1. First, both reactants A and B spontaneously bind to catalytic surface. Next, bound reactant A and B react and produce P on the surface surmounting the activation energy. Finally, the product P desorbs from the catalyst endothermically. A B P Metal surface Figure 1.1 Schematic illustration of a simple heterogeneous catalytic reaction. Here, it can be explained how heterogeneous catalysis can be governed by the catalytic reaction environment in terms of activity and selectivity. Catalysts provide an alternative reaction path and reduce an activation energy forming more stable complex with reactants than isolated reactants do. Catalytic activity refers to the increase in the rate of reaction for a specified chemical reaction in the presence of the catalyst and expressed in kinetics terms of reaction rate and activation energy. The selectivity of a reaction is the fraction of the starting material that is converted to the desired product and expressed by the ratio of the amount of desired product to the reacted quantity of a 3 reaction partner. The catalytic selectivity is of great importance in industrial catalysis to inhibit the undesirable side reactions and to magnify the production yield. Understanding heterogeneous catalysis requires kinetics, which enables to correlate the rate of a reaction mechanism to macroscopic limitation such as concentration, pressure and temperature. Kinetics is an important tool in catalysis to link the microscopic molecular reaction to the macroscopic industrial reactor design. Assuming the surface of a catalyst as an area comprised of a definite number of elementary active sites where some areas are vacant and other regions are covered with adsorbed atoms or molecules, Langmuir (1922) conceived an adsorption theory based on the relationship between coverage of an oxygen gas and its partial pressure over and above the surface, called Langmuir isotherm. Taylor (1925), developing the concept on the catalytic surfaces, proposed the concept of active sites, saying that only a small fraction of the surface is catalytically active. The catalytic surface had been ideally presumed with single crystal surfaces until modern spectroscopy technique of high-pressure scanning tunneling microscopy detected realistic surfaces of catalysts as a mixture of terraces, plateau, steps, and islands. Surface chemistry shows that catalytically important precious metals such as Pt and Pd possess a faced-cubic centered (fcc) bulk structure, whose low index faces are commonly studied in the quantum chemical calculations: the fcc(100), fcc(110) and fcc(111) surfaces. Actives sites of catalytic surfaces are often modeled by constructing a cluster of atoms or as periodic planer single crystal called slab. The fcc(111) surface provides three different types of active sites: On-top sites, Bridge sites between two 4 atoms, and Hollow sites between three atoms. The fcc(110) surface exhibits four kinds of active sites: On-top sites, Short-Bridge sites between two atoms in a single row, Long-Bridge sites between two atoms in a adjacent rows, and Higher coordination sites in the troughs, while the fcc(100) surface possesses three active sites: On-top sites, Bridge sites between two atoms, and Hollow sites between four atoms. The coordination number of each surface - the number of nearest neighboring atoms – is somewhat connected to the chemical reactivity of surfaces so that the most open fcc(110) surface shows high reactivity followed by the fcc(100) surface, and the closepacked fcc(111) surface gives the most stable surface plane. For example, the surface energy – the energy consumed to create a surface from a bulk – from DFT calculation is 2.299 J/m2, 2.734 J/m2 2.819 J/m2 for Pt(111), Pt(100) and Pt(111), respectively. (Vitos et al., 1998) Figure 1.2 Schematic illustration of low index surface of the fcc-structured heterogeneous catalyst: the fcc(100) is in the right, the fcc(110) in the middle, and the fcc(111) in the left. The heterogeneous catalytic reaction kinetics mechanism can be accounted either Langmuir-Hinshelwood mechanism or Eley-Rideal mechanism, where the former mechanism model is more widely adopted. The Langmuir-Hinshelwood mechanism presumes that all reactants are adsorbed on the catalytic surface, and then surface reactions take place in the chemisorbed states, while Eley-Rideal mechanism assumes that one of reactants chemisorbs first and reacts with another species in gas phase. Eq. 5 (1.1) demonstrates the simplest heterogeneous catalytic elementary reactions following Langmuir-Hinshelwood reaction mechanism, where the active site of a catalytic surface has been denoted by an asterisk mark. (Adsorption) A + * ↔ A* (1.1a) (Adsorption) B + * ↔ B* (1.1b) (Surface reaction) A* + B* ↔ AB* + * (1.1c) (Desorption) AB* ↔ AB + * (1.1d) The rate of elementary reactions of heterogeneous catalysis depends on the interaction between reactants and the catalytic surfaces at the molecular level. Adsorption is regarded as an important step because it is the initial step of heterogeneous catalysis. Comprehensive understanding on the reaction mechanism, electronic interaction and characteristics of adsorption is indispensable so that voluminous studies on adsorption, especially chemisorption, have been performed both experimentally and computationally. As the result of the interaction between the adsorbate and the surface of catalysts, chemical bonds are formed at active sites of transition metal catalysts, whose strength determines the activity of catalytic surfaces. Sabatier found the volcano-curve like relationship between the heat of adsorption and the rate of a catalytic reaction. If the chemical bond is too weak the catalyst is unable to start surface reaction by dissociating a bond, while the chemical bond is too strong, the adsorbate is unable to desorb from the surface. At the optimum rate of heterogeneous catalytic reaction the 6 catalyst activity shows the best performance. This principle can be used in the rational design for novel catalysts. 1.3 Application of First Principles-based Modeling This section is devoted to describe briefly how first principles-based modeling has contributed to the advancement of catalytic reactivity and selectivity in the design of new catalysts. The successful application of first principles-based modeling in the prediction of ammonia synthesis on the Ru(0001) surface will be presented and provide confidence for applications in other heterogeneous catalytic reactions, such as aromatic hydrogenation. Catalytic ammonia synthesis reaction mechanism is known to be comprised of the dissociation of N2 and H2, subsequent stepwise addition reactions to NH3, and desorption of ammonia, whose elementary reaction steps are illustrated in the Eq. (1.2). (Nitrogen Dissociation) N2 + 2* ↔ 2 N* (1.2a) (Hydrogen Dissociation) H2 + 2(*) ↔ 2 H(*) (1.2b) (Stepwise addition) NHx* + H(*) ↔ NHx+1* (x = 0, 1, 2) (1.2c) (Desorption) NH3* ↔ NH3 (1.2d) The rate determining step in the overall ammonia synthesis reaction with industrially used Fe catalyst is the initial dissociation of nitrogen gas molecule. The first principles study on the microscopic pathways of ammonia synthesis on Ru(0001) surface suggested that the stepwise addition elementary reaction step was probably the rate determining step. (Zhang et al., 2001) Logadóttir and Nørskov (2003) performed DFT 7 studies of all the elementary steps in the ammonia synthesis both over terrace sites and step sites of a ruthenium catalyst and found based on the calculated potential energy diagram for ammonia synthesis reaction pathway over the Ru(0001) surface that the active sites are located at steps rather than on flat terraces, that is, the reaction mainly occurs at the step sites. The surface step sites play roles as active sites to stabilize the reaction intermediate relative than the flat terrace sites and to reduce the activation energy of each elementary step. (Sholl, 2006) Figure 1.3 The calculated potential energy diagram for NH3 synthesis from N2 and H2 over closed-packed and stepped Ru surface. Adopted from Honkala et al. (2005). They additionally conclude that the N2 dissociation step is the rate determining step in the overall reaction comparing the rates of N2 dissociation reaction and stepwise addition reaction of NH* + H(*) ↔ NH2*. The rate constant can be expressed in terms of the activation energy ( Ea ,i ) for dissociation on local environment ( i ), where 8 ν denotes the prefactor, k B stands for the Boltzmann constant, and T for temperature in the following equation. ki = ν e − Ea ,i / k BT (1.3) Both because step sites are more reactive for N2 dissociation than flat terrace surface and because N2 dissociation step is the rate determining step, Honkala et al. (2005) could predict the overall reaction rate for ammonia synthesis over a nanoparticle ruthenium catalyst directly from first principles calculations. First, to describe approximately the real catalytic material covered by a complex arrangement of adsorbates, all possible local environments of N2 dissociation transition state configurations are proposed without no adsorbate in the neighboring sites, with nitrogen atom, hydrogen atom, NH or NH2 in the neighboring sites. Their relevant activation energies, including the co-adsorbate effects on the active sites of Ru(0001) step surface, are calculated with the help of DFT method. The probability of observing ( Pi ) each possible local adsorbates configuration has been predicted conducting the grand canonical Monte Carlo simulations. The total reaction rate was expressed as ( r T , p N2 , pH 2 , p NH3 ) 2 ⎛ p NH 3 = ⎜1 − 3 ⎜ pH p N K g 2 2 ⎝ ⎞ ⎟ ∑ Pi ki p N 2 ⎟ i ⎠ (1.4) where K g is the gas-phase equilibrium constant, and p N2 , pH 2 , and p NH3 are the partial pressure of N2, H2 and NH3, respectively. Eq. (1.4) assumes that all elementary reaction steps with the exception of the rate determining step happen in equilibrium and must stop once gas phase equilibrium is established. 9 In addition to clarify the number of active sites on a realistic catalyst quantitatively, experimental information on the number of active sites per gram of catalyst is required to compare with the DFT-based model, where the number of active site is expressed as a function of nanoparticle radius through analysis of the atomistic Wulff construction, which determines the global shape of crystal in equilibrium from local interaction. This first principles comparison showed that the experimentally observed rate was underestimated only by a factor of 3-20. The slight discrepancy between the measured and calculated productivity has been observed, which may be caused either by systematic errors in the bonding description of the different adsorbates or configurations, or the underestimation of the number of active sites. (Honkala et al., 2005) For more detail information, please refer to the review has been done by Sholl (2006). 1.4 Electronic Interaction in Heterogeneous Catalysis In this section, the basic concepts of chemical bonding to transition metal surfaces will be understood. First, carbon monoxide interaction with the metal surface will be illustrated; then, it will be extended to more complicated ethylene. Next, adsorbateadsorbate interaction will be briefly presented. Finally, electronic interaction of benzene adsorption on transition metal surface will be explained. 1.4.1 Surface-adsorbate interaction Chemical bonding constructed by the adsorption of a molecule on a transition metal surface can be initially understood by the Newns-Anderson model (Newns, 1969; Anderson, 1961), where both the bonding and the anti-bonding molecular orbitals of 10 the adsorbate contribute to the chemical bonding. Strong chemisorption bond can be generated when the originally empty anti-bonding orbital becomes filled and shifted above the Fermi level as the results of the interaction with the metal surface. When molecular bonds are shifted toward the surface to increase its overlap with the metal d states, the electron density of the metal is transferred to fill anti-bonding orbital, called “back donation”. The d-band model (Hammer et al., 1996) is the simplest one-electron quantum mechanical description of the interaction of atoms and molecules with a metal surface. The adsorbate coupling to the d states is a two-level problem evoking a bonding and an anti-bonding state. (Hammer and Nørskov, 2000) When a molecule interacts with a broad s band of a metal, the adsorbate state broadens and called “weak chemisorption”, whereas it splits off into bonding and anti-bonding states after the interaction with a narrow d-band, called “strong chemisorption”. (Fig. 1.4) As the chemisorption become stronger, the d-band energy becomes narrower and shifts up anti-bonding states above the Fermi level. (Fig 1.5) This results in the strong chemical bonding between adsorbate and surface. Conceptually, the d bands are characterized by the position of d band center, which can be used to compare the reactivity of various transition metal surfaces because the adsorption energy varies with the relative position of the d band center to the Fermi level. 11 Figure 1.4 The local density of states at an adsorbate in two limiting cases: (a) for a broad surface band relevant to the interaction with a metal s band; (b) for a narrow metal band representing the interaction with a transition metal d band. Adopted from Hammer and Nørskov (2000) Figure 1.5 The local density of states projected onto an adsorbate state interacting with the d bands at a surface. Adopted from Hammer and Nørskov (2000) 1.4.1.1 CO interaction with the Pt(111) surface The chemisorption of carbon monoxide can be analyzed initially by the Blyholder model (1964) or d-band model. In the Blyholder model the chemisorption of CO is energized by the combination of both σ-donation, which is an electron transfer from the highest occupied molecular orbital (HOMO) 5σ bonding orbital to the substrate, and π-back-donation, whose electron transfer from metal d-state to the CO causes the degeneration of the 2π* anti-bonding orbital. Similarly to the aforementioned d band model, Fig. 1.6 illustrates that the interaction with the metal s-states results in a downshift and broadening of both CO 2π* and 5σ states, whereas the coupling to the d 12 states of metal causes the split-off both bonding and anti-bonding states. This infers that the contribution of the filled 5σ orbital to the chemical bonding between CO to the Pt(111) surface is comparatively minimal but the attractive interaction between empty 2π* orbital to the metal surface becomes more dominant to the right in the periodic table. (Hammer and Nørskov, 2000) Figure 1.6 The self-consistent electronic DOS projected onto the 5σ and 2π* orbitals of CO: in vacuum and on Al(111) and Pt(111) surface. From Hammer et al. (1996) 1.4.1.2 C2H4 interaction with the Pd(111) surface The ethylene adsorption on a Pd(111) surface can be understood by the interaction between the frontier orbitals of ethylene with the sp-band and d-band of the metal. As illustrated in Fig 1.7, the frontier orbitals – π and π*, are downshifted and broadened upon the interaction with the sp-band, then renormalized frontier orbitals interact with the valence d-band of the metal so that bonding and anti-bonding orbitals are constructed as a result of electron donation and electron back-donation. 13 Figure 1.7 Frontier orbital interaction in the di-σ adsorption of ethylene on Pd(111). From Pallassana and Neurock (2000) Pallassana and Neurock (2000) studied the di-σ adsorption of ethylene on a metal surface analyzing the changes in the electronic properties of the metal surface layer. To compare the chemical reactivity of different metal surface, the d-band model of Hammer and Nørskov (2000) has been used calculating d-band center and d-band filling. Here, d-band center has been obtained taking the first moment of the projected density of d-states about the Fermi energy, and the d-band filling is determined by the fractional area for the projected d-band below the Fermi energy to the integrated area 14 of the d-band from -∞ to +∞. Finally, they found the linear relationship between the ethylene adsorption energy and the d-band center of the bare metal substrates. 1.4.2 Adsorbate-adsorbate interaction The surface coverage of reactants, intermediates and products on catalysts are depending on reaction conditions, such as temperature and partial pressure. If the coverage becomes large, adsorbate-adsorbate interactions are not negligible so that these interactions should be evaluated in the adsorption energy calculations. (Hammer and Nørskov, 2000) Both attractive and repulsive adsorbate-adsorbate interactions are manifested; (Mortensen et al., 1999) the attractive interactions are usually weak and observed at low coverage, (Bozso et al., 1977) while the repulsive interactions are prevalent at high coverage. (Stampfl et al., 1996; Stampfl et al., 1999) A sharp slope of heat of adsorption with coverage often experimentally observed is contributed from repulsive interaction at high coverage. (Brown et al., 1998) The four factors affecting interactions among adsorbates are listed as follow: first, direct interactions due to overlap of wavefunctions is dominated by the Pauli repulsion; second, indirect interactions resulted from the electronic structure of transition metal changes, i.e., a downshift of d states, on the adsorption of one adsorbate, lead to weaker interactions with other adsorbates; third, elastic interaction brought by local distortions of the surface lattice on adsorption contributes to repulsive interaction with other adsorbates; fourth, non-local electrostatic effect can be justified as dipole-dipole interaction. 15 The strong coverage dependence of adsorption energies are related to the reactivity of a surface. The more weakly an adsorbate binds, the more reactive the surface becomes. Furthermore, the reactivity of catalytic surface can be manipulated by the adjustment of reaction temperature or partial pressure. 1.4.3 Electronic interaction on benzene adsorption on Pt(111) The experimental studies, such as HREELS (High Resolution Electron Energy Loss Spectroscopy), LEED (Low Energy Electron Diffraction) and RAIRS (Reflection Absorption IR Spectroscopy), demonstrated that benzene is adsorbed parallel to the surface mainly due to the interaction of π electrons of the aromatic ring with d-orbitals of metal surface. (Haq and King, 1996; Lehwald et al., 1978; Wander et al., 1991) With the development of density functional theory, theoretical studies have been highlighted and the study on electronic interaction between aromatics and transition metals has been extensively investigated. First, Yamagishi et al. (2001) studied benzene adsorption on the Ni(111) surface of p(√7×√7)R19.1º unit cell. They approached at the level of molecular orbitals to analyze benzene molecular chemisorption on the surface. During a molecular adsorption on a metal surface, the frontier orbitals, such as highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO), interact with the s-, p- and d-state of the surface at the Fermi level. This may result in two cases, either LUMO is closest to the Fermi level so that net charge flows from surface to molecule, or HOMO is closest to the Fermi level so that net charge flows back to surface from molecule. (Fig 1.8) For the case of benzene chemisorption on Ni(111) surface, the 16 electrons flow from the Ni substrate to the benzene adsorbate so the bonding between benzene adsorbate and the surface can be deduced as the product of the mixing of surface state with the LUMO. Figure 1.8 Schematic orbital mixing diagrams for molecular adsorbate. Case (a) displays the Fermi level is closer to the LUMO than the HOMO; case (b) shows the HOMO is closer to the Fermi level. Adopted from Yamagishi et al. (2001). Next, Saeys et al. (2002) found that adsorption site preference can be understood by the analysis of the molecular orbitals formed on adsorption. On the adsorption, benzene σ orbitals in C-C bonds and C-H bonds are stabilized by the interaction with Pt orbitals so that around 5 % of the electron density of the benzene σ orbital is donated into empty Pt orbitals. The σ-interaction depends on the adsorption site and plays a role in the site preference. Further, the strongest interaction can be found at π orbitals. On the adsorption at the Bridge(30) site, doubly degenerate HOMO 1e1g π bonding molecular orbital is split into a low-lying orbital and a high-lying orbital, removing its degeneracy. (Fig 1.9) The low-lying orbital causes two C-atoms to form an σ-like bond with a Pt atom right below, whereas the high-lying orbital has πinteraction with the other C-atoms. To stabilize low-lying orbital energy, 36.2º tilted C-H bonds maximize the overlap of carbon pz orbitals with Platinum dyz and dz2 17 orbitals. This leads to a strong C-Pt bond and strong adsorption energy at the bridge site. On the other hand, the 1e2u π* anti-bonding molecular orbitals which are LUMO in isolated benzene are partly filled by back-donation from the Pt dz2 orbital, which causes C-C bonds length elongation. 18 Figure 1.9 Orbital energy diagram for benzene in the gas phase and adsorbed at the Bridge(30) and Hollow(0) sites. Adopted from Saeys et al. (2002). 19 1.5 Organization of the Thesis This thesis consists of four chapters. The first chapter briefly explains about firstprinciples based modeling, catalysis and aromatic chemisorption. Computational methods such as theoretical backgrounds for quantum calculations will be presented in the second chapter to assist the understanding of readers. The computational results on three different agenda will be covered in the third chapter: (i) how to achieve converged adsorption energy values for benzene on Pt(111); (ii) what is the reliable method to obtain an accurate adsorption energy; (iii) what is the coverage effect on benzene adsorption on Pt(111). The last chapter will contain the discussion and final conclusion. 20 1.6 References Anderson, P.W., “Localized Magnetic States in Metals”, Physical Review, 124, pp. 41. 1961. Blyholder, G., “Molecular Orbital View of Chemisorbed Carbon Monoxide”, Journal of Physical Chemistry, 68, pp. 2772. 1964. Bozso, F., Ertl, G., Grunze, M., and Weiss, M., “Interaction of nitrogen with iron surfaces I. Fe(100) and Fe(111)”, Journal of Catalysis, 49, pp. 18. 1977. Brown, W.A., Kose, R., and King, D.A., “Femtomole Adsorption Calorimetry on SingleCrystal Surfaces”, Chemical Reviews, 98, pp. 797. 1998. Hammer, B., Morikawa, and K Nørskov, J.K., “CO Chemisorption at Metal Surface and Overlayers”, Physical Review Letters, 76, pp. 2141. 1996. Hammer, B., and Nørskov, J.K., “Theoretical Surface Science and Catalysis – Calculations and Concepts”, Advances in Catalysis, 45, pp. 71. 2000. Haq, S., and King, D.A., “Configurational Transitions of Benzene and Pyridine Adsorbed on Pt{111} and Cu{110} Surfaces: An Infrared Study”, Journal of Physical Chemistry, 100, pp. 16957. 1996. Honkala, K., Hellman, A., Remediakis, I.N., Logadottir, A., Carlsson, A., Dahl, S., Christensen, C.H., and Norskov, J.K., “Ammonia Synthesis from First-Principles Calculations”, Science, 307, pp. 555. 2005. Jacobsen, C.J., Dahl, S., Clausen, B.S., Bahn, S., Logadottir, A., and Norskov, J.K., “Catalyst Design by Interpolation in the Periodic Table: Bimetallic Ammonia Synthesis Catalysts”, Journal of the American Chemical Society, 123, pp. 8404. 2001. Jandeleit, B., Schaefer, D.J., Powers, T.S., Turmer, H.W., and Weinberg, W.H., “Combinatorial Materials Science and Catalysis”, Angewandte Chemie International Edition, 38, pp. 2494. 1999. Langmuir, I., “The mechanism of the catalytic action of platinum in the reactions 2Co+O2= 2Co2 and 2H2+O2=2H2O”, Transactions of the Faraday Society, 17, pp. 621. 1922. Lehwald, S., Ibach, H., and Demuth, J.E., “Vibration Spectroscopy Of Benzene Adsorbed On Pt(111) And Ni(111)”, Surface Science, 78, pp. 577. 1978. Logadóttir, Á., and Nørskov, J.K., “Ammonia synthesis over a Ru(0001) surface studied by density functional calculations”, Journal of Catalysis, 220, pp. 273. 2003. Mittasch, A., and Frankenburg, W., “Early Studies of Multi-component catalysts”, Advances in Catalysis, 2, pp. 81. 1950. 21 Mortensen, J. J., Morikawa, Y., Hammer, B., and Nørskov, J.K., “Density Functional Calculations of N2 Adsorption and Dissociation on a Ru(0001) Surface”, Journal of Catalysis,169, pp. 85. 1997. Neurock, M., “Perspective on the First Principles Elucidation and the Design of Active Sites”, Journal of Catalysis. 216. pp. 73. 2003. Newns, D.M., “Self-Consistent Model of Hydrogen Chemisorption”, Physical Review, 178, pp. 1123. 1969. Pallassana, V., and Neurock, M., “Electronic Factors Governing Ethylene Hydrogenation and Dehydrogenation Activity of Pseudomorphic PdML/Re(0001), PdML/Ru(0001), Pd(111), and PdML/Au(111) Surfaces”, Journal of Catalysis, 191, pp. 301. 2000. Saeys, M., Reyniers, M.F., Marin, G.B., and Neurock, M., “Density Functional Study of Benzene Adsorption on Pt(111)”, Journal of Physical Chemistry B. 106, pp. 7489. 2002. Scholl, D.S., “Applications of Density Functional Theory to Heterogeneous Catalysis”, The Royal Society of Chemistry, 4, pp. 108. 2006. Stampfl, C., Schwegmann, S., Over, H., Scheffler, M., and Ertl, G., “Structure and Stability of a High-Coverage (1×1) Oxygen Phase on Ru(0001)”, Physical Review Letters, 77, pp. 3371. 1996. Stampfl, C., Kreuzer, H.J., Payne, S.H., Pfnür, H., and Scheffler, M., “First-Principles Theory of Surface Thermodynamics and Kinetics”, Physical Review Letters, 83, pp. 2993. 1999. Taylor, H.S., “A Theory of the Catalytic Surface”, Proceeding of the Royal Society of London, 108, pp. 105. 1925. Vitos, L., Ruban, A.V., Skriver, H.L., and Kollár, J., “The surface energy of metal”, Surface Science, 411, pp. 186. 1998. Wander, A., Held, G, Hwang, R., Blackman, G.S., Xu, M.L., Andres, P., Van Hove, M.A., and Somorjai, G.A., “A diffuse LEED study of the adsorption structure of disordered benzene on Pt( 111)”, Surface Science, 249, pp. 21. 1991. Xu, J., and Saeys, M., “First Principles Study of the Coking Resistance and the Activity of a Boron Promoted Ni Catalyst”, Chemical Engineering Society, 62, pp. 5039. 2007. Yamagishi, S., Jenkins, S.J., and King, D.A., “Symmetry and site selectivity in molecular chemisorption : Benzene on Ni{111}”, Journal of Chemical Physics, 114. pp. 5765. 2001. Zhang, C., Liu, Z-.P., and Hu, P., “Stepwise addition reaction in ammonia synthesis: A first principles study”, Journal of Chemical Physics, 115, pp. 609. 2001. 22 Chapter 2 Computational Method 2.1 First Principles Quantum Chemical Methods To understand the chemisorption step in a catalytic reaction at the electronic level, first principles quantum chemical methods are selected as a probe tool. Quantum chemical calculations enable to quantify the energy of atomic and molecular adsorption and identify the adsorption structures so that it is possible to explore the trend of catalyst activity and selectivity. 2.1.1 Time independent Schrödinger equation The fundamental equation for first principles quantum chemical calculations is the non-relativistic time-independent Schrödinger equation, which describes the quantum behavior of nuclei and electrons in molecules. Hˆ Ψ = E Ψ (2.1) Here Hˆ , a Hamiltonian, is a quantum mechanical operator for the total energy, which is the sum of kinetic energy and potential energy due to Coulomb interaction; Ψ is the wavefunction, and E is the eigenvalue of a particular stationary state. In the Schrödinger equation, the Hamiltonian Hˆ describes the sum of the kinetic energy of the electrons and nuclei, the electrostatic energy due to nuclei-electrons attraction, and the potential energy due to electron-electron and nucleus-nucleus repulsion. The Born-Oppenheimer approximation considers electrons as moving while 23 nuclei are fixed because nuclei are much heavier than electrons, for example, the mass ratio of nuclei to electrons in hydrogen atom is 1,800 and the mass ratio increases to 20,000 in carbon atom. Thus, ignoring the kinetic energy term due to nuclei, the Hamiltonian can be separated into the electronic Hamiltonian Hˆ elec expressed as Hˆ elec = Tˆe + Vˆext + Vˆee (2.2) where Tˆe is the kinetic energy of electrons, Vˆext is the external electrostatic potential due to nuclei, and Vˆee is the electron-electron interaction energy. The solution of the Schrödinger equation with the electronic Hamiltonian is the electronic wavefunction Ψelec and the electronic energy Eelec . Hˆ elec Ψelec = Eelec Ψelec (2.3) Since Dirac’s famous remark in 1929 that, “…application of these [fundamental] laws leads to equations that are too complex to be solved,” (Pople, 1999) securing the approximate solution of the Schrödinger partial differential equations has become the ultimate target of quantum chemistry and quantum mechanics with the assistance of variety theoretical disciplines, such as ab initio wavefunction based methods, ab initio density functional theory (DFT) methods and semi-empirical methods. Both ab initio wavefunction based methods and ab initio density functional theory methods are derived from first principles, which uses only the inherent physical constants to approximate the electron-electron interaction energy of many-electrons 24 system and does not use any parameters fitted to experimental data. Semi-empirical methods use fitted parameters to match experimental data so that they are less computationally demanding than first principles-based methods. In this chapter only first principles-based methods will be discussed in detail. 2.1.2 Hartree-Fock approximation Analytical solutions of the electronic time-independent Schrödinger equation are not possible for catalytic systems of interest. Rather, the solution of Eq. (2.3) can be obtained employing numerical approximation schemes in terms of one-electron orbitals. The Hartree-Fock (HF) approximation is a physically reasonable approach of the complicated N-electron wavefunction Ψelec , implementing the Slater determinant, Φ SD so that the N-electron Schrödinger equation is reduced to n single-electron problems. The Slater determinant is the anti-symmetric product of N one-electron wavefunctions and the Hartree-Fock energy EHF , is the lowest energy corresponding to the best Slater determinant and is determined operating the variational principles in Eq. (2.4). N Slater determinants are determined searching the best spin orbitals, which satisfy the constraint that the energy corresponding to a Slater determinant should be maintained as minimal. EHF = min E [Φ SD ] Φ SD → N (2.4) 25 The Hartree-Fock energy is a functional of the spin orbitals, and the best spin orbitals χ i , can be determined by the Hartree-Fock equations in Eq. (2.5) derived from Eq. (2.4). Here, the eigenvalue of the one-electron Hamiltonian ε i physically denotes individual molecular orbital energies. fˆi χ i = ε i χ i , i = 1, 2,K N . 1 fˆi = − ∇ 2 + VC (ri ) + VHF (ri ) 2 (2.5) (2.6) Similar to the electronic Hamiltonian in Eq. (2.2), the Fock operator in the HF equations (Eq. 2.6) is comprised of the kinetic energy and electrostatic potential energy due to electron-nucleus attraction VC , and the Hartree-Fock potential VHF equivalent to the electron-electron interaction energy. The HF potential is the average repulsive potential experienced by one electron due to the other N-1 electrons so that the individual electron-electron interactions are accounted in an average way neglecting an actual behavior of electrons. However, in reality, the motion of individual electrons should be corrected to the instantaneous positions of the other electrons both in the short range and in the long range, which is called electron correlation effect. Since the Fock operator depends on the spin orbitals, which is the very solution of the eigenvalue problem, the HF equation have to be solved as follow: (i) guess spin orbitals to solve the HF equation; (ii) determine the new spin orbitals based on the solution for the HF equation; (iii) compare the new spin orbitals with the previous one; 26 and (iv) continue the calculation loop until they converge within desired criteria. This procedure is called Hartree-Fock Self-Consistent Field (SCF) approximation. 2.1.3 Electron correlation methods Even though Hartree-Fock SCF theory is successful to determine molecular electronic wavefunctions and properties, the poor description of electron correlation in HartreeFock causes the failure of energetic prediction in a chemical accuracy level. HartreeFock scheme is a mean-field theory, in which each electron has its own wavefunction, which in turn obeys an effective one-electron Schrödinger equation. In the mean-field theory, when electron #2 changes position electron #1 has no idea on the instantaneous position of electron #2 so that their motion is uncorrelated. In reality, however, because the direct Coulomb repulsion of electrons, the instantaneous position of electron #2 forms the center of a region in space which electron #1 avoids. (Knowles et al., 2000) The difference between the exact ground state energy and the HartreeFock energy is the correlation energy expressed in Eq. (2.7). Ec = EHF − EExact (2.7) There are two types of correlation energy: dynamical correlation and static correlation. The former arises from the overestimation of short-range electron repulsion in HF wavefunctions, in other words, the HF calculation overestimates the probability of finding the two electrons in the cusp region, where the wavefunction increases linearly from the center of nuclei, contributing to the overestimation of bond length; however, the latter contributes to underestimation of bond length arising from correlation with electrons in long-range on molecular dissociation. With chemical bonds breaking, 27 dynamical correlation is mitigated. As a result of overlooking dynamical correlation, HF methods tend to overestimate bond length and underestimate binding energy. The dynamical and static correlations are often counteracted each other so that HF calculations are amended due to partial nullification of correlation error. (Knowles et al., 2000) For the high level electron correlation treatment, two kinds of ab initio wavefunction based methods are available: single-reference methods and multi-reference methods: The single-reference methods can be applied to represent dynamical correlation effect of optimized molecules at their ground states when non-dynamical correlation effect is weak. Even though various multi-reference methods are available, such as, multiconfiguration self-consistent field (MCSCF) wavefunction method, multi-reference configuration interaction (MRCI) calculation and so on, here three widely used singlereference methods are explained in brief: Configurational Interaction, Møller-Plesset Perturbation theory and Coupled-Cluster theory. 2.1.3.1 Configurational interaction In the configurational interaction (CI) method, the trial wavefunction is constructed as a linear combination of the ground-state wavefunction, singly excited wavefunction and doubly excited-state wavefunction, etc. Theoretically, full-CI method presents the exact energy within the given basis set. However, full-CI calculations are computationally demanding, so it is feasible only on very small systems. Alternatively, the CI expansion can be truncated at the level of singlet or doublet, termed CI with Singles (CIS) or CI with Doubles (CID). CIS is equal to HF for the ground-state energy because there is no difference in matrix elements between the HF wavefunction 28 and singly excited determinant so that CID should be applied to acquire any improvement over the HF results. Among CI methods, CISD method is effective for various systems providing the 80~90% of recovery in the available correlation energy. (Levine, 2000) Unfortunately, the truncated CI methods are not size-extensive. If the computed energy scales linearly with the number of non-interaction particles, a computational method is considered size-extensive. For CISD method, the larger molecule becomes, the lesser amount of correlation energy is recovered due to the lack of the disconnected quadruples. The lack of size-extensiveness leads to rare application of CI calculations due to no guarantee on the energy convergence. (Duch and Diercksen, 1994) 2.1.3.2 Møller-Plesset perturbation theory A perturbation theory considers that if a problem of interest differs slightly from the already solved one, the solution to the problem should be close to the known system. In correlation energy calculations, the perturbation theory is employed to obtain an approximate solution accompanying correction terms with perturbation. Eq. (2.8) demonstrates a mathematically expressed many-body perturbation theory, where a Hamiltonian operator Hˆ consists of a reference Hamiltonian Hˆ 0 , an excited Hamiltonian H ′ , and a parameter λ , which determines the perturbation degree. (Levine, 2000) Hˆ = Hˆ 0 + λ H ′ (2.8) 29 Møller-Plesset (MP) perturbation theory (Møller and Plesset, 1934) calculates the correlation energy by taking the reference Hamiltonian and an excited Hamiltonian. Analogously to the CI method, the first-order energy is exactly the HF energy and electron correlation energy can be described from the second-order method, MP2, which recovers 80~90 % of the correlation energy. However, it should be applied only to systems which have sufficiently large HOMO-LUMO gap, otherwise the perturbation expansion diverges. Even though MP4 method renders reliably accurate solution to most systems, MP2 level calculation has been conducted to examine catalysis owing to the system size and electronic complexity of catalytic system models. (Santen and Neurock, 2006) 2.1.3.3 Coupled-Cluster theory Coupled cluster (CC) theory (e.g. Jensen, 1999) employs an exponential wave operator to overcome the size-extensive deficiency of truncated CI methods. In CI methods, a series of excited-state wavefunction operators should be truncated in terms of single, double and triple excited states. It is said that CCSD(T) method is feasible to recover correlation energy within 4 kJ/mol, but it is only applicable to small systems up to 10 heavy atoms due to drastic increase of the computational cost with N7 order, where N is the number of atoms. In summary, ab initio wavefunction based quantum calculation methods for high level treatment of electron correlation energy can be lined up in terms of accuracy as follows; HF < MP2 < CISD < CCSD < MP4 < CCSD(T) < full-CI. 30 2.1.4 Density Functional Theory (DFT) The first principles quantum chemical methods so far explained are wavefunctionbased methods, whose key variable is the wavefunction determined by 3N spatial and N spin variables for N-electron system. Density functional theory uses an electron density, a function of three spatial variables, as a means to solve approximately the time-independent Schrödinger equation. The Hohenberg-Kohn theorem presumes that the total energy of the system at the ground state is an unique functional (i.e. a function which converts a function into a number) of the ground state electron density ρ (r ) , that is, the probability of finding electrons within the volume element at r in the N-electron system - E [ρ (r )] . By the application of the variational principles, the ground state energy of an atom or molecule can be written in a certain form including the unknown universal functional F [ρ ] , and the explicitly system-dependent potential energy VNe , due to the nuclei-electron attraction, E0 = min { F [ρ ] + ρ →N ∫ ρ (r ) V Ne dr } (2.9) The energy must be evaluated using the corresponding universal functional, that is, the summation of the true kinetic energy term T, the electron-electron interaction energy term Eee , which is divided into two parts: the known classical Coulomb interaction energy J, and the unknown non-classical Coulomb interaction energy Encl . 31 The Kohn and Sham (1965) approach to the unknown universal functional introduces a non-interacting reference system, where there is no electron-electron interactions so that the major part of the true kinetic energy, a non-interacting kinetic energy TS , is approximated to good accuracy with local potential Vs (r ) , and one-electron KohnSham equations (Eq. 2.10) have been formulated following the framework of the HF scheme replacing spin orbitals with Kohn-Sham (KS) orbitals, and the Fock operator with Kohn-Sham operator (Eq. 2.11). fˆi KS ϕi = ε i ϕi (2.10) 1 KS fˆi = − ∇ 2 + VS (ri ) 2 (2.11) As a result, the residual part of the true kinetic energy TC is taken into account in exchange-correlation energy term E XC , as well as the non-classical Coulomb interaction energy, Encl . Therefore, the universal functional can be revised as follow: F [ρ ] = Ts [ρ ] + J [ρ ] + E XC [ρ ] E XC [ρ ] = { T [ρ ] − TS [ρ ] } + (2.12) { E ee [ρ ] − J [ρ ] } = TC [ρ ] + Encl [ρ ] (2.13) Above equations clearly shows that the universal functional has only one unknown term, the exchange-correlation energy. 32 2.1.5 Exchange-Correlation functionals The residual part of the true kinetic energy and the non-classical Coulomb interaction energy are unknown so that the accurate method is essential to the approximation of the true exchange-correlation energy. The accuracy of the density functional theory calculation, therefore, depends on the quality of exchange-correlation (XC) energy approximation. There are four types of XC energy approximation: the local density approximation (LDA), generalized gradient approximation (GGA), meta-GGA and hybrid functionals. 2.1.5.1 Local density approximation (LDA) The homogenous electron gas, in which electrons are evenly distributed on a positive charge background to produce the electrically neutral total ensemble, is a good physical model for simple metals, but far from the correct description in the rapid density changes in atoms and molecules. However, this model is the only system in which we can perceive the form of the exchange and correlation energy exactly or at least very accurately. In the uniform electron gas, the exchange-correlation functional is expressed as follows: LDA [ρ ] = ∫ ρ (r )ε XC [ρ ]dr E XC (2.14) Here, ε XC is the exchange-correlation energy per particle of the uniform electron gas density, and can be divided into exchange and correlation parts, ε [ρ ] XC = ε [ρ ] X + ε [ρ ] C (2.15) 33 The frequently used LDA XC functional is SVWN, which is composed of the Slater exchange and the VWN correlation part. The VWN correlation energy was obtained in the random phase approximation, while the parametrization scheme of Ceperly and Alder was used for the VWN5 correlation part. These LDA functionals overestimate bond energies by 150 kJ/mol of atomization energy calculation with JGP test set, (Johnson et al., 1993) while Hartree-Fock (HF) method underestimates atomization energy by 360 kJ/mol in the same set. The latter is because HF method neglects the correlation energy contribution and describes chemical bonding poorly. 2.1.5.2 Generalized gradient approximation (GGA) The LDA concept uses the exchange-correlation energy for the uniform electron gas at every point in the system regardless of the inhomogeneity of the real charge density. For non-uniform charge densities the exchange-correlation energy may differ significantly from the uniform result. This deviation can be expressed in terms of the gradients and higher spatial derivatives of the total charge density. To correct the deviation and to enhance the shortcomings of LDA functionals, not only the density at a particular position but also the gradient of the charge density, Δρ (r ) , - differentiating α and β spins - are introduced to develop a gradient corrected exchange-correlation functional. GGA [ρα , ρ β ] = ∫ f (ρα , ρ β , Δρα , Δρ β E XC ) dr (2.16) GGA is separated into two parts – an exchange contribution part and a In practice, E XC correlation part. 34 GGA E XC = E XGGA + ECGGA (2.17) However, the construction on gradient corrected functionals is merely based on mathematical methods and does not have any physical meaning so that it produces much poorer results than LDA because the LDA is at least based on the physically relevant homogeneous gas model. Therefore, parametrization should be done based on either experimental data such as atomization energies, or first principles to obtain the exchange (X) part, E XGGA = E XLDA − ∑ ∫ F (Sσ ) ρσ 4 3 r dr (2.18) σ where, Sσ is the reduced density gradient for spin σ and also known as a local inhomogeneity parameter. With the same spirit but in a more complicated way, the correlation part also has been developed. These functionals are called generalized gradient approximation (GGA). According to the construction method of the functionals, GGAs can be roughly divided in two groups. The first GGA XC functional group is the “locally based” functionals constructed from the uniform electron gas, e.g., the PW91 functional (Perdew et al., 1992). The second group is semiempirical functionals with one or more parameters fitted to a particular class of systems, e.g., PBE functional (Perdew et al., 1996). 35 2.1.6 Basis sets The one-electron Kohn-Sham equations in Eq. (2.10) represent a complicated system holding a differential operator for the kinetic energy term and an integral operator for the Coulomb contributions. To solve the one-electron KS equations in the iterative self-consistent field with the application of computational schemes, the linear combination of atomic orbital (LCAO) expansion (Roothaan, 1951) of the KS molecular orbitals has been adopted. Here Kohn-Sham molecular orbitals ϕi are expanded as computationally accessible basis sets, that is, a set of predefined basis functions η n as displayed in Eq. (2.19). Through the LCAO expansion, a set of computationally demanding coupled integro-differential equations has been transformed into computationally accessible linear matrix algebra. (Koch and Holthausen, 2001) ϕi = ∑C ni ηn (2.19) n In the LCAO expansion, if basis function A reproduces unknown functions more closely than basis function B; less number of basis function A is required to achieve a desired level of accuracy than basis function B. (Levine, 2000) If the number of basis functions is increased to infinity, the exact molecular orbital is described accurately and the set of used basis functions is called “complete basis set”. Therefore, how to choose the type of basis functions and the basis set size is crucially important in DFT electronic structure calculations. The basis functions commonly used in first principles quantum chemical calculations with the LCAO expansion schemes are physically relevant Slater-type orbitals (STO) 36 known as atomic orbitals and computationally efficient Gaussian-type orbitals (GTO). The other quantum chemical computational codes, which do not adopt the LCAO scheme, use a plane wave basis set. An orbital is a contour of probability, in which at least 95 % of electrons can be found. In atoms, electrons occupy atomic orbitals, while they occupy molecular orbitals in molecules. Slater-type orbitals expressed in spherical polar coordinates as displayed in Eq. (2.20), where N is a normalization constant, Ylm stands for spherical harmonic functions, and ξ for an orbital exponent, can be a natural choice for basis functions because the similarity to the atomic orbital of hydrogen atom. The Slater-type orbitals behave precisely both in the short-range and in the long-range, in other words, it exhibits the correct cusp (linear increase of the wavefunction from the origin to spatial directions) at the nucleus with a discontinuous derivative and the exponential tail with the desired decay rate at the long-range. As a computational code employed Slatertype orbitals as basis functions, Amsterdam Density Functional code (ADF) is available. η STO = N r n−1 exp [− ξ r ] Ylm (Θ, φ ) (2.20) Different from the Slater-type orbitals, the Gaussian-type orbitals (GTO) in Cartesian coordinates displayed in Eq. (2.21) do not guarantee the correct reproduction of exponential tail at the long range or the linear increase of the wavefunction at the nucleus. Here, N is a normalization factor, α is a parameter for the orbital exponent, and the summation of the exponent l + m + n determines types of orbitals, such as, s-, p- or d-orbitals. 37 η GTO = [ N x l y m z n exp − α r 2 ] (2.21) In general, the Slater-type orbital describes the wavefunction behavior better than the Gaussian-type orbital and more Gaussian-type orbitals should be employed to achieve the same accuracy with the Slater-type orbitals. Alternatively, contracted Gaussian functions (CGF) have been designed with the linear combination of a set of primitives, where primitives refer to individual Gaussian functions, expressed as η CGF = ∑d a η aGTO (2.22) a Here, the contraction coefficients d a can be determined to mimic a single STO function. Now basis set can be explained as the set of exponents, such as, α in the Gaussian-type orbitals or ξ in the Slater-type orbitals, and contraction coefficients for a range of atoms. Gaussian03, one of the wavefunction based quantum chemical computational codes, facilitates CGF to expand the molecular orbitals. (Frish et al., 2003) From the chemist’s perspective, molecules are often regarded as a collection of slightly distorted atoms. Hence it is primarily required for the primitive basis sets to deliver an accurate description of the atoms. The minimum basis set whose number of functions used is the smallest is inadequate to describe the isolated atom so that additional primitive functions are constructed by optimization of the Hartree-Fock energy of the atom. For example, polarization functions should be augmented to better describe distortion of atomic orbitals in molecules and especially to reproduce precisely 38 chemical bonding, and diffuse functions should be supplemented for accurate description of anions and weak bonds. (Davidson and Feller, 1986) It is beneficial to mention the basis set superposition error (BSSE) resulted from the basis set incompleteness. The more incomplete the basis set is, the more the interaction energy is overestimated. The molecular calculations tend to utilize any available basis functions on neighboring centers to compensate the basis set deficiency. When the basis set is increased, the change in a relative energy seems to converge, called basis set effect. The counterpoise method (Boys and Bernardi, 1970) is commonly applied to estimate the BSSE. Suppose the simple reaction - A + B → AB , where basis sets of two molecules A and B will be denoted by subscripts ‘a’ and ‘b’, the combined basis set of the complex AB denoted by ‘ab’, and the complex geometry denoted by a superscript *. The reaction energy can be calculated from the complex energy minus the monomer energies expressed in Eq. (2.23). To estimate the BSSE contribution on this reaction energy, the energies of each of the two fragments are calculated with the geometry in the complex using the individual basis sets, and the same calculations are done with the combined basis set of ab, which results in the implementation of ghost orbitals, basis functions without its nuclei. The counterpoise correction energy can be computed by the formula expressed in Eq. (2.24). Therefore, the counterpoise corrected reaction energy can be obtained as ΔEreactioin − ΔECP . (Levine, 2000) ΔEreaction = E ( AB )ab − E ( A)a − E (B )b * (2.23) ΔECP = E ( A)ab − E ( A)a + E (B )ab − E (B )b * * * * (2.24) 39 2.1.7 Plane wave basis sets The plane wave basis set implements the LCAO scheme differently from the atomic orbital basis set. Plane wave is the solutions of the Schrödinger equation of a free particle employing a free-electron model of solid-state physics. Figure 2.1 Lattice model of one-dimension system The one-dimensional molecular system composed of a chain of hydrogen atoms is illustrated in Figure 2.1. The orbitals can be depicted in a lattice model, where n indicates an index for lattice point, χ n stands for a basis function for each lattice point, and a denotes the lattice spacing. (Hoffmann, 1988) According to Bloch’s theorem (Ashcroft and Mermin, 1976), under periodic boundary conditions the wavefunction of one electron ψ k can be expressed as a lattice periodic part - χ n , and a plane-wave-like part - eikna , where k denotes an index for nodes in a periodic system as shown in Eq. (2.25). ψk = ∑e ikna χn n At k = 0, ψ 0 = ∑ e 0 χ n = ∑ χ n n n At k = π n , ψ π = ∑ eπin χ n = ∑ (− 1) χ n n n n (2.25) n 40 The wavefunction for the s-orbital in a hydrogen atom corresponding to the value of k=0 demonstrates the strongest bonding with the most stable energy, but at the value of k=π/a the one-electron wavefunction for the s-orbital displays the anti-bonding with the most unstable energy. Further, there is a significant range of k value, called the first Brillouin zone, which efficiently presents a periodic system, such as a chain of hydrogen atoms: |k| ≤ π/a. This Brillouin zone can be reduced by symmetry operation into the irreducible Brillouin zone. Compared to the atomic orbital basis set, the plane wave basis set has many advantages. The plane wave basis set is independent of atomic positions, orthonormal by construction and no BSSE. The periodicity of the plane wave basis set may benefit for periodic systems, such as bulk or surface, while it might be disadvantageous for aperiodic molecules. The plane wave basis set requires much more basis functions than the basis set used in wavefunction based methods. To overcome this basis set size problem, the pseudopotential approximation has been introduced, which replace the complicated behavior of the core electrons of an atom with effective core potentials or pseudopotentials. 2.1.8 Pseudopotentials Figure 2.2 schematically illustrates the electron-nucleus potential − Z r of all-electron potentials diverges to -∞ as r is close to 0. The valence wavefunction, ψ V , oscillates rapidly in the core region to ensure that the valence electron wavefunction is orthogonal to the core electron wavefunction, which requires a huge number of plane wave at the cost of paramount computing time. Fortunately, this problem can be tackled by the introduction of pseudopotentials. 41 From the chemists view, a core electron is considered as an inert in the chemical environment, whereas a valence electron mainly contributes to chemical properties, such as chemical bonding and scattering properties. This assumption enables to limit the actual calculations to the valence electrons only with the indirect inclusion of the core electrons with the aids of effective core potentials. In reality, pseudopotentials replace electronic degrees of freedom in the Hamiltonian by an effective potential. Therefore, the introduction of pseudopotentials delivers several benefits. First, it speeds up computational calculations by reducing the basis set size tremendously; Secondly, it minimizes the number of electrons of interest and the degree of freedom as well. Lastly, relativistic effect can be included into effective potentials to some extent. Figure 2.2 A schematic illustration of all-electron (solid lines) and pseudo- (dashed lines) potentials and their corresponding wavefunctions. 42 2.1.8.1 Norm-conserving pseudopotentials To construct pseudopotentials, the following four general conditions should be fulfilled. First, the pseudo-wavefunction generated from the pseudopotential at each quantum number should be nodeless in the core region. Second, the normalized atomic radial pseudo-wavefunction with angular momentum should be equal to the normalized radial all-electron wavefunction beyond a cutoff radius rc , or converge rapidly to that value of the all-electron wavefunction. Third, in the region r < rc the charge of pseudo-wavefunction and the one of all-electron wavefunction must be equal. Lastly, the eigenvalue of the valence all-electron wavefunction and the pseudo potential eigenvalue must be equal. Pseudopotentials satisfying conditions mentioned above are called “norm-conserving” pseudopotentials. (Troullier and Martins, 1990) 2.1.8.2. Ultra-Soft (US) pseudopotentials Still, norm-conserving pseudopotentials require an expensive computational cost because its high cutoff energy results in the use of large number of plane wave. To overcome this computational constraint of norm-conserving pseudopotentials, Vanderbilt (1990) suggested a new approach – relaxing norm-conserving conditions and increasing the cutoff radius - to construct “ultra-soft” pseudopotentials, whose cutoff energy for plane wave calculation is drastically cut down by 30 Ry (≈ 400 eV) in average from those with all-electron method. The recipe to construct ultra soft pseudopotentials is well explained by Vanderbilt group (Vanderbilt, 1990; Laasonen et al., 1993). First, an all-electron calculation is performed on a free atom in some reference configuration. Then, a set of reference energies is chosen at which the good scattering properties are obtained for each 43 angular momentum. Next, the set of pseudo-wave-functions are constructed from the all-electron wave functions with no norm-conservation constraint. Finally, the cutoff radius is determined well beyond the radial wave-function maximum as displayed in Fig 2.2 and 2.3 satisfying the only constraint that pseudo-wave-functions should be matched to the all-electron wavefunctions at the cutoff radius. Kresse and Hafner (1994) developed accurate ultra soft pseudopotential construction schemes with good convergence properties for the first-row atoms and transition-metal systems. Figure 2.3 Comparison on pseudo-wavefunctions generated using the norm-conserving pseudopotential by Hamann, Schlüter and Chiang (dotted line) and US (dashed line) for the oxygen 2p orbital with regards to the oxygen 2p radial wavefunction (solid line). Adopted from Vanderbilt (1990) To visualize the difference between the norm-conserving pseudopotential method and the US pseudopotential method, the pseudo-wavefunction generated by Hamann, Schlüter and Chiang (HSC) recipe (1979) and US recipe for the oxygen 2p orbital are compared to the radial wavefunction in Figure 2.3. The cutoff radius applied in US recipe is larger than that of HSC recipe, implying that norm-conserving pseudopotential calculations require higher energy cutoff than US pseudopotential calculations need. 44 2.1.8.3 Projector augmented wave (PAW) method The projector augmented wave (PAW) method first proposed by Blöchl (1994) is an all-electron method for efficient ab initio electronic structure calculations. The PAW method is an electronic structure method with the combination of the augmented wave method and the pseudopotential approach. As depicted in Figure 2.2, the wavefunction oscillates rapidly close to the nucleus due to the large attractive potential of the nucleus, whereas in the bonding region the wavefunction is fairly smooth. This complicated behavior of wavefunction contributes to the computational difficulties in the highly accurate description of the bonding region, and explains the large variation of the atomic region. The augmented wave method tackles this issue by dividing the wavefunction into an atom-like partial-wave expansion in the atomic region and the envelop functions expanded into plane waves for the bonding description in the regions of between atoms. (Blöchl, 1994) The PAW method is built on a transformation, which maps the true wave functions with their complete nodal structure onto the numerically convenient auxiliary wavefunction. (Blöchl et al., 2003) The transformation makes it possible to expand a smooth auxiliary wavefunction into a rapidly converged plane wave, and to evaluate all physical properties from reconstructed physical wavefunction. An all-electron wavefunction for the valence state ψ is a full one-electron KohnSham wave function and can be retrieved from the auxiliary wavefunction ψ~ by the linear transformation determined by three quantities: all-electron partial wave φi , 45 ~ one auxiliary partial wave φi , and one projector function ~ pi as displayed in Eq. (2.26) Here, the project function is a fixed function introduced for the linear transformation of each auxiliary wavefunction. ~ ψ = ψ~ + ∑ ( φi − φi ) ~ pi ψ~ (2.26) i In the same way, the wavefunction for the core states ψ c can be decomposed into three contributions: an auxiliary core wave function ψ~ c , an all-electron core partial ~ wave φ c , and an auxiliary core partial wave φ c as expressed in Eq. (2.27). This is consistent with the frozen-core approximation. ~ ψ c = ψ~ c + φ c − φ c (2.27) Therefore, the all-electron wavefunction and auxiliary partial wave forms complete sets of functions expanded by the convenient plane wave basis sets within the augmentation region, also known as core region in the pseudopotential approach. 46 2.2 Computational Codes Both GAUSSIAN03 - molecular orbital code - and VASP - periodic plane wave code are employed to perform the total energy electronic structure calculations for the research of adsorption energies and electronic structures of adsorbate, transition metal surface and their complex. 2.2.1 Vienna Ab initio Simulation Package (VASP) The Vienna Ab initio Simulation Package (VASP) (version 4.6) (Kresse and Hafner, 1993; Kresse and Furthmüller, 1996) has been used for ab initio density functional theory calculations. VASP uses a plane wave basis set and the projector augmentedwave method or pseudopotentials to solve the Kohn-Sham equations. Two kinds of pseudopotential have been implemented in VASP. One is US pseudopotentials with the combination with both LDA and PW91 exchange-correlation functionals, the other is PAW pseudopotentials with one LDA XC functional and two GGA XC functionals – PW91 and PBE. The periodic plane wave DFT code, VASP provides periodic boundary conditions to perform surface calculations. For atoms and molecules, a simple cubic cell describes a molecule in gas phase. 2.2.2 Gaussian 03 Gaussian 03 (Frisch et al., 2003) is an electronic structure package used for atoms, molecules and reactive systems to predict molecular energies, structures, vibrational frequencies and electron densities. Gaussian code has been developed so that ab initio, 47 density functional theory, semi-empirical, molecular mechanics and various hybrid methods are implemented. The molecular orbital ab initio wavefunction-based code, Gaussian 03 enables to perform single-point energy calculations for any given structures, geometry optimization calculations to search the geometry at the stationary point on the potential energy surface, and vibrational frequency calculations to determine the second derivatives of the energy. 48 2.3 References Ashcroft, N.W., and Mermin, N.D., “Solid State Physics”, Harcourt College Publishers, 1976 Blöchl, P.E., “Projector augmented-wave method”, Physical Review B, 50, pp. 17953. 1994. Blöchl, P.E., Forst, C.J., and Schimpl, J., “Projector augmented wave method: ab initio molecular dynamics with full wavefunctions”, Bulletin of Material Science, 26, pp. 33. 2003. Boys, S.F., and Bernardi, F., “The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors”, Molecular Physics, 19, pp. 553. 1970. Davidson, E.R., and Feller, D., “Basis Set Selection for Molecular Calculations”, Chemical Reviews, 86, pp. 681. 1986. Duch, W., and Diercksen, G.H.F., “Size-extensivity corrections in Configuration-Interaction Methods”, Journal of Chemical Physics, 101, pp. 3018. 1994. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; and Pople, J. A.; GAUSSIAN 03. Gaussian, Inc., Pittsburgh PA, 2003 Hamann, D.R., Schlüter, M., and Chiang, C., “Norm-Conserving Pseudopotentials”, Physics Review Letters, 43, pp. 1494. 1979. Hoffmann, R. A., “Chemical and Theoretical way to look at bonding on surface”, Reviews of Modern Physics, 60(3), pp. 601. 1988. Jensen, F., “Introduction to Computational Chemistry”, New York, John Wiely & Sons. 1999. Johnson, B.G., Gill, P.M.W., and Pople, J.A., “The performance of a family of density functional methods”, Journal of Chemical Physics, 98, pp. 5612. 1993. Knowles, P., Schütz, M., and Werner, H.J., “Ab Initio Method for Electron Correlation in Molecules”, In Modern Methods and Algorithms of Quantum Chemistry, Vol. 1, ed by Grotendorst, J., pp. 69. Jülich, NIC Series. 2000. Kresse, G., and Hafner, J., “Ab initio molecular dynamics for liquid metals”, Physical Review B, 47, pp. 558. 1993. 49 Kresse, G., and Hafner, J., “Norm-conserving and ultra-soft pseudopotentials for first-row and transition elements”, Journal of Physics: Condensed Matter, 6, pp. 8245. 1994. Kresse, G., and Furthmüller, J., “Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set”, Computational Materials Science, 6, pp. 15. 1996. Koch, W., and Holthausen, M.C., “A Chemist’s Guide to Density Functional Theory”, Weinheim, WILEY-VCH. 2001. Kohn, W., and Sham, L., “Self-Consistent Equations Including Exchange and Correlation Effects”, Physical Review A, 140, pp. 1133. 1965. Laasonen, K., Pasquarello, A., Car, R., Lee, C., and Vanderbilt, D., “Car-Parrinello molecular dynamics with Vanderbilt ultrasoft pseudopotentials”, Physical Review B, 47, pp. 10142. 1993. Levine, I. N., “Quantum Chemistry”, New Jersey, Prentice Hall. 2000. Møller, C., and Plesset, M.S., “Note on an Approximation Treatment for Many-Electron Systems”, Physical Review, 46. pp. 618. 1934. Perdew, J.P., Chevary, J.A., Vosko, S.H., Jackson, K.A., Pederson, M.R., Singh, D.J., and Fiolhais, C., “Atoms, molecules, solids, and surfaces: Application of the generalized gradient approximation for exchange and correlation”, Physical Review B, 46, pp. 6671. 1992. Perdew, J.P., Burke, K., and Ernzerhof, M., “Generalized Gradient Approximation Made Simple”, Physical Review Letters, 77, pp. 3865. 1996. Pople, J. A., “Quantum Chemical Models”, Angewandt Chemie International Edition, 38, pp. 1894. 1999. Roothan, C.J., “New Developments in Molecular Orbital Theory”, Reviews of Modern Physics, 23, pp. 69. 1951. Santen, R.A., and Neurock, M., “Molecular Heterogeneous Catalysis: A Conceptual and Computational Approach”, Weinheim, WILEY-VCH. 2006. Troullier, N., and Martins, J.L., “Efficient Pseudopotentials for plane-wave calculations”, Physical Review B, 41, pp. 7892. 1990. Vanderbilt, D., “Soft self-consistent pseudopotentials in a generalized eigen value formalism”, Physical Review B, 41, pp. 7892. 1990. 50 Chapter 3 Benzene Chemisorption on Pt(111) In the study on the chemisorption of aromatic molecules on transition metals, benzene adsorption on the Pt(111) surface is a good starting system, since benzene and platinum typically represent aromatic molecules and transition metals, respectively. Benzene is the simplest aromatic molecule and has a relatively simple electronic structure due to its high symmetricity. Platinum is widely used as the industrial heterogeneous catalysts in petroleum reforming and in automotive exhaust gas converters. Benzene adsorption on Pt(111) has been extensively examined both experimentally and computationally. To understand benzene adsorption on the Pt(111) surface, two topical questions should be answered: (i) what is the preferential adsorption site, and (ii) what is the binding energy of chemisorbed benzene. Sholl (2006) suggested in his review paper on the application of density functional theory (DFT) to heterogeneous catalysis that three different sources of uncertainty in DFT calculations should be taken care of when comparing with experiments. First, the use of an approximate exchange-correlation functional contributes to the deviation of DFT results from the experimental values; second, DFT calculation must be numerically converged to eliminate mathematical uncertainties; last, the intrinsic complexity of heterogeneous catalysts should be accounted for when comparing DFT calculations with experimental results. 51 The objective of the current study is mainly to answer the two first questions to achieve a better understanding of electronic interaction of the benzene with a Pt(111) surface. The following sections consist of three parts. First, a converged low coverage DFT value for the benzene adsorption on the Pt(111) surface has been obtained through numerical convergence tests to remove mathematical uncertainty. Next, the accuracy of the DFT adsorption energy has been examined with the high level wavefunction based methods. Finally, the benzene adsorption study on Pt(111) has been extended to the industrially interesting region of high coverage to investigate coverage effects. 3.1 Converged DFT Benzene Adsorption Energy on Pt(111) The Pt(111) slab has high-symmetric sites: Atop, Bridge, Hollow-hcp and Hollow-fcc, while adsorbed benzene has two orientations on the Pt(111) surface: One orientation is 30º where symmetric C-C bonds are orthogonal to the horizontally arranged Pt atomic rows, and the other is 0º where symmetric C-C bonds are parallel with the horizontally layered Pt atoms. Hence, eight adsorption configurations are available for chemisorbed benzene on the Pt(111) surface. They are schematically illustrated in Fig. 3.1. Figure 3.1 Schematic of high-symmetry benzene adsorption sites on a Pt(111) surface (Saeys et al., 2002) 52 The adsorption energy of benzene on the Pt(111) surface is calculated from the following equation, where E denotes the total electronic energy of each component: ΔEads = Ebenzene Pt (111) − Ebenzene − EPt (111) (3.1) Like other adsorption properties, the adsorption energy also depends on surface coverage. Surface coverage can be defined differently in experimental work and in computational calculations. Experimental surface coverage is the number of adsorbed molecules on a surface divided by the number of molecules in a filled monolayer on that surface. 2.3×1014 molecules of adsorbed benzene per square centimeter of the Pt(111) surface at 300 K are experimentally measured as the saturated coverage. The theoretical definition of surface coverage is the number of adsorbed molecules divided by the number of surface atoms in a unit cell. For example, if one benzene molecule binds on the surface whose unit cell consists of nine Pt atoms, its theoretical surface coverage can be expressed as 1/9 ML, where ML stands for monolayer. In the current DFT study, the theoretical surface coverage term is preferred. However, experimental surface coverage, θ exp , can be converted to theoretical surface coverage, θ, according to Eq. 3.2. For example, experimental surface coverage of approximately 0.7 can be converted to 1/9 ML in theoretical surface coverage. θ = # of benzenes # of Pt atoms (2.3 × 10 ) × θ 14 = 2 exp 1cm area of a surface unit cell (3.2) 53 3.1.1 Review of the literature: benzene adsorption studies on Pt(111) Experimental studies of the benzene adsorption sites on the Pt(111) surface have been performed extensively with the development of surface spectroscopic techniques. High-resolution electron energy-loss spectroscopy (HREELS) vibrational analysis suggested that benzene may adsorb at two sites: Atop(0) and Hollow-hcp(0). (Lehwald et al., 1978; Cemie et al., 1995) Temperature Desorption Spectroscopy (TDS) studies also indicated that benzene adsorbs at more than one site. (Tsai and Muetterties, 1982) Angle Resolved Ultraviolet Photoelectron Spectroscopy (ARUPS) proposed adsorption at the Hollow-hcp(0) and the Hollow-fcc(0) sites. (Somers et al., 1987) Low Energy Electron Diffraction (LEED) supported Bridge site adsorption only. (Ogletree et al., 1987) At low coverage and at 4 K, Scanning Tunneling Microscopy (STM) studies by Weiss and Eigler (1993) showed several chemisorption sites occupied by benzene, which were identified as Hollow-hcp(0), Atop(0) and Bridge(30) by Sautet and Bocquet (1996). A Reflection-Adsorption Infra-Red Spectroscopy (RAIRS) study proposed both Bridge(30) and Hollow-hcp(0) sites were the dominant sites (Haq and King, 1996). Near Edge X-ray Adsorption Fine Structure (NEXAFS) supported Atop(0) and Atop(30) adsorption (Yimagawa and Fujikawa, 1996) Experimental results for the benzene adsorption sites on Pt(111) surface are summarized in Table 1.1. Unfortunately, these experimental methods fail to provide the conclusive answer to the most preferable adsorption site of benzene on the Pt(111) surface, though Hollow(0), Bridge(30) and Atop(0) are most cited. 54 Table 3.1 Summary of the experimental studies of the benzene adsorption sites on Pt(111) Atop Experimental Methods (0) HREELS √ Bridge (30) (0) Hollow (30) (30) √ ARUPS √ LEED STM (0) √ √ RAIRS NEXAFS √ √ Total 3 1 1 √ √ √ √ √ 3 4 0 Alternatively, quantum chemical calculations have been used to determine the benzene adsorption sites on Pt(111). The Pt(111) surface has been simulated either by the cluster approach or by a periodic slab method. In the cluster approach, a finite number of atoms are used to represent the catalyst surface, while the periodic slab method expands a surface unit cell repeatedly in the x and y directions with vacuum embedded between slabs to model an solid catalytic surface in contact with gas phase. Figure 3.2 Examples of a cluster model (left) and a periodic slab model (right) for benzene adsorption on Pt(111) 55 In quantum chemical computational methods, the adsorption energy at each site has been used to determine the preferred adsorption site. Benzene adsorption on the Pt(111) surface is an exothermic reaction so that the adsorption energy has a negative value if it forms the stable adsorption complex. The quantum chemically computed benzene adsorption energies at individual sites reported in the literatures are shown in Table 3.2. The semi-empirical atom superposition and electron delocalization molecular orbital (ASED-MO) calculations on benzene adsorption on Pt17 cluster (Anderson et al., 1984) indicated that Bridge(30) and Hollow-hcp(0) sites were most favored and the benzene adsorption at Atop sites seemed impossible. The extended Hückel study based on the experimental STM results (Sautet and Bocquet, 1994) whose adsorption energy values are the same as those the extended Hückel study with a periodic model by Minot et al. (1995) showed that the adsorption energy for both the Bridge(30) and the Hollow-hcp(0) were similar outweighing that of Atop(0) site. DFT total energy calculations with a Pt22 cluster (Saeys et al., 2002) as well as with a periodic slab model (Morin et al., 2003) proposed that Bridge(30) site was the preferred adsorption site for benzene on the Pt(111) surface followed by the Hollowhcp(0) site at low coverage, although a periodic slab calculation gave smaller adsorption energy values than a cluster calculation produced. Table 3.2 Adsorption energies at various benzene adsorption sites on Pt(111) in the literature Computational Methods Atop (kJ/mol) Bridge (kJ/mol) Hollow-hcp (kJ/mol) Hollow-fcc (kJ/mol) (0) (30) -65 -54 (0) (30) (0) (30) (0) (30) ASED-MO (Anderson et al., 1984) -9 -8 -71 -188 -158 -100 Extended Hückel (Minot et al., 1995) -87 -125 -125 DFT cluster (Saeys et al., 2002) -66 -102 -71 -51 DFT periodic slab (Morin et al., 2003) -43 -100 -73 -40 56 Using density functional theory both Saeys et al. (2002) and Morin et al. (2003) determined the preferred benzene adsorption site on the Pt(111) surface. Both papers provided vibrational frequency analysis of adsorbed benzene on Pt(111) similar to the HREELS results (Lehwald et al., 1978; Cemie et al., 1995) and RAIRS study (Haq and King, 1996) so that at low coverage most benzenes prefer to adsorb at Bridge(30) sites of the closed packed surface of platinum while some adsorbs at Hollow-hcp(0) and Hollow-fcc(0) sites. The electronic analysis based on the change of the orbital energies upon adsorption by Saeys et al. suggested that the Bridge(30) site was more preferred than Hollow-hcp(0) or Hollow-fcc(0) sites. This has been confirmed by Morin et al. who proposed that benzene molecules adsorbed at Hollow-hcp(0) and Hollow-fcc(0) site were minority species compared to those adsorbates at Bridge(30) sites with a ratio of 1 to 25 at low coverage. The binding energies of molecules on transition metal surfaces can be experimentally calibrated using either temperature programmed desorption (TPD) or single crystal adsorption calorimetry (SCAC), which is designed specifically to quantify the heats of adsorption of low vapor pressure molecules, such as benzene. TPD studies indicated the heat of adsorption of benzene on Pt(111) ranging from 117 to 129 kJ/mol at low coverage, (Xu et al., 1994; Campbell et al., 1989) while SCAC method proposed 197 kJ/mol of the heat of adsorption at 300 K. (Gottfried et al., 2006) Recently DFT periodic slab calculations for benzene adsorption energy on the Pt(111) surface at low coverage of 1/9 ML have been performed in VASP. Two types of pseudopotentials - ultrasoft (US) and projector augmented wave (PAW) are applied with the PW91 generalized gradient approximation (GGA) exchange-correlation 57 functional. The energy cutoff, which determines the plane wave basis set employed, has been set to 287 eV for US-PW91 and to 400 eV for PAW-PW91. The unit cell of p(3×3) is used to bind one benzene molecule resulting in a surface coverage of 1/9 ML. Four-layer-thick slabs have initially been used to determine the benzene adsorption energy, where the top two layers are relaxed upon the interaction with benzene and the bottom two layers are fixed at the Pt bulk-interlayer distance to model the bulk part of the catalyst. The six-layer-thick slab is also applied to examine the effect of the change in the Pt(111) slab thickness, where the top two layers are relaxed and the bottom four layers are fixed. The first Brillouin zone is a uniquely defined primitive cell of the reciprocal lattice, which has enough information for the electronic property calculations. To accurately reproduce the first Brillouin zone, VASP determines sets of special points for the Brillouin zone integration using the Monkhorst-Pack method. Table 3.3 Benzene adsorption energy on Pt(111) at 1/9 ML by DFT periodic slab calculation using VASP Adsorption Energy (kJ/mol) Source PP-XC # of layers in Slab k-point mesh Bridge(30) Hollow-hcp(0) Saeys et al. (2002) US-PW91 4 layers 2×2×1 -117 -75 Morin et al. (2003) US-PW91 4 layers 3×3×1 -77 -53 Morin et al. (2003) PAW-PW91 4 layers 3×3×1 -84 -60 Morin et al. (2003) PAW-PW91 4 layers 5×5×1 -87 -65 Morin et al. (2003) PAW-PW91 6 layers 3×3×1 -123 -94 Morin et al. (2003) PAW-PW91 6 layers 5×5×1 -100 -73 58 DFT adsorption energy results for benzene on Pt(111) at 1/9 ML coverage in Table 3.3 are almost conclusive. Still a few points should be clarified. First of all, the adsorption energy results at Bridge(30) site reported in both DFT papers show the variation of 40 kJ/mol despite using the same US pseudopotential with the PW91 XC functional. Second of all, the adsorption energy of -87 kJ/mol at the 4-layered-thick slab with PAW pseudopotential with the PW91 XC functional is far from both TPD experimental data of 117 ~ 129 kJ/mol and 197 kJ/mol from SCAC experiment though it is regarded as the most stable one (Morin et al., 2004). This indicates that the DFT periodic slab calculation has a significant deviation from experimentally measured heat of adsorption for benzene on Pt(111). Third of all, the -100 kJ/mol of adsorption energy at the 6-layered-thick Pt(111) slab with the k-points mesh of 5×5×1 seems converged with the number of k-points as well as with the slab thickness. To clarify aforementioned issues of previous DFT calculations, therefore, the converged DFT adsorption energy should be obtained. 3.1.2 Convergence test for DFT benzene adsorption energy on Pt(111) In this section, systematic convergence tests for the benzene adsorption energy on the Bridge(30) site of Pt(111) have been performed to obtain the converged DFT adsorption energy. In the following DFT calculations, a force convergence criterion of 0.01 eV/Å is imposed for the geometry optimization and an energy convergence criterion of 10-5 eV is imposed for the ground state energy minimization. First, the k-point convergence test has been performed increasing the number of kpoints till the adsorption energy difference converges within 4 kJ/mol. The number of 59 k-points has been increased with Monkhorst pack grid of 3×3×1 initially, intermediately increased grid of 5×5×1 and finally up to 7×7×1. Next, the convergence test for the vacuum thickness has been performed increasing the vacuum thickness by one bulk-interlayer distance (a0/√3) from around 9 Å equivalent to four bulk-interlayer distance to approximately 20 Å equivalent to nine bulkinterlayer distance. The function of the vacuum space between neighboring slabs is to prevent the dipole-dipole interaction between adjacent transition metal slabs. (Fig 3.3) dvacuum Top layer a0/√3 2nd Top layer a0/√3 Bottom layer dslab Figure 3.3 Schematic illustration of the Pt(111) slab (left) and the super cell including slab and vacuum (right). Lastly, the convergence test on the slab thickness has been conducted increasing the number of layers of Pt atoms from 3-layers to 6-layers as illustrated in Figure 3.4. The slab model consists of a surface part and a bulk part, where the two top-most layers are relaxed to describe the surface part and the rest are fixed at the bulk-interlayer distance for the description of the bulk part. The creation of a surface leads to an outward relaxation of the surface atoms for the Pt(111) slab and the distance between the surface and the subsurface layer increases by 60 0.8 to 1.0 %. The outward relaxation is typical for metals with a (nearly) filled d-band and is caused by a repulsive force. The d-d repulsion overcompensates the inward relaxation caused by an inward electrostatic force created by spreading of sp electrons into the vacuum (R. Smoluchowski, Phys. Rev. 60, 661, 1941). For the 4, 5, and 6layer slabs, the subsurface atoms move inward and the distance between the first and second subsurface layer decreases by 0.4 to 0.8 %. The inward relaxation is caused by charge redistribution. The 3-layer slab however does not follow this trend and the subsurface atoms move outward because of symmetry. Relaxation reduces the surface energy by 0.04 J/m2 for the 3-layer slab and by 0.004 J/m2 for thicker slabs. Upon benzene adsorption the surface atoms in direct contact with benzene relax outward by 0.125 to 0.164 Å. This relaxation contributes about 40 kJ/mol to the adsorption energy and should hence not be neglected. The surface relaxation upon adsorption is not uniform, the atoms in direct contact with benzene move up, while the other surface atoms move down. Table 3.4 Surface relaxation upon benzene adsorption for various Pt(111) slab models Pt Slab Activea Non-activeb a 3-layer 4-layer 5-layer 6-layer Surface 0.164 Å 0.131 Å 0.125 Å 0.148 Å Subsurface 0.047 Å 0.018 Å 0.015 Å 0.029 Å Surface -0.009 Å -0.033 Å -0.012 Å -0.006 Å Subsurface 0.009 Å -0.015 Å -0.014 Å -0.003 Å Pt atoms in direct contact with benzene, or directly below benzene. b Other atoms. 61 3-layered slab 4-layered slab 5-layered slab 6-layered slab 1st Top layer 2nd Top layer Fixed layer Fixed layer Fixed layer Fixed layer Figure 3.4 Schematic illustration of the Pt(111) slab models with various slab-thickness. The cutoff energy convergence test has been omitted first because the energy cutoff of PAW potentials provided in VASP has already been optimized and second because it is believed that increasing the number of plane waves in VASP has negligible effect on the adsorption energy only causing huge computational cost. For the following DFT calculations, PAW pseudopotentials with the PBE exchangecorrelation (XC) functional have been used following the recommendation of a VASP guide because the PBE implementation follows strictly the PBE prescription, while the PW91 implementation does not because the parameterization of Perdew and Zunger is used for the LDA part instead of Perdew-Pade approximation. (Kresse, 2004) 62 3.1.2.1 Convergence test on vacuum thickness To check the effect of vacuum thickness on the adsorption energy, the adsorption energy calculation at the Bridge(30) site on a Pt(111) 4-layered slab at 5×5×1 k-point mesh has been performed varying the vacuum thickness. Table 3.5 shows that when vacuum thickness is thinner than 14 Å equivalent to a 6-bulk-inter-layer distance, dipole-dipole interactions between slabs contribute to the adsorption energy overestimation. Hence, it is required to ensure sufficient vacuum space to exclude dipole-dipole interaction effect on adsorption energy term. On the other hand, increasing the vacuum thickness above 14 Å in the 4-layered slab is not efficient because it has no effect on the adsorption energy only increasing computation time. Therefore, setting 14 Å for vacuum thickness can be the best choice to exclude the dipole-dipole interaction and to reduce computational cost. Table 3.5 Convergence test with various vacuum thickness for benzene adsorption energy on Pt(111) at the Bridge(30) site Vacuum thickness 9Å 12 Å 14 Å 16 Å 18 Å 20 Å Adsorption Energy (kJ/mol) -90 -84 -81 -81 -79 -80 CPU time (hour) 99 97 106 138 141 212 3.1.2.2 Convergence test on the number of k-points and the slab thickness Fixing the vacuum thickness at 14 Å, the convergence test on slab thickness has been performed for the benzene adsorption energy at the Bridge(30) site on Pt(111). In each slab calculation, the k-point mesh is increased from 3×3×1 grid to 7×7×1 grid to ensure the k-point convergence. The adsorption energy calculation results are listed in Table 3.6 and graphically presented in Figure 3.5. 63 Table 3.6 Slab thickness convergence test for benzene adsorption energy on Pt(111) at Bridge(30) site along with k-point convergence test Number of slab layers Adsorption Energy (kJ/mol) 3 layers 4 layers 5 layers 6 layers 3×3×1 -134 -78 -119 -119 5×5×1 -132 -82 -107 -100 7×7×1 -125 -87 -107 -102 Benzene Adsorption Energy on Pt(111) at 1/9ML -75 Adsorption Energy -85 -95 k3x3x1 k5x5x1 -105 k7x7x1 -115 -125 -135 3L 4L 5L 6L Figure 3.5 Benzene adsorption energy at the Bridge(30) site with various slab thickness of Pt(111) slab models at different k-points mesh conditions at low coverage Table 3.6 and Figure 3.5 indicate that a numerically converged benzene adsorption energy in a p(3×3) unit cell is obtained for a 5×5×1 k-point mesh for the different slab models. Second, the benzene adsorption energy tends to converge for a 5-layer slab to a value of -107 kJ/mol. Third, the widely used 4-layer slab model produces the weakest adsorption energy while the 3-layer slab model gives the most strongest adsorption energy. 64 3.1.3 Electronic analysis of DFT benzene adsorption on Pt(111) Benzene adsorption on the Pt(111) surface can be understood by the electronic analysis using the d-band model and the projected density of states (PDOS). The d-band model is a simple concept to relate changes in the electronic structure of transition metal surfaces to changes in chemical reactivity and to help understand the variation in adsorption energies in heterogeneous catalysis. (Hammer & Nørskov, 2000) The first moment of the density of states projected onto the metal d-band states is called the center of the d-band, ε d , which describes variations in the adsorption energies quite well. If the d-band center is close to the Fermi energy, the d-band interacts with the LUMO than the HOMO resulted in the strong electron backdonation (Fig 3.6). The variation of d-band center energy can occur when the surface structure changes, such as (100) low index surface, step or kink. Much stronger electron back-donation effect may be expected in other Pt surface structures. Figure 3.6 Schematic illustration of the electron donation (left) and electron back-donation (right) in the d-band model. From Bligaard and Nørskov (2007). 65 First, the d-band model has been applied to explain the variation of the benzene adsorption energy for different slab models from 3-layered slab to 6-layered slab in terms of the changes in the electronic properties of the metal surface. The workfunction, d-band filling degree and the d-band center energy of Pt(111) slabs with and without chemisorbed benzene at various conditions are summarized in Table 3.7 and 3.8. Since Pt has 10 valence electrons per atom, the d-bands of Pt(111) should be nearly full similarly to the Pd(111) surface with 0.96 of d-band filling. (Pallassana and Neurock, 2000) But, the d-band filling of 3-layered and 4-layered slab models show significant deviation by 0.28 and 0.08, while the d-band filling of 5-layered and 6layered slab converges to 0.94 and 0.95 as illustrated in Fig 3.8. In the same way, the d-band center energy converges to -2.36 eV at 5-layered and 6-layered slab model. The work function of Pt(111) slab shows 5.70 ± 0.07 eV at most slab models except the 3layered slab model, which is close to the experimental data of 5.70 eV (Kiskinova et al., 1983). Furthermore, the change in d-band filling and d-band center upon benzene adsorption is very different at the 3-layered slab model and the other thicker slabs. It is presumed that electron donation effect plays a main role in the surface-adsorbate interaction at 3layered slab. On the other hand, the converged change of d-band filling and d-band center as slab goes thicker implies that balanced electron donation and back-donation have occurred upon benzene adsorption on the Pt(111) surface. 66 The workfunction change of Pt(111) upon benzene adsorption, ΔΦ , is related to the charge redistribution upon adsorption so that it can be interpreted in terms of donation and back-donation effect. (Mittendorfer et al., 2003) Negative value means stronger donation effect from the molecule, which makes the center of d-band downshifted, whereas positive value indicates back-donation effect from metal d-states, which pushes up the d-band center energy level to the Fermi energy as shown in Fig 3.6. ΔΦ = Φ C 6 H 6 / Pt (111) − Φ Pt (111) (3.3) Abon et al. (1985) experimentally measured a decrease in the workfunction of benzene by -1.52 eV upon adsorption of a monolayer of benzene. Such a decrease indicates a strong dominance of electron donation from benzene to Pt(111) over back-donation to the benzene LUMO. However, in the DFT-PBE calculations, donation and backdonation are calculated to be more or less balanced. The calculated workfunction change is largest for the 3-layered slab, where donation from the benzene π orbital (HOMO) to the surface d-states dominates. For the 4-layer slab the change in workfunction is very small and donation and back-donation have a similar magnitude. For the 5- and 6-layer slab back-donation is more pronounced than donation effect. This stronger donation can explain the highest benzene adsorption energy for the 3layered slab. The trend from donation to back-donation with increase in slab thickness further follows the change in binding energy since both donation and back-donation lead to stronger adsorption. The balance between donation and back-donation is further somewhat influenced by the grid size. For the 6-layer slab, back-donation is more pronounced for a 3×3×1 k-point grid, while donation dominates for the 5×5×1 and 7×7×1 grids. 67 Table 3.7 Surface electronic properties of various Pt(111) slabs Electronic Properties 3-layered slab 4-layered slab 5-layered slab 6-layered slab d-band filling 3×3×1 5×5×1 7×7×1 0.68 0.68 0.69 0.88 0.87 0.88 0.93 0.94 0.94 0.95 0.95 0.95 d-band center (eV) 3×3×1 5×5×1 7×7×1 -1.82 -2.12 -2.15 -2.19 -2.23 -2.25 -2.32 -2.36 -2.35 -2.34 -2.37 -2.36 Work function (eV) 3×3×1 5×5×1 7×7×1 5.86 5.85 5.87 5.71 5.71 5.71 5.64 5.67 5.68 5.69 5.77 5.74 Table 3.8 Surface electronic properties of various Pt(111) slabs with benzene Electronic Properties 3-layered complex 4-layered complex 5-layered complex 6-layered complex d-band filling 3×3×1 5×5×1 7×7×1 0.94 0.91 0.92 0.94 0.94 0.94 0.96 0.96 0.96 0.97 0.97 0.97 d-band center (eV) 3×3×1 5×5×1 7×7×1 -2.06 -2.29 -2.30 -2.31 -2.36 -2.36 -2.41 -2.43 -2.43 -2.42 -2.46 -2.45 Work function (eV) 3×3×1 5×5×1 7×7×1 5.72 5.75 5.74 5.71 5.70 5.72 5.68 5.70 5.70 5.71 5.69 5.71 Table 3.9 Changes of surface electronic properties of Pt(111) upon adsorption Electronic Properties 3-layered slab 4-layered slab 5-layered slab 6-layered slab d-band filling 3×3×1 5×5×1 7×7×1 0.26 0.23 0.23 0.06 0.07 0.06 0.03 0.02 0.02 0.02 0.02 0.02 d-band center (eV) 3×3×1 5×5×1 7×7×1 -0.24 -0.17 -0.15 -0.11 -0.13 -0.11 -0.09 -0.07 -0.08 -0.08 -0.09 -0.09 Work function (eV) 3×3×1 5×5×1 7×7×1 -0.14 -0.10 -0.13 0.00 -0.01 0.01 0.04 0.03 0.02 0.02 -0.08 -0.03 68 E − E fermi (eV ) Figure 3.7 Electronic density of states (DOS) projected to d-bands of the Pt(111) surface at various slab models. 69 Next, the density of states projected to the carbon 2pz orbital of benzene and Pt(111) surface d-bands is obtained to explain the electron interaction upon benzene adsorption on various Pt(111) slab models. Figure 3.8 Electronic density of states (DOS) projected to d-bands of Pt(111) (solid line) and C 2pz orbital for benzene (dotted line) chemisorbed at the Pt(111) surface. 70 As displayed in Fig 3.8, the molecule-surface interaction can be understood analyzing the interaction between frontier orbitals and the sp-band and d-band of the metal. The 1e1g(π) orbital (HOMO) and 1e2u(π*) orbital (LUMO) downshift to lower energy level, interact with the metal surface, and split into both bonding and anti-bonding orbitals. The 1e1g(π) orbital interacts with the metal sp-state so that it provides electrons to the surface, called “electron donation”, whereas the 1e2u(π*) orbital interacts with the metal d-band so that it becomes filled, called “electron back-donation”. Due to the electron donation, the d-band of metal surface becomes filled, while the empty LUMO of benzene becomes filled as the results of electron back-donation. Therefore, it is suggested that the 5-layered or 6-layered slab model be used for the accurate description of electronic interaction for benzene adsorption on the Pt(111) slab. 71 Figure 3.9 Electronic density of states (DOS) projected to d-bands of Pt(111) (solid line) and C 2pz orbital for benzene (dotted line) chemisorbed at the Pt(111) surface at various slab models. 72 3.1.4 Summary From the adsorption energy convergence tests and electronic structure analysis, it is possible to conclude that the 5-layered slab model satisfies every convergence criteria and provides numerically converged adsorption energy correctly describing electron donation and back-donation effects within the limitation of DFT calculations. Therefore, it is believed that the adsorption energy at the 5-layered slab with k-points mesh of 5×5×1, -107 kJ/mol, is acceptable as a converged value in DFT calculation with PAW-PBE pseudopotential for benzene adsorption on the Pt(111) slab at the low coverage of 1/9 ML. One of the main objectives of the current study is to obtain numerically converged DFT-PBE adsorption energies for benzene on Pt(111). Note that those values might still be very different from the “exact” adsorption energy – i.e. the exact solution of the Schrödinger equation. DFT-PBE is only one approximation to the Schrödinger equation, and other DFT methods such as LDA, GGA, meta-GGA, or hybrid functionals, are expected to show a different convergence behavior. However, we believe that our conclusions remain valid for related calculations at the DFT-PBE level, since the electronic properties of the surface are converged. 73 3.2 Accuracy of DFT Adsorption Energy on Pt(111) The achievement of an accurate result in DFT calculations for transition metal systems are really challenging firstly because the use of an approximate XC functional contributes to the deviation of DFT results from the true solution, secondly because DFT calculation must be numerically converged with basis set size, number of kpoints and unit cell size, etc., and thirdly because heterogeneous catalysts are in reality complex so that it is difficult to know among fcc(111), fcc(110), fcc(100), steps or defects, which sites involve in the most important catalytic reaction step. (Sholl, 2006) The objective of this section is to compare DFT-PBE adsorption energy for benzene on the Pt(111) surface with other computational chemistry methods, such as MP2, CCSD(T) and B3LYP. In the following sections, first, computational studies on CO adsorption on the Pt(111) surface are reviewed. Next, DFT adsorption energy results will be validated against wavefunction-based total energy calculations. Then, the DFT adsorption energy will be corrected with the high-level electron correlation methods, such as MP2 calculations. Finally, the HOMO (Highest Occupied Molecular Orbitals) - LUMO (Lowest Unoccupied Molecular Orbitals) gap in the gas phase molecules and the workfunction of the Pt(111) surface will be combined to relate DFT adsorption energy values with those of other computational chemistry methods. 74 3.2.1 Review of computational studies of CO adsorption on Pt(111) The chemisorption of carbon monoxide on transition metal surfaces is an important elementary step in catalytic reactions such as CO oxidation, CO hydrogenation and Fischer-Tropsch synthesis. (Somorjai, 1994) One technological application example is the catalyst in the car exhaust to promote the oxidation of CO to CO2. The CO/Pt(111) system has become a benchmark system to test the accuracy of first principles calculations. Experimental studies of CO adsorption on a Pt(111) surface clearly indicate that, under ultrahigh vacuum condition, carbon monoxide binds at the Atop sites exclusively at low coverage and at the Atop site and the Bridge sites at high coverage. Low-energy electron diffraction (LEED) analysis indicates that CO molecules adsorb both Atop sites and Bridge sites at one-half monolayer coverage at 150 K (Ogletree et al., 1986). Using LEED and EELS (Electron Energy Loss Spectroscopy), Steininger et al. (1982) found that for p(√3×√3)R30º structure, carbon monoxide adsorbed at the Atop sites exclusively, half of the CO occupied Bridge sites and the other half Atop sites is found where at a high coverage a c(4×2)-2CO/Pt(111) structure. Scanning tunneling microscopy (STM) studies reconfirmed the Atop-Bridge configuration at the coverage of θ CO = 0.5 ML. (Stroscio and Eigler, 1991; Bocquet and Sautet, 1996; Pedersen et al., 1999) The experimentally measured CO chemisorption energy on the Pt(111) surface using single crystal adsorption calorimetry (SCAC) is 116 kJ/mol at a coverage of 0.5 ML and 183 kJ/mol at low coverage. (Yeo et al., 1996). The binding energy difference 75 between the two different sites at high coverage has empirically been calibrated as 6 kJ/mol by constructing a potential energy surface for CO/Pt(111). (Schweizer et al., 1989) This narrow energy gap functions as a stringent test for any quantum calculation schemes. On the other hand, DFT calculations have faced qualitative accuracy problems in the prediction of the chemisorption energy and of the most preferred adsorption site since the CO/Pt(111) puzzle reported by Feibelman et al. (2001), which highlighted the discrepancy between DFT predictions and experimental results for the preferred CO adsorption site on Pt(111) surface. All DFT calculations with various computer codes in their paper produced more stable CO binding energies at Hollow-fcc site of the Pt(111) surface rather than at Atop site by 0.10 ~ 0.45 eV. In addition, the same discrepancy in the preferred adsorption site has been identified at the Cu(111) surface contradicting angle-resolved photoemission extended fine structure (ARPEFS) results of Atop site binding at the coverage of 1/3 ML. Since the paper of Feibelman et al. (2001) on the CO/Pt(111) puzzle, the influence of the exchange-correlation functional error both on the CO binding energy and on site preference has been explored and a few compensation remedies have been proposed. Grinberg et al. (2002) investigated the influence of the pseudopotential and the exchange-correlation functional on the CO binding energy and site preference. They discovered that the pseudopotential error was negligible once converged, and that the DFT-GGA calculation showed bond energy error in molecules with bond-order change. For instance, bond energies for single bond molecules such as C-H, N-H and O-H 76 were accurately calculated with the error of less than 0.2 eV in DFT-GGA methods, but the bond energies for double bonded molecules, e.g. C=C and O=O were significantly overestimated showing 0.4 eV and 1.0 eV of error respectively, while the bond energies for triple bond molecules NO, CO and N2 were underestimated with errors of 0.8, 0.4 and 0.6 eV, respectively. Finally, it was concluded that DFT-GGA treated different bond orders with unequal accuracy whenever CO bonds were broken and formed on the Pt(111) surface, which caused a significant error in DFT-GGA description of the surface interaction with a CO molecule. The CO-metal surface interaction can be conceptually understood with the Blyholder (1964) model, which rationalizes a simple molecular orbital diagram of the nature of the metal-carbon-oxygen bond in terms of electron donation from the CO 5σ HOMO to the metal and electron back-donation from the metal to the CO 2π* LUMO. Another view introduced by Föhlisch et al. (2000) is that CO chemisorption on metal surface is the results of a balance between the σ repulsive interaction and π attractive interaction, where both interactions are increasing with higher coordination of adsorption sites, such as Atop < Bridge < Hollow. Note that the π attractive interaction causes the internal CO bond weakening. Hammer et al. (1996) proposed a “d band model” to describe the interaction between the metal d states and the CO 2π* and 5σ states. Kresse et al. (2003) found that CO-metal surface bond was formed as a result of the metal d states interactions with both CO 5σ and 2 π* orbitals, and electron donation pushed CO to the top site, whereas the back-donation forced CO to the Hollow site. They argued that LDA and GGA XC functionals put CO 2π* LUMO in the lower energy level, which causes the overestimation of the interaction with the metal, leading 77 to the wrong prediction for CO preferred adsorption site on the Pt(111) surface. In order to reduce the interaction between the LUMO and the metal d states, LUMO has been shifted to higher energies with the modified density functional, GGA+U. It has been designed with the introduction of a strong intra-atomic interaction in a screened Hartree-Fock manner for the description of on-site Coulomb interaction. DFT calculation results using a GGA+U type functional presented a linear dependence of CO chemisorption energy on the energetic position of the CO 2π* orbital as a result of correction the energy level of LUMO for gas-phase CO to higher energies. The XC functional modification study by Kresse et al. (2003) ignited more extensive studies to focus on the improvement of the exchange-correlation functionals. Doll (2004) succeeded to predict correctly the most favorable adsorption site for CO on Pt(111) with the B3LYP hybrid functional which produces a wider HOMO-LUMO gap than PW91 gradient corrected functional by 2.4 eV analyzing the HOMO-LUMO gap of the chemisorbed CO both fcc and top site on the Pt(111) surface. Gil et al. (2003) supported the superiority of B3LYP to PW91 in the description of HOMOLUMO gap, whose values are 9.5 eV and 6.8 eV, respectively. However, their DFTB3LYP cluster calculations with Pt4, Pt10, Pt13, Pt19 and Pt52 and DFT-GGA periodic slab calculation failed to predict the correct adsorption site. Mason et al. (2004) proposed an empirical adsorption energy correction scheme based on the internal CO stretch vibrational frequency. First, a simple first-principles correction term is determined from the difference between the singlet-triplet excitation energy of the DFT-GGA calculation and the value of a Configuration Interaction calculation. It was perceived that the DFT-GGA reproduced the unrealistically small 78 singlet-triplet excitation energy, whereas coupled-cluster quantum chemical calculation accurately reproduced the experimental values. The singlet-triplet excitation energy is closely related the HOMO-LUMO gap of CO and the energy level of the LUMO because the CO triplet state is the product of an excitation of an electron from the 5σ orbital to the 2π* orbital. Next, the chemisorption energy correction term is determined using the slope of the chemisorption energies plotted against the singlettriplet excitation energy difference, the singlet-triplet excitation energy difference between DFT-GGA calculations and high-level wavefunction based methods, such as CI method, as graphically shown in Figure 3.10. The difference of the chemisorption energy correction values at different adsorption sites implies that the DFT-GGA treats a π attractive back-donation unequally hence it describes the CO bond weakening effect differently at the Atop and the Hollow sites. Finally, the correction energy term can be obtained based on the internal CO stretch frequency because they have a negative linear relationship. Using PBE GGA XC functional, this empirical adsorption energy correction reproduced experimental adsorption site preference for the (100) and (111) surfaces of Pt, Rh, Pd and Cu as well as the chemisorption energy values in excellent agreement. Furthermore, Abild-Pedersen and Anderson (2007) applied the same correction scheme to more extended systems with the RPBE functional and achieved changes of the adsorption site preference for 7 of the metals – Co(0001), Cu(111), Pt(111), Rh(111), Ru(0001), Fe(110) and Mo(110). 79 Figure 3.10 First-principles extrapolation procedure based on the plot of adsorption energy for CO on Pt(111) Hollow-hcp site versus the singlet-triplet excitation energy difference. From Mason et al. (2004). Hu et al. (2007) developed a systematic approach to correct XC errors of the DFT slab calculations exploring the CO/Cu(111) and the CO/Ag(111) systems. First, DFT slab calculations with LDA or GGA XC functionals are performed in a super-cell with the relaxed five-layer slabs modeling CO adsorption at low coverage of 1/9 ML. Next, DFT cluster-calculations are run with the same XC functional and the same geometry as the previous step, then, corresponding cluster-calculation for exactly the same cluster is performed employing an improved exchange correlation treatment methods, such as, B3LYP functional, HF plus MP2 or CCSD(T), etc. Here, the XC correction energy, E XCcorr , can be obtained by subtracting the adsorption energy with LDA or GGA functional from the improved XC treatment calculation result. E XCcorr cluster cluster = Eads ( XC − better ) − Eads ( XC − less ) (3.3) 80 The XC correction energy difference between Atop site and Hollow site converges as the cluster size increases up to 16 atoms for the case of Cu cluster. Finally, using the converged XC correction energy difference term, the DFT-LDA or GGA error of the slab calculation can be corrected. Hu et al. (2007) succeeded to correctly predict the preferred CO adsorption site on Cu(111) at low coverage applying the XC energy correction scheme at the B3LYP level and for the case of Ag(111) both GGA level and B3LYP level XC energy correction scheme were able to predict the experimentally preferred adsorption site. 3.2.2 Exchange-Correlation (XC) correction approach In this section, a systematic approach proposed by Hu et al. (2007) will be applied to determine the benzene adsorption energy on Pt(111) slab. First, the DFT periodic slab calculation for benzene adsorption on the Pt(111) surface has been obtained from the literature and the benzene adsorption energy on a small Pt3 cluster system has been obtained in VASP with the PBE exchange-correlation functional. Next, the same cluster calculation has been run in Gaussian03 with the same PBE functional minimizing the basis set superposition error. Then the DFT adsorption energy has been calculated with the B3LYP hybrid XC functional in order to obtain the XC correction energy term. MP2 and CCSD(T) calculations are also performed using the same geometry to treat electron correlation more accurately. Finally, the adsorption energy values from different computational chemistry methods will be combined and the accurate adsorption energy will be predicted. 81 3.2.2.1 DFT adsorption energy in a periodic slab calculation To compare with benzene adsorption energy, various molecules are selected as testing systems, such as methyl, carbon monoxide, ethylene and 1,3-butadene. CH3 can give a simple view of carbon adsorption on the metal surface, carbon monoxide and ethylene are extensively studied molecules in the adsorption on the Pt(111) surface, and 1,3butadiene is also an interesting unsaturated hydrocarbon. The DFT molecular adsorption energies on a 5-layer Pt(111) slab obtained from the DFT-PBE calculation with a k-point grid of 5×5×1 are summarized in the Table 3.10 Table 3.10 DFT adsorption energy results for molecules on Pt(111) slab Molecules CO CH3 C2H4 C4H6 C6H6 Method DFT-PBE DFT-PBE DFT-PBE DFT-PBE DFT-PBE Coverage (ML) 0.11 0.11 0.11 0.11 0.11 Adsorption Site Hollow-fcc Atop Bridge(di-σ) Bridge(di-σ/π) Hollow-hcp(0) Eads (kJ/mol) -176 -199 -117 -169 -71 3.2.2.2 DFT-PBE adsorption energy on a small cluster in VASP To compare two different types of DFT methods - the plane wave periodic calculation and the molecular orbital basis set method, the adsorption energy on the Pt3 cluster of CO, CH3, C2H4, C4H6 and C6H6, has been calculated in both VASP and Gaussian03. Here the VASP calculation results will be presented and the adsorption energy computed in Gaussian03 will be given in the next section. In VASP, PAW pseudopotentials with the PBE XC functional have been used for all platinum, carbon, oxygen and hydrogen atoms with a cutoff energy of 400 eV. 82 Molecules, the Pt3 cluster and the complex are placed in a cubic cells of 14 ×14 × 14 Å3 and only a single k-point is used for all geometry optimization calculations. For molecules, spin-polarized geometry optimized calculations have been performed with Gaussian smearing method with 0.01 of sigma setting the force convergence criteria of 0.01 eV/Å. Regarding the Pt3 cluster calculation, spin-polarized geometry optimization has been done using the same convergence criteria as for molecules. The Pt3 cluster has two spin states: a singlet spin state and a triplet spin state. The triplet spin sate of the Pt3 cluster has been found -8 kJ/mol more stable with the nearest Pt-Pt distance of 2.49 Å than the singlet spin state of the Pt3 cluster. To find out the optimized adsorption complexes of CO, CH3, C2H4, C4H6 and C6H6 on Pt3 cluster, spinpolarized geometry optimization calculations have been run with the force convergence criteria of 0.01 eV/Å. The DFT-PBE molecular adsorption energy values on the Pt3 cluster are listed in Table 3.11 and optimized structures of adsorption complexes are illustrated in Figure 3.11. The molecular bond length of adsorbed molecules are elongated upon binding to the Pt3 cluster by 0.07 Å for the C=O double bond, by 0.18 Å for the C=C double bond in ethylene, by 0.07~0.08 Å for the C-C bonds in benzene as results of electron backdonation from the Pt3 cluster. For the Pt3 cluster, the Pt-Pt distance becomes longer as the size of adsorbates become larger, because large molecules have more electrons to transfer to the surface. For example, the Pt-Pt distance of the Pt3 cluster with CO adsorbate is 2.63 Å, while it elongates to 2.87 Å for benzene adsorption. 83 Table 3.11 Adsorption energy results for molecules on Pt3 cluster in VASP CO CH3 C2H4 C4H6 C6H6 Adsorption Site Hollow Atop Bridge(di-σ) di-Bridge Hollow(0) Eads (kJ/mol) -214 -241 -155 -206 -70 1.11 Molecules 1.10 2. 50 50 2. 2. 02 1.99 1.21 1.10 2.63 2.52 CO on Hollow site of Pt3 cluster CH3 on Atop site of Pt3 cluster 0 1.1 1.49 5 1.100 2.10 2.03 2.03 1.51 2.0 0 0 2.10 1.1 1.455 5 1.49 1.100 2. 65 2.65 2.52 2.70 C4H6 on di-Bridge site of Pt3 cluster C2H4 on Bridge site of Pt3 cluster 1.4 9 1. 09 09 1. 1.0 1.462 75 1.4 75 2.8 2.8 7 2. 08 2.08 2. 0 8 2.08 8 2. 0 2.08 1.475 7 2.87 C6H6 on Hollow site of Pt3 cluster Figure 3.11 Adsorption structures of molecules adsorbed on a Pt3 cluster, optimized in VASP. 84 3.2.2.3 DFT-PBE adsorption energy on a small cluster in G03 In Gaussian03, the DFT-PBE molecular adsorption energy calculations on the Pt3 cluster have been performed using the adsorption geometries optimized in VASP. The single point energy calculation has been performed checking the stability of DFT wavefunction to ensure the true minimum energy solutions. For carbon, oxygen and hydrogen atom, correlation consistent basis sets (Dunning, 1989) and balanced basis sets (Weigend and Ahlrichs, 2005) are used. For platinum, the SDD basis set, which uses primitive Gaussians for the valence electrons of heavy atoms implementing the Stuttgart/Dresden effective core potential (ECP) including relativistic effects within Dirac-Fock theory, is adopted due to its popularity and accuracy. Generally, the SDD valence basis set does not include polarization functions. To model the outer valence electron and its behavior, three relativistic basis sets have been designed adding (2s2p1d) diffuse functions denoted as “++” and three different polarization functions: (1f1g), (2f1g), and (3f2g). (Dyall, 2004) In addition, one polarized triple zeta valence basis set (Def2-TZVPP) and two polarized quadruple zeta valence basis sets (Def2-QZVP and Def2-QZVPP) are used for Pt with the small-core effective core potential, denoted as Def2-ECP (Metz et al., 2000). The number of primitives and contractions of the basis sets are explicitly specified: the parentheses () displays the number of primitives that are given in the order of angular momentum quantum number and square brackets [] are used to specify the number of resulting contractions. 85 Table 3.12 Details of the basis sets used in the molecular calculations Basis sets H C&O cc-pVDZ (4s,1p) -> [2s,1p] (9s,4p,1d) -> [3s,2p,1d] aug-cc-pVDZ (5s,2p) -> [3s,2p] (10s,5p,2d) -> [4s,3p,2d] cc-pVTZ (5s,2p,1d) -> [3s,2p,1d] (10s,5p,2d,1f) -> [4s,3p,2d,1f] aug-cc-pVTZ (6s,3p,2d) -> [4s,3p,2d] (11s,6p,3d,2f) -> [5s,4p,3d,2f] Def2-TZVPP (5s,2p,1d) -> [3s,2p,1d] (11s,6p,2d,1f) -> [5s,3p,2d,1f] Def2-QZVP (7s,3p,2d,1f) -> [4s,3p,2d,1f] (15s,8p,3d,2f,1g) -> [7s,4p,3d,2f,1g] Def2-QZVPP (7s,3p,2d,1f) -> [4s,3p,2d,1f] (15s,8p,3d,2f,1g) -> [7s,4p,3d,2f,1g] Table 3.13 Details of the valence basis sets and effective core potentials for the Pt atoms Valence basis sets SDD (8s,7p,6d) -> [6s,5p,3d] SDD++(1f1g) (10s,9p,7d,1f,1g) -> [8s,7p,4d,1f,1g] SDD++(2f1g) (10s,9p,7d,2f,1g) -> [8s,7p,4d,2f,1g] SDD++(3f2g) (10s,9p,7d,3f,2g) -> [8s,7p,4d,3f,2g] Def2-TZVPP (8s,7p,6d,2f,1g) -> [6s,4p,3d,2f,1g] Def2-QZVP (10s,8p,6d,3f,1g) -> [7s,5p,4d,3f,1g] Def2-QZVPP (10s,8p,6d,4f,2g) -> [7s,5p,4d,4f,2g] Effective Core Potentials SDD ECP PT 0 PT-ECP 5 60 H POTENTIAL 1 2.000000 1.00000000 S-H POTENTIAL 2 2.000000 13.42865100 2.000000 6.71432600 P-H POTENTIAL 2 2.000000 10.36594400 2.000000 5.18297200 D-H POTENTIAL 2 2.000000 7.60047900 2.000000 3.80024000 F-H POTENTIAL 1 2.000000 3.30956900 G-H POTENTIAL 1 2.000000 5.27728900 Def2 ECP 0.00000000 579.22386100 29.66949100 280.86077400 26.74538200 120.39644400 15.81092100 24.31437600 PT 0 PT-ECP 3 60 f-ul potential 1 2.000000 3.30956857 s-ul potential 3 2.000000 13.42865130 2.000000 6.71432560 2.000000 3.30956857 p-ul potential 3 2.000000 10.36594420 2.000000 5.18297210 2.000000 3.30956857 d-ul potential 3 2.000000 7.60047949 2.000000 3.80023974 2.000000 3.30956857 24.31437573 579.22386092 29.66949062 -24.31437573 280.86077422 26.74538204 -24.31437573 120.39644429 15.81092058 -24.31437573 -24.21867500 86 To ensure the accuracy of the employed basis set, the basis set superposition error (BSSE) is calculated following the procedure proposed by Boys and Bernardi (1970). BSSE results from the overestimation in the chemical bonding due to the incomplete basis set for each fragment. Take benzene on the Pt3 cluster, the wavefunction of benzene or Pt3 cluster is expanded in much fewer basis functions than the wavefunction of the benzene adsorption complex so that it makes the complex more stabilized resulting in an overestimation of the adsorption energy. Therefore, to estimate an overestimation of BSSE from the adsorption energy, the energy should be corrected for the BSSE. The BSSE-corrected adsorption energy has been calculated as follow: First, to calibrate the adsorbate basis set effect on Pt3 cluster, a single point energy calculation is performed for the Pt3 cluster after removing the adsorbate atoms but keeping the corresponding basis functions at the positions from the optimized adsorption structure. Next, to determine the effect of Pt basis set effect on the stability of the benzene molecule, the single point energy calculation is performed removing Pt atoms but keeping the corresponding basis functions, where adsorbates and Pt3 cluster are fixed at the same position in the optimized adsorption complex. [ ΔEads ( A − Pt3 ) BSSE = E a ( A) A + E b ( Pt3 ) Pt3 − E a&b ( A − Pt3 ) A− Pt3 [ + [E + E a&b ( A) A− Pt3 − E a ( A) A− Pt3 a&b ] ( Pt3 ) A− Pt3 − E b ( Pt3 ) A− Pt3 ] ] (3.4) 87 Here, ΔEads ( A − Pt3 ) BSSE in left hand side means the BSSE-corrected adsorption energy of the adsorbate on the Pt3 cluster. The superscript indicates the basis set: a for the basis set for the adsorbate and b for the valence basis set for the Pt atoms. The subscript denotes the geometry: A stands for the optimized adsorbate, Pt3 for the optimized Pt3 cluster, and A-Pt3 for the optimized the adsorption complex. In the Eq. (3.4), the first term is the adsorption energy of molecules on Pt3 cluster without BSSE correction, while the second term for the Pt3 cluster basis set effect on the adsorbate, and the final term accounts for the adsorbate basis set effect on the Pt3 cluster. All three terms contribute to the BSSE-corrected adsorption energy. The DFT-PBE molecular binding energies on the Pt3 cluster with and without BSSE correction are listed in Table 3.13 and Table 3.14. The BSSE corrected binding energy results are consistent with the well-known basis size effect on BSSE. The BSSE contribution of molecule fragment with aug-cc-pVDZ basis set is larger than 4 kJ/mol, but it decreases below 2 kJ/mol with aug-cc-pVTZ basis set. (Table 3.14) In the same way, the larger the Pt3 cluster basis set becomes, the smaller its contribution to BSSE becomes, reducing from 5 kJ/mol to below 3 kJ/mol. The BSSE contribution of Pt basis with the Def2-ECP basis is reduced to below 2 kJ/mol. (Table 3.15) Both SDD++(3f2g) basis set for Pt with SDD ECP and Def2_QZVP basis set for Pt with Def2 ECP give similar BSSE magnitude and BSSE-corrected DFT-PBE molecular adsorption energies. In general, DFT-PBE calculations in VASP and Gaussian03 give a similar adsorption energy on the Pt3 cluster, as expected. 88 Table 3.14 Comparison of the DFT-PBE binding energies on a Pt3 cluster with different basis sets in Gaussian03 Basis set combination Adsorbate CO CH3 C2H4 C4H6 C6H6 Adsorption Energy (kJ/mol) BSSE (kJ/mol) Pt basis set Molecular basis set BSSE corrected BSSE uncorrected Pt3 contribution Molecular contribution SDD++(1f1g) aug-cc-pVDZ -220 -227 2.38 4.38 SDD++(2f1g) aug-cc-pVDZ -219 -226 2.39 4.46 SDD++(3f2g) aug-cc-pVDZ -221 -228 1.22 6.42 SDD++(1f1g) aug-cc-pVTZ -221 -226 4.28 0.82 SDD++(2f1g) aug-cc-pVTZ -221 -226 4.35 0.84 SDD++(3f2g) aug-cc-pVTZ -214 -217 1.77 1.12 SDD++(1f1g) aug-cc-pVDZ -249 -253 1.01 2.80 SDD++(2f1g) aug-cc-pVDZ -250 -254 1.01 2.85 SDD++(3f2g) aug-cc-pVDZ -252 -257 0.69 4.45 SDD++(1f1g) aug-cc-pVTZ -247 -249 1.82 0.12 SDD++(2f1g) aug-cc-pVTZ -248 -250 1.82 0.12 SDD++(3f2g) aug-cc-pVTZ -248 -249 1.14 0.16 SDD++(1f1g) aug-cc-pVDZ -160 -167 1.90 5.02 SDD++(2f1g) aug-cc-pVDZ -160 -167 1.90 5.14 SDD++(3f2g) aug-cc-pVDZ -161 -169 1.25 7.68 SDD++(1f1g) aug-cc-pVTZ -158 -162 3.12 0.34 SDD++(2f1g) aug-cc-pVTZ -158 -162 3.11 0.35 SDD++(3f2g) aug-cc-pVTZ -158 -160 1.95 0.49 SDD++(1f1g) aug-cc-pVDZ -206 -216 2.91 7.36 SDD++(2f1g) aug-cc-pVDZ -206 -216 2.91 7.49 SDD++(3f2g) aug-cc-pVDZ -206 -219 1.74 11.15 SDD++(1f1g) aug-cc-pVTZ -203 -209 4.80 0.66 SDD++(2f1g) aug-cc-pVTZ -203 -209 4.80 0.68 SDD++(3f2g) aug-cc-pVTZ -202 -206 2.75 1.03 SDD++(1f1g) aug-cc-pVDZ -83 -94 3.42 8.08 SDD++(2f1g) aug-cc-pVDZ -83 -95 3.44 8.19 SDD++(3f2g) aug-cc-pVDZ -83 -97 2.28 12.06 SDD++(1f1g) aug-cc-pVTZ -79 -85 5.16 1.12 SDD++(2f1g) aug-cc-pVTZ -79 -86 5.18 1.15 SDD++(3f2g) aug-cc-pVTZ -76 -81 3.27 1.72 89 Table 3.15 Comparison of the DFT-PBE binding energies on a Pt3 cluster with different valence basis sets for Pt in Gaussian03 Basis set combination Adsorbate CO CH3 C2H4 C4H6 C6H6 Adsorption Energy (kJ/mol) BSSE (kJ/mol) Pt basis set Molecular basis set BSSE corrected BSSE uncorrected Pt3 contribution Molecular contribution SDD++(1f1g) aug-cc-pVTZ -221 -226 4.28 0.82 SDD++(2f1g) aug-cc-pVTZ -221 -226 4.35 0.84 SDD++(3f2g) aug-cc-pVTZ -214 -217 1.77 1.12 Def2-TZVPP aug-cc-pVTZ -217 -223 4.53 0.75 Def2-QZVP aug-cc-pVTZ -217 -221 2.34 0.96 Def2-QZVPP aug-cc-pVTZ -217 -219 1.26 1.03 SDD++(1f1g) aug-cc-pVTZ -247 -249 1.82 0.12 SDD++(2f1g) aug-cc-pVTZ -248 -250 1.82 0.12 SDD++(3f2g) aug-cc-pVTZ -248 -249 1.14 0.16 Def2-TZVPP aug-cc-pVTZ -247 -249 2.08 0.10 Def2-QZVP aug-cc-pVTZ -247 -248 1.18 0.14 Def2-QZVPP aug-cc-pVTZ -247 -248 0.48 0.16 SDD++(1f1g) aug-cc-pVTZ -158 -162 3.12 0.34 SDD++(2f1g) aug-cc-pVTZ -158 -162 3.11 0.35 SDD++(3f2g) aug-cc-pVTZ -158 -160 1.95 0.49 Def2-TZVPP aug-cc-pVTZ -157 -161 3.66 0.33 Def2-QZVP aug-cc-pVTZ -157 -160 1.95 0.43 Def2-QZVPP aug-cc-pVTZ -157 -158 0.94 0.50 SDD++(1f1g) aug-cc-pVTZ -203 -209 4.80 0.66 SDD++(2f1g) aug-cc-pVTZ -203 -209 4.80 0.68 SDD++(3f2g) aug-cc-pVTZ -202 -206 2.75 1.03 Def2-TZVPP aug-cc-pVTZ -201 -207 5.11 0.67 Def2-QZVP aug-cc-pVTZ -202 -205 2.57 0.84 Def2-QZVPP aug-cc-pVTZ -200 -202 1.20 1.02 SDD++(1f1g) aug-cc-pVTZ -79 -85 5.16 1.12 SDD++(2f1g) aug-cc-pVTZ -79 -86 5.18 1.15 SDD++(3f2g) aug-cc-pVTZ -76 -81 3.27 1.72 Def2-TZVPP aug-cc-pVTZ -75 -82 5.56 1.17 Def2-QZVP aug-cc-pVTZ -75 -79 2.93 1.40 Def2-QZVPP aug-cc-pVTZ -72 -76 2.01 1.83 90 Table 3.16 Number of basis functions used for each calculation Basis set combination Adsorbate CO CH3 C2H4 C4H6 C6H6 # of basis functions used Pt basis set Molecular basis set Basis function Primitive Gaussians Cartesian basis function Total SDD++(1f1g) aug-cc-pVTZ 287 458 344 1,089 SDD++(2f1g) aug-cc-pVTZ 308 488 374 1,170 SDD++(3f2g) aug-cc-pVTZ 356 563 449 1,368 Def2-TZVPP aug-cc-pVTZ 260 446 323 1,029 Def2-QZVP aug-cc-pVTZ 308 491 383 1,182 Def2-QZVPP aug-cc-pVTZ 356 566 458 1,380 SDD++(1f1g) aug-cc-pVTZ 310 466 364 1,140 SDD++(2f1g) aug-cc-pVTZ 331 496 394 1,221 SDD++(3f2g) aug-cc-pVTZ 379 571 469 1,419 Def2-TZVPP aug-cc-pVTZ 283 454 343 1,080 Def2-QZVP aug-cc-pVTZ 331 499 403 1,233 Def2-QZVPP aug-cc-pVTZ 379 574 478 1,431 SDD++(1f1g) aug-cc-pVTZ 379 566 444 1,389 SDD++(2f1g) aug-cc-pVTZ 400 596 474 1,470 SDD++(3f2g) aug-cc-pVTZ 448 671 549 1,668 Def2-TZVPP aug-cc-pVTZ 352 554 423 1,329 Def2-QZVP aug-cc-pVTZ 400 599 483 1,482 Def2-QZVPP aug-cc-pVTZ 448 674 558 1,680 SDD++(1f1g) aug-cc-pVTZ 517 766 604 1,887 SDD++(2f1g) aug-cc-pVTZ 538 796 634 1,968 SDD++(3f2g) aug-cc-pVTZ 586 871 709 2,166 Def2-TZVPP aug-cc-pVTZ 490 754 583 1,827 Def2-QZVP aug-cc-pVTZ 538 799 643 1,980 Def2-QZVPP aug-cc-pVTZ 586 874 718 2,178 SDD++(1f1g) aug-cc-pVTZ 609 912 714 2,235 SDD++(2f1g) aug-cc-pVTZ 630 942 744 2,316 SDD++(3f2g) aug-cc-pVTZ 678 1,017 819 2,514 Def2-TZVPP aug-cc-pVTZ 582 900 693 2,175 Def2-QZVP aug-cc-pVTZ 630 945 753 2,328 Def2-QZVPP aug-cc-pVTZ 678 1,020 828 2,526 91 The number of basis functions used for each calculation is compared to discuss the relationship between the BSSE and the basis set size as well as the computational cost. The aug-cc-pVTZ basis set is used for C, H and O atoms with different ECP and basis set for Pt. Based on the basis functions displayed in Table 3.16, the 6 combinations of basis set and ECP for Pt can be ordered according to the basis function size as follow: Def2_TZVPP with the small core ECP < SDD++(1f1g) basis set plus the large core ECP < SDD++(2f1g) basis set plus the large core ECP ≈ Def2_QZVP with the small core ECP < SDD++(3f2g) basis set plus the large core ECP ≈ Def2_QZVPP with the small core ECP. The less number of basis functions are employed in Def2_QZVP with the small core ECP than SDD++(3f2g) basis set plus the large core ECP, but the magnitude of BSSE of the former is smaller than that of the latter. This implies the basis set is better suited for DFT-PBE calculations. Compared to the Def2_QZVPP basis set utilizing the largest number of basis functions, Def2_QZVP basis set seems more efficient due to little difference in BSSE magnitude with less number of basis functions. 3.2.2.4 DFT-B3LYP adsorption energy on a small cluster in G03 In this section, DFT-B3LYP adsorption energy calculations on the Pt3 cluster are reported with the same geometry as in the DFT-PBE calculation in VASP to evaluate the XC correction energy term using hybrid XC functional compared with PBE generalized gradient approximation functional. E XCcorr cluster cluster = Eads ( B3LYP) − Eads ( PBE ) (3.5) 92 For the adsorption energy calculations, the aug-cc-pVTZ basis set is chosen for H, C and O atoms, and Def2-QZVP valence basis set for Pt atoms is used with the smallcore Def2-ECP. DFT-PBE molecular adsorption energies both in VASP and in Gaussian03 show consistency with a difference of 2 ~ 6 kJ/mol. However, DFTB3LYP calculations produce lower adsorption energies than DFT-PBE results by 68, 20, 43, 67 and 87 kJ/mol for CO, CH3, C2H4, C4H6 and C6H6, respectively. The -149 kJ/mol DFT-B3LYP adsorption energy for CO is acceptable because it gives less strong binding energy than DFT-PBE values, which is consistent with -141 ~ -144 kJ/mol of from Pt18 cluster calculation (Gil et al., 2003). For benzene adsorption, DFTB3LYP produces a repulsive interaction. Table 3.17 Comparison of DFT-B3LYP and DFT-PBE binding energies on a Pt3 cluster in Gaussian03 Adsorbate Methods Adsorption Energy (kJ/mol) BSSE (kJ/mol) BSSE corrected BSSE uncorrected Pt3 contribution Molecular contribution CO DFT-B3LYP DFT-PBE EXCcorr -149 -217 68 -152 -221 2.07 2.31 0.92 0.95 CH3 DFT-B3LYP DFT-PBE EXCcorr -227 -247 20 -229 -248 1.21 1.18 0.14 0.14 C2H4 DFT-B3LYP DFT-PBE EXCcorr -114 -157 43 -116 -159 1.45 1.76 0.39 0.42 C4H6 DFT-B3LYP DFT-PBE EXCcorr -135 -202 67 -128 -205 -7.32 2.57 0.77 0.84 C6H6 DFT-B3LYP DFT-PBE EXCcorr 12 -75 87 17 -79 -6.85 2.93 1.28 1.40 93 3.2.2.5 Adsorption energy on a small cluster validated with correlated wavefunction based methods Wavefunction-based methods are applied to treat electron correlation accurately, especially using high level electron correlation treatment methods, such as MP2 and CCSD(T) methods. Benzene adsorption on a Pt3 cluster has been examined because it is the system of interest. The calculation results are shown in Table 3.18. To obtain the accurate benzene adsorption energy on a Pt3 cluster, the complete basis set (CBS) limit of coupled cluster theory with single and double excitations and a quasi-perturbative treatment of triple excitations, CCSD(T), estimation has been used as follows: (Sponer et al., 2004; Sinnokrot and Sherrill, 2004) [ CCSD (T ) / CBS MP 2 CCSD (T ) MP 2 ΔEads = ΔEads ( IB) + ΔEads ( SB) − ΔEads ( SB ) ] (3.6) where IB denotes an infinite-basis-set calculation and SB denotes an small-basis-set calculation. Three basis sets have been chosen for Hartree-Fock Self Consistent Field (HF-SCF) calculation and for the MP2 correlation energy calculation. For HF-SCF calculation, Def2 ECP with Def2-QZVPP basis set for platinum and benzene are sufficient to obtain a converged value with negligible BSSE. However, MP2 calculation shows a BSSE of 76 kJ/mol. Schultz et al. (2006) used a much larger basis sets for a CCSD(T) calculation for the Pd/CO system. Their basis set is denoted as MTZ; a (9s8p7d3f2g/7s6p4d3f2g) valence electron basis set for Pd and a aug-cc-pVTZ basis set for C and O, and MQZ; a (12s11p9d5f4g3h /8s7p7d5f4g3h) valence electron basis 94 set for Pd and a aug-cc-pVQZ basis set for C and O. The Def2-QZVPP valence basis set used in this work (10s,8p,6d,4f,2g/7s,5p,4d,4f,2g) for Pt is similar to that in MTZ, but not as large as the MQZ basis set. The larger basis set, such as QZ or 5Z basis set, are expected to be required for converged MP2 calculation results. Lastly, for the CCSD(T) calculation, SDD ECP for Pt and SDD valence basis set for Pt and cc-pVTZ basis set for C and H atoms are used, which can be considered as small basis set. The use of a different HF wave function as an initial guess for the HF and MP2 calculations contributes to the deviation of adsorption energies despite using the same basis set. The restricted HF wavefunctions have been used for the CCSD(T) calculation for the small Pt cluster despite causing the stability problem, because the open shell unrestricted HF calculation makes a CCSD(T) calculation improbable. MP2 and CCSD(T) are found to predict a significantly higher adsorption energy for benzene on a Pt3 cluster than DFT-PBE. The best MP2/Def2-QZVPP value of –280 kJ/mol is 205 kJ/mol stronger than the DFT-PBE/aug-pVTZ value of -75 kJ/mol. The CCSD(T)/SDD adsorption energy is 154 kJ/mol weaker than the corresponding MP2/SDD value, but still 232 kJ/mol stronger than the DFT-PBE value. All these values have been corrected for basis set superposition error. One should note that the basis set requirements increase exponentially with the number of electrons for correlated wave-function based methods, and the basis superposition error is still 76 kJ/mol for the MP2/Def2-QZVPP calculation, despite using a large double polarized quadruple zeta basis set. Unfortunately, calculating numerically converged benzene adsorption energies at the MP2 and CCSD(T) level of theory is beyond current computational capabilities. However, based on the MP2 and CCSD(T) calculations it 95 is expected that the adsorption energy will increase when back-donation is accurately accounted for. Regardless of the convergence problem, a -134 kJ/mol of benzene adsorption energy on Pt3 cluster has been estimated from the Eq. 3.6, using SDD ECP and valence basis set for Pt and cc-pVTZ basis set for benzene. Table 3.18 Benzene binding energies on Pt3 cluster with correlated wavefunction based methods in Gaussian03 Adsorption Energy (kJ/mol) Combination BSSE (kJ/mol) Pt Basis set Benzene Basis set BSSE corrected BSSE uncorrected Pt3 contribution Molecular contribution SDD cc-pVTZ 194 184 7.10 2.47 Def2-TZVPP Def2-TZVPP 221 216 4.15 1.12 Def2-QZVPP Def2-QZVPP 220 219 0.86 0.26 SDD cc-pVTZ -199 -440 228.26 12.84 Def2-TZVPP Def2-TZVPP -242 -353 96.96 14.81 Def2-QZVPP Def2-QZVPP -280 -356 69.83 6.54 HF SDD cc-pVTZ 139 - - - MP2 SDD cc-pVTZ -423 - - - CCSD SDD cc-pVTZ -226 - - - CCSD(T) SDD cc-pVTZ -277 - - - Methods HF-SCF MP2 96 3.2.2.6 Comparison with wave function based methods: binding energy of di-σ and π ethylene on a small Pt2 cluster To study the accuracy of DFT-PBE to describe di-σ and π type adsorption, the adsorption of ethylene on a small Pt2 cluster was studied using DFT-PBE and different wavefunction-based methods such as MP2, CCSD, and CCSD(T) using the Def2 Effective Core Potential and the corresponding Def2-QZVPP basis set (Table 3.19). Di-σ adsorption is predicted to be preferred over π-adsorption by all computational methods except HF theory. However, the high level CCSD(T) method predicts significantly stronger binding energies than DFT-PBE, by 96 kJ/mol and by 41 kJ/mol for the di-σ and π-adsorption structures respectively. Though the CCSD(T) binding energies are likely to be an overestimation due to remaining basis set superposition error, it seems reasonable to predict that DFT-PBE underestimates the ethylene adsorption energy. This is consistent with the underestimation of back-donation by the DFT-PBE method, caused by the high lying LUMO. Table 3.19 Ethylene binding energies on Pt2 cluster with DFT-PBE and correlated wavefunction based methods in Gaussian03 Adsorption Energy (kJ/mol) Combination Pt Basis set Ethylene Basis set di-σ type π type Adsorption Energy Difference (kJ/mol) SDD cc-pVTZ -230 -178 52 Def2-QZVPP Def2-QZVPP -219 -174 45 UHF Def2-QZVPP Def2-QZVPP -28 -54 -26 MP2 Def2-QZVPP Def2-QZVPP -431 -254 177 CCSD Def2-QZVPP Def2-QZVPP -275 -192 83 CCSD(T) Def2-QZVPP Def2-QZVPP -315 -215 100 Methods DFT-PBE 97 3.2.3 Electronic structure based correction approach So far, various levels of calculations have been performed to examine the accuracy of DFT-PBE adsorption energy of benzene on Pt(111) slab and Pt3 cluster. The DFT-PBE adsorption energy of –71 kJ/mol with a plane wave calculation in VASP has been validated by a molecular orbital calculation in Gaussian03 with the BSSE correction, The DFT-B3LYP calculation predicts a lower adsorption energy, while the high-level of electron correlation methods give strong binding energy of the -126 ~ -134 kJ/mol with convergence problem. It is expected that the difference in the calculated adsorption energy deviation at different levels of theory may arise from a different treatment of the electronic structure of adsorbates. It is thought that the DFT description of the electronic structures of the gas phase molecule and of the transition metal substrate could have a significant discrepancy with the experimental data, in particular the energy level of the HOMO and the LUMO of molecule and workfunction of heterogeneous catalytic surfaces. Therefore, once the discrepancy between the DFT description of the electronic structure and experimental data can be systematically quantified, the DFT calculations might be linked to experimental results as close as possible. In the current section, the workfunction of the Pt(111) surface and the HOMO and LUMO gap of benzene described in the DFT-PBE calculation will be compared with DFT-B3LYP as well as CCSD(T) calculations and compared with experimental data in order to elucidate the difference in the electronic structure description. 98 3.2.3.1 HOMO-LUMO gap of gas phase molecules From the literature review on the CO/Pt(111) puzzle, it follows that the DFT-GGA calculation suffers from the poor description on the HOMO and LUMO of gas phase CO, resulted in the 0.4 eV stronger CO chemisorption energy compared to experimental values. Mason et al. (2004) used the CO singlet-triplet excitation energy deviation to map the HOMO-LUMO gap in gas-phase CO and found that a CCSD(T) calculation could accurately reproduce the experimental CO singlet-triplet excitation energy difference of 6.095 eV. Following the approach of Mason et al. (2004), ionization energies and electron affinities were calculated to indicate the energy level of HOMO and LUMO as well as its gap according to the Koopmans’ theorem, namely that “the first ionization energy of a molecule is equal to the energy of the highest occupied molecular orbital, and the electron affinity is the negative energy of the lowest unoccupied orbital.” (Phillips, 1961) The ionization energy can be evaluated as the difference in the total energy at 0 K of the cation and the corresponding neutral, and the electron affinity can be calculated as the difference in the total energy at 0 K of the anion and the corresponding neutral (Curtiss et al., 1998). The CBS-QB3 model available in Gaussian03, a high level compound method approximating the CCSD(T) result with a complete basis set, is chosen as a benchmark to compute ionization energies and electron affinities of the test molecules. In addition, DFT calculations with the PBE and the B3LYP functional with a aug-cc-pVTZ basis set have been performed to obtain the ionization energy and electron affinity. These computed electron affinity and ionization energy values are 99 compared with the experimental results and the corresponding HOMO-LUMO energy gap has been given in Tables 3-20 to 22. The HOMO-LUMO gap comparison displays a general trend that CBS-QB3 method gives the closest results to the experiment, followed by DFT-B3LYP and DFT-PBE methods. Table 3.20 Calculated ionization potential for molecules in the gas-phase Molecule Experiment a CO a CH3 C2H4 C4H6 C6H6 -14.01 eV -9.84 eV -10.51 eV -9.07 eV -9.24 eV CBS-QB3 -14.06 eV -9.80 eV -10.55 eV -9.07 eV -9.34 eV B3LYP/aug-cc-pVTZ -14.18 eV -9.95 eV -10.33 eV -8.73 eV -9.05 eV PBE/aug-cc-pVTZ -13.87 eV -10.04 eV -10.38 eV -8.72 eV -9.04 eV Reference (Lias, 2005) Table 3.21 Calculated electron affinity for molecules in the gas-phase Molecule Experiment CO i 1.33 eV CH3 ii 0.08 eV C2H4 iii 1.78 eV C4H6 iii 0.62 eV C6H6 iv 1.14 eV CBS-QB3 1.61 eV 0.01 eV 1.68 eV 0.66 eV 1.14 eV B3LYP/aug-cc-pVTZ 1.01 eV -0.11 eV 0.77 eV 0.38 eV 0.39 eV PBE/aug-cc-pVTZ 0.92 eV -0.16 eV 0.50 eV 0.59 eV 0.32 eV Reference (Repaey and Franklin, 1976); ii Reference (Ellison et al., 1978); Jordan, 1975) iv Reference (Hill and Squires, 1998). i iii Reference (Burrow and Table 3.22 Comparison of gap between ionization potential and electron affinity for molecules in the gas-phase Molecule CO CH3 C2H4 C4H6 C6H6 Experiment 12.68 eV 9.76 eV 8.73 eV 8.45 eV 8.10 eV CBS-QB3 12.45 eV 9.79 eV 8.87 eV 8.41 eV 8.20 eV B3LYP/aug-cc-pVTZ 13.17 eV 10.06 eV 9.56 eV 8.35 eV 8.66 eV PBE/aug-cc-pVTZ 12.95 eV 10.20 eV 9.88 eV 8.13 eV 8.72 eV 100 As an analogous term with the ionization energy of gas phase molecules, the workfunction can be calculated for transition metals and surfaces, and is defined as the smallest energy needed to extract an electron at 0 K. The workfunction can be obtained from the difference between the Fermi level and vacuum level. The DFT-PBE calculation in VASP gives a workfunction of 5.67 eV for the Pt(111) surface. This value compares well with the experimental value of 5.70 eV (Kiskinova et al., 1983). The value of surface workfunction also corresponds to energy level of highest filled d state of the metal, so that the electronic interaction between the surface and the gas phase molecules can be understood with the HOMO-LUMO energy diagram displayed in Figure 3.12. It shows that the various computational methods accurately predict the HOMO energy level relative to the workfunction, but display a large variation in the description on the LUMO energy level. For CH3 and C4H6, DFT calculations predict a LUMO energy close to the CCSD(T) and experimental value. However, DFT shows a big deviation in the LUMO energy for ethylene and benzene as well as carbon monoxide, which might lead to an underestimation of the effect of back-donation. The adsorption energy is a result of the surface interaction with the frontier orbitals of the adsorbate, which can be related to the relative position of the LUMO and the HOMO of gas phase molecules relative to the Fermi energy and the d-band center of the metal surface. If the LUMO is closer to the Fermi level, electrons will be backdonated from the surface to molecules, whereas if the HOMO is closer to the Fermi level, electrons will be transferred from the molecule to the surface as illustrated in Figure 3.13. 101 Figure 3.12 HOMO-LUMO energy diagram of the molecules and Pt(111). (Thick solid line: experiment, dashed lines: CBS-QB3, dotted lines: B3LYP and thin solid lines: PBE) 102 SSLUMO LUMO SSHOMO HOMO SSLUMO LUMO SSHOMO HOMO Figure 3.13 Schematic illustration of the surface interaction with the front orbitals: left-side describes electron back-donation and right-side illustrates electron donation. From Yamagishi et al. (2001). To combine the simple surface interaction model in Fig 3.13 and the HOMO-LUMO energy diagram of molecules with Pt(111) in Fig 3.12, the distance of the LUMO to the Fermi energy (SLUMO) and the distance of the HOMO to the Fermi energy (SHOMO) have been calculated to obtain the electron interaction strength from the difference between SLUMO and SHOMO. Here, a negative value means back-donation is dominant, while a positive value implies electron donation is dominant. Furthermore, to find out which computational method can accurately describe the electronic interaction for molecular adsorption on the Pt(111) surface, the SLUMO - SHOMO is compared to the experimental value in Table 3.23. 103 Table 3.23 The electronic interaction strength parameters related to the HOMO and LUMO of molecules compared to the Fermi energy of Pt(111) Parameters (SLUMO - SHOMO) (SLUMO - SHOMO) comparison to Exp. Methods CO CH3 C2H4 C4H6 C6H6 Experiment -3.94 eV 1.48 eV -0.89 eV 1.71 eV 1.02 eV CBS-QB3 -4.27 eV 1.59 eV -0.83 eV 1.67 eV 0.92 eV DFT-B3LYP -3.79 eV 1.56 eV 0.30 eV 2.29 eV 1.96 eV DFT-PBE -3.39 eV 1.52 eV 0.52 eV 2.09 eV 2.68 eV Experiment 0.00 eV 0.00 eV 0.00 eV 0.00 eV 0.00 eV CBS-QB3 -0.33 eV 0.11 eV 0.06 eV -0.04 eV -0.10 eV DFT-B3LYP 0.15 eV 0.08 eV 1.19 eV 0.58 eV 0.94 eV DFT-PBE 0.55 eV 0.04 eV 1.41 eV 0.38 eV 1.66 eV The electron interaction strength, (SLUMO - SHOMO), shows that CO adsorption on Pt(111) is governed by strong electron back-donation, while CH3 adsorption is driven mainly by electron donation. Ethylene adsorption is interesting because experimental data indicates that electron back-donation drives the adsorption energy, but DFT calculations show electron-donation is dominant. 1,3-Butadiene and benzene adsorption on Pt(111) are mainly governed by electron donation. However, the LUMO of benzene in DFT methods is much further from the Fermi level than the experimental value, implying the underestimation of electronic interaction between metal surface and benzene molecules, possibly explaining the adsorption energy deviation between the SCAC experimental value of -197 kJ/mol and the calculated DFT-PBE value of -107 kJ/mol. DFT methods are expected to be accurate for the CH3 adsorption energy, however, they are expected to be less accurate for ethylene and benzene. 104 3.3 Coverage Effects on the Benzene Adsorption energy and Site Preference on Pt(111) So far most theoretical studies of benzene adsorption on Pt(111) have been conducted at low coverage. It is widely recognized that adsorption properties are coveragedependent, such as the preferred adsorption site and the adsorption energy. (Lehwald, et al., 1978) As the coverage of benzene on Pt(111) increases, what kinds of changes can be expected for the adsorption properties, such as, the preferential adsorption mode, the geometry of chemisorbed benzene and its binding energy? In the following section, a review of the experimental and theoretical literature on the coverage effects on benzene adsorption on Pt(111) will be discussed and DFT-PBE computational results for high coverages will be presented. 3.3.1 Review of the experimental and theoretical literature on the coverage effects on benzene adsorption on Pt(111) Most studies of the coverage effect on benzene adsorption on Pt(111) have used experimental methods to detect changes in the preferred adsorption sites, chemisorbed benzene structure and adsorption energies. 3.3.1.1 Preferential adsorption mode for benzene at high coverage High Resolution Electron Energy Loss Spectroscopy (HREELS) studies at 300 K performed by Lehwald et al. (1978) indicate that benzene on Pt(111) shows two dominant IR peaks at 830 and 920 cm-1, whose relative intensity is 1:1 at very low coverages and 4:1 at higher coverages. 105 The effect of temperature was also studied by Lehwald et al. (1978) between 140 K and 300 K. Based on the independent changes in peak intensities, benzene was proposed to adsorb at the Atop and at Hollow sites. Haq and King (1996) used Reflection Adsorption IR Spectroscopy (RAIRS) to compare vibrational frequencies of adsorbed benzene on Pt(111) at different coverages and different temperatures. They also observed two bands with frequencies of 900 and 830 cm-1 at low coverages. At higher coverages and at 220 K, the 830 cm-1 band split into two bands at 820 and 829 cm-1, while band at 900 cm-1 disappeared. They attribute the difference between the two bands a change in the occupation of the Hollow and Bridge adsorption sites. The DFT frequency calculations by Saeys et al. (2002) successfully reproduced experimental HREELS spectrum from Lehwald et al. (1978) and found that the peaks observed at 1420 and 920 cm-1 are fingerprints for the Bridge site adsorbed benzenes. They also agreed with the finding of Haq and King (1996) that the peak at 830 cm-1 was the fingerprint for the Hollow site adsorbed benzene and believed that the Hollow site adsorption at high coverage might become more favorable. 3.3.1.2 Chemisorbed benzene structure as a function of coverage Using Surface Enhanced Raman Spectroscopy (SERS), Liu et al. (2006) examined the influence of the surface morphology and benzene concentration on the electrochemical adsorption behavior on a roughened Pt(111) surface. Raman spectra have been obtained at a potential of -0.5 V for the roughened Pt electrode in 0.1 M NaF 106 solution, varying benzene concentration from 0.1 mM to 9 mM. A broad band at 340 cm-1 is observed for a benzene concentration of 0.1 mM, which is mapped to a fingerprint for benzene chemisorbed parallel to the surface. At a 0.4 mM of benzene concentration, a peak at 1012 cm-1 is detected with a weak shoulder at 310 cm-1, where both peaks are considered to arise from the same species of perpendicularly adsorbed benzene to the surface. The peak at 991 cm-1 increases linearly with increasing benzene concentration from 0.4 mM up to a saturation concentration of 9 mM, and is believed to be a trace for weakly physisorbed benzene on the Pt(111). Based on these findings, Liu et al. proposed a schematic model for possible configuration of benzene adsorption on Pt(111) at various concentrations: parallel-chemisorbed benzene at low concentration up to 0.1 mM, vertically or tilted chemisorbed benzene on the surface at higher concentration and the physisorbed benzene above a concentration of 4 mM. 3.3.1.3 Benzene adsorption energy as a function of coverage The coverage effect on the benzene adsorption energy has been experimentally studied, using Thermal Desorption mass Spectroscopy (TDS) and Single Crystal Adsorption Calorimetric (SCAC) methods. TPD studies by Xu et al. (1994) found a desorption energy of 45 kJ/mol for a low temperature state at 178 K. According to a TDS study by Campbell et al. (1989), 43 ~ 51 kJ/mol of desorption energy for benzene adsorbed on the clean Pt(111) surface at 195 K. Ihm et al. (2004) fitted the varieties of the heat of adsorption at 300 K with coverage and proposed the second-order polynomial equation. ( ΔH ads = 197 − 48 θ θsat − 83(θ θsat ) 2 ) kJ mol (3.7) 107 where θ is the coverage and θsat the experimentally saturated coverage is 2.3×1014 benzenes at a unit surface area of square centimeter at 300 K. 3.3.2 DFT study of the coverage effect on the benzene adsorption energy At moderate coverage of 1/7 ML, the theoretical study for benzene adsorption on Pt(111) has not been extensively studied like at low coverage of 1/9 ML. Mittendorfer et al. (2003) studied adsorption of unsaturated hydrocarbons on Pt(111) and Pd(111) and computed benzene adsorption energy of -82 kJ/mol at Bridge(30) site on Pt(111) p(√7×√7)R19.1° surface, compared to the 85 kJ/mol of benzene heat of adsorption at 1/7 ML and 66 kJ/mol at 1/6 ML from Eq. 3.7. So far the DFT benzene adsorption energy at 1/6 ML has not reported yet. In this section, plane wave periodic DFT-PBE calculations are performed to investigate the coverage effects on a benzene adsorption energy, a chemisorbed benzene structure and a preferred adsorption site. The coverages of interest are studied using different surface unit cells: one benzene per p(3×3) unit cell for 1/9 ML coverage, one benzene per p(√7×√7)R19.1° unit cell for 1/7 ML coverage and two benzene molecules per c(2√3×3)R90º unit cell for 1/6 ML coverage. (Fig 3.14) 108 (a) p(3×3) – C6H6 unit cell for 1/9 ML (b) p(√7×√7)R19.1° - C6H6 unit cell for 1/7 ML (c) c(2√3×3)R90º – 2C6H6 unit cell for 1/6 ML Figure 3.14 Various coverage simulation with surface unit cells for benzene on Pt(111) 109 3.3.2.1 Preferred benzene adsorption sites at high coverage Various adsorption configurations were considered as displayed in Fig 3.15. Preliminary DFT calculations using a 1-layered slab with a minimum k-points grid have been used to the most preferred adsorption configuration, presuming the relative energy difference may not change as DFT adsorption energy converges. Among the three potential adsorption configurations in Fig. 3.15, case (a) produced the most stable structure with a total adsorption energy of -137 kJ/mol or -68 kJ/mol per benzene. Case (b) where benzene adsorbs at Hollow(0) sites shows the weakest binding energy of -6 kJ/mol for each adsorbed benzene. The case (c) where benzene adsorbs at Hollow(30) sites also shows weak adsorption energy of -12 kJ/mol per benzene molecule, but it seems a transition state because the binding geometry shows only three carbon-Pt bonds. This transition structure with 3 σ-type C-Pt bonding is expected to move to the more stable adsorption structure at the Bridge(30) sites. (a) Bridge(30)-Bridge(30); ΔEads = -137 kJ/mol 110 (b) Hollow(0)-Hollow(0); ΔEads = -12 kJ/mol (c) Hollow(30)-Hollow(30); ΔEads = -25 kJ/mol Figure 3.15 Possible adsorption configurations at high coverage of 1/6 ML for benzene on Pt(111) with their adsorption energy. Top-views are presented in left-side, side-views in rightside. 3.3.2.2 Coverage effect on the benzene adsorption energy It is well known that the adsorption energy is a coverage-dependent property. To determine the coverage effect, the adsorption energy for the preferred Bridge(30) and Hollow-hcp(0) adsorption sites with increasing coverage has been calculated and is summarized in Table 3.24. 111 (a) Bridge(30)-Bridge(30); ΔEads = -168 kJ/mol (b) Bridge(30)-Hollow(0); ΔEads = -102 kJ/mol Figure 3.16 Adsorption energy calculation results at high coverage of 1/6 ML using 6-layered slab with k-point grid of 5×5×1. Table 3.24 Benzene adsorption energy results on Pt(111) at low coverage Adsorption Energy (kJ/mol) Source PP-XC Coverage Slab thickness k-point mesh Bridge(30) Hollow-hcp(0) Saeys et al. (2002) US-PW91 1/9 ML 4-layered 2×2×1 -117 -75 Morin et al. (2003) PAW-PW91 1/9 ML 6-layered 5×5×1 -100 -73 Saeys et al. (2004) BP86 1/9 ML Pt22 cluster - -102 -75 Current PAW-PBE 1/9 ML 5-layered 5×5×1 -107 -71 Mittendorfer et al. (2004) PAW-PW91 1/7 ML 4-layered 7×7×1 -82 n.c. Current PAW-PBE 1/7 ML 5-layered 5×5×1 -90 -60 Current PAW-PBE 1/6 ML 1-layered 2×2×1 -68 n.c. Current PAW-PBE 1/6 ML 5-layered 5×5×1 -88 -54 Note. n.c. : not calculated The calculations indicate that at high coverage Hollow(0) adsorption might not be favorable (Fig. 3.13), therefore, the adsorption energy for two benzenes where one adsorbs at the Bridge(30) site and the other at the Hollow(0) site has been determined for coverage of 1/6 ML, as shown in Fig. 3.14. 112 The DFT-PBE benzene adsorption energy on Pt(111) at low coverage published in literature ranges from -102 kJ/mol to -117 kJ/mol, while DFT-PBE calculation in the current work gives -107 kJ/mol as a numerically converged value. The DFT adsorption energy at moderate coverage of 1/7 ML has been computed as -82 kJ/mol by Mittendorfer et al. (2003) and the current DFT-PBE adsorption energy calculation converges to -90 kJ/mol for a 5-layered slab with a k-points mesh of 5×5×1. The adsorption energy at high coverage of 1/6 ML at the Bridge(30) site varies from -68 to -88 kJ/mol, whereas adsorption energy for the Bridge(30)-Hollow(0) mode converges to -54 kJ/mol, which is well consistent with TPD results (Campbell et al., 1989; Xu et al., 1994). 3.3.2.3 Coverage effect on the benzene adsorption structure The adsorption geometry of benzene adsorbed at the Bridge(30) and the Hollowhcp(0) sites determining in the current work is compared with literature data in Tables 3.25 and 3.26. The notations for geometric characteristics are explained from Fig. 3.17. The reported geometric characteristics in the literature are comparable with the results in the current work. The bonding between carbon and platinum atoms is 0.02 Å shorter as well as the perpendicular distance of benzene ring from the Pt(111) surface implying a stronger interaction between benzene and the Pt(111) surface. The C-H tilted angle between two carbons and a Platinum atom has been widened by 2°, while the titled angle between one carbon and a Pt atom has been sharpened by 2°. These geometric differences support that the electronic interactions in the 5-layered slab is more pronounced than in the 4-layered slab, which results in a larger adsorption energy. 113 Table 3.25 Benzene adsorption geometry at the Bridge(30) of Pt(111) at various coverage Coverage PP-XC C-C bond length (Å) C-H tilted angle (°) C-Pt bond length (Å) Perpendicular distance (Å) r1 r2 α β R1 R2 d⊥ 0.11 ML US-PW91 1.43 1.45 35.5 16.1 2.20 2.27 2.09 0.11 ML PAW-PW91 1.43 1.47 37.2 15.5 2.18 2.22 2.04 0.11 ML PAW-PBE 1.43 1.47 35.0 17.3 2.16 2.20 2.02 0.14 ML PAW-PBE 1.43 1.47 36.3 17.5 2.16 2.20 2.01 0.17 ML PAW-PBE 1.43 1.47 37.3 16.3 2.18 2.20 2.05 Table 3.26 Benzene adsorption geometry at the Hollow-hcp(0) of Pt(111) slab at various coverage Coverage PP-XC C-C bond length (Å) C-H tilted angle (°) C-Pt bond length (Å) Perpendicular distance (Å) r1 r2 α R1 d⊥ 0.11 ML US-PW91 1.44 1.46 18.1 2.22 2.11 0.11 ML PAW-PW91 1.43 1.46 18.7 2.22 2.06 0.11 ML PAW-PBE 1.43 1.46 19.2 2.20 2.04 0.14 ML PAW-PBE 1.43 1.46 19.3 2.20 2.03 0.17 ML PAW-PBE 1.43 1.46 18.8 2.24 2.08 Figure 3.17 Benzene on the Pt(111) Bridge(30) site (left) and Hollow-hcp(0) site (right). Adopted from Saeys et al. (2002). 114 3.4 References Abild-Pedersen, F., and Anderson, M.P., “CO adsorption energies on metals with correction for high coordination adsorption sites–A density functional study”, Surface Science, 601, pp. 1747. 2007. Abon, M., Bertolini, J.C., Billy, J., Massardier, J., and Tardy, B., “Adsorption States of Benzene and Toluene on Pt(111): A Vibrational EELS, ΔΦ, AES and TDS Study”, Surface Science, 162, pp. 395. 1985. Anderson, A.B., McDevitt, M.R., and Urbach, F.L., “Structure and Electronic Factors in Benzene Coordination to Cr(CO)3 and to Cluster Models of Ni, Pt, and Ag(111) Surfaces”, Surface Science, 146, pp. 80. 1984. Bligaard, T., and Nørskov, J.K., “Ligand effects in heterogeneous catalysis and electrochemistry”, Electrochimica Acta, 52, pp. 5512. 2007. Blyholder, G., “Molecular Orbital View of Chemisorbed Carbon Monoxide”, Journal of Physical Chemistry, 68, pp. 2772. 1964. Bocquet, M.L., and Sautet, P., “STM and chemistry: a qualitative molecular orbital understanding of the image of CO on a Pt surface”, Surface Science, 360, pp. 128. 1996. Boys, S.F., and Bernardi, F., “The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors”, Molecular Physics, 19, pp. 553. 1970. Burrow, P.D., and Jordan, K.D., “On the electron affinities of ethylene and 1,3-butadiene”, Chemical Physics Letters, 36, pp. 594. 1975. Campbell, J.M., Seimanides, S.G., and Campbell, C.T., “Probing Ensemble Effects In Surface Reactions. 2. Benzene Adsorption on Clean and Bismuth-covered Pt(111)”, Journal of Physical Chemistry, 93, pp. 815. 1989. Cemie, F., Dippel, O., and Hasselbrink, E., “Resonant electron scattering from benzene chemisorbed on Pt(111)”, Surface Science, 342, pp. 101. 1995. Curtiss L.A., Redfern, P.C., Raghavachari, K., and Pople, J.A., “Assessment of Gaussian-2 and density functional theories for the computation of ionization potentials and electron affinities”, Journal of Chemical Physics, 109, pp. 42. 1998. Doll, K., “CO adsorption on the Pt(111) surface: a comparison of a gradient corrected functional and a hybrid functional”, Surface Science, 573, pp. 464. 2004. Dunning, T.H., “Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen”, Journal of Chemical Physics, 90, pp. 1007. 1989. Dyall, K.G., “Relativistic double-zeta, triple-zeta, and quadruple-zeta basis sets for the 5d elements Hf-Hg”, Theoretical Chemistry Accounts, 112, pp. 403. 2004. Ellison, G.B., Engelking, P.C., and Lineberger, W.C., “An experimental determination of the geometry and electron affinity of CH3”, Journal of the American Chemical Society, 100, pp. 2556. 1978. 115 Feibelman, P.J., Hammer, B., Nørskov, J.K., Wagner, F., Scheffler, M., Stumpf, R., Watwe, R. and Dumesic, J., “The CO/Pt(111) Puzzle”, Journal of Physical Chemistry B, 105, pp. 4018. 2001. Föhlisch, A., Nyberg, M., Hasselström, J., Karis, O., Pettersson, L.G.M., and Nilsson, A., “How Carbon Monoxide Adsorbs in Different Sites”, Physical Review Letters, 85, pp. 3309. 2000. Gil, A., Clotet, A., Ricart, J.M., Kresse, G., Garcia-Hernandez, M., Rösch, N., and Sautet, P., “Site preference of CO chemisorbed on Pt(111) from density functional calculations”, Surface Science, 530, pp. 71. 2003. Gottfried, J.M., Vestergaard, E.K., Bera, P., and Campbell, C.T., “Heat of Adsorption of Naphthalene on Pt(111) Measured by Adsorption Calorimetry”, Journal of Physical Chemistry B, 110, pp. 17539. 2006. Grinberg, I., Yourdshahyan, Y., and Rappe, A.M., “CO on Pt(111): A possible solution”, Journal of Chemical Physics, 117, pp. 2264. 2002. Hammer, B., Morikawa, Y., and Nørskov, J.K., “CO Chemisorption at Metal Surface and Overlayers”, Physical Review Letters, 76, pp. 2141. 1996. Hammer, B., and Nørskov, J.K., “Theoretical Surface Science and Catalysis – Calculations and Concepts”, Advances in Catalysis, 45, pp. 71. 2000. Haq, S., and King, D.A., “Configurational Transitions of Benzene and Pyridine Adsorbed on Pt{111} and Cu{110} Surfaces: An Infrared Study”, Journal of Physical Chemistry, 100, pp. 16957. 1996. Hill, B.T., and Squires, R.R., “Synthesis and Characterization of the Negative Ion of nonKekule Benzene”, Journal of the Chemical Society, Perkin Transactions II, 5, pp. 1027. 1998. Hu, Q.M., Reuter, K., and Scheffler, M., “Towards an Exact Treatment of Exchange and Correlation in Materials: Application to the “CO Adsorption Puzzle” and Other Systems”, Physical Review Letters, 98, pp. 176103. 2007. Ihm, H., Ajo, H.M., Gottfried, M., Bera, P., and Campbell, C.T., “Calorimetric Measurement of the Heat of Adsorption of Benzene on Pt(111)”, Journal of Physical Chemistry B, 108, pp. 14627. 2004. Kiskinova, M., Pirug, G., and Bonzel, H.P., “Coadsorption of Potassium and CO on Pt(111)”, Surface Science, 133, pp. 321. 1983. Kresse, G., Gil, A., and Sautet, P., “Significance of singlet-electron energies for the description of CO on Pt(111)”, Physical Review B, 68, pp. 073401. 2003. Kresse, G., “The PAW and US-PP database, Institut für Materialphysik and Center for Computational Material Science”, Universität Wien, Wien, Austria, 2004. Lehwald, S., Ibach, H., and Demuth, J.E., “Vibration Spectroscopy of Benzene Adsorbed on Pt(111) and Ni(111)”, Surface Science, 78, pp. 577. 1978. 116 Lias, S.G., "Ionization Energy Evaluation" in NIST Chemistry WebBook, NIST Standard Reference Database Number 69, Eds. P.J. Linstrom and W.G. Mallard, June 2005, National Institute of Standards and Technology, Gaithersburg MD, 20899 (http://webbook.nist.gov). Liu, G.K., Ren, B., Wu, D.Y., Duan, S., Li, J.F., Yao, J.L., Gu, R.A., and Tian, Z.Q., “Effect of Intrinsic Properties of Metals on the Adsorption Behavior of Molecules: Benzene Adsorption on Pt Group Metals”, Journal of Physical Chemistry B, 110, pp. 17498. 2006. Mason, S.E., Grinberg, I., and Rappe, A.M., “First-principles extrapolation method for accurate CO adsorption energies on metal surfaces”, Physical Review B, 69, pp. 161401(R). 2004. Metz, B., Stoll, H., and Dolg, M., “Small-core multiconfiguration-Dirac-Hartree-Fockadjusted pseudopotentials for post-d main group elements: Application to PbH and PbO”, Journal of Chemical Physics, 113, pp. 2563. 2000. Minot, C., Van Hove, M.A., and Somorjai, G.A., “Benzene Adsorption on Pt(111): A Theoretical Study”, Surface Review and Letters, 3, pp. 285. 1995. Mittendorfer, F., Thomazeau, C., Raybaud, P., and Toulhoat, H., “Adsorption of Unsaturated Hydrocarbons on Pd(111) and Pt(111): A DFT Study”, Journal of Physical Chemistry B, 107, pp. 12287. 2003. Morin, C., Simon, D., and Sautet, P., “Density-Functional Study of the Adsorption and Vibration Spectra of Benzene Molecules on Pt(111)”, Journal of Physical Chemistry B, 107, pp. 2995. 2003. Morin, C., Simon, D., and Sautet, P., “Chemisorption of Benzene on Pt(111), Pd(111), and Rh(111) Metal Surfaces: A Structural and Vibrational Comparison from First Principles”, Journal of Physical Chemistry B, 108, pp. 5653. 2004. Morni, C., Simon, D., and Sautet, P., “Intermediates in the hydrogenation of benzene to cyclohexene on Pt(111) and Pd(111): A comparison from DFT calculations”, Surface Science, 600, pp. 1339. 2006. Ogletree, D.F., Van Hove, M.A., and Somorjai, G.A., “LEED Intensity Analysis of the Structures of Clean Pt(111) and of CO adsorbed on Pt(111) in the c(4×2) Arrangement”, Surface Science, 173, pp. 351. 1986. Ogletree, D.F., Van Hove, M.A., and Somorjai, G.A., “Benzene and Carbon Monoxide Adsorbed on Pt(111): A LEED Intensity Analysis”, Surface Science, 183, pp. 1. 1987. Pallassana, V., and Neurock, M., “Electronic Factors Governing Ethylene Hydrogenation and Dehydrogenation Activity of Pseudomorphic PdML/Re(0001), PdML/Ru(0001), Pd(111),and PdML/Au(111) Surfaces”, Journal of Catalysis, 191, pp. 301. 2000. Pedersen, M.O., Bocquet, M.L., Sautet, P., Laegsgaard, E., Stensgaad, I., and Basenbacher, F., “CO on Pt(111): binding site assignment from the interplay between measured and calculated STIM images”, Chemical Physics Letters, 299, pp. 403. 1999. Phillips, J.C., “Generalized Koopmans’ Theorem”, Physical Reviews, 123, pp. 420. 1961. 117 Refaey, K.M.A., and Franklin, J.L., “Endoergic ion-molecule-collision process of negative ions. III. Collisions of I- on O2, CO and CO2”, Internatioma Journal of Mass Spectrometry and Ion Process, 20, pp. 19. 1976. Saeys, M., Reyniers, M.F., Marin, G.B., and Neurock, M., “Density Functional Study of Benzene Adsorption on Pt(111)”, Journal of Physical Chemistry B, 106, pp. 7489. 2002. Sautet, P., and Bocquet, M.–L., “A theoretical analysis of the site dependence of the shape of a molecule in STM images”, Surface Science Letters, 304, pp. L445. 1994. Sautet, P., and Bocquet, M.–L., “Shape of molecular adsorbates in STM images: A theoretical study of benzene on Pt(111)”, Physical Review B, 53, pp. 4910. 1996. Scholl, D.S., “Applications of Density Functional Theory to Heterogeneous Catalysis”, The Royal Society of Chemistry, 4, pp. 108. 2006. Schultz, N.E., Gherman, B.F., Cramer, C.J., and Truhlar, D.G., “PdnCO(n=1,2): Accurate Ab Initio Bond Energies, Geometries, and Dipole Moments and the Applicability of Density Functional Theory for Fuel Cell Modeling”, Journal of Physical Chemistry B, 110, pp. 24030. 2006. Schweizer, E., Persson, B.N.J., Tushanus, M., Hoge, D., and Bradshaw, A.M., “The Potential Energy Surface, Vibrational Phase Relaxation and the Order–Disorder Transition in the Adsorption System Pt{111}–CO”, Surface Science, 213, pp. 49. 1989. Sinnokrot, M.O., and Sherrill, C.D., “Highly Accurate Coupled Cluster Potential Energy Curves for the Benzene Dimer: Sandwich, T-Shaped, and Parallel-Displaced Configurations”, Journal of Physical Chemistry A, 108, pp. 10200. 2004. Somers, J., Bridge, M.E., Lloyd, D.R., and McCabe, T., “An ARUPS Investigation of Benzene on Pt(111): Evidence for Distortion”, Surface Science Letters, 181, pp. L167. 1987. Somorjai, G.A., “Introduction to Surface Chemistry and Catalysis”, John Wiley & Sons Inc, New York, 1984. Sponer, J., Jurecha, P., and Hobza, P., “Accurate Interaction Energies of Hydrogen-Bonded Nucleic Acid Base Pairs”, Journal of the American Chemical Society, 126, pp. 10142. 2004. Steininger, H., Lehwald, S., and Ibach, H., “On the Adsorption of CO on Pt(111)”, Surface Science, 123, pp. 264. 1982. Stroppa, A., Termemtzidis, K., Paier, J., Kresse, G., and Hafner, J., “CO adsorption on metal surface: A hybrid functional study with plane-wave basis set”, Physical Review B, 76, pp. 195440. 2007. Stroscio, J.A., and Eigler, D.M., “Atomic and Molecular Manipulation with the Scanning Tunneling Microscope”, Science, 254, pp. 1319. 1991. Tsai, M.-C., and Muetterities, E.L., “Platinum Metal Surface Chemistry of Benzene and Toluene”, Journal of the American Chemical Society, 104, pp. 2534. 1982. 118 Weigend, F., and Ahlrichs, R., “Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy”, Physical Chemistry Chemical Physics, 7, pp. 3297, 2005. Weiss, P.S., and Eigler, D.M., “Site dependence of the apparent shape of a molecule in scanning tunneling micoscope images: Benzene on Pt{111}”, Physical Review Letters, 71, pp. 3139. 1993. Xu, C., Tsai, Y.L., and Koel, B.E., “Adsorption of Cyclohexane and Benzene on Ordered Sn/Pt(111) Surface Alloys”, Journal of Physical Chemistry, 98, pp. 585. 1994. Yamagishi, S., Jenkins, S.J., and King, D.A., “Symmetry and site selectivity in molecular chemisorption: Benzene on Ni{111}”, Journal of Chemical Physics, 114, pp. 5765. 2001. Yeo, Y., Vattuone, L., and King, D.A., “Calorimetric investigation of NO and CO adsorption on Pd{100} and the influence of preadsorbed carbon”, Journal of Chemical Physics, 106, pp. 1990. 1997. Yimagawa, M., and Fujikawa, T., “Multiple scattering approach to carbon XANES from benzene on Pt(111) surface”, Surface Science, 357-358, pp. 131. 1996. 119 Chapter 4 Conclusions The adsorption and desorption of aromatic molecules on a transition metal catalyst surface is a key step in many industrially important catalytic processes. To better understand this reaction, and to begin to shed light on differences between experimental and first principles data in the literature, a systematic theoretical study was carried out. Over the past decade, density functional theory (DFT) has been applied successfully to various heterogeneous catalytic reactions, providing insight into the electronic factors governing the adsorption energy, as well as providing a molecular level understanding of the reaction mechanisms and the kinetics. When DFT results are compared with experimental data, different sources of uncertainty should be considered. First, one must ensure that the DFT calculations are numerically converged to eliminate mathematical uncertainties; second, current state-of-the-art DFT provides an approximate solution to the Schrödinger equation, and in particular the calculation of electron affinities and the corresponding LUMO energy have been found to be problematic in some cases; and finally, coverage effects and interactions between adsorbed molecules should be considered. This study was initiated by the discrepancy between experimentally observed benzene adsorption energies (between -130 and -200 kJ/mol at low coverage), and calculated values (between -80 and -130 kJ/mol). First, we obtained a numerically converged low coverage adsorption energy of -107 kJ/mol at the DFT-PBE level of theory. This value is significantly weaker than experimental numbers. To address this point, the adsorption of the smaller but related ethylene on a small Pt2 cluster was studied. It was 120 found that DFT-PBE likely underestimates the adsorption energy, when compared to state-of-the-art, but extremely expensive CCSD(T) calculations. The underestimation was attributed to the underestimation of the electron affinity, leading to an underestimation of electron back-donation. In a first set of calculations, the numerically converged benzene adsorption energy at the DFT-PBE level of theory was determined for a low coverage of 1/9 monolayer (ML). The periodic slab calculations were performed with the Vienna Ab initio Simulation Package (VASP), using Projector Augmented Wave (PAW) pseudopotentials to replace the inner shell electrons. It was found that a 5-layer slab is required to accurately describe the surface d-band and the surface relaxation. The surface d-band filling is significantly underestimated when a 3-layer slab or a 4-layer slab is used by 30 % and 8 %, respectively. The position and width of the electronic dband are key parameter determining the electron donation and back-donation between benzene and Pt(111). A 5×5×1 Monkhorst Pack k-point grid and 14 Å vacuum layer to separate repeating slabs led to a converged DFT-PBE adsorption energy of -107 kJ/mol for the bridge(30) site and –71 kJ/mol for the hcp-hollow(0) site for a p(3×3) unit cell. To validate the accuracy of the DFT-PBE method for this system, the ionization potential (HOMO) and the electron affinity (LUMO) were calculated at various levels of theory for 5 characteristic molecules: CO, methyl, ethylene, butadiene, and benzene, and compared with the workfunction of the Pt(111) surface. DFT-PBE was found to predict the position of the HOMO and the LUMO of methyl and butadiene within 0.25 eV of experimental data and hence an accurate description of the adsorption process 121 can be expected. DFT-PBE puts the 2π* LUMO of CO 0.4 eV too low and is hence expected to overestimate the effect of back-donation on the adsorption energy. The DFT-B3LYP HOMO and LUMO are within 0.3 eV of experimental values and DFTB3LYP is expected to accurately predict the CO adsorption energy and relative stability at different adsorption sites. For ethylene and benzene, both DFT-PBE and DFT-B3LYP significantly underestimate the electron affinity by -0.7 to -1.3 eV and are hence expected to underestimate back-donation. Since the effect of donation and back-donation are expected to change with adsorption site, DFT might not accurately describe the relative stability. To investigate the effect of the error in the LUMO energy on the adsorption energy, the benzene adsorption energy was recalculated on a small Pt3 cluster using correlated wave-function-based methods such as MP2 and CCSD(T). The latter method is found to accurately predict the electron affinity of benzene. MP2 and CCSD(T) are found to predict a significantly higher adsorption energy for benzene on a Pt3 cluster than DFTPBE. The MP2/Def2-QZVPP value of –280 kJ/mol is 205 kJ/mol stronger than the DFT-PBE/aug-pVTZ value of -75 kJ/mol. The CCSD(T)/SDD adsorption energy is 154 kJ/mol weaker than the corresponding MP2/SDD value, but still 232 kJ/mol stronger than the DFT-PBE value. All these values have been corrected for basis set superposition error. One should note that the basis set requirements increase exponentially with the number of electrons for correlated wave-function based methods, and the basis superposition error is still 76 kJ/mol for the MP2/Def2-QZVPP calculation, despite using a large double polarized quadruple zeta basis set. Unfortunately, calculating numerically converged benzene adsorption energies at the MP2 and CCSD(T) level of theory is beyond current computational capabilities. 122 However, based on the MP2 and CCSD(T) calculations it is expected that the adsorption energy will increase when back-donation is accurately accounted for. Finally, to elucidate the experimentally observed change in the preferred adsorption site at higher coverages, the adsorption of benzene was studied for coverages of 1/9, 1/7 and 1/6 monolayer. The latter coverage corresponds to the experimentally observed saturation coverage for a monolayer of benzene on Pt(111). DFT-PBE calculations did not predict a change in the preferred adsorption site with coverage. In summary, the calculations in this work indicate that a 5 layer slab and a 5×5×1 kpoint mesh are required to obtain a converged description of the electronic structure of the surface of Pt(111) in a p(3×3) unit cell at the DFT-PBE level of theory. The comparison with experimental ionization energies and electron affinities, as well as the calculation of ethylene binding energies on Pt2 with a wide variety of approximate theoretical methods, indicate that DFT-PBE puts the anti-bonding 2π* LUMO too high in energy – it underestimates the electron affinity – and hence underestimates backdonation and underestimates the adsorption energy of ethylene and, likely, benzene. These findings can likely be transferred to other systems. As a first step, one should calculate the electronic structure of the adsorbing molecule. If the selected method does not predict the electronic structure accurately (e.g. electron affinity and ionization potential), then the adsorption energy might be over- or underestimated. Furthermore, our calculations indicate that a 3 or 4 layer slab might not provide a converged description of a transition metal surface, and at least a 5 layer slab should be considered to obtain numerically converged values. 123 A number of measures provide an indication to ensure a sufficiently thick slab geometry: first, the surface relaxation should be converged and consistent with experimental observations, if available; second, electronic properties, such as the degree of d-band filling, d-band center, and workfunction should be close to the experimental data. Considering the similarity between different transition metals, 5 layers might be required for most DFT-PBE calculations. Note that 3 and 4 layer slabs are most commonly used in the literature, and the results might hence not be fully converged. This study stresses the importance of selecting a correct model to obtain a numerically converged description of the surface electronic structure of the metal surface. The results are important to begin to understand the different electronic effects influencing benzene and ethylene adsorption, a key step in the mechanism of aromatic hydrogenation, and future studies could explore the effect of benzene and hydrogen co-adsorption, as well as the effect of substituents as in toluene and xylene, on the aromatic adsorption energy on Pt(111). In addition, the study points toward deficiencies in the DFT-PBE description of the electronic structure of aromatics and olefins. The first observation will help select a correct model for future studies. The second observation indicates that one should be careful interpreting adsorption energies, and DFT-PBE is probably not sufficiently accurate for predictive kinetics. However, relative binding energies are probably predicted more accurately, in particular if the interaction is dominated by the interaction between the HOMO and the surface electronic structure. It should be stressed that the strength of DFT-PBE is to guide the qualitative discovery of improved 124 catalysts and the qualitative understanding and elucidation of reaction mechanisms. Theoretical observations and predictions can then guide experimental studies to confirm the theoretical predictions. This observation led to two conclusions: 1. DFT-PBE is not yet sufficiently accurate to quantitatively predict chemisorptions energies of aromatics, olefins on transition metal surfaces. DFT-PBE should therefore be used to understand trends in reactivity and to guide the design of new catalytic materials and provide understanding. Such ideas should subsequently be tested experimentally. This procedure is followed extensively in the Saeys group. DFT-PBE should not be used for predictive kinetic modeling. 2. The development of new and better functional is probably required to reach predictive accuracy. Such developments are underway, but slow. A key test will likely be the accuracy of the adsorbate electronic structure, in particular the HOMO-LUMO gap. Hybrid functionals and possibly meta-GGA functionals might lead to significant improvements, as shown in selected Gaussian calculations. CCSD(T) type methods, though accurate, require prohibitively large basis sets and CPU time. Even if CPU speeds keep doubling every 18 months, it will be many years before systems containing more than a few transition metal atoms can be treated accurately with such methods. 125 Appendix A DFT-PBE Calculation Data In this appendix, DFT-PBE total energy calculation results in VASP for benzene adsorption on Pt(111) study have been provided. Table A.1 DFT-PBE total energy calculation results for convergence test with various vacuum thickness for benzene adsorption energy on Pt(111) at the Bridge(30) site Benzene (eV) Clean Pt(111) slab (eV) Pt(111) with benzene (eV) 9 Å (4-bulk-interlayer distance) -75.9877 -206.8226 -283.7452 12 Å (5-bulk-interlayer distance) -76.0072 -206.8240 -283.6995 14 Å (6-bulk-interlayer distance) -76.0137 -206.8250 -283.6753 16 Å (7-bulk-interlayer distance) -76.0032 -206.8230 -283.6624 18 Å (8-bulk-interlayer distance) -76.0076 -206.8233 -283.6534 20 Å (9-bulk-interlayer distance) -76.0052 -206.8142 -283.6453 Vacuum thickness Table A.2 DFT-PBE total energy calculation results for slab thickness convergence test for benzene adsorption energy on Pt(111) at Bridge(30) site along with k-point convergence test at low coverage of 1/9 ML 3-layers 4-layers 5-layers 6-layers (eV) (eV) (eV) (eV) 3×3×1 -151.9974 -207.2445 -261.3427 -316.0469 5×5×1 -151.7799 -206.9064 -261.0708 -315.7266 7×7×1 -151.7667 -206.7408 -260.9905 -315.5731 3×3×1 -229.3948 -284.0623 -338.5868 -393.2940 5×5×1 -229.1575 -283.7685 -338.1974 -392.7196 7×7×1 -229.0752 -283.6558 -338.1171 -392.6393 Number of slab layers Pt(111) slab Pt(111) with benzene Note. The total energy of benzene calculated in a 14 × 14 × 14 Å3 cubic cell has been used for the adsorption energy calculation. 126 Table A.3 DFT-PBE total energy calculation results for molecular adsorption energy on Pt3 cluster in VASP Molecules CO (eV) CH3 (eV) C2H4 (eV) C4H6 (eV) C6H6 (eV) Adsorbate only -14.7918 -18.1907 -31.9635 -57.0034 -76.0079 Pt3 cluster with adsorbate -25.9724 -29.6492 -42.5263 -68.0946 -85.6907 Note. The total energy of Pt3 cluster with triplet state is -8.9538 eV and used for the adsorption energy calculation. Table A.4 DFT-PBE total energy calculation results for slab thickness convergence test for benzene adsorption energy on Pt(111) at Bridge(30) site along with k-point convergence test at moderate coverage of 1/7 ML 3-layers 4-layers 5-layers 6-layers (eV) (eV) (eV) (eV) Pt(111) slab 3×3×1 5×5×1 7×7×1 9×9×1 -118.0863 -117.8968 -117.9512 -117.9509 -160.9630 -160.7581 -160.7688 -160.7786 -203.0040 -202.9470 -202.9541 -202.9605 -245.6229 -245.3669 -245.3949 -245.3880 Pt(111) with benzene 3×3×1 5×5×1 7×7×1 9×9×1 -195.2124 -195.0859 -195.0976 -195.1019 -237.6655 -237.5665 -237.5627 -237.5714 -280.0792 -279.8984 -279.9170 -279.9109 -322.4626 -322.2886 -322.3055 -322.2994 Number of slab layers Note. The total energy of benzene calculated in a 14 × 14 × 14 Å3 cubic cell has been used for the adsorption energy calculation. Table A.5 DFT-PBE total energy calculation results for slab thickness convergence test for benzene adsorption energy on Pt(111) along with k-point convergence test at moderate coverage of 1/6 ML Number of slab layers 5-layers (eV) 6-layers (eV) Pt(111) slab 3×3×1 5×5×1 -347.7357 -347.6333 -420.5326 -420.3564 Pt(111) with benzene at Bridge(30)-Bridge(30) site 3×3×1 5×5×1 -501.6788 -501.4877 -574.4115 -574.1332 Pt(111) with benzene at Bridge(30)-Hollow(0) site 3×3×1 5×5×1 -500.9669 -500.7768 -573.7266 -573.4446 Note. The total energy of benzene calculated in a 14 × 14 × 14 Å3 cubic cell has been used for the adsorption energy calculation. 127 [...]... Electronic Interaction in Heterogeneous Catalysis In this section, the basic concepts of chemical bonding to transition metal surfaces will be understood First, carbon monoxide interaction with the metal surface will be illustrated; then, it will be extended to more complicated ethylene Next, adsorbateadsorbate interaction will be briefly presented Finally, electronic interaction of benzene adsorption on. .. on transition metal surface will be explained 1.4.1 Surface-adsorbate interaction Chemical bonding constructed by the adsorption of a molecule on a transition metal surface can be initially understood by the Newns-Anderson model (Newns, 1969; Anderson, 1961), where both the bonding and the anti-bonding molecular orbitals of 10 the adsorbate contribute to the chemical bonding Strong chemisorption bond... analysis of the molecular orbitals formed on adsorption On the adsorption, benzene σ orbitals in C-C bonds and C-H bonds are stabilized by the interaction with Pt orbitals so that around 5 % of the electron density of the benzene σ orbital is donated into empty Pt orbitals The σ-interaction depends on the adsorption site and plays a role in the site preference Further, the strongest interaction can be... – π and π*, are downshifted and broadened upon the interaction with the sp-band, then renormalized frontier orbitals interact with the valence d-band of the metal so that bonding and anti-bonding orbitals are constructed as a result of electron donation and electron back-donation 13 Figure 1.7 Frontier orbital interaction in the di-σ adsorption of ethylene on Pd(111) From Pallassana and Neurock (2000)... reaction, particularly benzene adsorption on Pt(111), to pave the way for the first principle-based modeling for benzene hydrogenation on catalytic surfaces 2 1.2 Principles in Heterogeneous Catalysis Heterogeneous catalysis reaction is simply comprised of adsorption, surface reaction, and desorption in terms of elementary steps Suppose a simple chemical reaction ( A + B → P ) happens in the presence of. .. transition metal changes, i.e., a downshift of d states, on the adsorption of one adsorbate, lead to weaker interactions with other adsorbates; third, elastic interaction brought by local distortions of the surface lattice on adsorption contributes to repulsive interaction with other adsorbates; fourth, non-local electrostatic effect can be justified as dipole-dipole interaction 15 The strong coverage... the study on electronic interaction between aromatics and transition metals has been extensively investigated First, Yamagishi et al (2001) studied benzene adsorption on the Ni(111) surface of p(√7×√7)R19.1º unit cell They approached at the level of molecular orbitals to analyze benzene molecular chemisorption on the surface During a molecular adsorption on a metal surface, the frontier orbitals, such... Comprehensive understanding on the reaction mechanism, electronic interaction and characteristics of adsorption is indispensable so that voluminous studies on adsorption, especially chemisorption, have been performed both experimentally and computationally As the result of the interaction between the adsorbate and the surface of catalysts, chemical bonds are formed at active sites of transition metal catalysts,... used in the rational design for novel catalysts 1.3 Application of First Principles- based Modeling This section is devoted to describe briefly how first principles- based modeling has contributed to the advancement of catalytic reactivity and selectivity in the design of new catalysts The successful application of first principles- based modeling in the prediction of ammonia synthesis on the Ru(0001)... design of Co-Mo bimetallic catalyst for ammonia synthesis analyzing the volcano-shaped correlation among ammonia synthesis activity 1 and nitrogen adsorption energy of potential catalysts with the help of density functional theory (DFT) calculations (Jacobsen et al., 2001) First principles calculations has been employed in the search for novel catalysts demonstrating that a reaction rate of ammonia synthesis .. .FIRST PRINCIPLES STUDY OF BENZENE ADSORPTION ON TRANSITION METAL SURFACES HONG WON KEON (B Eng.(Hons.), Sungkyunkwan University, Republic of Korea) A THESIS SUBMITTED FOR THE DEGREE OF MASTER... Electron volt fcc Face-centered cubic hcp Hexagonal closed-packed ΔEads Adsorption energy ΔHads Heat of Adsorption ΔΦ Change in workfunction of metal surface upon adsorption Φ Workfunction of metal. .. Interaction in Heterogeneous Catalysis In this section, the basic concepts of chemical bonding to transition metal surfaces will be understood First, carbon monoxide interaction with the metal surface

Ngày đăng: 06/10/2015, 21:06

Từ khóa liên quan

Mục lục

  • Thesis-Chapter 0.pdf

  • Thesis-Intro Ch 1.pdf

  • Thesis- Method Ch 2.pdf

  • Thesis-R&D Ch 3.pdf

  • Thesis-Conclusion Ch 4.pdf

  • Thesis-Appendix.pdf

Tài liệu cùng người dùng

  • Đang cập nhật ...

Tài liệu liên quan