PREDICTION OF STRUCTURAL AND THERMOCHEMICAL PROPERTIES: COMPUTATIONAL STRATEGIES FOR SMALL MOLECULES TO PERIODIC SYSTEMS By Zainab H. A. Alsunaidi A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Chemistry – Doctor of Philosophy 2017 ABSTRACT PREDICTION OF STRUCTURAL AND THERMOCHEMICAL PROPERTIES: COMPUTATIONAL STRATEGIES FOR SMALL MOLECULES TO PERIODIC SYSTEMS By Zainab H. A. Alsunaidi The prediction of thermochemical properties is central in chemistry and is essential in industry to predict the stability of materials and to gain understanding about properties of reactions of interest such as enthalpies of formation, activation energies and reaction enthalpies. Advances in state-of-the-art computing and algorithms as well as high-level ab initio methods have accounted for the generation of a considerable amount of thermochemical data. Today, the field of thermochemistry is largely dominated by computational methods, particularly with their low cost relative to the cost of experiment. There are many computational approaches used for the prediction of thermochemical properties. In selecting an approach, considerations about desired level of accuracy and computational efficiency need to be made. Strategies that have shown utility in the prediction of thermochemical properties with high accuracy and lower computational cost than high-level ab initio methods are ab initio composite approaches, or model chemistries, such as the correlation consistent Composite Approach (ccCA). ccCA has been shown to predict enthalpies of formation within “chemical accuracy”, which is considered to be 1 kcal mol-1, on average, for main group elements with respect to well-established experiments. In this dissertation, ccCA and the commonly used Gn composite methods have been utilized to establish effective routes for the determination of structural and thermochemical properties of oxygen fluorides species and for organoselenium compounds. To assess the reliability of these approaches, enthalpies of formation were calculated and compared to experimental data. Density functionals have also been employed in these projects to examine their performance in comparison to experiments as well as to composite methods. The impact of several thermochemical approaches on the accuracy of the predicted enthalpies of formation via various computational methods has been also considered such as the traditional atomization approach and molecular reaction approaches. Additionally, in this dissertation, the reaction of a direct amination of benzene to produce aniline on the Ni(111) surface was investigated to identify possible reaction intermediates and to determine the thermodynamically preferred reaction pathways. The adsorption behaviors and energetics of all species involved in this reaction are presented. Periodic density functionals were used to consider this heterogeneous catalytic process. Because DFT is based on the uniform electron gas model, which in principle resembles the band theory of metallic systems, DFT is particularly good at modeling metallic systems and thus well suited for the study of heterogeneous catalysts at the molecular level. Copyright by ZAINAB H. A. ALSUNAIDI 2017 This dissertation is dedicated to my husband Uthman and children Riadh, Sahar, and Yousef. Thank you for always believing in me. v ACKNOWLEDGMENT I would first like to praise and thank Allah for all the countless blessings and for giving me the strength and courage to complete this dissertation. The faith in his support, glory, and mercy has always held me strong. Upon completion, I would like to acknowledge many people who have been supportive during my graduate career. First, I would like to thank my research advisor Professor Angela K. Wilson for continuous support and patience during my PhD study at both the University of North Texas and Michigan State University. Her mentorship, guidance, motivation, and leadership have contributed to my growth as an independent scientist. I am very grateful to her for encouraging me to get publications and to present at conferences. Her words, advice, and actions have all granted me an incredible academic experience at the U.S. I acknowledge financial support from the University of Dammam in Saudi Arabia. I believe that this scholarship has contributed successfully to my personal and academic development. It also gave me the opportunity to meet and learn from outstanding scientists from around the world. Without their financial support, none of this would have been possible. I acknowledge Professor Thomas Cundari for collaboration on the work presented in Chapter 5 and for his support and enthusiastic sharing of knowledge. I thank my committee members here at Michigan State University Professor James “Ned” Jackson, Professor Gary Blanchard, and Professor Katharine Hunt for their time and suggestions. I also thank my former committee members at the university of North Texas Professor Thomas Cundari, Professor Paul Marshall, and Professor Mohammad Omary for their time and suggestions during my qualifying exams. vi Thank you to my amazing colleagues in the Wilson group, past and present, for their continuous sharing of knowledge and techniques. Thank you to Dr. Jiaqi Wang and Michael Jones for support and helpful conversation. I thank Dr. Kameron Jorgensen, who introduced me to ab initio composite methods, Dr. Charlie Peterson for fruitful discussion and for help with my solid state calculations, and Dr. George Schoendorff and Dr. Inga Ulusoy for thoughtful and interesting discussions. I thank Dr. Deborah Penchoff, for her friendship, many enriching conversations, and help in managing challenges and stress along the way. I thank my friends who have been there for me during my Ph.D. work. All of my friends in Denton have helped me to keep focus and maintain balance through these hectic years. I thank them for constant cheering and for making our stay in Denton a pleasant experience. The support from my friends in Saudi Arabia has always been substantial for my success. Their words, prayers, and encouragement always have been like fuel to continue and to succeed. I am especially grateful to my immediate family for their love, prayers, words and encouragements that have been sources of strength and inspiration. I express my greatest gratitude to my parents, Sajedah and Dr. Husain, for their endless support, prayers, and love, and who made me what I am today. I am very grateful to my mother- and father-in-law, Khaledah and Riadh, for their support, love, and encouragement. My deepest gratitude goes to all my siblings for consistent motivation, support, love, prayers and always believing in me. Finally, I am endlessly grateful to the love of my life, my husband, Uthman Alhumaidan, for his patience, understanding, and keeping me calm through all of my stressful challenges. Your wisdom, genius, kindness, love, and trust have contributed the most in my success. No such words can describe my gratitude and appreciation. You have been always there for me either to celebrate a little achievement or to overcome obstacles. I cannot remember any time that you vii were disappointed in me even with all of my frustration and absence. Thank you for reminding me and keeping me involved in my kids’ childhood. Without your support, none of this would have been possible. To my precious gems Riadh, Sahar, and Yousef, thank you for being such responsible, supportive, and respectful kids. Thank you for always bringing joy and love to my life. You are my heroes and my everything. viii TABLE OF CONTENTS LIST OF TABLES ............................................................................................................................... xii LIST OF FIGURES ............................................................................................................................. xv LIST OF SCHEMES.......................................................................................................................... xvii CHAPTER 1 INTRODUCTION ........................................................................................................... 1 CHAPTER 2 THEORETICAL BACKGROUND ................................................................................. 6 2.1 The Schrödinger Equation .................................................................................................... 6 2.1.1 Born-Oppenheimer Approximation ............................................................................................ 8 2.1.2 Slater Determinant....................................................................................................................... 9 2.1.3 Variational Principle ................................................................................................................. 10 2.2 Ab initio Methods................................................................................................................ 11 2.2.1 Hartree-Fock Approximation .................................................................................................... 11 2.2.2 Post-HF Methods....................................................................................................................... 13 2.2.2.1 Configuration Interaction ...................................................................................... 14 2.2.2.2 Perturbation Theory .............................................................................................. 15 2.2.2.3 Coupled Cluster Theory ........................................................................................ 17 2.3 Density Functional Theory ................................................................................................. 19 2.3.1 Local Density Approximation (LDA) ....................................................................................... 21 2.3.2 Generalized Gradient Approximation (GGA) ........................................................................... 22 2.3.3 Meta-GGA................................................................................................................................. 22 2.3.4 Hybrid- GGA............................................................................................................................. 23 2.4 Basis Sets ............................................................................................................................ 23 2.4.1 Atom-Centered Basis Sets ......................................................................................................... 24 2.4.1.1 Pople Basis Sets .................................................................................................... 26 2.4.1.2 Correlation Consistent Basis Sets ......................................................................... 27 2.4.2 Plane Wave Basis Sets .............................................................................................................. 29 2.4.2.1 Pseudopotentials ................................................................................................... 30 2.5 Composite Methods ............................................................................................................ 30 2.5.1 Gaussian-n (Gn) Theory ............................................................................................................ 31 2.5.2 Correlation Consistent Composite Approach (ccCA) ............................................................... 33 REFERENCES………………………………………………………………………………….……37 CHAPTER 3 DFT AND AB INITIO COMPOSITE METHODS: INVESTIGATION OF OXYGEN FLUORIDE SPECIES1 ........................................................................................................................ 48 3.1 Introduction ......................................................................................................................... 48 3.2 Computational Methodology .............................................................................................. 54 3.3 Results and Discussion ....................................................................................................... 55 3.3.1 Structures................................................................................................................................... 55 3.3.2 Enthalpies of Formation (ΔH°f, 298) ........................................................................................... 65 3.1 Conclusion .......................................................................................................................... 76 ix REFERENCES ......................................................................................................................... 78 CHAPTER 4 ENTHALPIES OF FORMATION FOR ORGANOSELENIUM COMPOUNDS VIA SEVERAL THERMOCHEMICAL SCHEMES ................................................................................. 86 4.1. Introduction ........................................................................................................................ 86 4.2 Computational Details ........................................................................................................ 93 4.2.1 Methods ..................................................................................................................................... 93 4.2.2 Thermochemistry ...................................................................................................................... 94 4.2.2.1 Atomization Approach (RC0) ............................................................................... 94 4.2.2.2 Isodesmic (RC2) and Hypohomodesmotic (RC3) Reaction Schemes .................. 95 4.2.3 Organoselenium Compounds .................................................................................................... 96 4.2.4 Reference Data .......................................................................................................................... 99 4.3 Results and Discussion ..................................................................................................... 102 4.3.1 Atomization Approach Using Composite Methods and B3PW91 .......................................... 104 4.3.2 Homodesmotic Approach Using Composite Methods and B3PW91...................................... 107 4.3.3 Atomization Approach Using Single Point Energies within the ccCA Method...................... 113 4.3.4 Homodesmotic Approach Using Single Point Energies within the ccCA Method ................. 116 4.4 Conclusion ........................................................................................................................ 120 APPENDIX ............................................................................................................................. 124 REFERENCES ....................................................................................................................... 133 CHAPTER 5 TOWARDS A MORE RATIONAL DESIGN OF THE DIRECT SYNTHESIS OF ANILINE: A DFT STUDY ............................................................................................................... 141 5.1 Introduction ....................................................................................................................... 141 5.2 Computational Details ...................................................................................................... 145 5.2.1 Method .................................................................................................................................... 145 5.2.2. Surface Model ........................................................................................................................ 146 5.2.3 Thermochemistry .................................................................................................................... 147 5.2.4 Method Calibration ................................................................................................................. 148 5.3 Results And Discussion .................................................................................................... 152 5.3.1 Adsorption Geometries and Energies ...................................................................................... 152 5.3.1.1 NH3*, NH2*, NH*, N*, and H*.............................................................................. 152 5.3.1.2 C6H6*, C6H5*, C6H5NH2*, and C6H5NH* .......................................................... 156 5.3.2 Decomposition Reactions ........................................................................................................ 165 5.3.2.1 Decomposition of Ammonia ............................................................................... 165 5.3.2.2 Decomposition of Benzene ................................................................................. 167 5.3.3 Production of Aniline .............................................................................................................. 167 5.3.3.1 Langmuir-Hinshelwood Mechanism .................................................................. 168 5.3.3.2 Rideal-Eley Mechanism ...................................................................................... 169 5.3.4 Desorption of Aniline .............................................................................................................. 170 5.3.5 Density of States...................................................................................................................... 171 5.4. Conclusions ...................................................................................................................... 175 APPENDIX ............................................................................................................................. 176 REFERENCES ....................................................................................................................... 188 CHAPTER 6 CONCLUDING REMARKS....................................................................................... 196 6.1 DFT and Composite Methods Investigations ................................................................... 196 6.1.1 Enthalpies of Formation for Oxygen Fluoride Species via Atomization Approach................ 196 x 6.1.2 Enthalpies of Formation for Organoselenium Compounds via Reactions Schemes ............... 197 6.1.3 Future Interest ......................................................................................................................... 198 6.2 Periodic DFT for the Direct Amination of Benzene on the Ni(111) ................................ 199 6.2.1 Future Interest ......................................................................................................................... 200 REFERENCES ....................................................................................................................... 201 xi LIST OF TABLES Table 3.1 Structural parameters of the oxygen fluoride species at different level of theories, bond lengths are in angstroms and bond angles and dihedral angles in degree. ............................ 56 Table 3.2 The expectation value of the total spin . ............................................................... 62 Table 3.3 Enthalpies of formation for chlorine oxides and related hydrides. ............................... 67 Table 3.4 Enthalpies of formation for the oxygen fluoride species using M06 and M06-2X paired with the correlation consistent basis sets. ............................................................................. 68 Table 3.5 Enthalpies of formation for oxygen fluoride species using the different variants of ccCA method. ....................................................................................................................... 69 Table 3.6 Calculated enthalpies of formation for the oxygen fluoride species using all methods and the MADs of these methods with respect to the reference data. .................................... 70 Table 3.7 Calculated enthalpies of formation for the oxygen fluoride species using M06 and M06-2X methods based on B3LYP/aug-cc-pVTZ geometries. ........................................... 74 Table 3.8 Calculated Enthalpies of formation for the oxygen fluoride species using all methods and the MADs of these methods with respect to the ATcT values. ...................................... 75 Table 4.1 The definition of the homodesmotic reaction schemes according to Wheeler.57,58 ...... 91 Table 4.2 Experimental and estimated enthalpies of formation in kcal mol-1 for organoselenium compounds.a,b ...................................................................................................................... 101 Table 4.3 Calculated enthalpies of formation (in kcal mol-1) for the elemental products and reactants of organoselenium compounds via the atomization approach (RC0). ................. 102 Table 4.4 Calculated enthalpies of formation (in kcal mol-1) for the hydrocarbon fragments via the atomization approach (RC0). ........................................................................................ 103 Table 4.5 Calculated enthalpies of formation (in kcal mol-1) for target organoselenium molecules via atomization approach (RC0). ........................................................................................ 106 Table 4.6 Calculated enthalpies of formation (in kcal mol-1) for target molecules via the isodesmic approach (RC2). ................................................................................................. 110 Table 4.7 Calculated enthalpies of formation (in kcal mol-1) for target molecules via the hypohomodesmotic approach (RC3). ................................................................................. 111 xii Table 4.8 Overall differences between RC0, RC2, and RC3 using composite methods and B3PW91. ............................................................................................................................. 112 Table 4.9 Calculated enthalpies of formation (in kcal mol-1) for target organoselenium molecules via the atomization approach (RC0) using the specified level of theory at the B3LYP/ccpVTZ geometry.a ................................................................................................................ 114 Table 4.10 Calculated enthalpies of formation (in kcal mol-1) for target organoselenium molecules via isodesmic reaction approach (RC2) using the level of theories specified in the table, at the B3LYP/cc-pVTZ geometry.a ........................................................................... 117 Table 4.11 Caculated enthalpies of formation (in kcal mol-1) for target organoselenium molecules via hypohomodesmotic reaction approach (RC3) using the level of theories specified in the table, at the B3LYP/cc-pVTZ geometry.a ........................................................................... 118 Table 4.12 Overall differences between RC0, RC2, and RC3 using the level of theories specified in the table.a......................................................................................................................... 119 Table 5.1 Bond dissociation energies (BDE) of gaseous molecules in kcal mol-1. .................... 149 Table 5.2 Adsorption energies (Ead) for ammonia and benzene adsorbed on a Ni(111) surface in kcal mol-1. ........................................................................................................................... 151 Table 5.3 Adsorption sites and adsorption energies (Ead) (in kcal mol-1) for NHx (x = 0 - 3) species and H adsorbed on a Ni(111) surface. .................................................................... 153 Table 5.4 Bond lengths (r) in Å and bond angles (a) in degree of free and adsorbed NHx (x = 1 3) species, adsorbed N, and adsorbed H. ............................................................................ 155 Table 5.5 Vibrational frequencies (in cm-1) of free and adsorbed NH3, NH2, and NH. ............. 156 Table 5.6 Adsorption sites and adsorption energies (Ead) (in kcal mol-1) of the aromatic species adsorbed on a Ni(111) surface. ........................................................................................... 157 Table 5.7 Bond lengths (r) in Å of free and adsorbed benzene and phenyl. ............................... 162 Table 5.8 Bond lengths (r) in Å and bond angles (a) in degrees of free and adsorbed aniline and anilide. ................................................................................................................................. 163 Table 5.9 C-H, C-C, C-N, and N-H vibrational frequencies (in cm-1) of gas phase and adsorbed C6H6, C6H5, C6H5NH2, and C6H5NH. ................................................................................. 164 Table 5.10 Reaction energies ΔE, enthalpies ΔH, and free energies ΔG of the adsorption and decomposition reactions of ammonia on the Ni(111) surface, all in kcal mol-1.a ............... 166 xiii Table 5.11 Reaction energies ΔE, enthalpies ΔH, and free energies ΔG of the adsorption and decomposition reactions of benzene on the Ni(111) surface, all in kcal mol-1.a................. 167 Table 5.12 Reaction energies ΔE, enthalpies ΔH, and free energies ΔG of the proposed reaction processes of the production of aniline on the Ni(111) surface, all in kcal mol-1.a .............. 169 Table 5.13 Reaction energies ΔE, enthalpies ΔH, and free energies ΔG of the desorption reaction of aniline on the Ni (111) surface, all in kcal mol-1.a .......................................................... 170 Table 5.14 Bond lengths (r) in Å and bond angles (a) in degree of free and adsorbed benzene and phenyl. ................................................................................................................................. 183 Table 5.15 Bond lengths (r) in Å and bond angles (a) in degree of free and adsorbed aniline and anilide. ................................................................................................................................. 184 Table 5.16 Calculated vibrational frequencies of gas phase and adsorbed aromatic species, in cm-1. ................................................................................................................................ 185 xiv LIST OF FIGURES Figure 3.1 B3LYP/aug-cc-pVTZ structures of the oxygen fluoride species included in this study. ............................................................................................................................................... 64 Figure 4.1 Selenium-containing elemental reactants and products. ............................................. 97 Figure 4.2 Isodesmic (RC2) reaction schemes. ........................................................................... 98 Figure 4.3 Hypohomodesmotic (RC3) reaction schemes. ............................................................ 98 Figure 4.4 The mean absolute deviations (MADs) of the calculated enthalpies of formation using three thermochemical schemes, RC0, RC2, RC3, with composite methods and single point energy calcuations utilized within ccCA methodology. ..................................................... 123 Figure 4.5 The optimized structures for organoselenium molecules at the B3LYP/cc-pVTZ level of theory. a. Dimethyl selenide. b. Dimethyl diselenide. c. Divinyl selenide. d. Diethyl selenide. e. Diethyl diselenide. f. Diisopropyl selenide. g. Dipropyl selenide. h. Dibutyl selenide ............................................................................................................................... 125 Figure 5.1 Adsorption structure of ammonia (a) and benzene (b) on the Ni(111) surface. (*) Indicates an adsorbed molecule. ......................................................................................... 151 Figure 5.2 Adsorption sites of NH2 (a), NH (b), N (c), and H (d) on the Ni(111) surface. (*) Indicates an adsorbed molecule..................................................................................... 154 Figure 5.3 Adsorption sites of phenyl (a), aniline (b), and anilide (c) on the Ni(111) surface. (*) Indicates an adsorbed molecule..................................................................................... 160 Figure 5.4 A top view of aniline adsorbed on the p(3x3) surface unit cell of Ni(111) on bridge site A (a), on the p(5x5) surface unit cell on bridge site B (b), and a side view of aniline adsorbed on the large 5x5x4 Ni(111) slab model. .............................................................. 161 Figure 5.5 The local density (LDOS) of N in NH (a), N (b), and Ni (d) in NH/Ni(111) system and the density of states (DOS) of clean Ni(111) surface (c). ................................................... 174 Figure 5.6 Spin density of the NH/Ni(111) system. ................................................................... 174 Figure 5.7 Lattice constant convergence test for bulk fcc Ni using variable Ecutoff and fixed kpoints (9x9x9). The DFT-optimized lattice constant is 3.52 Å .......................................... 177 Figure 5.8 K-points convergence test for the Ni(111) surface using two values of Ecutoff (400 and 500 eV). The optimal set of k-points for our systems is 5x5x1. ......................................... 177 xv Figure 5.9 Energy cut off convergence test for the Ni(111) surface using 5x5x1 k-points. The optimal energy cut off for our systems is 400 eV................................................................ 178 Figure 5.10 Adsorption sites and adsorption energies (Ead) of NH2 adsorbed on an fcc site (a), and NH2 adsorbed on an atop site (b) of the Ni(111) surface. (*) Indicates an adsorbed molecule. ............................................................................................................................. 179 Figure 5.11 Adsorption sites and adsorption energies (Ead) of C6H5 adsorbed on an hcp site (a), and C6H5 adsorbed on a bridge site (b) of the Ni(111) surface. (*) Indicates an adsorbed molecule. ............................................................................................................................. 179 Figure 5.12 The local density of states (LDOS) of N in NH2 (doublet) (a), N (b), in NH2/Ni(111) system. ................................................................................................................................ 180 Figure 5.13 The density of states (DOS) of clean Ni(111) surface (a) and the local density of states (LDOS) of Ni in NH2/Ni(111) system (b). ............................................................... 181 Figure 5.14 Spin density of the NH2/Ni(111) system. ................................................................ 182 xvi LIST OF SCHEMES Scheme 5.1 Synthesis of aniline. Adapted from Reference 1. .................................................... 142 Scheme 5.2 The proposed direct route of the production of aniline. .......................................... 143 Scheme 5.3 The overall modeled reaction for the direct production of aniline. ......................... 144 Scheme 5.4 The Ni(111) surface model. .................................................................................... 146 Scheme 5.5 Possible reactions pathways for the production of aniline and the change in the reaction free energies for each process. ................................................................... 171 xvii CHAPTER 1 INTRODUCTION Computational chemistry is a rapidly growing field of chemistry that uses fundamental physics, mathematics, and computer simulation to investigate chemical systems. Advances in high computing performance have played an important role in improvements in computational methods and algorithms that, in turn, facilitate the elucidation of complex chemical systems, such as solids, biological macromolecules, polymers, and ionic liquids. Information such as electronic structure, reactivity of atoms and molecules, thermochemical and spectroscopic properties, just to name a few, can all be reliably provided from high-level theoretical calculations. This information helps chemists to gain in-depth understanding of a particular chemical problem and develop and design new reactions or systems. Computational applications also aid in interpreting available experimental measurements and provide predictions for systems/properties that are unknown or unexplored, or that are difficult to address experimentally. Nowadays, applications in critical areas such as anti-cancer drugs, energy development, astronomy, and green chemistry have been explored theoretically. Two main classes of computational chemistry methods are molecular mechanics and electronic structure methods and within each class are numerous approaches. Each family of approaches differs in the type of approximation, attainable accuracy, and computing resources required (disk space, time, and memory). Knowing the applicability and limitations of a theory allow suitable methods to be selected for a given chemical problem while taking into account the system size, the target accuracy, the available computing resources, and the property of interest. Molecular mechanics (MM) are classical mechanics approaches that solve for Newton’s equations of motion using pre-determined empirical force fields to describe potential energy 1 surface (PES). MM are highly useful approaches for providing molecular and thermodynamic properties of large molecules, as large as >10000 atoms, such as enzymes, from a macroscopic perspective. These approaches are used intensively in biochemical research. The reliability of a MM calculation depends on how accurate the empirically derived force fields are and how suitable/transferable they are for a system of interest and on the correct sampling of the phase space. Electronic structure methods, on the other hand, are quantum mechanical approaches that are based on solving the time-independent Schrödinger equation and are increasingly important in chemistry, physics and material science. They can be categorized into three classes: ab initio molecular orbital (MO) theory, density functional theory (DFT), and semi-empirical methods. Both semi-empirical and ab initio methods are wave function based; however, unlike ab initio methods, semi-empirical methods, as expected from the name, use fitting parameters typically derived from experimental data (e.g. enthalpy of formation, ionization potentials, structural parameters, and dipole moments) to compensate for approximations made to the quantum mechanical model, such as neglecting two-electron integrals, i.e. electron correlation is approximated using empirical parameters. Although semi-empirical methods are computationally faster compared to ab initio methods, and can be applied to systems containing >1000 atoms, they are less accurate than ab initio methods and their accuracies are dependent on the experimental database for which they were parameterized. In this dissertation, ab initio molecular orbital (MO) theory and density functional theory (DFT), discussed in more detail in Chapter 2, were employed to investigate systems spanning from small inorganic molecules to larger organic compounds to extended periodic systems. The aim is to utilize various computational approaches and strategies to predict accurate (within “or 2 near “chemical accuracy”) thermochemical properties and to evaluate the performance of these methods relative to experiment, where available. Another aspect of this work is to use density functional theory as well as statistical thermodynamic to identify and study possible adsorbed intermediates involved in a heterogeneous reaction process. Ab initio methods are developed from first principles (physical constants), i.e. do not use empirically obtained force fields or parameters, and use several approximations to solve for the many-body Schrödinger equation. If these approximations are small in magnitude (such as in Configuration Interaction (CI)), accurate solutions to the Schrödinger equation can be achieved yet at high computational cost. These less approximated methods can be only applied to small molecules. Density functional theory, however, can be applied to larger systems (~100 - 1000 atoms) at a moderate computational cost, i.e. DFT scales as N4, where N represents the number of basis functions. DFT solves for the Schrödinger equation in terms of the total electron density instead of a wave function. The solution for the Schrödinger equation provides a real numerical value of the energy of a system and its electron distribution. Calculating the second derivative of the energy (Hessian matrix) with respect to atomic positions determines force constants and vibrational frequencies. This information allows predicting thermochemical properties including enthalpies of formation, bond dissociation energies, ionization potentials, and electron affinities. There are strategies that have been established to address the accuracy and high computational cost demands of high-level ab initio methods. For example, a number of ab initio composite approaches, or model chemistries, have been developed for predicting thermochemical properties with accuracies that mimic those possible from high-level ab initio methods, such as coupled cluster with single, double, and perturbative triple excitations (CCSD(T)), but with reduced computational cost. Composite methods utilize a series of steps 3 combining lower level methods and basis sets to replicate results with higher-level methods. Gaussian-n (Gn) composite methods and the correlation consistent Composite Approach (ccCA) both are discussed in Chapter 2 and were utilized in Chapter 3 and Chapter 4. Understanding the utility of theoretical methods for predicting energetic properties of various classes of molecules is important, and can lead to the development of effective strategies for future studies. In Chapter 3, composite methods including ccCA, G3, and G3B3 are utilized for the prediction of enthalpies of formation (∆ ° ′ ) for oxygen fluoride species, which is compared to reliable experimental data. In addition, the performance of several density functionals including M06 and M06-2X, for determining structures and ∆ ° ′ of the same set of oxygen fluorides also was examined. The set includes fluorides, difluorides, dioxides, trioxides, and the corresponding hydrides. These molecules have been a great challenge for the computational community particularly due to the unusual geometry of molecules such FOO and FOOF. The research in Chapter 4 includes the prediction of accurate enthalpies of formation for a set of organoselenium compounds using several thermochemical schemes via composite methods and density functionals. Selenium compounds play a substantial role in organic synthetic reactions, in the semiconductor industry, and in biochemistry. Thermochemical properties, particularly ∆ ° ′ , are of great importance in providing predictions and insight about chemical reactions including molecular stability, reaction enthalpy, and bond dissociation energy. Isodesmic and hypohomodesmotic reaction schemes have been developed to cancel errors arising from differential correlation effects and size extensivity (which will be discussed herein), which, in turn, can result in a better prediction of ∆ ° 4 as compared with the conventional atomization approach. In Chapter 5, plane-wave DFT is employed to investigate possible reaction mechanisms for the direct amination of benzene on the Ni(111) surface. The direct amination of benzene to produce aniline is of significant interest from a green chemistry perspective, although, it is very challenging because of the strong C-H bond of benzene and the N-H bond of ammonia. In this research, a full detailed study of this reaction including adsorption behaviors (structures, energetics, frequencies, electronic interactions) of all species on the Ni(111) surface and thermochemical properties is done and suggestions of reliable computational models are provided. The study uses the PBE and PBE-D3 functionals to determine adsorption structures and energetics for a variety of adsorption systems including adsorbed NHx species and adsorbed aromatic species to consider possible reaction intermediates relevant to the direct amination of benzene. Statistical thermochemistry is employed to calculate reaction enthalpies and reaction free energies of the proposed reaction pathways between imide and benzene. The study also aims to compare and correlate a heterogeneous Ni(111)–imide model with the corresponding homogenous nickel-imide model for the C-H amination reaction. In Chapter 6, all of the research projects are summarized and future interests are discussed. 5 CHAPTER 2 THEORETICAL BACKGROUND 2.1 The Schrödinger Equation In quantum mechanics, electrons are described using a wave function (Ψ). The wave function conveys all information of the state of the system and can be used with an operator to measure an observable property of a given system. A cornerstone wave equation of quantum mechanics is the non-relativistic time-dependent Schrödinger equation,1–6 Eq. 2.1, − where ℏ Ψ , , = Ψ , , is the Hamiltonian operator (the energy operator), Ψ , , (2.1) is the wave function of electronic coordinates, nuclear coordinates, and time, respectively. equals to √−1, and ℏ equals to ℎ ⁄ 2 , where ℎ is the Planck’s constant. The time-dependent Schrödinger equation describes how a wave function Ψ , , for a system evolves with time when governed by a Hamiltonian. In other words, it describes the dynamics of a given system and provides the probability distribution of the state of the particle at each time. In quantum chemistry, many applications are concerned with obtaining the constant energy of stationary states, i.e. standing waves that have no time dependence, of a chemical system. Hence, the non-relativistic time-independent Schrödinger equation, Eq. 2.2, is widely used, which can be derived from Eq. 2.1 using the separation of variables principle in which the wave function is factored into a function that depends only on spatial coordinates and a function that depends only on time. The time-independent Schrödinger equation, Eq. 2.2, is employed in this dissertation and will be referred to as the Schrödinger equation henceforth. It is a second order differential equation in electronic and nuclear coordinates. The Schrödinger equation is 6 also an eigenvalue equation, where the Hamiltonian operator (H) is applied to a wave function , ( ) that describes the chemical system to give the total energy (E) of that system.7 , = , (2.2) The Hamiltonian is shown in Eq. 2.3 in atomic units and is expressed as the sum of the kinetic energy ( ) and potential energy ( ) operators of N electrons (i and j) and M nuclei (A and B) within a system: % 1 1 % % % 1 1 ∇/$ 3/ 1 $ = − ! ∇# − ! −!! + !! | / − #| 2)*))+ 2 0/ ())))*) 7 # − 87 ()*)+ () )))+ () #&' /&' /&' #&' #&') 89# ))*) )))+ ,- 1 ,2 1 3/ 3: +!! | / − :| ())))*))))+ /&' :9/ 52- 5-- (2.3) 522 where MA is the mass of the nuclei, 3/ and 3: are the charges of nuclei A and B, respectively. | / − # | is the distance between an electron and a nucleus, 7 # − 8 7 is the distance between two electrons, | / − :| is the distance between two nuclei, and ∇$ is the Laplacian operator. The first two terms represent the kinetic energy operator of the electrons ( ; ) and the nuclei ( < ). The following terms correspond to the potential energy operator of the attraction between electrons and nuclei ( <; ), the repulsion between electrons ( ;; ), and the repulsion between nuclei ( << ). 8 The Schrödinger equation, Eq. 2.2, elucidates that certain energy levels (eigenvalues) are allowed for a given system and correspond to wave functions that are the stationary state solutions (eigenfunctions) of this given system. The wave function itself is not measureable, however it can be squared ( ∗ ) to give the probability, | $ |of finding an electron within a given region in space, according to Max Born,9 i.e. it can describe the electron density of the 7 system. The integral of | normalized wave function. $| over all the volume space >? must be equal to one for a @| $| >? = 〈 | 〉 = 1 (2.4) 2.1.1 Born-Oppenheimer Approximation The Schrödinger equation cannot be solved exactly except for one electron systems such as hydrogen atom or hydrogenic cations. Thus, solving for the total energy for multi-electron systems requires the use of approximations due to the many body problems arising from the correlation of the motion of many electrons in the Schrödinger equation. One fundamental approximation is based on the fact that the masses of the nuclei are much larger than the mass of the electrons, which, in turn, means that the nuclei move more slowly than the electrons for a given kinetic energy. Thus, the nuclei can be considered fixed with respect to the motion of the electrons. This is the essence of the so-called Born-Oppenheimer approximation10 that allows a decoupling of the nuclear and electronic motions. With the assumption that the electrons move in the field of fixed nuclei, the Hamiltonian can be reduced to an expression independent of the nuclear motion8 as shown in Eq. 2.5, ;C;D % 1 % % % 1 3/ 1 = − ! ∇$E − ! ! +!! | / − #| 2)*))+ ())))*) 7 # − 87 () )))+ () #&' /&' #&' #&')89# ))*) )))+ ,- 52- 5-- (2.5) where the kinetic energy term of the nuclei ( < ) is zero when applying the Born-Oppenheimer approximation. The repulsion between the nuclei ( << ) becomes a constant that is obtained classically via Coulomb’s law. Since constant terms in an operator do not have an effect on the wave function, the ( << ) term is not included in the electronic Hamiltonian. When the electronic 8 Hamiltonian ( is obtained ;C;D ) is applied to a wave function, the electronic Schrödinger equation, Eq. 2.6, ;C;D where ;C;D = ;C;D ;C;D ;C;D is obtained from a separation of variables from (2.6) , . The addition of the potential energy of the repulsion of nuclei ( ;C;D energy results in the total energy FGF = FGF ;C;D << ) to the electronic of the system, as shown in Eq. 2.7. 1 1 3/ 3: +!! | / − :| ())))*))))+ /&' :9/ 522 (2.7) 2.1.2 Slater Determinant The electronic wave function ;C;D in the electronic Schrödinger equation (Eq. 2.6) must be a product of both the spatial and spin functions in order to fully describe an electron. As fermions, electrons cannot share the same set of quantum numbers, i.e. they must obey the Pauli exclusion principle.11,12 Furthermore, the electronic wave function is required to be antisymmetric with respect to the interchange between spatial or spin functions. To achieve both conditions a determinant can be utilized to construct an antisymmetric wave function which was introduced first by Slater.13 For an N-electron system x' , x$ , … , x% occupying N spin orbitals (J# , J8 , … , J% (2.8), the Slater determinant satisfies the requirement of an antisymmetric wave function, 9 Ψ x' , x$ , … , x% where ' √%! J# x' 1 J# x$ MM = ⋮ √K! J# x% J8 x' J8 x$ ⋮ J8 x% ⋯ ⋯ ⋱ ⋯ J% x' J% x$ M M ⋮ J% x% (2.8) is a normalization factor. The rows in a Slater determinant represent electrons and the columns represent spin orbitals. The interchange between two electrons (two rows) or two spin states (two columns) changes the sign of the determinant. In addition, if the same electron occupies the same spin orbital, then two columns will be equal and the determinant is zero.8 2.1.3 Variational Principle The variational principle states that for a well-behaved wave function the expectation value of the Hamiltonian must result in an energy ground state energy ( energy. Q) ) that is greater than or equal to the exact of that system, as shown in Eq. 2.9, i.e. there is a lower bound to the = 〈Ψ7H7Ψ〉 ≥ 〈Ψ|Ψ〉 Q (2.9) The wave function Ψ here is a trial wave function and the expectation value is the resultant trial energy . For a normalized wave function, the denominator 〈Ψ|Ψ〉, as mentioned earlier, equals 1 and thus the energy will equal to 〈Ψ7H7Ψ〉. The variational principle allows the determination of the best wave function that gives the most accurate energy, i.e. the lowest energy of a given system. The more parameters the trial wave function contains, the closer the trial energy is to the exact ground state energy. 10 In order to solve the Schrödinger equation, approximate forms of the electronic Hamiltonian, termed methods, are used in combination with basis sets, which are mathematical functions used to construct the wave function. There are now several methods utilized for this purpose, such as ab initio methods, semiempirical methods, and density functional theory (DFT). Here the focus is only on the ab initio and density functional methods. 2.2 Ab initio Methods Ab initio means “from the beginning,” indicating that these calculations are based on quantum mechanics, physical constants, and laws of physics without reference to empirical data or parameters. The lack of experimental parameters makes ab initio methods flexible to apply to different systems and problems without showing any degree of bias. Various ab initio methods have been developed based on the type of approximation employed to solve the Schrödinger equation. 2.2.1 Hartree-Fock Approximation The Hartree-Fock (HF) approximation14–18 is among the simplest ab initio methods and is often the first approximation made when solving the Schrödinger equation for multi-electron systems. It can be used to provide a well-defined starting point for more advanced wave function based methods. The electron-electron repulsion in HF theory is computed using the mean-field approximation that treats the motion of each electron in an average potential field of other electrons. The Hartree-Fock wave function ΨTU is a single Slater determinant and for an N electron system occupying N spin orbitals (J# , J8 , … , J% is 11 ΨTU = |J' J$ ⋯ J% V (2.10) The variational principle is used to select the “best” spin orbitals that will give the lowest energy when using the HF equation, Eq. 2.10. The eigenvalue HF equation can be written as where J# WX# J# = Y# J# (2.11) represents the one-electron orbital wave function, Y# is the energy of that orbital, and WX# is the Fock operator of the form $ > = W }, ~, • . v⋯> % (2.29) In the Kohn-Sham (KS) theory, the kinetic energy is calculated from an N independent non-interacting electron system that is constructed from a single determinant wave function of a ‚|ƒ for a fully interacting fully interacting system.31 The total Kohn-Sham DFT energy system is defined as in Eq. (2.30)20 where „ ‚|ƒ ‚|ƒ = „ ‚|ƒ + <; ‚|ƒ + …‚|ƒ + †t ‚|ƒ is the kinetic energy of the non-interacting electrons, (2.30) <; ‚|ƒ is the potential energy of the attraction between the nuclei and the electrons, …‚|ƒ is the Coulomb term representing the electron-electron repulsion, and †t ‚|ƒ is the exchange-correlation energy that accounts for all the remaining terms of the electron-electron interactions, e.g. the kinetic energy caused by interacting electrons. The DFT energy expression, Eq. (2.30), is within the Born-Oppenheimer approximation as there is no nuclear kinetic energy and the internuclear repulsions are still solved classically. The exchange-correlation energy †t ‚|ƒ is a functional of the electron density and can be divided into two approximate functionals; the exchange functional ( same-spin electron interactions and the correlation functional ( mixed-spin electron interactions as 20 † ‚|ƒ t ‚|ƒ corresponding to the corresponding to the †t ‚|ƒ = † ‚|ƒ + t ‚|ƒ (2.31) Because the exact form of the exchange-correlation functional is unknown, a variety of approximations of the †t formulations have been developed. These approximations can be categorized into two main classes: the local density approximation (LDA), which utilizes local functionals that depend only on the electron density | and the generalized gradient approximation (GGA) that uses gradient-corrected functionals that depend not only on the electron density | but also on its gradient ( ∇‡ ). Perdew42 presented a hierarchy of DFT functionals and named it the “Jacob’s Ladder” of DFT, which includes LDA (first rung), GGA (second rung), meta-GGA (third rung), hybrid-GGA (fourth rung), and double hybrid GGA (fifth rung). The most commonly used functionals are LDA up to the hybrid GGA, which will be introduced herein. 2.3.1 Local Density Approximation (LDA) The LDA32 is the simplest approximation in which the density is treated locally as a non- interacting uniform electron gas. In LDA, the density is defined as | = K/ , where N is the number of electrons and V is the volume of the gas. This form of electron density is only used in the LDA. Generally, since electrons have spin ‰ or spin Š, LDA is replaced by the local spin density approximation (LSDA). In LSDA, the total electron density | is replaced by the spin electronic density, |‹ and |Œ . The LSDA approximation generally provides better molecular geometries and vibrational frequencies than the HF approximation; however, it tends to overestimate the chemical bonding.33 The LDA approximation is commonly used in solid-state 21 physics for studying metals where the electron density varies slowly.34 One of the commonly used LSDA functionals is VWN.35 2.3.2 Generalized Gradient Approximation (GGA) Since in LDA the electron density is treated as a local property that does not reflect the spatial variation in densities, the GGA method36–41 corrects for this shortcoming by including the electron density and its first derivative ∇‡ . One of the commonly used GGA corrected exchange functionals is B88 developed by Becke36 in which a correction parameter is added to the LDA exchange energy. GGA corrected correlation functionals were also developed, such as the LYP38 functional and the P8642 functional. Because the exchange-correlation energy is comprised from two terms, the exchange and the correlation terms, the gradient-corrected exchange functional can be combined with the gradient-corrected correlation functional. The most widely used combinations are BLYP,36,38 PBE,40,41 and BP86.36,42 The PBE functional is a non-empirical functional and has numerous applications on metallic systems largely because it correctly describes the slowly varying electron densities of metals. 2.3.3 Meta-GGA Additional improvement over the GGA method is the meta-GGA (MGGA) approach in which more inhomogeneity property is added to the electron density by including the second derivative of the electron density ∇$ | and/or, local kinetic energy density ?| . These additional terms can offer a better performance than LDA and GGA particularly in the 22 description of chemical interactions. TPSS43 and M06-L44 are commonly used meta-GGA functionals. 2.3.4 Hybrid- GGA In DFT, the description of the exchange energy suffers from the self-interaction problem that arises from the spurious interaction of an electron with itself. The idea in hybrid functionals is to use the Hartree-Fock exact exchange in addition to the Kohn-Sham correlation. This can help improve the performance of the exchange functional of the DFT. However, the inclusion of 100% of the HF exchange term can in some cases worsen the performance of the functional because it can result in a poor description of the total exchange-correlation hole of DFT. Combination of HF exchange with DFT exchange, however, largely improves the description of molecularly properties.33,45 Two types of the hybrid functionals are developed. The first one is the hybrid GGA (HGGA) in which the HF exchange function is added to a pure GGA functional. B3LYP is one of the most popular hybrid-GGA functional containing 20% of HF exchange.33,38,46 In addition, B3PW91 is another popular hybrid-GGA functional that also includes 20% of HF exchange.33,47 The other type of hybrid functional is hybrid MGGA (HMGGA). In this type, a percentage of HF exchange energy is added to the MGGA functional. The Minnesota functionals M06, M062X, and M06-HF are all useful HMGGA functionals,48 in which M06 contains 27%, M06-2X contains 54%, and M06-HF contains 100% HF exchange. 2.4 Basis Sets In the previous sections, the focus was on using several computational methods, such as ab initio methods and density functional theory, to approximate the Hamiltonian operator in the 23 Schrödinger equation. However, the accuracy of all the methods depends on the number and quality of the basis functions (basis sets) that are used to describe the wave function. Basis sets are basically vectors that are used to define a certain region of space. In quantum chemistry, basis sets are mathematical functions that describe the molecular orbitals (MOs) and can be expanded as a linear combination of atomic orbitals or basis functions (J\ Eq. (2.32), •# = ! Ž\# J\ \ (2.32) where • represent a MO and Ž\# are the weighting coefficients indicating the relative importance of each basis function (or atomic orbital) J\ . A finite number of basis functions is always used in conjunction with an electronic structure method to construct the wave function ( Ψ ) which is used as a solution to the Schrödinger equation. This finite set can be a source of error in the calculation and is usually referred to as incomplete basis set error. Types of basis sets commonly used are atom-centered basis sets, such as the Pople basis sets49–52 and the correlation consistent basis sets developed by Dunning and coworkers.53–59 These basis sets are constructed using basis functions localized in a region of space. Another kind of basis set is constructed with plane waves. Plane waves are delocalized basis function basis sets that are usually used for describing periodic systems, such as crystals.60 2.4.1 Atom-Centered Basis Sets The atomic orbital basis sets are sets of localized functions used to construct atomic orbitals. The linear combination of these atomic orbitals produces the MOs of a given system. 24 Two types of these basis sets have commonly been used: Slater-type orbitals (STOs)7,61 and Gaussian-Type orbitals (GTOs)62–64. STOs are developed to resemble the hydrogen atomic orbitals and are expressed as: Ψ r, θ, ϕ = K ’C“ θ, ϕ <”' u ”•\ (2.33) where K is the normalization factor and ’C“ is the angular function and is defined by the angular – and magnetic — quantum numbers. is the distance between the electron and the nucleus and ˜ is the principle quantum number. ™ is the exponent and it determines the spatial extent of the function, i.e. small ™ gives diffuse functions while large ™ gives tight function. STOs can successfully describe short range and long range behavior to correctly describe the cusp at the nucleus and the tail, respectively. While STOs correctly describe these regions, electronic integrals cannot be solved analytically and thus must be computed numerically. This results in additional computational expense while also making the quality of the computed results dependent on the size of the integration grid. Alternatives to STOs are GTOs which have the following functional form: Ψ r, θ, ϕ = K ’C“ θ, ϕ where the radius exponent here is squared u ”•\ š $<”$”C u ”•\ š (2.34) compared to the exponent in the STOs u ”•\ . This formulation is easier to compute than the STOs since integrals over GTOs have analytic solutions. However, the performance of the GTOs at the cusp and the tail of a radial function is poor. GTOs have a zero gradient at the r=0 rather than a cusp in addition to decaying too slowly in the tail region. To improve the performance of the GTOs, multiple GTOs are used in a linear combination to reproduce the shape of STOs. This requires many functions, thus contracted 25 GTOs are often used to minimize the computational expense. As the number of GTOs included in the linear combination increases, the description of the atomic orbital using the GTOs become more accurate. A minimal basis set is the smallest number of basis functions that can describe all electrons of an atom, i.e. one function per atomic orbital. For example, the minimal basis set of hydrogen is one basis function representing the 1s atomic orbital. For the carbon atom, and other first row elements, the minimal basis set includes two functions for the 1s and 2s orbitals and three functions for the px, py, and pz. Although the minimal basis set can describe atomic orbitals, it is not enough to describe the molecular orbitals where atomic orbitals can undergo distortion. Furthermore, minimal GTO basis sets cannot reproduce the cusp and tail regions correctly. To overcome this shortcoming, additional basis functions can be added to the valence orbitals since they are essential for chemical bonding. These basis sets are called split-valence basis sets. To illustrate, a double-™ basis set has two GTO functions for each valence orbitals, a triple-™ basis set has three GTO functions for each valence orbitals, and so on. 2.4.1.1 Pople Basis Sets Pople and coworkers49–52 developed the Gaussian split-valence basis sets that employ the contracted GTOs concept. A general notation of a Pople double-™ basis set is X-YZG, where X represents the number of the primitive GTOs used to construct the contracted GTO for core orbitals. YZ represent the number of primitive GTOs used to make two contracted GTOs for the valence orbitals. For Pople triple-™ basis set, a number will be added to the split valence function to show how many primitive GTOs used to make the third contracted GTO for the valence orbital, e.g. X-YZTG. Common examples of the Pople double-™ basis set are 6-31G, 6-21G, 4- 26 31G, 3-21G, etc. and Pople triple-™ basis set can be written as 6-311G. Polarization functions can be added to the basis set to better describe distortion or polarization that occurs when a bond is formed. In order to allow for this polarization, the angular momentum of the polarizing function must increase by one compared with the type of function being polarized, e.g. p-functions to polarize s-functions, d-functions to polarize p-functions, etc. These functions are represented by an asterisk “*” as in 6-31G*. They also can be denoted by adding d or p, such as 6-31G(d,p), in which a set of d functions are added to non-hydrogen atoms and a set of p functions are added to the hydrogen atom. Diffuse functions that have small ™ exponents are added to account for long range interaction, such as if an electron is far away from the nucleus as happens in anionic systems. Diffuse functions are represented as a plus sign “+” as in 6-31+G which means s and p functions are added to non-hydrogen atoms. 2.4.1.2 Correlation Consistent Basis Sets The correlation consistent (cc) basis sets introduced by Dunning and coworkers53–59 were designed in order to systematically and predictably recover the correlation energy of the valence electrons by increasing the number of the basis functions per atoms.56 Specifically, these basis sets built on the idea that functions recovering same amount of correlation energy are added to the same shell. The correlation consistent (cc) basis sets are denoted as cc-pVnZ, where pV indicates polarized valence functions, Z is zeta, and n is related to the number of functions used to describe the valence orbitals. The n also can be an indication of the maximum angular momentum function contained within the basis set. The zeta level differs for different values of n. The values of n can be D(2), T(3), Q(4), 5, and 6 and so on. For example, for first row atoms, in which the minimal basis set is 2s1p, the cc-pVDZ set includes 3s2p1d, where a d polarized 27 function is added. The cc-pVTZ includes 4s3p2d1f, where the second d and the first f recover the same amount of correlation energy. The cc-pVQZ includes 5s4p3d2f1g, in which the third d, the second f, and the first g recover the same amount of correlation energy. The reason for the inclusion of these higher basis functions is to systematically recover more correlation energy of a molecule. The systematic recovery of the correlation energy built in these basis sets allows for the extrapolation of the correlation energy to the complete basis set (CBS) limit where the error arising from the use of a finite basis set vanishes, yet the intrinsic error from the chosen method remains. Various extrapolation techniques estimate the CBS limit, including the Peterson extrapolation scheme65 and the Schwartz extrapolation scheme.66 Further discussion of the extrapolation schemes of the correlation consistent basis set is presented in Section 2.5.2. Additive basis functions such as diffuse functions (aug)54,57 and polarized core-valence correlating functions (CV)56,59 can also be added without affecting the convergent behavior of the basis sets, such as aug-cc-pCVTZ. The diffuse functions are often added to help describe long-range interactions, while the polarized core-valence correlating functions are used when the subvalence or core electrons are included in the correlated method in addition to the electrons in the valence space. Additionally, a relativistic re-contraction of the Hartree-Fock set of functions is often used in conjunction with relativistic calculations.67 This re-contraction often provides faster convergence of the relativistic wave function as well as aiding in optimization to the correct electronic state. 28 2.4.2 Plane Wave Basis Sets Plane wave (PW) basis sets are basis sets used to describe valence electrons in extended systems, such as solids and crystals, and are used in conjunction with periodic boundary conditions (PBC). Since the valence electrons in periodic systems behave as free electrons, using the atom-centered basis sets to describe these systems is impractical. Rather, plane wave basis sets are used and are delocalized across the entire periodic system. The discrete energy levels vanish in periodic systems, thus bands are formed rather than localized atomic and molecular orbitals. For a periodic system, plane wave basis functions u #›œ are used to construct the periodic wave function Ψ› r , Eq. 2.35. This wave function obeys the Bloch theorem68 that states that “the eigenfunction of the wave equation for a periodic potential is the product of a plane wave u #›.œ times a function u› r with the periodicity of the crystal lattice”. Thus, Ψ› r represents a numerical solution to the Schrӧdinger equation. Ψ› r = u #›œ (2.35) In Eq. 2.35, ž is the wave vector. The size of a plane wave basis set is controlled by the highest value of the k vector.69 Plane wave basis sets are very large compared with all-electron basis sets. Since these basis sets are not spatially bound, an energy cut-off has to be set to allow for an SCF convergence. A common energy cut off value is 200 eV, which corresponds to about 20,000 basis functions depending on the size of the unit cell being modeled.20 29 2.4.2.1 Pseudopotentials The core electrons in the metallic systems are strongly localized near the nuclei, which in turn requires a large number of PW functions in order to describe this behavior. Plane wave basis sets work well in describing the valence electrons in extended systems but are inefficient at describing core electrons. Since core electrons are not significantly involved in chemical bonding and reaction, the effect of core electrons can be implicitly described using pseudopotentials, which smear the nuclear charges and model the effect of the core electrons. Different types of pseudopotentials have been developed, such as norm-conserving pseudopotentials,70 ultra-soft pseudopotentials,71 and the projector augmented-wave (PAW) pseudopotentials.72,73 The PAW method developed by Blӧchl is known to be an efficient method for predicting the electronic structure of materials because it is capable to include the upper core electrons as well as the valence electrons in the self-consistent iteration of the Kohn equations. In the PAW method, the exact wave function of the core electrons is mapped onto an auxiliary pseudo-wave function in the core region using projectors and retaining the correct nodal structure of the pseudo-wave function, in contrast, to standard pseudopotentail methods.72 2.5 Composite Methods In order to reduce the computational cost of high level ab initio calculations, composite methods or model chemistries have been developed. These methods are constructed of sequential additive steps of a combination of high level methods with a small basis sets and low level methods with large basis sets resulting in a high level of accuracy with lower computational cost. Composite methods that have been developed include the Gaussian-n (Gn) methods by Pople et 30 al.,74–83 the Weizmann-n (Wn) methods by Martin et al.,84–89 the High accuracy Extrapolated Ab initio Thermochemistry (HEAT) method by Stanton et al.,90–92 the complete basis set (CBS) methods by Petersson et al.,93–98 and the correlation consistent Composite Approach (ccCA) by Wilson et al.99–110 Herein the Gn methods and the ccCA are introduced and are used in Chapter 3 and Chapter 4. 2.5.1 Gaussian-n (Gn) Theory Gaussian-n methods are composite methods that perform sequential ab initio molecular calculations to predict accurate thermochemical properties at low computational cost. The original Gn method is the Gaussian-1 (G1) developed by Pople and coworkers74–83 with a target accuracy of ± 2 kcal mol-1 for compounds containing first-row elements and ± 3 kcal mol-1 for compounds containing second-row elements.74 The G1 method performs poorly for the dissociation energies of ionic species, triplet states, some hydrides, and hypervalent species.75 Therefore, a series of developments have been implemented in G1 and other versions of the Gn theories have been developed including G2,75 G3,78 and G482 with higher target accuracy of ~1 kcal mol-1. In addition, each Gn theory has several variants, for example, the G3 family of composite methods includes G3(MP2),80 G3B3,79 and G3-RAD.111 The most commonly used version of Gn methods is G3, which is used in Chapter 3 and Chapter 4 of this dissertation. In G3, a sequence of ab initio molecular orbital calculations is performed to predict accurate energetics. The following computational steps are used in G3:78 1. Initial geometry optimization and frequency calculations are carried out at the HF/631G(d) level of theory. A scaling factor of 0.8929 is used to correct for the anharmonicity in the vibrations. These frequencies are used to calculate the zero point energy, E(ZPE). 31 2. Equilibrium geometries are then obtained using the MP2(full)/6-31G(d) level of theory. This equilibrium structure is used for all the single point energy calculations that follow. 3. A single point energy calculation is performed at the MP4/6-31G(d) level. The energy obtained from this step is the reference energy of the G3 method. 4. A series of single point energy calculations are carried out at high level of theories and are used to improve the reference energy as follows: a. A correction for the correlation effect beyond the MP4 method using the quadratic configuration interaction method (QCISD(T)),112 ΔE(QCI): ΔE(QCI) = E[QCISD(T)/6-31G(d)]-E[MP4/6-31G(d)] (2.36) b. A correction for diffuse functions, ΔE(+): ΔE(+) = E[MP4/6-31+G(d)]-E[MP4/6-31G(d)] (2.37) c. A correction for higher polarization function, ΔE(2df,p): ΔE(2df,p) = E[MP4/6-31G(2df,p)]-E[MP4/6-31G(d)] (2.38) d. A correction for a larger basis set effect, ΔE(G3large): ΔE(G3large) = E[MP2(full)/G3large]-E[MP2/6-31G(2df,p)]- E[MP2/6(2.39) 31+G(d)]+E[MP2/6-31G(d)] 5. Atomic spin orbit correction is included, ΔE(SOa). 6. An empirical “high level correction” (E(HLC)) is added to the reference energy to recover any remaining correlation energy, such as correlation of valence electron pairs 32 and correlation of unpaired electrons in molecules. HLC also aims to minimize the difference between theory and experiment in the predicted enthalpies of formation and atomization energies. 7. A combination of the reference energy and all other contribution results in the G3 energy at 0 K, E0(G3): E0(G3) = E[MP4/6-31G(d)] + ΔE(QCI) + ΔE (+) + ΔE(2df,p) + (2.40) ΔE(G3large) +E(HLC) +ΔE(SOa) +E(ZPE) The goal of the G3 method is to achieve target accuracy within the QCISD(T)/G3Large level of theory. G3B3 method79 is similar to G3 in that both perform single point energy calculations at the same level and procedure but at different structure. The geometry optimization in G3 is at MP2(full)/6-31G(d) level of theory; in contrast, G3B3 uses the B3LYP/6-31G(d) level of theory to predict structures and to calculate zero point energy. G3B3 actually was developed for openshell systems where the unrestricted MP2 method, used in G3, suffers from spin contamination of the HF reference wavefunction.79 The inclusion of the empirical parameters within the HLC term introduces a bias in the Gn methods. 2.5.2 Correlation Consistent Composite Approach (ccCA) The ccCA method, developed by Wilson and coworkers,99–110 avoids the use of the empirical high-level correction (HLC) found in the Gn methods and utilizes the convergence behavior of the correlation consistent basis sets to eliminate the basis set incompleteness error. The ccCA method is designed to approach the [CCSD(T,FC1)-DK/aug-cc-pCV∞Z-DK] energy at a reduced computational cost, which in turn, can achieve the chemical accuracy of 1 kcal/mol 33 of reliable experiments for energetic properties of main group species, on average. The detailed computational steps in ccCA are as follows: 1. A geometry optimization and frequency calculation are performed at the B3LYP/ccpVTZ. The optimized structure is determined to be a minimum on the potential energy surface via a Hessian calculation. Harmonic vibrations are scaled by a factor of 0.989 in order to account for anharmonicity and are used to calculate the zero point energy,E(ZPE). The equilibrium structure predicted in this step is used for all the single point energy calculations that follow. 2. A single point energy calculation is carried out at the MP2/cc-pVnZ level, where n= D, T, and Q and the energies are extrapolated to the CBS limit to obtain the ccCA reference energy. The HF reference energies (resulted from this step) are extrapolated to the CBS limit using the Feller extrapolation scheme:113,114 ˜ = Where TU”t:„ + Ÿ exp −1.63˜ ˜ is the energy at the nth zeta-level of the used basis set, (2.41) TU”t:„ is the extrapolated HF energy to the CBS limit, and B is a fitting variable. The energy resulting from this extrapolation is the E(HF/CBS). The MP2 valence correlation energies are then extrapolated using one of the following extrapolation schemes. One option is the Peterson extrapolation scheme which is a mixed Gaussian and exponential formula:65 ˜ = t:„ + Ÿ exp‚− ˜ − 1 ƒ + £ exp‚− ˜ − 1 $ ƒ 34 (2.42) ˜ is the energy at the nth zeta-level of the used basis set, where t:„ is the energy at the CBS limit, and B and C are fitting parameters. The ccCA variant that utilizes the Peterson extrapolation scheme is referred to as ccCA-P. Other alternative extrapolation schemes are the Schwartz extrapolation schemes,66 which are based on the cubic or quartic inverse power of the highest order angular momentum (–“^] ) function included in the basis set: –“^] = t:„ + Ÿ –“^] + 1 2 ] (2.43) where at x = 3 ccCA is designated as ccCA-S3 and is designated as ccCA-S4 at x = 4. The energy resulting from this extrapolation is the E(MP2)/CBS. 3. A series of single point energy calculations are carried out at high level of theories and are used as additive terms that added to the ccCA reference energy. These calculations as follows: a. A contribution of a scalar relativistic effect is accounted for using the spin-free one-electron Douglas-Kroll Hamiltonian at the MP2 level with the corresponding relativistically re-contracted basis set, ΔE(DK): ΔE(DK) = E[MP2-DK/cc-pVTZ-DK] – E[MP2/cc-pVTZ] (2.44) b. A contribution of the core-core and the core-valence correlation effects is determined by employing an MP2 calculation that correlates the outer core as well as the valence in conjunction with a core-valence basis set, ΔE(CV): 35 ΔE(CV) = E[MP2(FC1)/aug-cc-pCVTZ] – E[MP2/aug-cc-pVTZ] (2.45) where FC1 indicates that outer core electrons are correlated in addition to the valence. c. A contribution of high levels of electron correlation that are not described by MP2 is determined using CCSD(T) with the cc-pVTZ basis set, ΔE(CC): ΔE(CC) = E[CCSD(T)/cc-pVTZ] – E[MP2/cc-pVTZ] (2.46) 4. Atomic spin orbit correction is included, ΔE(SOa). 5. Finally, a combination of the reference energy and all other contribution results in the ccCA energy at 0 K, E0(ccCA): E0(ccCA) = E(HF/CBS) + E(MP2/CBS) + ΔE(CC) + ΔE(CV) + ΔE(DK) + (2.47) ΔE(SOa) + E(ZPE) Various variants of ccCA have been developed including the ccCA-TM115 for first row transition metal, the relativistic-pseudopotential (rp)-ccCA108 for transition metals heavier than 3d elements, multireference (MR)-ccCA106 for molecules with multireference characters, and ONIOM-ccCA107 for large chemical system. Other implementations in ccCA include methods to decrease the computational cost such as the resolution-of-the-identity (RI)-ccCA105 and the ccCA-F12.109 All of which have been found to be efficient and accurate, overall, within the target accuracy. 36 REFERENCES 37 REFERENCES (1) Schrödinger, E. Quantisierung Als Eigenwertproblem. Ann. Phys. 1926, 386 (18), 109– 139. (2) Schrödinger, E. Quantisierung Als Eigenwertproblem. Ann. Phys. 1926, 384 (6), 489–527. (3) Schrödinger, E. Quantisierung Als Eigenwertproblem. Ann. Phys. 1926, 80 (13), 437–490. (4) Schrödinger, E. Über Das Verhältnis Der Heisenberg‐Born‐Jordanschen Quantenmechanik Zu Der Meinen. Ann. Phys. 1926, 384 (8), 734–756. (5) Schrödinger, E. An Undulatory Theory of the Mechanics of Atoms and Molecules. Phys. Rev. 1926, 28 (6), 1049–1070. (6) Kragh, H. Erwin Schrödinger and the Wave Equation: The Crucial Phase. Centaurus 1982, 26 (2), 154–197. (7) Levine, I. N. Quantum Chemistry, 7th ed.; Pearson Education Inc.: New York, U. S., 2014. (8) Szabo, A.; Ostlund, N. S. Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory, 1st ed.; Dover Publications, Inc.: Mineola, New York, 1996. (9) Born, M. Zur Quantenmechanik Der Stoßvorgänge. Z. Physik. 1926, 37 (12), 863–867. (10) Born, M.; Oppenheimer, R. Zur Quantentheorie Der Molekeln. Ann. Phys. 1927, 389 (20), 457–484. (11) Pauli, W. Über Den Zusammenhang Des Abschlusses Der Elektronengruppen Im Atom Mit Der Komplexstruktur Der Spektren. Z. Physik. 1925, 31 (1), 765–783. (12) Pauli, W. Exclusion Principle and Quantum Mechanics. In Writings on Physics and Philosophy; Springer Berlin Heidelberg: Berlin, Heidelberg, 1994; pp 165–181. (13) Slater, J. C. The Theory of Complex Spectra. Phys. Rev. 1929, 34 (10), 1293–1322. (14) Hartree, D. R. The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part III. Term Values and Intensities in Series in Optical Spectra. Math. Proc. Cambridge 38 Philos. Soc. 1928, 24 (3), 426–437. (15) Hartree, D. R. The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part II. Some Results and Discussion. Math. Proc. Cambridge Philos. Soc. 1928, 24 (1), 111– 132. (16) Hartree, D. R. The Wave Mechanics of an Atom with a Non-Coulomb Central Field. Part I. Theory and Methods. Math. Proc. Cambridge Philos. Soc. 1928, 24 (1), 89–110. (17) Fock, V. Näherungsmethode Zur Lösung Mehrkörperproblems. Z. Physik. 1930, 61 (1), 126–148. (18) Fock, V. "Selfconsistent Field" Mit Austausch Für Natrium. Z. Physik. 1930, 62 (11), 795–805. (19) Combs, L. L.; Holloman, M. Semiempirical Calculations of Internal Barriers to Rotation and Ring Puckering. J. Mol. Struct. 1976, 33 (2), 289–305. (20) Jensen, F. Introduction to Computational Chemistry, 2nd ed.; John Wiley & Sons Ltd.: Chichester, U.K., 2007. (21) Møller, C.; Plesset, M. S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46 (7), 618–622. (22) Jensen, F. Introduction to Computational Chemistry, 2nd Ed; John Wiley & Sons, Ltd: West Sussex, England, 2007. (23) Čížek, J. On the Correlation Problem in Atomic and Molecular Systems. Calculation of Wavefunction Components in Ursell‐Type Expansion Using Quantum‐Field Theoretical Methods. J. Chem. Phys. 1966, 45 (11), 4256–4266. (24) Purvis, G. D.; Bartlett, R. J. A Full Coupled‐cluster Singles and Doubles Model: The Inclusion of Disconnected Triples. J. Chem. Phys. 1982, 76 (4), 1910–1918. (25) Raghavachari, K.; Trucks, G. W.; Pople, J. A.; Head-Gordon, M. Reprint of: A FifthOrder Perturbation Comparison of Electron Correlation Theories. Chem. Phys. Lett. 2013, 589, 37–40. (26) Bartlett, R. J.; Musia, M. Coupled-Cluster Theory in Quantum Chemistry. Rev. Mod. Phys. 2007, 79 (1), 291–352. (27) Koch, W.; Holthausen, M. C. A Chemist’s Guide to Density Functional Theory, 2ed ed.; Wiley-VCH Verlag GmbH: Weinheim, Germany, 2002. 39 Des Quantenmechanischen (28) Sholl, D. S.; Steckel, J. A. Density Functional Theory : A Practical Introduction, 1st ed.; John Wiley & Sons, Inc.: New Jersey, U.S., 2009. (29) Tsuneda, T. Density Functional Theory in Quantum Chemistry; Springer Japan: Tokyo, Japan, 2014. (30) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136 (3), B864– B871. (31) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140 (4), A1133–A1138. (32) Politzer, P.; Abu-Awwad, F. A Comparative Analysis of Hartree-Fock and Kohn-Sham Orbital Energies. Theor. Chem. Acc. 1998, 99 (2), 83–87. (33) Becke, A. D. A New Mixing of Hartree–Fock and Local Density-Functional Theories. J. Chem. Phys. 1993, 98 (2), 1372–1377. (34) Lewars, E. G. Computational Chemistry: Introduction to the Theory and Applications of Molecular and Quantum Mechanics, 2nd ed.; Springer Science, Business Media B.V.: Dordrecht, Netherlands, 2011. (35) Vosko, S. H.; Wilk, L.; Nusair, M. Accurate Spin-Dependent Electron Liquid Correlation Energies for Local Spin Density Calculations: A Critical Analysis. Can. J. Phys. 1980, 58 (8), 1200–1211. (36) Becke, A. D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38 (6), 3098–3100. (37) Becke, A. D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A, Gen. Phys. 1988, 38 (6), 3098–3100. (38) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37 (2), 785–789. (39) Miehlich, B.; Savin, A.; Stoll, H.; Preuss, H. Results Obtained with the Correlation Energy Density Functionals of Becke and Lee, Yang and Parr. Chem. Phys. Lett. 1989, 157 (3), 200–206. (40) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. 40 (41) Perdew, J. P.; Burke, K.; Ernzerhof, M. Erratum: Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1997, 78 (7), 1396–1396. (42) Perdew, J. P. Density-Functional Approximation for the Correlation Energy of the Inhomogeneous Electron Gas. Phys. Rev. B 1986, 33 (12), 8822–8824. (43) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Climbing the Density Functional Ladder: Nonempirical Meta–Generalized Gradient Approximation Designed for Molecules and Solids. Phys. Rev. Lett. 2003, 91 (14), 146401–146401. (44) Zhao, Y.; Truhlar, D. G. A New Local Density Functional for Main-Group Thermochemistry, Transition Metal Bonding, Thermochemical Kinetics, and Noncovalent Interactions. J. Chem. Phys. 2006, 125 (19), 194101–194118. (45) Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98 (7), 5648. (46) Burke, K. Perspective on Density Functional Theory. J. Chem. Phys. 2012, 136 (15), 150901–150909. (47) Perdew, J. P.; Wang, Y. Accurate and Simple Analytic Representation of the Electron-Gas Correlation Energy. Phys. Rev. B 1992, 45 (23), 13244–13249. (48) Zhao, Y.; Truhlar, D. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120 (1), 215–241. (49) Hehre, W. J.; Stewart, R. F.; Pople, J. A. Self‐Consistent Molecular‐Orbital Methods. I. Use of Gaussian Expansions of Slater‐Type Atomic Orbitals. J. Chem. Phys. 1969, 51 (6), 2657–2664. (50) Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self-Consistent Molecular Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular Orbital Studies of Organic Molecules. J. Chem. Phys. 1972, 56 (5), 2257–2261. (51) Binkley, J. S.; Pople, J. A.; Hehre, W. J. Self-Consistent Molecular Orbital Methods. 21. Small Split-Valence Basis Sets for First-Row Elements. J. Am. Chem. Soc. 1980, 102 (3), 939–947. (52) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self‐consistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72 (1), 650–654. 41 (53) Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90 (2), 1007–1023. (54) Woon, D. E.; Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. III. The Atoms Aluminum through Argon. J. Chem. Phys. 1993, 98 (2), 1358–1371. (55) Woon, D. E.; Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. IV. Calculation of Static Electrical Response Properties. J. Chem. Phys. 1994, 100 (4), 2975–2988. (56) Woon, D. E.; Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. V. Core‐valence Basis Sets for Boron through Neon. J. Chem. Phys. 1995, 103 (11), 4572–4585. (57) Wilson, A. K.; Woon, D. E.; Peterson, K. A.; Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. IX. The Atoms Gallium through Krypton. J. Chem. Phys. 1999, 110 (16), 7667–7676. (58) Dunning, T. H.; Peterson, K. A.; Wilson, A. K. Gaussian Basis Sets for Use in Correlated Molecular Calculations. X. The Atoms Aluminum through Argon Revisited. J. Chem. Phys. 2001, 114 (21), 9244–9253. (59) Peterson, K. A.; Dunning, T. H. Accurate Correlation Consistent Basis Sets for Molecular Core–valence Correlation Effects: The Second Row Atoms Al–Ar, and the First Row Atoms B–Ne Revisited. J. Chem. Phys. 2002, 117 (23), 10548–10560. (60) Kohanoff, J. Electronic Structure Calculations for Solids and Molecules: Theory and Computational Methods; Cambridge University Press: Cambridge, UK, 2006. (61) Cook, D. B. Handbook of Computational Quantum Chemistry; Dover Publications Inc.: Mineola, U.S., 2005. (62) Boys, S. F. Electronic Wave Functions. I. A General Method of Calculation for the Stationary States of Any Molecular System. Proc. R. Soc. London 1950, A200 (1063), 542–554. (63) Boys, S. F. Electronic Wave Functions. II. A Calculation for the Ground State of the Beryllium Atom. Proc. R. Soc. London 1950, A201 (1064), 125–137. (64) O-ohata, K.; Taketa, H.; Huzinaga, S. Gaussian Expansions of Atomic Orbitals. J. Phys. Soc. Japan 1966, 21 (11), 2306–2313. 42 (65) Peterson, K. A.; Woon, D. E.; Dunning, T. H. Benchmark Calculations with Correlated Molecular Wave Functions. IV. The Classical Barrier Height of the H+H2→H2+H Reaction. J. Chem. Phys. 1994, 100 (10), 7410–7415. (66) Schwartz, C. Importance of Angular Correlations between Atomic Electrons. Phys. Rev. 1962, 126 (3), 1015–1019. (67) de Jong, W. A.; Harrison, R. J.; Dixon, D. A. Parallel Douglas–Kroll Energy and Gradients in NWChem: Estimating Scalar Relativistic Effects Using Douglas–Kroll Contracted Basis Sets. J. Chem. Phys. 2001, 114 (1), 48–53. (68) Bloch, F. Über Die Quantenmechanik Der Elektronen in Kristallgittern. Z. Physik. 1929, 52 (7), 555–600. (69) Kittel, C. Introduction to Solid State Physics, 8th ed.; John Wiley & Sons, Inc.: New Jersy, USA, 2005. (70) Zunger, A.; Cohen, M. L. First-Principles Nonlocal-Pseudopotential Approach in the Density-Functional Formalism: Development and Application to Atoms. Phys. Rev. B 1978, 18 (10), 5449–5472. (71) Vanderbilt, D. Soft Self-Consistent Pseudopotentials in a Generalized Eigenvalue Formalism. Phys. Rev. B 1990, 41 (11), 7892–7895. (72) Blöchl, P. E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50 (24), 17953– 17979. (73) Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector AugmentedWave Method. Phys. Rev. B 1999, 59 (3), 1758–1775. (74) Pople, J. A.; Head‐Gordon, M.; Fox, D. J.; Raghavachari, K.; Curtiss, L. A. Gaussian‐1 Theory: A General Procedure for Prediction of Molecular Energies. J. Chem. Phys. 1989, 90 (10), 5622–5629. (75) Curtiss, L. A.; Raghavachari, K.; Trucks, G. W.; Pople, J. A. Gaussian‐2 Theory for Molecular Energies of First‐ and Second‐row Compounds. J. Chem. Phys. 1991, 94 (11), 7221–7230. (76) Curtiss, L. A.; Carpenter, J. E.; Raghavachari, K.; Pople, J. A. Validity of Additivity Approximations Used in Gaussian‐2 Theory. J. Chem. Phys. 1992, 96 (12), 9030–9034. (77) Curtiss, L. A.; Raghavachari, K.; Pople, J. A. Gaussian‐2 Theory Using Reduced Møller– 43 Plesset Orders. J. Chem. Phys. 1993, 98 (2), 1293–1298. (78) Curtiss, L. A.; Raghavachari, K.; Redfern, P. C.; Rassolov, V.; Pople, J. A. Gaussian-3 (G3) Theory for Molecules Containing First and Second-Row Atoms. J. Chem. Phys. 1998, 109 (18), 7764–7776. (79) Baboul, A. G.; Curtiss, L. A.; Redfern, P. C.; Raghavachari, K. Gaussian-3 Theory Using Density Functional Geometries and Zero-Point Energies. J. Chem. Phys. 1999, 110 (16), 7650–7657. (80) Curtiss, L. A.; Redfern, P. C.; Raghavachari, K.; Rassolov, V.; Pople, J. A. Gaussian-3 Theory Using Reduced Møller-Plesset Order. J. Chem. Phys. 1999, 110 (10), 4703–4709. (81) Curtiss, L. A.; Raghavachari, K. Gaussian-3 and Related Methods for Accurate Thermochemistry. Theor. Chem. Acc. 2002, 108 (2), 61–70. (82) Curtiss, L. A.; Redfern, P. C.; Raghavachari, K. Gaussian-4 Theory. J. Chem. Phys. 2007, 126, 084108–084112. (83) Curtiss, L. A.; Redfern, P. C.; Raghavachari, K. Gaussian-4 Theory Using Reduced Order Perturbation Theory. J. Chem. Phys. 2007, 127 (12). (84) Martin, J. M. L.; de Oliveira, G. Towards Standard Methods for Benchmark Quality Ab Initio thermochemistry-W1 and W2 Theory. J. Chem. Phys. 1999, 111 (5), 1843–1856. (85) Martin, J. M. L. Heat of Atomization of Sulfur Trioxide, SO3: A Benchmark for Computational Thermochemistry. Chem. Phys. Lett. 1999, 310 (3–4), 271–276. (86) Parthiban, S.; Martin, J. M. L. Assessment of W1 and W2 Theories for the Computation of Electron Affinities, Ionization Potentials, Heats of Formation, and Proton Affinities. J. Chem. Phys. 2001, 114 (14), 6014–6029. (87) Boese, A. D.; Oren, M.; Atasoylu, O.; Martin, J. M. L.; Kállay, M.; Gauss, J. W3 Theory: Robust Computational Thermochemistry in the kJ/mol Accuracy Range. J. Chem. Phys. 2004, 120 (9), 4129–4141. (88) Karton, A.; Rabinovich, E.; Martin, J. M. L.; Ruscic, B. W4 Theory for Computational Thermochemistry: In Pursuit of Confident Sub-kJ/mol Predictions. J.Chem. Phys. 2006, 125, 144108–144117. (89) Karton, A.; Martin, J. M. L. Performance of W4 Theory for Spectroscopic Constants and Electrical Properties of Small Molecules. J. Chem. Phys. 2010, 133 (14), 144102–144117. 44 (90) Tajti, A.; Szalay, P. G.; Császár, A. G.; Kállay, M.; Gauss, J.; Valeev, E. F.; Flowers, B. A.; Vázquez, J.; Stanton, J. F. HEAT: High Accuracy Extrapolated Ab Initio Thermochemistry. J. Chem. Phys. 2004, 121 (23), 11599–11613. (91) Bomble, Y. J.; Vázquez, J.; Kállay, M.; Michauk, C.; Szalay, P. G.; Császár, A. G.; Gauss, J.; Stanton, J. F. High-Accuracy Extrapolated Ab Initio Thermochemistry. II. Minor Improvements to the Protocol and a Vital Simplification. J. Chem. Phys. 2006, 125 (6), 064108–064115. (92) Harding, M. E.; Vázquez, J.; Ruscic, B.; Wilson, A. K.; Gauss, J.; Stanton, J. F. HighAccuracy Extrapolated Ab Initio Thermochemistry. III. Additional Improvements and Overview. J. Chem. Phys. 2008, 128 (11), 114111–114115. (93) Petersson, G. A.; Bennett, A.; Tensfeldt, T. G.; Al‐Laham, M. A.; Shirley, W. A.; Mantzaris, J. A Complete Basis Set Model Chemistry. I. The Total Energies of Closed‐ shell Atoms and Hydrides of the First‐row Elements. J. Chem. Phys. 1988, 89 (4), 2193– 2218. (94) Montgomery, J. A.; Ochterski, J. W.; Petersson, G. A. A Complete Basis Set Model Chemistry. IV. An Improved Atomic Pair Natural Orbital Method. J. Chem. Phys. 1994, 101 (7), 5900–5909. (95) Ochterski, J. W.; Petersson, G. A.; Montgomery, J. A. A Complete Basis Set Model Chemistry. V. Extensions to Six or More Heavy Atoms. J. Chem. Phys. 1996, 104 (7), 2598–2619. (96) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. A Complete Basis Set Model Chemistry. VI. Use of Density Functional Geometries and Frequencies. J. Chem. Phys. 1999, 110 (6), 2822–2827. (97) Montgomery, J. A.; Frisch, M. J.; Ochterski, J. W.; Petersson, G. A. A Complete Basis Set Model Chemistry. VII. Use of the Minimum Population Localization Method. J. Chem. Phys. 2000, 112 (15), 6532–6542. (98) Wood, G. P. F.; Radom, L.; Petersson, G. A.; Barnes, E. C.; Frisch, M. J.; Montgomery, J. A. A Restricted-Open-Shell Complete-Basis-Set Model Chemistry. J. Chem. Phys. 2006, 125 (9), 094106–094122. (99) DeYonker, N. J.; Grimes, T.; Yockel, S.; Dinescu, A.; Mintz, B.; Cundari, T. R.; Wilson, A. K. The Correlation-Consistent Composite Approach: Application to the G3/99 Test Set. J. Chem. Phys. 2006, 125, 104111–104115. (100) DeYonker, N. J.; Cundari, T. R.; Wilson, A. K. The Correlation Consistent Composite Approach (ccCA): An Alternative to the Gaussian-N Methods. J. Chem. Phys. 2006, 124, 45 114104–114121. (101) DeYonker, N. J.; Cundari, T. R.; Wilson, A. K.; Sood, C. A.; Magers, D. H. Computation of Gas-Phase Enthalpies of Formation with Chemical Accuracy: The Curious Case of 3Nitroaniline. J. Mol. Struct. THEOCHEM 2006, 775 (1–3), 77–80. (102) DeYonker, N. J.; Mintz, B.; Cundari, T. R.; Wilson, A. K. Application of the Correlation Consistent Composite Approach (ccCA) to Third-Row (Ga−Kr) Molecules. J. Chem. Theory Comput. 2008, 4 (2), 328–334. (103) DeYonker, N. J.; Wilson, B. R.; Pierpont, A. W.; Cundari, T. R.; Wilson, A. K. Towards the Intrinsic Error of the Correlation Consistent Composite Approach (ccCA). Mol. Phys. 2009, 107 (8–12), 1107–1121. (104) Grimes, T. V; Wilson, A. K.; DeYonker, N. J.; Cundari, T. R. Performance of the Correlation Consistent Composite Approach for Transition States: A Comparison to G3B Theory. J. Chem. Phys. 2007, 127 (15), 154117–154125. (105) Prascher, B. P.; Lai, J. D.; Wilson, A. K. The Resolution of the Identity Approximation Applied to the Correlation Consistent Composite Approach. J. Chem. Phys. 2009, 131 (4), 044130–044142. (106) Oyedepo, G. A.; Wilson, A. K. Multireference Correlation Consistent Composite Approach [MR-ccCA]: Toward Accurate Prediction of the Energetics of Excited and Transition State Chemistry. J. Phys. Chem. A 2010, 114 (33), 8806–8816. (107) Das, S. R.; Williams, T. G.; Drummond, M. L.; Wilson, A. K. A QM/QM Multilayer Composite Methodology: The ONIOM Correlation Consistent Composite Approach (ONIOM-ccCA). J. Phys. Chem. A 2010, 114 (34), 9394–9397. (108) Laury, M. L.; DeYonker, N. J.; Jiang, W.; Wilson, A. K. A Pseudopotential-Based Composite Method: The Relativistic Pseudopotential Correlation Consistent Composite Approach for Molecules Containing 4d Transition Metals (Y–Cd). J. Chem. Phys. 2011, 135 (21), 214103–214113. (109) Mahler, A.; Wilson, A. K. Explicitly Correlated Methods within the ccCA Methodology. J. Chem. Theory Comput. 2013, 9 (3), 1402–1407. (110) Riojas, A. G.; Wilson, A. K. Solv-ccCA: Implicit Solvation and the Correlation Consistent Composite Approach for the Determination of pKa. J. Chem. Theory Comput. 2014, 10 (4), 1500–1510. (111) Henry, D. J.; Sullivan, M. B.; Radom, L. G3-RAD and G3X-RAD: Modified Gaussian-3 46 (G3) and Gaussian-3X (G3X) Procedures for Radical Thermochemistry. J. Chem. Phys. 2003, 118 (11), 4849–4860. (112) Pople, J. A.; Head‐Gordon, M.; Raghavachari, K. Quadratic Configuration Interaction. A General Technique for Determining Electron Correlation Energies. J. Chem. Phys. 1987, 87 (10), 5968–5975. (113) Feller, D. Application of Systematic Sequences of Wave Functions to the Water Dimer. J. Chem. Phys. 1992, 96 (8), 6104–6114. (114) Feller, D. The Use of Systematic Sequences of Wave Functions for Estimating the Complete Basis Set, Full Configuration Interaction Limit in Water. J. Chem. Phys. 1993, 98 (9), 7059–7071. (115) Jiang, W.; DeYonker, N. J.; Determan, J. J.; Wilson, A. K. Toward Accurate Theoretical Thermochemistry of First Row Transition Metal Complexes. J. Phys. Chem. A 2012, 116 (2), 870–885. 47 CHAPTER 3 DFT AND AB INITIO COMPOSITE METHODS: INVESTIGATION OF OXYGEN FLUORIDE SPECIES1 3.1 Introduction The oxygen fluorides have attracted interest because they can be employed as propellants in the rocket industry and can be used as strong fluorinating and oxidizing agents. In addition, oxygen fluorides play a role as intermediates in atmospheric chemistry and are believed to make a minor contribution to the destruction of ozone.1-6 The source of fluorine in the atmosphere originates from the decomposition of chlorofluorocarbons (CFCs) and their radical fragments, but most of the atmospheric fluorine is in the form of hydrogen fluoride (HF). Hydrogen fluoride is formed from the fast reaction of a fluorine atom with methane and water vapor.5,7 Although the role of fluorine in ozone depletion is minor, the percentage of fluorine in the atmosphere has been reported to be increasing with time.8-10 Thus, accurate thermochemical properties are required for modeling fluorine compounds in the study of atmospheric reactions. Due to the unstable nature of oxygen fluorides, experimental measurements of the energetic properties have been limited. Computational approaches can aid in understanding such systems. Investigating the structural properties of the oxygen fluorides has been a challenge to the computational chemical community, particularly for FOO and FOOF. The F-O bond in oxygen fluorides is a covalent bond between two highly electronegative atoms where both atoms contain lone pair electrons. Therefore, the F-O bond exhibits strong electron lone pair – lone pair repulsion and can become very long in molecules such as FOO and FOOF (~0.2 Å longer than the F-O bond in FOF),11,12 requiring consideration of high-level electron correlation methods. In FOO and FOOF, the O-O bond length is similar to that in the O2 molecule but ~0.2 Å shorter This entire chapter has reprinted from Z. H. A. Alsunaidi, and A. K. Wilson, “DFT and ab initio composite methods: Investigation of oxygen fluoride species” Computational and Theoretical Chemistry. 2016, 1095, 71-82, with permission of Elsevier. 1 48 than the O-O bond length in HOOH.11 The unusual geometry of FOO and FOOF presented a computational difficulty for electronic structure methods, which led to numerous investigations of oxygen fluorides using a variety of methods to study their structures and energetic properties.13-10 Many methods have been unsuccessful in predicting the right structure for oxygen fluorides, such as FOO, FOOO, and FOOF, with respect to the experimental geometries, as will be seen in the following sections. As a full literature review of these efforts is outside the scope of this paper, a number of significant and recent investigations are highlighted. The FOO structure and enthalpy of formation (ΔH°f, 298) have been computed by Francisco et al.16 using Møller-Plesset perturbation theory (MP2, MP3, and MP4), complete active space self-consistent field (CASSCF), and quadratic configuration interaction [QCISD(T)] in conjunction with Pople’s basis sets. The study found that all MPn methods underestimated the F-O bond length by > 0.2 Å. QCISD(T)/631G(d) yielded the best FO bond length that is only shorter by 0.002 Å from the experimental length (expt. re (F-O) = 1.649 ± 0.013 Å,31 where re indicates an equilibrium structure), whereas the best CASSCF description of the F-O bond length is 0.8 Å shorter than experiment. Francisco’s study reported an enthalpy of formation at 0 K for FOO of 8.9 ± 3 kcal mol-1 by using isodesmic and isogyric reaction schemes using QCISD(T)/6-311G(d,p) results.16 Ventura and Kieninger’s26 study on FOO concluded that B3LYP/6-311++G(3df, 3pd) is a reliable method to describe structures and predict reaction enthalpies for molecules involving F-O bonds. Studies by Denis30,32,33 found that the inclusion of the full treatment of the triple excitation [CCSDT instead of CCSD(T)] overcame the spin contamination problem presented in UCCSD(T), hence CCSDT predicted an accurate structure and energetics of the FOO molecule. Karton et al.20 reported the ΔH°f, 298 of FOO of 5.87 ± 0.16 kcal mol-1 in excellent agreement with experiment 49 (6.1 ± 0.5 kcal mol-1) using the high-level computationally demanding W4 method.34 A recent theoretical study by Feller et al.15 obtained a correct structure of FOO using R/UCCSD(T)/augcc-pVTZ level of theory and with a calculated value of ΔH°f, 298 of 6.4 ± 0.7 kcal mol-1 using a composite approach that is based on coupled cluster theory with up to quadruple excitations. The difference in the uncertainties estimated by Karton (5.87 ± 0.16 kcal mol-1)20 and Feller (6.4 ± 0.7 kcal mol-1)15 imputes to their different approaches. While W4 estimates uncertainties based on the performance of a set of 25 small molecules,20 Feller’s approach uses molecule-bymolecule criteria to calculate the estimated uncertainties.15 The structure and the ΔH°f, 298 of FOOF have been studied extensively with a broad variety of quantum chemical methods. The computational challenge in the FOOF structure arises from the anomeric delocalization effect that exists in FOOF between the oxygen lone pair and the antibonding orbital of the F-O bond.29 Although CCSD(T)/aug-cc-pVTZ15 and B3LYP/6311++G(2d)18 can provide a qualitatively correct geometry for FOOF, very few methods used in previous work reproduced the experimental FOOF structure. In fact, a local density functional (LDF) paired with numerical and Gaussian basis sets35 and the local SVWN functional paired with 6-311++G(2d)18 were two methods that predicted the closest re of the F-O bond compared to experiment, with LDF being superior. LDF predicted F-O and O-O bond lengths that are 0.01 Å and 0.001 Å off from experimental geometries (rs(F-O) =1.575 ± 0.003 Å and rs(O-O) = 1.217 ± 0.003 Å),11 respectively. LDF predicted this good description for FOOF likely due to the high and evenly distributed electron density in FOOF, as justified in Ref. 34. MP2, MP3, and MP4 with different size and type of basis sets, on the other hand, predicted incorrect geometries for FOOF with respect to experiment, however MP6 at the complete basis set (CBS) limit predicted an accurate geometry.29 Not only is the structure of FOOF problematic but its ΔH°f, 298 has also 50 been difficult to predict. The calculated ΔH°f, 298 of FOOF using even high-level ab initio methods has a large deviation from the experimental value reported in the NIST-JANAF thermochemical table (4.6 ± 0.5 kcal mol-1).36-37 The ΔH°f, 298’s reported by Karton et al.20 is 7.84 ± 0.18 kcal mol-1 and 8.21 ± 0.18 kcal mol-1 using W4 energies at the CCSD(T)/cc-pVQZ and experimental geometries, respectively. The most recent value of the ΔH°f, 298 of FOOF is 6.4 ± 0.7 kcal mol-1 and was calculated using a coupled cluster-base composite approach.15 These are only a few examples of this large deviation from experiment. Predicting the conformational structure of FOOO has also been challenging. Frecer et al.38 investigated the FOOO that formed by F + O3 reaction and found that FO(O)2 is the most stable structure with F-O being a weak bond (3.671 Å). However, FOOO was observed later as a stabilized intermediate in dilute mixtures of F2 and O3 in solid argon by FT-IR spectroscopy.39 Based on the reported frequencies of FOOO, it is characterized as a FO-O2 complex, and it cannot be a weak van der Waals complex,39 which differs from Frecer’s stable structure,38 mentioned above. Quantum chemical studies by Li et al.40 and Peiró-García et al.41 of the F + O3 reaction mechanisms using MP2/6-31G(d) and QCISD/6-311+G(d,p), respectively, also could not predict the FOOO ground state complex observed experimentally in the argon matrix.39 A geometry optimization and frequency calculations of FOOO were performed by Roohi et al.25 at the CCSD/aug-cc-pVDZ, CCSD/6-311+G(d), and QCISD/aug-cc-pVDZ levels of theory. Roohi’s study25 showed that the planar FOOO with dihedral angle of 0.0o is the most stable structure, with its calculated frequencies agreeing well with the reported experimental frequencies.39 No ΔH°f, 298 for FOOO has been previously reported, to our knowledge. The structure of the corresponding hydride FOOOH has only been studied previously by MINDO42 and by MP2/6-3lG(d).43 As MP2 theory has encountered difficulty for calculating the F-O and 51 the O-O bond lengths, an additional investigation of the FOOOH structure has to be done. While methods such as CCSDT and QCISD are computationally demanding, ab initio composite methods have been developed to circumvent the computational demands of such methods. One such approach, the correlation consistent Composite Approach (ccCA)44-46, is a method that has been demonstrated to be practical and reliable for the prediction of thermochemical properties, such as enthalpies of formation, ionization potentials, and electron affinities. The targeted accuracy of ccCA for main group molecules is to yield a mean absolute deviation of approximate chemical accuracy, 1 kcal mol-1 at reduced computational cost. This is in contrast to coupled cluster-based composite methods, which generally strive for a chemical accuracy of ± 0.24 kcal mol-1, such as the W4 composite method and the approach used by Feller et al.15 Because the performance of ccCA for a variety of halogen oxides and their related hydrides has not been examined in detail, it is of our interest to consider the utility of ccCA in describing oxygen fluorides, such as FOO, FOOF, and FOOO. Density functionals provide another option, as, overall, functionals have a lower formal computational scaling than post-HF methods such as CCSD(T) and CCSDT, though DFT predictions such as for enthalpies of formation, in general, do not reach the accuracies achievable by composite methods, such as ccCA. Thus, they are system dependent methods and are worth considering for each system. For example, the Minnesota density functionals M0647 and M062X47 were used by Meyer and Kass,48 in conjunction with the correlation consistent basis sets49-51 to assess their performance for predicting the ΔH°f, 298’s of a set of chlorine oxides and related hydrides (ClOx and ClOxH, where x =1-4) with respect to the ΔH°f, 298’s of W4 method.34 The capability of G352 and G3B353 for calculating ΔH°f, 298’s of chlorine oxides were also investigated in the same study.48 The main findings from the Meyer and Kass study are that M06 52 ΔH°f, 298’s were found to differ by an average of 1.3 kcal mol-1 from the W4 ΔH°f, 298’s, while M06-2X resulted in larger error (6.2 kcal mol-1). G3 and G3B3 ΔH°f, 298’s yielded an average error of 4.6 kcal mol-1 and 6.5 kcal mol-1, respectively. The author attributed the large errors of G3 and G3B3 ΔH°f, 298’s to their poor predicted geometries. Although M06 and M06-2X functionals were examined for the prediction of the ΔH°f, 298’s of chlorine oxides, the capability of M06 and M06-2X to predict structures and ΔH°f, 298’s has not been assessed for other halogen oxides. Thus, it is of interest to evaluate these functionals for oxygen fluoride species, as well as examine the performance of G3 and G3B3 methods for these systems. G3 and G3B3 are used here, largely, as they were included in the Meyer and Kass study.48 Though G4 is a more modern method, beginning with an MP4 reference energy is a costly start, and as shown in previous studies, G4 predicts very similar energies as G3.54-55 In the present study, the reliability of ccCA, G3, and G3B3 for the prediction of the ΔH°f, 298’s of oxygen fluoride species was evaluated. In addition to these composite methods, the performance of M06 and M06-2X was also examined for predicting the structures and enthalpies of formation of oxygen fluoride species. A set of various oxygen fluorides were considered in this study, including FO, FOO, FOOO, and the related hydrides (FOH, FOOH, and FOOOH) and difluorides (FOF, FOOF, and FOOOF), where the ΔH°f, 298’s of FOOO and FOOOH have not been reported previously. The effects of basis set size and spin contamination were also considered. For comparison, the ΔH°f, 298’s of chlorine oxides and related hydrides have been provided, whereas full theoretical investigations for chlorine oxides and related hydrides can be found in the Meyer and Kass study and references therein.48 53 3.2 Computational Methodology All calculations were performed using the Gaussian 09 software package.56 The hybridmeta-generalized gradient approximation (HMGGA) Minnesota functionals (M0647 and M062X)47 in conjunction with the augmented correlation consistent polarized valence basis sets (augcc-pVnZ), where n = D, T, Q,49-51 were used to optimize the structures of all molecules under investigation. The tight-d correlation consistent basis set of Dunning et al.,57 aug-cc-pV(n+d)Z, where n = D, T, Q, were used for chlorine. Frequency calculations were performed to ensure that the structure is a stationary point. The enthalpies of formation for these structures were determined at the same level of theory. ccCA, G3, and G3B3, were also applied to predict the enthalpies of formation. In ccCA, geometry optimization and vibrational frequencies calculations are performed using B3LYP/augcc-pVTZ.44 The harmonic vibrational frequencies are then corrected using a scale factor of 0.989 as recommended in Ref. 44. The remaining steps in ccCA involve a series of single point energy calculations performed using the B3LYP/aug-cc-pVTZ geometry (for details see Chapter2 Section 2.5.2). Several variants of ccCA can be used, which vary by the means used to extrapolate the MP2 energy (MP2/aug-cc-pV∞Z) to the CBS limit.44 ccCA-S3 and ccCA-S4 utilize Schwartz’s inverse cubic and quartic extrapolation scheme, respectively.58 ccCA-P utilizes Peterson’s mixed Gaussian/exponential extrapolation scheme59 and ccCA-PS3 is an average of ccCA-P and ccCA-S3. G352 and G3B353 both are composite methods that involve the same series of single point energy calculations though based upon different geometries (see Chapter 2 Section 2.5.1). The geometry optimization in G3 is at the MP2(full)/6-31G(d) level of theory, while G3B3 uses the B3LYP/6-31G(d) level of theory. 54 For open-shell systems (radicals) such as FOO and FOOO, which show a degree of spin contamination, restricted open shell (RO) calculations are used, i.e. ROM06, ROM06-2X. In addition, RO-ccCA60 and G3-RAD61 energies were obtained and used to calculate their ΔH°f, 298’s. Gaussian 0956 was used to determine these energies. The mean absolute deviations (MADs) of the calculated ΔH°f, 298’s were determined for all of the utilized methods with respect to the experimental values, unless otherwise noted. 3.3 Results and Discussion 3.3.1 Structures The structural parameters obtained by M06 and M06-2X for all of the species are listed in Table 3.1. The optimized structures at the B3LYP/aug-cc-pVTZ, MP2(full)/6-31G(d), and B3LYP/6-31G(d) levels, which are used for geometry optimizations in ccCA, G3, and G3B3, respectively, were also considered (shown in Table 3.1). Experimental structural parameters of FO,22 FOO,31 FOF,12 FOOF,11 and FOH,17 have been reported and were utilized as reference data to determine the performance of the considered methods. To the authors’ knowledge no experimental observations have been reported for the geometric parameters of FOOH, FOOO, and FOOOF. Thus, for these molecules theoretical results from rigorous methods such as coupled cluster are presented to calibrate the considered methods (CCSD(T)/TZ2P structure for FOOH,21 CCSD/6-311+G(d) structure for FOOO,25 and CCSD(T)/cc-pVTZ for FOOOF).19 55 Table 3.1 Structural parameters of the oxygen fluoride species at different level of theories, bond lengths are in angstroms and bond angles and dihedral angles in degree. FO Method/Basis set M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) MP2(full)/6-31g(d) Experimenta r (FO) 1.332 1.328 1.324 1.329 1.329 1.325 1.351 1.354 1.344 1.354 FOH Method/Basis set M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) MP2(full)/6-31g(d) Experimentb r (FO) 1.408 1.405 1.401 1.401 1.400 1.397 1.430 1.434 1.444 1.4350 ± 0.0031 r (OH) 0.973 0.969 0.966 0.971 0.968 0.967 0.971 0.977 0.979 0.9657 ± 0.0016 FOF Method/Basis set M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ r (FO) 1.380 1.377 1.374 1.376 1.374 a (FOF) 103.5 103.6 103.7 102.9 103.0 56 a (FOH) 99.1 99.2 99.3 99.3 99.4 99.5 98.6 97.8 97.1 97.54 ± 0.50 Table 3.1 Continued. M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) MP2(full)/6-31g(d) Experimentc 1.371 1.403 1.409 1.423 1.412 103.1 103.9 103.9 102.6 103.1 FOO Method/Basis set M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ ROM06/aug-cc-pVDZ ROM06/aug-cc-pVTZ ROM06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ ROM062X/aug-cc-pVDZ ROM062X/aug-cc-pVTZ ROM062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) MP2(full)/6-31g(d) Experimentd Method/Basis set M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) r (FO) 1.811 1.752 1.747 1.589 1.580 1.572 2.090 1.520 1.519 1.508 1.500 1.500 1.618 1.571 1.649±0.013 r (FO) 1.447 1.438 1.434 1.419 1.415 1.412 1.470 1.465 r (OO) 1.343 1.346 1.343 1.358 1.361 1.357 1.365 1.376 r (OO) 1.177 1.171 1.170 1.182 1.178 1.178 1.184 1.190 1.187 1.194 1.194 1.191 1.188 1.211 1.200±0.013 FOOH r (OH) 0.977 0.972 0.970 0.973 0.970 0.968 0.974 0.981 57 a (FOO) 110.7 110.6 110.7 110.6 110.6 110.7 111.2 110.0 110.1 109.9 110.0 110.1 111.2 111.0 111.2±0.36 a (FOO) 105.9 105.9 106.0 105.3 105.4 105.5 106.3 106.1 a (OOH) 104.0 103.9 104.0 103.6 103.6 103.7 103.7 102.9 d(FOOH) 85.5 85.2 85.4 84.6 84.6 84.7 85.0 83.1 Table 3.1 Continued. MP2(full)/6-31g(d) CCSD(T)/TP2Ze 1.468 1.481 1.39 1.393 0.981 0.969 105.0 105.4 102.0 101.9 83.1 84.5 FOOF Method/Basis set M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) MP2(full)/6-31g(d) Experimentf Method/Basis set M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ ROM06/aug-cc-pVDZ ROM06/aug-cc-pVTZ ROM06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ ROM062X/aug-cc-pVDZ ROM062X/aug-cc-pVTZ ROM062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) ROB3LYP/aug-cc-pVTZ ROB3LYP/6-31g(d) r (FO) 1.506 1.494 1.485 1.426 1.420 1.419 1.523 1.497 1.496 1.575±0.003 r (FO) 1.338 1.335 1.331 1.357 1.354 1.350 1.331 1.333 1.327 1.360 1.359 1.356 1.354 1.360 1.379 1.380 r (OO) 1.217 1.219 1.221 1.285 1.289 1.286 1.227 1.266 1.291 1.217±0.003 FOOO r (FO-O), r (FOO-O) 2.469, 1.196 2.408, 1.191 2.449, 1.189 1.704, 1.192 1.687, 1.189 1.681, 1.188 2.598, 1.191 2.537, 1.187 2.702, 1.187 1.589, 1.204 1.575, 1.205 1.573, 1.203 2.715, 1.204 2.504, 1.211 1.709, 1.201 1.710, 1.214 58 a (FOO) 108.6 108.5 108.6 106.6 106.7 106.8 109.3 108.3 106.9 109.5±0.5 a (FOO), a (OOO) 97.8, 110.8 93.0, 106.5 92.4, 105.9 100.2, 110.1 100.9, 110.5 101.1, 110.7 88.5, 103.4 90.4, 105.6 91.0, 107.8 103.3, 112.1 103.9, 112.4 104.0, 112.5 101.1, 113.8 91.0, 103.3 101.2, 111.2 100.3, 110.2 d (FOOF) 87.2 86.9 87.0 85.9 85.7 85.8 88.1 86.7 85.8 87±0.5 d (FOOO) 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 Table 3.1 Continued. MP2(full)/6-31g(d) CCSD/6-311+G(d)g 1.378 1.745, 1.200 100.5, 111.5 0.0 FOOOH Method/Basis set r (FO) r (FO-O), r (O-OH) r (OH) M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) MP2(full)/6-31g(d) 1.483 1.471 1.466 1.427 1.423 1.421 1.502 1.485 1.481 1.290, 1.427 1.292, 1.428 1.291, 1.424 1.331, 1.406 1.333, 1.407 1.330, 1.403 1.303, 1.461 1.331, 1.454 1.358, 1.450 0.974 0.970 0.968 0.972 0.967 0.968 0.972 0.979 0.981 a (FOO), a (OOO), a (OOH) 106.5, 109.0, 101.1 106.5, 109.0, 101.1 106.6, 109.1, 101.2 105.3, 107.9, 101.8 105.5, 108.1, 102,1 105.6, 108.2, 102.2 107.3, 109.6, 100.7 106.3, 108.8, 99.9 104.9, 107.5, 99.7 d (FOOO) d (OOOH) -85.8 -85.7 -85.6 -84.0 -84.0 -83.9 -87.4 -84.9 -83.2 91.7 91.4 91.8 91.4 90.1 90.3 96.0 89.0 87.0 FOOOF Method/Basis set r (FO) r (OO) a (FOO), a (OOO) d (FOOO) M06/aug-cc-pVDZ 1.419 1.355 106.2, 109.0 ± 93.0 M06/aug-cc-pVTZ 1.413 1.355 106.2, 109.1 ±92.5 M06/aug-cc-pVQZ 1.410 1.353 106.3, 109.2 ±92.4 M062X/aug-cc-pVDZ 1.403 1.360 105.3, 108.1 ±90.9 M062X/aug-cc-pVTZ 1.400 1.360 105.7, 108.3 ±90.8 M062X/aug-cc-pVQZ 1.399 1.357 105.7, 108.5 ±91.0 B3LYP/aug-cc-pVTZ 1.440 1.377 106.6, 109.3 ±93.6 B3LYP/6-31g(d) 1.440 1.386 105.8, 108.7 ±90.5 MP2(full)/6-31g(d) 1.365 1.340 105.6, 108.2 ±90.1 CCSD(T)/cc-pVTZh 1.444 1.385 105.5, 108.3 ±91.3 a b c d e f g Reference 22. Reference 17. Reference 12. Reference 31. CCSD(T)/TZ2P: Reference 21. Reference 11. CCSD/6-311+G(d): Reference25. h CCSD(T)/cc-pVTZ: Reference 19. 59 The F-O bond length calculated by M06 and M06-2X for all set of molecules showed a systematic decrease in length when using the correlation consistent basis set family going from aug-cc-pVDZ, aug-cc-pVTZ, to aug-cc-pVQZ (Table 3.1). However, this decrease in the bond length while increasing basis set size does not occur for the O-O bond length as shown in Table 3.1, and can be dependent on the density functional being utilized. Both M06 and M06-2X underestimated the F-O and O-O bond lengths with respect to the corresponding experimental values. M06-2X predicted shorter F-O bond lengths and longer FO-O bond lengths than M06. The average difference between the aug-cc-pVTZ F-O bond lengths and the aug-cc-pVQZ F-O bond lengths is 0.005 Å for M06 and it is 0.002 Å for M06-2X, and for the FO-O bond length it is 0.003 Å for both functionals whereas the aug-cc-pVDZ bond lengths are generally longer. Yet, the M06/aug-cc-pVDZ and M06-2X/aug-cc-pVDZ predicted the closest geometries to the reference data with average errors in the F-O bond length of 0.032 Å and 0.062 Å, respectively. Thus, increasing the basis set size will not always give the better structures for the systems under investigation. For open-shell species, such as FO, FOO, and FOOO, the degree of spin contamination resulting from the mixing of higher spin states into the wavefunction was calculated, as high spin contamination results in incorrect geometries and energies. Previous studies showed a high degree of spin contamination when studying FOO using other methods, such as MP2.16,26 No spin contamination was found when calculating FO, but it was present when calculating FOO and FOOO. The expectation values of the total spin, , are listed in Table 3.2, where the optimal value for these radicals is = 0.75. The effect of spin contamination becomes appreciable as the deviation of from 0.75 increases, and this deviation decreases with increasing the size of the basis sets, as shown in Table 3.2. Both UM06 and UM06-2X suffer from spin contamination associated with FOO (small) and FOOO (severe), 60 as shown in Table 3.2. This error was corrected by using the restricted open-shell density functionals as shown in Table 3.1. ROM06 and ROM06-2X provided better structural parameters with respect to the reference data for FOO and FOOO than what UM06 and UM062X predicted. Thus, the ROM06 and ROM06-2X geometries were used later to calculate the enthalpies of formation. FOOOF was found to exist in two conformers. Both have very similar geometries but differ in the dihedral angles. One conformer has d(OOOF) = 91.3° and the other one has d(OOOF) = 82.0°.19 Because the energy difference between the two conformers at the CCSD(T)/cc-pVTZ level is very small, 0.24 kcal mol-1, with the d(OOOF) = 82.0° conformer having the lowest energy,19 only one conformer was included in the molecule test set. M06-2X described the structure of FOOOF better than M06 when compared to the CCSD(T) structure. Overall the geometries predicted by M06 are in better agreement with the reference data than M06-2X, although both functionals generally underestimated the F-O and the O-O bond lengths. Thus, for this set of molecules, doubling the amount of Hartree-Fock exchange from M06 to M06-2X does not improve the results. As shown in Table 3.1, the F-O bond lengths obtained by B3LYP/aug-cc-pVTZ are in very good agreement (MAD of 0.01 Å) with respect to the reference data, with the exception of the F-O bond distance(s) in FOO and FOOF. The difference in F-O bond lengths is more than 0.01 Å (0.018 Å in FOO and 0.049 Å in FOOF) compared to experimental values. Although B3LYP/aug-cc-pVTZ did not predict very accurate bond lengths for FOO and FOOF, it provided a qualitatively correct description of the structures of these two molecules. The capability of B3LYP to describe the geometries has been noticed previously for some of the oxygen fluorides.26,35 Quite similarly, the calculated F-O bond lengths at the B3LYP/6-31G(d) level were in very good agreement with respect to the reference data; however, not only is the F-O bond 61 length underestimated in FOO and in FOOF (by 0.065 Å in FOO and 0.075 Å in FOOF), but also the F-O bond distance in FOOH was underestimated by about 0.016 Å using B3LYP/6-31G(d) in Table 3.2 The expectation value of the total spin . Molecule Method/Basis set FOO M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) 0.756 0.754 0.754 0.782 0.750 0.75 0.75 0.7509 FOOO M06/aug-cc-pVDZ M06/aug-cc-pVTZ M06/aug-cc-pVQZ M062X/aug-cc-pVDZ M062X/aug-cc-pVTZ M062X/aug-cc-pVQZ B3LYP/aug-cc-pVTZ B3LYP/6-31g(d) 0.892 0.782 0.782 0.803 0.780 0.807 0.790 0.776 comparison to the reference data. Additionally, when using the 6-31G(d) basis set, the average error for B3LYP increased from 0.01 Å (obtained when using an aug-cc-pVTZ basis set) to 0.02 Å. Thus, again B3LYP with the small basis set 6-31G(d) is not enough to describe peroxide systems, such as FOO, FOOF, and FOOH. B3LYP in conjunction with aug-cc-pVTZ also resulted in the least average error of 0.01 Å (from the reference data) for the O-O bond length compared to other methods. The effect of spin contamination for open shell systems using UB3LYP has been tested and the total spin operators are presented in Table 3.2, for each level of theory. The values for FOO show that the use of UB3LYP at either basis set levels 62 does not result in spin contamination, whereas for FOOO, both UB3LYP/aug-cc-pVTZ and UB3LYP/6-31G(d) levels resulted in large deviation from the optimal value = 0.75. Consequently, the ROB3LYP is used to predict the geometry for FOOO, which is in agreement with the structure predicted by CCSD,25 as shown in Table 3.1. For FOOOF, B3LYP was able to predict the two conformers with d(OOOF) = 91.3° and with d(OOOF) = 82.0° and the B3LYP FO and O-O bond lengths agree well with the CCSD(T) bond lengths. The predicted geometries for FOF, FOH, and FOOH by MP2(full)/6-31g(d) agree well with the reference data within a 0.01 Å difference. However, MP2(full)/6-31g(d) predicted a 0.1 Å shorter F-O bond and a 0.1 Å longer O-O bond for FOOF compared to experimental bond lengths, this large difference in the bond lengths was also found previously using MP2/631g**.43 Likewise, the F-O and O-O bond lengths predicted by MP2(full)/6-31g(d) for FOOOF are 0.08 Å and 0.05 Å shorter than CCSD(T) results, indicating that this level of theory is not enough to describe the peroxide’s geometry. For the open-shell systems FOO and FOOO, the UMP2 method could not provide converged geometries due to large spin contamination. As mentioned earlier for FOOOH, no reliable theoretical or experimental geometries have been reported. Based on the success of B3LYP/aug-cc-pVTZ method in predicting the FOOOF geometry compared to the CCSD(T)/cc-pVTZ, the B3LYP/aug-cc-pVTZ geometry for FOOOH is considered as the most reliable structure. For this compound, only one conformer d(FOOO) = -87.4°) is found to be a stable structure. The B3LYP/6-31G(d) structure of FOOOH is quite similar to the structure at the B3LYP/aug-cc-pVTZ level, with a large difference of 0.03 Å in the FO-O bond length. M06 and M06-2X generally underestimated the bond lengths of 63 FO FOO FOOO FOF FOH FOOF FOOH FOOOF FOOOH Figure 3.1 B3LYP/aug-cc-pVTZ structures of the oxygen fluoride species included in this study. 64 FOOOH as compared to the B3LYP/aug-cc-pVTZ geometries. MP2 (full)/6-31g(d) predicted shorter F-O bond length and longer FO-O bond length than the B3LYP/aug-cc-pVTZ. As a result of the above discussion, M06 and M06-2X are not recommended methods for predicting the geometries of oxygen fluorides and related hydrides and difluorides. MP2 performed well for most of the closed-shell compounds with the exception of the peroxides. The geometries obtained by B3LYP/aug-cc-pVTZ, shown in Figure 3.1, result in the lowest deviation from the reference data among the considered methods. As shown in Figure 3.1, using B3LYP/aug-cc-pVTZ supports conclusions from previous studies12,17,31 that FOO, FOF, and FOH are bent with bond angles of 111.2°, 103.9°, and 98.6°, respectively. Different bond angles indicate that the bond angle opens more as the repulsion between bonds increases. Similarly, FOOF and FOOH have dihedral angles of 88.2° and 85.0°, respectively, in order to minimize the repulsion between bonds. FOOOF and FOOOH display a zigzag shape with d(FOOO) = 93.6°, and d(FOOO) = -87.4°, respectively. In contrast to all the peroxide systems, the stable conformer of FOOO is when d(FOOO) = 0.0°. Intermolecular dispersion forces might be the cause of the stable FOOO structure. 3.3.2 Enthalpies of Formation (ΔH°f, 298) To provide a comparison between the performance of the utilized methods on chlorine oxide and oxygen fluoride species, the ΔH°f, 298’s of chlorine oxides and related hydrides were determined using M06/aug-cc-pVQZ, M06-2X/aug-cc-pVQZ, ccCA-S3, G3, and G3B3 and are listed in Table 3.3 The MADs of the considered methods with respect to experiment for predicting ΔH°f, 298’s are provided in Table 3.3 as well. These MADs are 1.1 (ccCA-S3), 2.1 (M06), 2.5 (M06-2X), 2.3 (G3), and 3.5 (G3B3) kcal mol-1. The MAD of ccCA-S3 indicates that 65 ccCA is a reliable method in predicting energetics for chlorine oxides. M06 resulted in a MAD of 2.1 kcal mol-1 for the calculated ΔH°f, 298’s of chlorine oxides, which is 0.9 kcal mol-1 greater than the MAD (1.2 kcal mol-1) that resulted from the calculations done by Meyer and Kass.48 This small difference arises from the use of a temperature correction approach that reduces the energy contributed from the low vibrational frequency modes as pointed out in Meyer and Kass study.48,62 In the present study, however, scale factors of 0.9853 (M06) and 0.9733 (M06-2X)63 were used for the correction of the vibrational frequencies in the computations. The ΔH°f, 298’s calculated by M06 for chlorine oxides in the present study are in good agreement with the experimental values with the exception of ClO3, which is known to be problematic for not only computational methods but also for experiments, as demonstrated by the error bar associated with the experimental ΔH°f, 298 (± 3 kcal mol-1),64-66 as compared with smaller uncertainties for many main group species. Only ccCA-S3 and G3 predict the ΔH°f, 298 of ClO3 within the experimental uncertainty. The ΔH°f, 298’s obtained by G3 are in relatively good agreement with the experiments (MAD = 2.3 kcal mol-1), yet the ClO2 ΔH°f, 298 predicted by G3 overestimated the experimental value by ~4 kcal mol-1. Similar to G3, M06-2X achieved a MAD of 2.5 kcal mol-1, while the MAD for G3B3 was larger (MAD of 3.2 kcal mol-1). Therefore, ccCA results in the lowest MAD with respect to experiment for the prediction of energetic properties of chlorine oxides, followed by M06. 66 Table 3.3 Enthalpies of formation for chlorine oxides and related hydrides. ΔH°f, 298 K (kcal mol-1) Compd. ccCA-S3 M06a M06-2Xa G3 G3B3 Expt. ClO 25.3 23.1 23.5 25.9 26.7 24.29 ±0.03b ClO2 23.9 21.2 26.8 26.8 27.6 22.6 ± 0.3b,c ClO3 44.0 40.5 50.2 48.3 51.0 46 ± 3d ClO4 57.7 52.6 65.4 66.0 65.5 - -19.5 -18.0 -19.3 -17.4 -16.9 -18.4 ± 0.03e HOCl 3.5 4.6 6.5 6.4 7.3 HOClO -2.8 -3.9 3.3 2.0 3.8 HOClO2 -0.8 -3.4 5.5 5.6 8.7 HOClO3 1.1 2.1 2.5 2.3 3.5 MAD a M06 and M06-2X in conjunction with aug-cc-pV(Q+d)Z for chlorine and aug-cc-pVQZ for oxygen and hydrogen. bReference 70. cReference 71. dReferences 64-66. eReference 72. The ΔH°f, 298’s for all of the oxygen fluoride species included in this study were calculated using M06 and M06-2X in conjunction with aug-cc-pVDZ, aug-cc-pVTZ, and aug-ccpVQZ at 298 K using the atomization energy approach and the results are shown in Table 3.4. A systematic decrease in the ΔH°f, 298 values as the size of the basis set increases is observed, since the basis set with larger zeta (ξ) level recovers more energy, as shown in Table 3.4. The differences between energies determined using M06/aug-cc-pVTZ ΔH°f, 298’s and those determined using M06/aug-cc-pVQZ ΔH°f, 298’s is 0.1 - 1.1 kcal mol-1 with an average difference of 0.5 kcal mol-1, whereas the differences between energies determined using M06-2X/aug-ccpVTZ ΔH°f, 298’s and those determined using M06-2X/aug-cc-pVQZ ΔH°f, 298’s is 0.3 – 1.9 kcal mol-1 with an average difference of 1.2 kcal mol-1. To evaluate the reliability of the utilized methods in calculating ΔH°f, 298’s, the following reference data was used: experimental ΔH°f, 298 values for FO, FOO, FOF, FOOF and FOH; the CCSD(T)/ANO4 ΔH°f, 298 value for FOOH; and 67 the extrapolated CCSD(T)/aug-cc-pV(T,Q)Z ΔH°f, 298 value for FOOOF. For FOOO and FOOOH neither experimental nor theoretical ΔH°f, 298 values are available. The MADs of the calculated ΔH°f, 298’s with respect to the reference data were computed. Because of the wellknown challenges of FOOF,13,20,21,28 the MAD was also calculated without FOOF. For ccCA, the ccCA-S3 variant was selected as it results in the lowest MAD for the ΔH°f, 298’s of oxygen fluoride species with respect to the reference data as shown in Table 3.5 as compared with the other ccCA variants. Table 3.4 Enthalpies of formation for the oxygen fluoride species using M06 and M06-2X paired with the correlation consistent basis sets. ΔH°f, 298 K (kcal mol-1) Compd. M06 Reference dataa M06-2X aDZ aTZ aQZ aDZ aTZ aQZ FO 29.6 28.5 28.1 29.1 28.3 27.5 26.1±2.4 FOH -16.9 -17.3 -17.7 -16.4 -18.3 -18.6 -23.16±1.2 FOF 12.9 9.7 9.6 11.5 9.6 8.7 5.9±0.5 FOO 10.6 9.5 9.0 18.3 16.0 14.5 6.1±0.5 FOOH -6.0 -6.9 -7.6 -5.6 -7.6 -8.4 -10.4±1.0b FOOF 14.1 11.3 11.2 18.6 16.9 15.5 4.6±0.5 FOOO 40.3 37.8 36.7 43.4 43.2 41.3 - FOOOH 4.9 3.5 2.6 6.2 4.1 2.9 - FOOOF 35.3 31.5 30.9 34.4 32.7 30.9 26.6c MAD 6.3 4.4 4.0 7.7 6.0 5.2 5.7 4.0 3.5 6.7 5.0 4.3 MAD w/o FOOF aDZ: aug-cc-pVDZ aTZ: aug-cc-pVTZ aQZ: aug-cc-pVQZ a NIST-JANAF Tables: References 36-37. b CCSD(T)/ANO4: Reference 21. c Extrapolated CCSD(T)/augcc-pV(T,Q)Z: Reference 19. 68 Table 3.5 Enthalpies of formation for oxygen fluoride species using the different variants of ccCA method. ΔH°f, 298 K (kcal mol-1) Compd. ccCA-P ccCA-S3 ccCA-PS3 ccCA-S4 Reference dataa FO 27.3 27.0 27.2 27.3 26.1±2.4 FOH -21.1 -21.6 -21.4 -21.1 -23.16±1.2 FOF 6.6 6.2 6.4 6.7 5.9±0.5 FOO 8.1 7.5 7.8 8.1 6.1±0.5 FOOH -11.0 -11.7 -11.3 -10.9 -10.4±1.0b FOOF 9.4 8.7 9.0 9.4 4.6±0.5 FOOO 31.8d 31.0d 31.4d 31.8d - FOOOH -0.7 -1.6 -1.1 -0.6 - FOOOF 27.8 26.8 27.3 27.8 26.6c MAD 1.8 1.4 1.6 1.8 - MAD w/o FOOF 1.3 0.9 1.1 1.3 - a NIST-JANAF Tables: References 36-37. b CCSD(T)/ANO4: Reference 21. c Extrapolated CCSD(T)/augcc-pV(T,Q)Z: Reference 19. dUsing RO-ccCA The calculated ΔH°f, 298’s of FO by ccCA, M06, M06-2X, G3, and G3B3 are within the reported experimental uncertainty (± 2.4 kcal mol-1), as shown in Table 3.6. The ΔH°f, 298’s of FO calculated by G3 and G3B3 are the nearest to the reported experimental value, while the ΔH°f, 298 calculated by M06 deviated the most by 2.0 kcal mol-1, but is still within the experimental uncertainty. The FO ΔH°f, 298’s predicted by ccCA is also in very good agreement with experiment. For FOH, the ΔH°f, 298’s calculated by ccCA is found to be the closest to the reported experimental value, while G3 and G3B3 provide ΔH°f, 298’s that are 1.6 and 1.9 kcal mol-1, respectively, less than the experimental uncertainty. M06 and M06-2X underestimate the ΔH°f, 298’s of FOH by 5.46 and 4.56 kcal mol-1, respectively, with respect to the experimental value. 69 The ccCA ΔH°f, 298 of FOF is within the experimental error bar. G3 and G3B3 predict ΔH°f, 298’s of 6.5 and 6.8 kcal mol-1, which are greater than the experimental uncertainty by 0.1 and 0.4 kcal mol-1, respectively. M06 predicts a ΔH°f, 298 of FOF that is 3.2 kcal mol-1 outside of the experimental uncertainty, while the ΔH°f, 298 calculated by M06-2X is 2.3 kcal mol-1 outside of the experimental uncertainty. Table 3.6 Calculated enthalpies of formation for the oxygen fluoride species using all methods and the MADs of these methods with respect to the reference data. ΔH°f, 298 K (kcal mol-1) Compd. ccCA-S3 M06a M06-2Xa G3 G3B3 Reference datab FO 27.0 28.1 27.5 26.1 26.5 26.1±2.4 FOH -21.6 -17.7 -18.6 -20.4 -20.1 -23.16±1.2 FOF 6.2 9.6 8.7 6.5 6.8 5.9±0.5 FOO 7.5e 9.0f 14.5f 7.1g 7.0 6.1±0.5 FOOH -11.7 -7.6 -8.4 -10.3 -10.2 -10.4±1.0c FOOF 8.7 11.2 15.5 9.3 8.9 4.6±0.5 FOOO 31.0e 36.7f 41.3f 30.1g 30.1g - FOOOH -1.6 2.6 2.9 0.2 -0.3 - FOOOF 26.8 30.9 30.9 27.8 27.3 26.6d MAD 1.4 4.0 5.2 1.5 1.5 - MAD w/o FOOF 0.9 3.5 4.3 0.9 0.9 - a M06 and M06-2X in conjunction with aug-cc-pVQZ. b NIST-JANAF Tables: Reference 36-37. c CCSD(T)/ANO4: Reference 21. d Extrapolated CCSD(T)/aug-cc-pV(T,Q)Z: Reference 19. e Using ROccCA. f Using ROM06 and ROM06-2X. g Using G3-RAD. For FOO, ccCA and G3B3 give ΔH°f, 298’s that are outside the experimental error by 0.9 and 0.4 kcal mol-1, respectively. These results demonstrate the utility of ccCA and G3B3 in predicting the FOO ΔH°f, 298. G3-RAD, one of the G3 versions developed for open-shell systems, is used instead of G3 to calculate the ΔH°f, 298’s of the radicals FOO and FOOO. The G3-RAD 70 ΔH°f, 298 of FOO is also in good agreement with experiment. The calculated ΔH°f, 298 for FOO using ROM06 is above experimental value by 2.9 kcal mol-1, whereas the ΔH°f, 298 calculated by ROM06-2X is 10.6 kcal mol-1 greater than experiment. This large deviation can be explained by the poor geometry obtained using these ROM06 and ROM06-2X. For FOOH, the CCSD(T)/ANO4 ΔH°f, 298 which is -10.4 ± 1.0 kcal mol-1,21 is being used as a reference value for this molecule. G3 and G3B3 predict ΔH°f, 298’s of FOOH of -10.3 and 10.2 kcal mol-1, respectively, which are in excellent agreement with the CCSD(T) value. Conversely, the ccCA ΔH°f, 298 value of FOOH is -11.7 kcal mol-1, which is more negative than the CCSD(T) value by 1.3 kcal mol-1. Based on the fact that oxygen fluoride molecules are considered highly correlated molecules, accounting for the core correlation correction in the method is highly important. Thus, ccCA ΔH°f, 298’s can be more accurate than the CCSD(T) energies because ccCA energies includes core-valence and core-core correction terms, whereas the reported CCSD(T) results used the frozen-core approximation.21 G3B3 includes the high level correction term (HLC) in the energy. The HLC is calculated based on empirical parameters and it is added to the G3 energy to reduce the error between theory and experiment. Without the HLC the G3B3 ΔH°f, 298 deviates by ~7.0 kcal mol-1 from the CCSD(T) value. M06 and M06-2X both underestimate the ΔH°f, 298 of FOOH by 2.6 and 1.0 kcal mol-1, respectively. For FOOF, the ccCA ΔH°f, 298 value lies 3.6 kcal mol-1 outside the error bar of the value reported by NISTJANAF.36,37 G3 and G3B3 predict ΔH°f, 298’s that are greater than experimental error bar by 4.2 and 3.8 kcal mol-1, respectively. M06 and M06-2X give ΔH°f, 298’s larger than experiment by 6.6 and 10.9 kcal mol-1, respectively. Several previous high-level theoretical studies have pointed out the discrepancy between experimental and calculated ΔH°f, 298’s, and suggested that the experiment to be revisited.15,20,21,28 For example, high-level methods predicted ΔH°f, 298 of FOOF 71 of 9.6 ± 0.9 kcal mol-1 (iCAS-CI+Q),15 8.7 ± 2.0 kcal mol-1 (CCSD(T)/ANO4),21 7.84 ± 0.18 kcal mol-1 (W4),20 and 7.3 kcal mol-1 (B3PW91/aug-cc-pVQZ).28 In addition, from this work the ccCA value of the ΔH°f, 298 of FOOF is 8.7 kcal mol-1. Thus, the ΔH°f, 298’s provided by theoretical approaches is between 7-9 kcal mol-1, whereas the experimental value is 4.6 ± 0.5 kcal mol-1. Because of this large and consistent discrepancy, the MAD of the ΔH°f, 298’s is calculated with and without FOOF. RO-ccCA, G3-RAD, ROM06, and ROM06-2X were employed to calculate the ΔH°f, 298 for FOOO and the predicted ΔH°f, 298’s are listed in Table 3.6. Due to the lack of reference data for FOOO, the ΔH°f, 298 of ccCA for FOOO (31.0 kcal mol-1) is considered the most accurate based on the previous successes of ccCA in predicting energetic properties (MAD of 1.01 kcal mol-1 using ccCA-PS3 for main group molecules (G03/05 test set)).44 G3-RAD also predicts a very close value to the ccCA value (30.1 kcal mol-1). However, including empirical parameters makes G3-RAD a system-dependent method. Thus, the RO-ccCA value is recommended. The computed ΔH°f, 298 for FOOO using ROM06 is 5.7 kcal mol-1 greater than the RO-ccCA value. M06-2X predicts very high ΔH°f, 298’s for FOOO compared to the other methods. For FOOOH, the calculated ΔH°f, 298’s of ccCA is considered the reference data for the same reason previously mentioned. The ΔH°f, 298 of ccCA for FOOOH (-1.6 kcal mol-1) is chemically sensible since formation of hydride is usually exothermic with fluoride lowering its stability. In addition, G3B3 predicts as exothermic ΔH°f, 298 as ccCA, but greater by 1.3 kcal mol-1. G3 and M06 and M06-2X overestimated the ΔH°f, 298 of FOOOH by 1.8, 4.2, and 4.5 kcal mol-1, respectively, with respect to the ccCA value. Finally, ccCA successfully predicts the ΔH°f, 298 of FOOOF with only a 0.2 kcal mol-1 deviation from the extrapolated CCSD(T)/aug-cc-pV(T,Q)Z value calculated previously by Huang et al.,19 where other methods such as G2 and G96PW91/D95(3df) predicted 72 ΔH°f, 298’s that are > 7.0 kcal mol-1 higher than the CCSD(T) value.67 Here G3B3 ΔH°f, 298 is also in good agreement with the CCSD(T) value, with a deviation of only 0.7 kcal mol-1 between the two, but G3 ΔH°f, 298 differs by 1.2 kcal mol-1. This difference between G3B3 and G3 results typically comes from geometry optimization, with MP2 found to be insufficient to describe the structure of dioxygen fluoride species. M06 and M06-2X again overestimated the enthalpy of formation of FOOOF by 4.3 kcal mol-1 with respect to the CCSD(T) value. The MADs of the calculated ΔH°f, 298’s for all the utilized methods with respect to the reference data were computed and provided in Table 3.6. The MADs of M06 and M06-2X as shown in Table 3.6 are 4.0 and 5.2 kcal mol-1 (3.5 and 4.3 kcal mol-1 without FOOF), respectively. Thus, compared to the chlorine oxides in Table 3.3, M06 and M06-2X do not perform well for oxygen fluoride species. That can be attributed to the poor geometries obtained by M06 and M06-2X. Because of the deficiencies in the M06 and M06-2X computed geometries, single point M06 and M06-2X energy calculations were performed using the B3LYP/aug-cc-pVTZ geometries to examine the performances of M06 and M06-2X for oxygen fluorides energy calculations. ΔH°f, 298 and MAD’s were calculated and listed in Table 3.7. However, no improvement is noticed in the M06 and M06-2X ΔH°f, 298’s when using B3LYP geometries. That supports our conclusion that M06 and M06-2X are not considered reliable methods to predict structures and energetic properties for oxygen fluorides. G3 and G3B3 (including the G3-RAD values) perform well with MAD’s of 1.5 kcal mol-1 for both methods (0.9 kcal mol-1 for both methods without FOOF). This performance is expected taking into account the inclusion of the HLC term in the G3 and G3B3 energy. For example, the G3B3 ΔH°f, 298 of FO without the HLC term is 30.3 kcal mol-1, a value that is greater by 3.8 kcal mol-1 than the G3B3 ΔH°f, 298 listed in Table 3.6. The MAD of the G3 and G3B3 methods (1.5 kcal mol-1) is 73 much smaller for the oxygen fluorides than for the chlorine oxides (2.3 kcal mol-1 for G3 and 3.5 kcal mol-1 for G3B3), so G3 and G3B3 perform better for oxygen fluorides than for chlorine oxides. The MAD of ccCA-S3 is 1.4 kcal mol-1 (0.9 kcal mol-1 without FOOF) for oxygen fluorides and is 1.1 kcal mol-1 for chlorine oxides. This indicates a capability of ccCA to predict reliable energetic properties for halogen oxides. Table 3.7 Calculated enthalpies of formation for the oxygen fluoride species using M06 and M06-2X methods based on B3LYP/aug-cc-pVTZ geometries. ΔH°f, 298 K (kcal mol-1) Compd. B3LYPa//M06b Reference datac FO 28.4 27.7 26.1±2.4 FOH -17.5 -18.5 -23.16±1.2 FOF 10.2 9.3 5.9±0.5 FOO 9.3 15.0 6.1±0.5 FOOH -6.9 -7.7 -10.4±1.0d FOOF 11.8 19.9 4.6±0.5 FOOO 34.7 47.5 - FOOOH 4.9 4.5 - FOOOF 31.8 32.1 26.6e MAD 4.5 6.0 - 4.0 4.4 - MAD w/o FOOF a B3LYPa//M06-2Xb b B3LYP/aug-cc-pVTZ. M06/aug-cc-pVQZ and M06-2X/aug-cc-pVQZ NIST-JANAF Tables: Reference 36-37. d CCSD(T)/ANO4: Reference 21. e Extrapolated CCSD(T)/aug-cc-pV(T,Q)Z: Reference 19. c In addition, a comparison between the calculated ΔH°f, 298’s in this study and the recent Active Thermochemical Tables (ATcT)68,69 values of the ΔH°f, 298’s of FO, FOH, FOF, FOO, FOOF, and FOOOF are examined. The ATcT tables use artificial intelligence algorithms to reduce the uncertainties in the experimentally measured ΔH°f, 298’s by combining experimental 74 and highly accurate theoretical thermochemical data.68,69 As shown in Table 3.8, the uncertainties of the ATcT values of FO, FOH, FOF, and FOO decreased in comparison to the NIST-JANAF values (Table 3.6). Thus, to evaluate our methods against the ATcT values, the MADs of the calculated ΔH°f, 298’s for all the utilized methods with respect to the ATcT values were computed and provided in Table 3.8. The MADs of ccCA-S3, M06, M06-2X, G3, and G3B3 were all lowered by 0.5-0.9 kcal mol-1 in comparison with the MADs shown in Table 3.6 (with ccCA-S3 providing the lowest MAD, at 0.6 kcal mol-1). The large difference in the MADs between NIST-JANAF table and ATcT is likely attributed to the large difference between the ATcT value of the ΔH°f, 298 of FOOF (8.04±0.09 kcal mol-1) and the NIST-JANAF value (4.6 ± 0.5 kcal mol-1). Table 3.8 Calculated Enthalpies of formation for the oxygen fluoride species using all methods and the MADs of these methods with respect to the ATcT values. ΔH°f, 298 K (kcal mol-1) Compd. ccCA-S3 M06a M06-2Xa G3 G3B3 ATcTb FO 27.0 28.1 27.5 26.1 26.5 26.51±0.04 FOH -21.6 -17.7 -18.6 -20.4 -20.1 -20.85±0.05 FOF 6.2 9.6 8.7 6.5 6.8 5.91±0.06 FOO 7.5e 9.0f 14.5f 7.1g 7.0 5.99±0.06 FOOH -11.7 -7.6 -8.4 -10.3 -10.2 - FOOF 8.7 11.2 15.5 9.3 8.9 8.04±0.09 FOOO 31.0c 36.7d 41.3d 30.1e 30.1e - FOOOH -1.6 2.6 2.9 0.2 -0.3 - FOOOF 26.8 30.9 30.9 27.8 27.3 26.63±1.86 MAD 0.6 3.1 4.7 0.8 0.7 a d M06 and M06-2X in conjunction with aug-cc-pVQZ. bATcT: Reference 68-69. c Using RO-ccCA. Using ROM06 and ROM06-2X. e Using G3-RAD. 75 3.1 Conclusion The capability of ccCA, G3, and G3B3 for the prediction of the enthalpies of formation of oxygen fluoride species was evaluated. In addition, the performance of M06 and M06-2X in conjunction with the correlation consistent basis sets (aug-cc-pVnZ), where n = D, T, Q, was also examined for predicting the structures and enthalpies of formation of oxygen fluoride species. An important finding from this study is that though M06 and M06-2X are useful functionals for many main group species (including chlorine oxides), M06 and M06-2X were less successful in the prediction of reasonable structures and ΔH°f, 298’s for oxygen fluorides, oxygen difluorides, and related hydrides. This could be generalized to systems containing F-O and/or O-O bonds (e.g. peroxides and polyoxides). Geometries predicted by B3LYP/aug-ccpVTZ are generally in good agreement with the listed reference data. When calculating the enthalpies of formation, ccCA-S3 provides the lowest MAD (1.4 kcal mol-1 with respect to the reference data; 0.9 kcal mol-1 excluding FOOF) without any parameterized energies. While ccCA-S3 provides the smallest MAD of the four different CBS extrapolation formulas considered for this set of molecules, other ccCA variants, such as ccCA-PS3, have been found to be useful in previous studies.44,45 G3 and G3B3 achieved a MAD that is greater than ccCA-S3 by only 0.1 kcal mol-1 with respect to the reference data while incorporating an empirical parameter that is intended to reduce the overall MAD of G3 and G3B3 (or MAD of 0.9 kcal mol-1 excluding FOOF, which is identical to ccCA-S3 in this case). The performance of G3 and G3B3 for chlorine oxides is not as good as for oxygen fluorides. ccCA-S3 predictions of the ΔH°f, 298’s of both the oxygen fluorides and of the chloride oxides species are in good agreement with the experimental values. In addition, when comparing the calculated ΔH°f, 298’s to the ATcT values, ccCA-S3 provides the lowest MAD (0.6 kcal mol-1) in comparison to the other methods included 76 in this study. The enthalpies of formation for FOOO and FOOOH are predicted to be 31.0 and -1.6 kcal mol-1, respectively, by the ccCA-S3 method. Overall, the use of the correlation consistent Composite Approach (ccCA) is recommended for such systems, with promise for other halogen systems and peroxides. 77 REFERENCES 78 REFERENCES (1) Prather, M. J.; McElroy, M. B.; Wofsy, S. C. Reductions in Ozone at High Concentrations of Stratospheric Halogens. Nature. 1984, pp 227–231. (2) Prather, M. J.; Watson, R. T. Stratospheric Ozone Depletion and Future Levels of Atmospheric Chlorine and Bromine. Nature. 1990, pp 729–734. (3) Sander, S. P.; Friedl, R. R.; Yung, Y. L. Rate of Formation of the ClO Dimer in the Polar Stratosphere: Implications for Ozone Loss. Science. 1989, pp 1095–1098. (4) Solomon, S. Progress towards a Quantitative Understanding of Antarctic Ozone Depletion. Nature. 1990, pp 347–354. (5) Ravishankara, A. R.; Turnipseed, A. A.; Jensen, N. R.; Barone, S.; Mills, M.; Howard, C. J.; Solomon, S. Do Hydrofluorocarbons Destroy Stratospheric Ozone? Science. 1994, pp 71–75. (6) Parrish, A.; De Zafra, R. L.; Solomon, P. M.; Barrett, J. W.; Carlson, E. R. Chlorine Oxide in the Stratospheric Ozone Layer: Ground-Based Detection and Measurement. Science. 1981, pp 1158–1161. (7) Colussi, A. J.; Grela, M. A. Rate of the Reaction between Oxygen Monofluoride and Ozone: Implications for the Atmospheric Role of Fluorine. Chem. Phys. Lett. 1994, 229 (1–2), 134–138. (8) Zander, R.; Rinsland, C. P.; Mahieu, E.; Gunson, M. R.; Farmer, C. B.; Abrams, M. C.; Ko, M. K. W. Increase of Carbonyl Fluoride (COF2) in the Stratosphere and Its Contribution to the 1992 Budget of Inorganic Fluorine in the Upper Stratosphere. J. Geophys. Res. Atmos. 1994, 99 (D8), 16737–16743. (9) Nassar, R.; Bernath, P. F.; Boone, C. D.; McLeod, S. D.; Skelton, R.; Walker, K. A.; Rinsland, C. P.; Duchatelet, P. A Global Inventory of Stratospheric Fluorine in 2004 Based on Atmospheric Chemistry Experiment Fourier Transform Spectrometer (ACEFTS) Measurements. J. Geophys. Res. Atmos. 2006, 111 (D22), D22313. (10) Brown, A. T.; Chipperfield, M. P.; Richards, N. A. D.; Boone, C.; Bernath, P. F. Global Stratospheric Fluorine Inventory for 2004–2009 from Atmospheric Chemistry Experiment Fourier Transform Spectrometer (ACE-FTS) Measurements and SLIMCAT Model Simulations. Atmos. Chem. Phys. 2014, 14 (1), 267–282. 79 (11) Jackson, R. H. The Microwave Spectrum, Structure, and Dipole Moment of Dioxygen Difluoride. J. Chem. Soc. 1962, 4585–4592. (12) Pierce, L.; Di Cianni, N.; Jackson, R. H. Centrifugal Distortion Effects in Asymmetric Rotor Molecules. I. Quadratic Potential Constants and Average Structure of Oxygen Difluoride from the Ground-State Rotational Spectrum. J. Chem. Phys. 1963, 38 (3), 730– 739. (13) Feller, D.; Dixon, D. A. Coupled Cluster Theory and Multireference Configuration Interaction Study of FO, F2O, FO2, and FOOF. J. Phys. Chem. A 2003, 107 (45), 9641– 9651. (14) Alcami, M.; Mó, O.; Yáñez, M.; Cooper, I. L. The Performance of Density-Functional Theory in Challenging Cases: Halogen Oxides. J. Chem. Phys. 2000, 112 (14), 6131– 6140. (15) Feller, D.; Peterson, K. A.; Dixon, D. A. Refined Theoretical Estimates of the Atomization Energies and Molecular Structures of Selected Small Oxygen Fluorides. J. Phys. Chem. A 2010, 114 (1), 613–623. (16) Francisco, J. S.; Zhao, Y.; Lester, W. A.; Williams, I. H. Theoretical Studies of the Structure and Thermochemistry of FO2 Radical: Comparison of Møller–Plesset Perturbation, Complete‐active‐space Self‐consistent‐field, and Quadratic Configuration Interaction Methods. J. Chem. Phys. 1992, 96 (4), 2861–2867. (17) Halonen, L.; Ha, T. K. Equilibrium Structure and Anharmonic Force Field of Hypofluorous Acid (HOF). J. Chem. Phys. 1988, 89 (8), 4885–4888. (18) Jursic, B. S. The Density Functional Theory Investigation of the Equilibrium Structures of OOF, FOOF, OOF2, and FOOOF. J. Mol. Struct. THEOCHEM 1996, 366 (1–2), 97–101. (19) Huang, M.-J.; Watts, J. D. Theoretical Characterization of the F2O3 Molecule by Coupled Cluster Methods. J. Phys. Chem. A 2010, 114, 10197. (20) Karton, A.; Parthiban, S.; Martin, J. M. L. Post-CCSD(T) Ab Initio Thermochemistry of Halogen Oxides and Related Hydrides XOX, XOOX, HOX, XOn, and HXOn (X = F, Cl), and Evaluation of DFT Methods for These Systems. J. Phys. Chem. A 2009, 113 (16), 4802–4816. (21) Lee, T. J.; Rice, J. E.; Dateo, C. E. The Varying Nature of Fluorine Oxygen Bonds. Mol. Phys. 1996, 89 (5), 1359–1372. 80 (22) Miller, C. E.; Drouin, B. J. The X1 2Π3/2 and X2 2Π1/2 Potential Energy Surfaces of FO. J. Mol. Spectrosc. 2001, 205 (2), 312–318. (23) Minyaev, R. M.; Gribanova, T. N. Structure and Stability of Halogen Oxoacids XOnH (X = F, Cl; n = 1-4): A Quantum Chemistry Study. Russ. J. Inorg. Chem. 2004, 49, 579–586. (24) O’Hare, P. A. G.; Wahl, A. C. Oxygen Monofluoride (OF, 2Π): Hartree–Fock Wavefunction, Binding Energy, Ionization Potential, Electron Affinity, Dipole and Quadrupole Moments, and Spectroscopic Constants. A Comparison of Theoretical and Experimental Results. J. Chem. Phys. 1970, 53 (6), 2469–2478. (25) Roohi, H.; Mackiabadi, B. Conformations of O3–F 1:1 Complexes. An Ab Initio Study. Bull. Chem. Soc. Jpn. 2007, 80 (10), 1914–1919. (26) Ventura, O. N.; Kieninger, M. The FO2 Radical: A New Success of Density Functional Theory. Chem. Phys. Lett. 1995, 245 (4–5), 488–497. (27) Zhao, Y.; Francisco, J. S. Ab Initio Studies of the Structure and Thermochemistry of FO Radicals. Chem. Phys. Lett. 1990, 167 (4), 285–290. (28) Kieninger, M.; Segovia, M.; Ventura, O. N. A Discrepancy between Experimental and Theoretical Thermochemical Characterization of Some Oxygen Fluorides. Chem. Phys. Lett. 1998, 287 (5–6), 597–600. (29) Kraka, E.; He, Y.; Cremer, D. Quantum Chemical Descriptions of FOOF: The Unsolved Problem of Predicting Its Equilibrium Geometry. J. Phys. Chem. A 2001, 105 (13), 3269– 3276. (30) Denis, P. A.; Ventura, O. N. CCSDT Study of the Fluoroperoxyl Radical, FOO. Chem. Phys. Lett. 2004, 385 (3–4), 292–297. (31) Yamada, C.; Hirota, E. The Infrared Diode Laser Spectrum of the ν2 Band of the FO2 Radical. J. Chem. Phys. 1984, 80 (10), 4694–4700. (32) Denis, P. A.; Ventura, O. N. Corrigendum to “CCSDT Study of the Fluoroperoxyl Radical, FOO” [Chem. Phys. Lett. 385 (2004) 292–297]. Chem. Phys. Lett. 2004, 395 (4– 6), 385–386. (33) Denis, P. A. On the Performance of CCSD(T) and CCSDT in the Study of Molecules with Multiconfigurational Character: Halogen Oxides, HSO, BN and O3. Chem. Phys. Lett. 2004, 395 (1–3), 12–20. (34) Karton, A.; Rabinovich, E.; Martin, J. M. L.; Ruscic, B. W4 Theory for Computational 81 Thermochemistry: In Pursuit of Confident Sub-kJ/mol Predictions. J. Chem. Phys. 2006, 125, 144108–144117. (35) Dixon, D. A.; Andzelm, J.; Fitzgerald, G.; Wimmer, E. Density Functional Study of a Highly Correlated Molecule, Oxygen Fluoride (FOOF). J. Phys. Chem. 1991, 95 (23), 9197–9202. (36) Chase, M. W.; Davies, C. A.; Downey, J. R.; Frurip, D. J.; McDonald, R. A.; Syverud, A. N. JANAF Thermochemical Tables, 3rd Ed. J. Phys. Chem. Ref. Data 1985, 14. (37) Chase, M. W. NIST‐JANAF Thermochemical Tables, 4th Ed. J. Phys. Chem. Ref. Data 1998, 2. (38) Frecer, V.; Jain, D. C.; Sapse, A.-M. Ab Initio Calculations for FO3, FO3+ , and FO3Complexes Formed by Fluorine with Ozone. Struct. Chem. 1998, 9 (1), 9–13. (39) Misochko, E. Y.; Akimov, A. V; Wight, C. A. Infrared Spectroscopic Observation of the Stabilized Intermediate Complex FO3 Formed by Reaction of Mobile Fluorine Atoms with Ozone Molecules Trapped in an Argon Matrix. J. Phys. Chem. A 1999, 103 (40), 7972– 7977. (40) Li, L.-C.; Wang, J.; Wang, X.; Tian, A.-M.; Wong, N.-B. Quantum Chemical Study of the Reaction Mechanism of Ozone and Methane with Fluorine and Chlorine Atoms. Int. J. Quantum Chem. 2002, 87 (5), 288–292. (41) Peiró-Garcı́a, J.; Nebot-Gil, I. An Ab Initio Study on the Mechanism of the F+O3→FO+O2 Reaction: Comparative Reactivity Study along the Isoelectronic NH2, OH and F Radicals Series. Chem. Phys. Lett. 2004, 391 (1–3), 195–199. (42) Glidewell, C. Structure and Conformation in Molecular Peroxides. J. Mol. Struct. 1980, 67, 35–44. (43) Gimarc, B. M.; Zhao, M. Oxygen Ring Strain Energies Revisited: Effects of Terminal Atoms of the Chain Reference Structure. J. Phys. Chem. 1994, 98 (6), 1596–1600. (44) DeYonker, N. J.; Wilson, B. R.; Pierpont, A. W.; Cundari, T. R.; Wilson, A. K. Towards the Intrinsic Error of the Correlation Consistent Composite Approach (ccCA). Mol. Phys. 2009, 107 (8–12), 1107–1121. (45) DeYonker, N. J.; Grimes, T.; Yockel, S.; Dinescu, A.; Mintz, B.; Cundari, T. R.; Wilson, A. K. The Correlation-Consistent Composite Approach: Application to the G3/99 Test Set. J. Chem. Phys. 2006, 125, 104111–104115. 82 (46) DeYonker, N. J.; Cundari, T. R.; Wilson, A. K. The Correlation Consistent Composite Approach (ccCA): An Alternative to the Gaussian-n Methods. J. Chem. Phys. 2006, 124, 114104–114121. (47) Zhao, Y.; Truhlar, D. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120 (1), 215–241. (48) Meyer, M. M.; Kass, S. R. Experimental and Theoretical Gas-Phase Acidities, Bond Dissociation Energies, and Heats of Formation of HClOx, X = 1-4. J. Phys. Chem. A 2010, 114 (12), 4086–4092. (49) Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90 (2), 1007–1023. (50) Woon, D. E.; Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. III. The Atoms Aluminum through Argon. J. Chem. Phys. 1993, 98 (2), 1358–1371. (51) Dunning, T. H.; Peterson, K. A.; Wilson, A. K. Gaussian Basis Sets for Use in Correlated Molecular Calculations. X. The Atoms Aluminum through Argon Revisited. J. Chem. Phys. 2001, 114 (21), 9244–9253. (52) Curtiss, L. A.; Raghavachari, K.; Redfern, P. C.; Rassolov, V.; Pople, J. A. Gaussian-3 (G3) Theory for Molecules Containing First and Second-Row Atoms. J. Chem. Phys. 1998, 109 (18), 7764–7776. (53) Baboul, A. G.; Curtiss, L. A.; Redfern, P. C.; Raghavachari, K. Gaussian-3 Theory Using Density Functional Geometries and Zero-Point Energies. J. Chem. Phys. 1999, 110 (16), 7650–7657. (54) Jorgensen, K. R.; Wilson, A. K. Enthalpies of Formation for Organosulfur Compounds: Atomization Energy and Hypohomodesmotic Reaction Schemes via Ab Initio Composite Methods. Comput. Theor. Chem. 2012, 991, 1–12. (55) Wilson, B. R.; DeYonker, N. J.; Wilson, A. K. Prediction of Hydrocarbon Enthalpies of Formation by Various Thermochemical Schemes. J. Comput. Chem. 2012, 33 (25), 2032– 2042. (56) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian 09, 2009. 83 (57) Peterson, K. A.; Dunning, T. H. Accurate Correlation Consistent Basis Sets for Molecular Core–valence Correlation Effects: The Second Row Atoms Al–Ar, and the First Row Atoms B–Ne Revisited. J. Chem. Phys. 2002, 117 (23), 10548–10560. (58) Schwartz, C. Importance of Angular Correlations between Atomic Electrons. Phys. Rev. 1962, 126 (3), 1015–1019. (59) Peterson, K. A.; Woon, D. E.; Dunning, T. H. Benchmark Calculations with Correlated Molecular Wave Functions. IV. The Classical Barrier Height of the H+H2→H2+H Reaction. J. Chem. Phys. 1994, 100 (10), 7410–7415. (60) Williams, T. G.; DeYonker, N. J.; Ho, B. S.; Wilson, A. K. The Correlation Consistent Composite Approach: The Spin Contamination Effect on an MP2-Based Composite Methodology. Chem. Phys. Lett. 2011, 504 (1–3), 88–94. (61) Henry, D. J.; Sullivan, M. B.; Radom, L. G3-RAD and G3X-RAD: Modified Gaussian-3 (G3) and Gaussian-3X (G3X) Procedures for Radical Thermochemistry. J. Chem. Phys. 2003, 118 (11), 4849–4860. (62) Meyer, M. M. Probing the Structure and Reactivity of Gaseous Ions. Ph.D. Dissertation, University of Minnesota, Minneapolis, MN, 2010. (63) Laury, M. L.; Carlson, M. J.; Wilson, A. K. Vibrational Frequency Scale Factors for Density Functional Theory and the Polarization Consistent Basis Sets. J. Comput. Chem. 2012, 33 (30), 2380–2387. (64) Colussi, A. J.; Grela, M. A. Kinetics and Thermochemistry of Chlorine- and NitrogenContaining Oxides and Peroxides. J. Phys. Chem. 1993, 97 (15), 3775–3779. (65) Rathmann, T.; Schindler, R. N. Ab Initio Calculations on the Geometries and Thermodynamic Stabilities of Chlorine Trioxides. Chem. Phys. Lett. 1992, 190 (6), 539– 542. (66) Rauk, A.; Tschuikow-Roux, E.; Chen, Y.; McGrath, M. P.; Radom, L. The Possible Role of Chlorine Trioxide Isomers in Relation to Stratospheric Ozone. J. Phys. Chem. 1993, 97 (30), 7947–7954. (67) Ju, X.-H.; Wang, Z.-Y.; Yan, X.-F.; Xiao, H.-M. Density Functional Theory Studies on Dioxygen Difluoride and Other Fluorine/oxygen Binary Compounds: Availability and Shortcoming. J. Mol. Struct. THEOCHEM 2007, 804 (1–3), 95–100. (68) Ruscic, B.; Pinzon, R. E.; Morton, M. L.; von Laszevski, G.; Bittner, S. J.; Nijsure, S. G.; Amin, K. A.; Minkoff, M.; Wagner, A. F. Introduction to Active Thermochemical 84 Tables: Several “Key” Enthalpies of Formation Revisited. J. Phys. Chem. A 2004, 108 (45), 9979–9997. (69) Ruscic, B.; Pinzon, R. E.; Laszewski, G. Von; Kodeboyina, D.; Burcat, A.; Leahy, D.; Montoy, D.; Wagner, A. F. Active Thermochemical Tables: Thermochemistry for the 21st Century. J. Phys. Conf. Ser. 2005, 16, 561–570. (70) Sander, S. P.; Friedl, R. R.; Golden, D. M.; Kurylo, M. J.; Moortgat, G. K.; Keller-Rudek, H.; Wine, P. H.; Ravishankara, A. R.; Kolb, C. E.; Molina, M. J.; et al. Chemical Kinetics and Photochemical Data for Use in Atmospheric Studies, Evaluation N. 15; JPL Publication 06-2, Jet Propulsion Laboratory: Pasadena, 2006. (71) Nickolaisen, S. L.; Friedl, R. R.; Sander, S. P. Kinetics and Mechanism of the Chlorine Oxide ClO + ClO Reaction: Pressure and Temperature Dependences of the Bimolecular and Termolecular Channels and Thermal Decomposition of Chlorine Peroxide. J. Phys. Chem. 1994, 98 (1), 155–169. (72) Joens, J. A. The Dissociation Energy of OH(X2 Π3/2) and the Enthalpy of Formation of OH(X2 Π3/2), ClOH, and BrOH from Thermochemical Cycles. J. Phys. Chem. A 2001, 105 (49), 11041–11044. 85 CHAPTER 4 ENTHALPIES OF FORMATION FOR ORGANOSELENIUM COMPOUNDS VIA SEVERAL THERMOCHEMICAL SCHEMES 4.1. Introduction Organoselenium compounds have been a subject of interest due to their potential applications in areas including organic synthesis,1,2 green chemistry,3,4 biochemistry,5-8 redox chemistry,9,10 and synthesis of conducting materials, semiconductors, and optoelectronic materials.11-15 Organoselenium reagents such as selenoxide, carbonyl selenide, isoselenocyanates, and selenones play an important role in organic reactions involving transformation mechanisms and typically result in high yields overall.1,2 The generation of complex alkenes via the stereospecific syn-elimination of selenoxides is another important and successful applications of organic selenium compounds.16-18 Although selenium is a chalcogen like oxygen and sulfur, it exhibits quite different chemistry. The selenium-carbon bond (bond length (r) = 1.98 Å and bond dissociation energy (BDE) = 55.93 kcal mol-1) is longer and weaker than that of the sulfur-carbon bond (r = 1.81 Å and BDE = 65.01 kcal mol-1), which, in turn, is longer and weaker than the carbon-oxygen bond (r = 1.41 Å and BDE = 85.56 kcal mol-1). Organoselenium reagents, as a result, are more active and involved in wide range of chemical applications than their corresponding organosulfur and organo-oxygen compounds.19 Insight about the potential utility of organoselenium compounds can be gained, in part, by knowing about their thermochemical properties including enthalpies of formation, Gibbs free energies, and bond dissociation energies at certain temperatures. Due to the weak Se-C and SeSe bonds, some organoselenium compounds tend to be relatively unstable during 86 thermochemical measurements. In addition to their instability, toxicity and difficulties in purification also contribute to the limited availability of thermochemical experimental data for organoselenium species as well as numerous discrepancies among the available experimental results.20,21 For example, Voronkov et al.22 reported an enthalpy of formation (∆ ° ,$¤¥ ) of diethyl selenide (C2H5SeC2H5) of -11.78 ± 0.96 kcal mol-1 while Tel’noi et al.23 reported an ∆ ° ,$¤¥ of -5.02 ± 0.96 kcal mol-1 for the same molecule. The most recent review of the thermochemistry of organoselenium compounds was in 2011 by Liebman and Slayden21 in which they reviewed and assessed the available ∆ ° ,$¤¥ ′ of organoselenium compounds that were mostly reported in three different review publications.22-24 The lack of reliable thermochemical properties of organoselenium species increases the need for high-level ab initio quantum chemical calculations for the prediction of energetic properties of organoselenium compounds and for the validation of available experimental data. Boyd et al.25,26 evaluated the performance of density functionals for the prediction of the geometries and bond dissociation energies of several biologically relevant organoselenium compounds with respect to the quadratic configuration interaction (QCISD) in conjunction with the cc-pVTZ basis set. The authors found that B3PW91 in conjunction with 6-311G(2df,p) performed the best.25,26 Another study by Maung et al.27 found that MP2/6-311G(d,p) predicted the most accurate BDE of HSe-H (H2Se) with a deviation of only 0.3 kcal mol-1 off from experiment, while B3LYP/6-311G(d,p) predicted a BDE that is 4.7 kcal mol-1 higher than the experimental value. Maung also calculated BDEs of other organoselenium compounds using several density functionals. Due to the absence of reliable reference data, according to Maung, no definitive conclusion can be reached from their study,27 although the non-local BHandHLYP functional was determined to be the most useful choice for the prediction of BDEs of 87 organoselenium compounds at an inexpensive computational cost with respect to its prediction of the BDEs of HSe-H and H-Se.27 Overall, most of the theoretical applications of quantum mechanics on organoselenium compounds involved the prediction and evaluation of BDEs rather than ∆ ° ,$¤¥ ′ .25-30 Although QCISD and coupled cluster with single, double, and perturbative triple excitations (CCSD(T)) have been reliable for predicting energetic properties, it is essential that the energies are extrapolated to the complete basis set (CBS) limit, i.e. the basis set incompleteness error is eliminated. However, CBS calculations with highly correlated methods are computationally demanding in terms of computer time, memory, and disk space. A number of ab initio composite approaches have been developed for modeling thermochemical properties with accuracy similar to that possible with CCSD(T)/CBS, but with reduced computational cost.31-38 Composite methods utilize a series of steps combining lower level methods and basis sets to replicate results possible with higher level methods. The most commonly used composite methods are the Gaussian-n (Gn) methods developed by Pople et al.31-34 Other successful composite methods are the correlation consistent Composite Approaches (ccCA) developed by Wilson et al.35-41 The Gn methods, detailed in Chapter 2 Section 2.5.2 and elsewhere,42 were extended to include molecules containing third-row main group elements.34,43,44 The developers of the Gn methods also introduced molecule sets, such as G3/05 set,45 that can be used to gauge the utility of computational approaches. The only selenium-containing compounds in the G3/05 set are SeH and SeH2. The deviations of the calculated atomization energies with respect to experiments for SeH are 0.1 kcal mol-1 (G2), -1.1 kcal mol-1 (G3), and -0.7 kcal mol-1 (G4), and the deviations for SeH2 are 1.1 kcal mol-1 (G2), 0.9 kcal mol-1 (G3), and 1.1 kcal mol-1 (G4).34,44 The ∆ ° 88 ,$¤¥ of SeH2 was computed using G2 and was only 1.1 kcal mol-1 off from the experimental value.46 ccCA, detailed in Chapter 2 Section 2.5.2 and elsewhere,47 also was applied to the G3/05 training set.37 The deviation of the calculated atomization energies when using ccCA with respect to experiment was -0.7 kcal mol-1 for SeH and -0.4 kcal mol-1 for SeH2. These deviations were reduced to -0.2 kcal mol-1 for both molecules when including the theoretical second-order atomic spin-orbit corrections, calculated by Blaudeau et al.48 using the configuration interaction (CI) method, to the ccCA energy,37 displaying a superior performance over the Gn methods for these molecules. The ccCA mean absolute deviation (MAD) for the molecules containing third-row atoms (Ga-Kr) included in the G3/05 set was 0.95 kcal mol-1 (0.88 kcal mol-1 when the secondorder atomic spin-order corrections was included) for a group of thermochemical properties including 19 atomization energies (D0), 11 enthalpies of formation (∆Hf), 15 ionization potentials (IP), 4 electron affinities (EA), and 2 proton affinities (PA).37 The aforementioned examples illustrate the rigor of these composite approaches. There has been much less investigation of the ∆ ° ,$¤¥ ′ of selenium-containing organic compounds than for oxygen- or sulfur-containing organic compounds. The most common and simplest method to determine the ∆ ° ,$¤¥ ′ in calculations is through the use of the atomization energy approach. This approach employs the difference in energy between the target molecule and its constituent atoms. Although the atomization approach (RC0) has been successfully applied to predict ∆ ° ,$¤¥ ′ using high levels of theory and/or model chemistries, differential electron correlation effects and size extensivity can be two problems associated with using this approach. Differential electron correlation effects come from the difference in correlation energy between the molecules (can involve conjugation, polarization, and strain) and the isolated atoms.49-51 Size extensivity is the ability of a method to scale linearly with increasing the number 89 of electrons, i.e. the method becomes independent of the size of the system.41,52 Typically, these two factors become increasingly important as the size of the target molecule is increased. Additionally, since the atomization approach involves relative energies of a molecule and its constituent atoms, any error associated with the approximation of the Schrödinger equation (i.e. Born-Oppenheimer approximation, lack of correlation correction, relativistic effects, and incomplete basis sets) will not be balanced resulting in an accumulation of errors.41,53 In order to reduce these errors, molecular reaction schemes such as the isogyric reaction, isodesmic reaction, and homodesmotic reaction schemes have been developed to calculate thermochemical properties based on bond interaction energies.54,55 In thermochemical reaction schemes, relative energies are calculated between the target molecule and its constituent molecules (known as elemental reactants and products) rather than between the target molecule and its constituent atoms as is the case when using the atomization approach. This allows for the cancellation of errors arising from differential electron correlation and size extensivity; thus, in principle, the isodesmic reaction scheme can provide more accurate energetic properties than the atomization approach depending on the accuracy of the calculated reaction enthalpy of the bond separation reaction (BSR) of the target molecule and the experimental ∆ ° molecules.40,41,52,56-58 ∆ ° ,$¤¥ ′ of the constituent Wheeler et al.57,58 defined a hierarchy of homodesmotic reactions for the calculation of ,$¤¥ ′ for hydrocarbon compounds. The hierarchy includes: isogyric (RC1), isodesmic (RC2), hypohomodesmotic (RC3), homodesmotic (RC4), and hyperhomodesmotic (RC5) reactions.57 As the homodesmotic reactions hierarchy increases from RC1 to RC5, the degree of error-balanced between reactants and products increases as well as the accuracy of computed enthalpies. For acyclic, closed-shell hydrocarbons, even low-level computational methods such 90 as Hartree-Fock can provide chemical accuracy when it is used with RC4 or RC5 reaction schemes.58 The definition of these reaction schemes, according to Wheeler,57,58 are shown in Table 4.1. In RC1 the number of electron pairs is maintained while in RC2 the number and type of carbon-carbon are both maintained. Not only the number and the bond type of C-C bonds are conserved in RC3, but also the hybridization state and the number of carbons with an equal number of hydrogen atoms attached are both maintained. RC4 and RC5 balance more chemical interactions than the lower scheme, see Table 4.1. This homodesmotic reaction hierarchy provides a clear classification and definition of homodesmotic reactions that make it possible to extend the scheme to a variety of organic molecules including those containing oxygen and sulfur atoms. Table 4.1 The definition of the homodesmotic reaction schemes according to Wheeler.57,58 Reaction scheme RC1 RC2 RC3 Name Reaction scheme constraints isogyric isodesmic hypohomodesmotic The number of electron pairs and unpaired electrons The number and bond type of carbon-carbon The number and the bond type of C-C bond, and The hybridization state of each carbon atom and the number of carbons with an equal number of hydrogen atoms attached “The number of each type of carbon-carbon bond [C§¨© − C§¨© , C§¨© − C§¨š , C§¨© − C§¨ , C§¨š − C§¨š , C§¨š − C§¨ , C§¨ − C§¨ , C§¨š = C§¨š , C§¨š = C§¨ , C§¨ = C§¨ , C§¨ ≡ C§¨]in reactants and products, and The numbers of each type of carbon atom (sp3, sp2, sp) with zero, one, two, and three hydrogens attached in reactants and products” The number of carbon-carbon bond types [ Hv C − CH$ , Hv C − CH, H$ C − CH$ , Hv C − C, H$ C − CH, H$ C − C, HC − CH, HC − C, C − C, H$ C = CH, HC = CH, H$ C = C, HC = C, C = C, HC ≡ C, and C ≡ C] in reactants and products, and The number of each type of carbon atom (sp3, sp2, sp) with zero, one, two, and three hydrogens attached in reactants and products” RC4 homodesmotic RC5 hyperhomodesmotic Wilson et al.41 extended this definition to include larger molecules and found that for the 91 ccCA and G3 methods it is necessary to use higher level reaction schemes to gain high accuracy when calculating the ∆ ° ,$¤¥ ′ of aromatic hydrocarbons. However, this statement does not hold for the G4 method. The ccCA, G3, and G4 methods used in the Wilson study resulted in MAD’s for the ∆ ° ,$¤¥ ′ of aromatic hydrocarbon of 2.88, 1.60, and 1.04 kcal mol-1, respectively for RC2 and 0.79, 1.55, and 1.78 kcal mol-1, respectively for RC3. These results show that RC3 is recommended for larger molecules, especially when using ccCA. Engelkemier and Windus59 extended this hierarchy to calculated ∆ ° ,$¤¥ ′ of oxygen-containing organic molecules and showed the effectiveness of these reaction schemes in the cancellation of errors and in providing more accurate enthalpies even for large molecules such β-D-glucopyranose-gg. Jorgensen and Wilson40 published a detailed study of the effect of the hypohomodesmotic reaction scheme (RC3) when used in combination with ccCA, G3, and G4 for the prediction of ∆ ° ,$¤¥ ′ of organosulfur species and compared it with the atomization approach (RC0). In general, the RC3 reaction scheme did decrease the overall MAD of the ∆ ° ,$¤¥ ′ for ccCA, G3, and G4 as compared to the RC0 approach, though not significantly. For example, the MAD of the ∆ ° ,$¤¥ ′ of organosulfur species computed with ccCA-P is 0.98 kcal mol-1 using RC0 and 0.54 kcal mol-1 using RC3.40 Thus, RC0 is still an effective method when using composite approaches such as ccCA, G3, and G4,40,41 particularly for light atoms or/and small molecules but not necessarily for heavy elements or/and large molecules. However, the decrease in the MAD when using the RC3 scheme compared to the RC0 for the component methods that are used for the additive terms in the ccCA methodology is very significant (2.39 kcal mol-1 when using MP2/aug-cc-pVQZ to 78.11 kcal mol-1 when using MP2/aug-cc-pVDZ).40 92 In the present study, the homodesmotic hierarchy is used to predict ∆ ° ,$¤¥ ′ for selenium-containing organic compounds. The effect of RC2 and RC3 schemes on predicting the ∆ ° ,$¤¥ ′ of organoselenium molecules via ccCA, G3, G4, and B3PW91/aug-cc-pVTZ is investigated and compared to RC0 results. Due to the lack of reliable thermochemical properties of organoselenium species, the molecule set in this study is smaller than for the organosulfur study,40 mentioned earlier. Additionally, the quality of the experimental ∆ ° ,$¤¥ ′ is also assessed due to the significant discrepancies between the reported experimental results. The performance of each individual step of ccCA, using RC2 and RC3 schemes, is also discussed. The study compares trends of the computational approaches used to calculate the ∆ ° chalcogen-containing hydrocarbon molecules (group 16 of the periodic table). ,$¤¥ ′ of 4.2 Computational Details 4.2.1 Methods All calculations have been carried out using the Gaussian 09 software package.60 The correlation consistent Composite Approach (ccCA), detailed in Chapter 2 Section 2.5.2 and elsewhere35,36,38 and has been applied widely in prior studies,40,41,47,61-63 were used to predict the ∆ ° ,$¤¥ ′ of organoselenium species. For molecules containing third row atoms (Ga-Kr), the atomic second-order spin orbit coupling is also added to the total ccCA atomic energies.37 The valence correlation space includes the 3s, 3p, 4s, 3d, and 4p orbitals for selenium in all single point energy calculations within the ccCA methodology and the FC1 correlation space include all the electrons except the 1s electrons for selenium.37 93 The most commonly used composite methods, G344 and G434 were used for comparison. The hybrid density functional B3PW9164,65 in conjunction with the aug-cc-pVTZ basis set was also used as it was found to provide accurate geometries and energies for organoselenium compounds when compared to results obtained using QCISD/cc-pVTZ.26 The ∆ ° ,$¤¥ ′ of organoselenium compounds were also calculated using each level of theory that are components of the ccCA methodology. The mean absolute deviation (MAD) and the mean signed deviation (MSD) were both calculated for each method. The geometries and structural parameters of all molecules of interest are presented in the Appendix. The frequency calculations and the wave function stability tests were both used to ensure that the predicted closed-shell structures are all a stable minimum. 4.2.2 Thermochemistry The ∆ ° ,$¤¥ ′ of organoselenium compounds were calculated using ccCA, G3, G4, and B3PW91 methods. Three thermochemical approaches were used in the calculations of the ∆ ° ,$¤¥ ′ : the atomization approach (RC0), the isodesmic reaction scheme (RC2), and the hypohomodesmotic reaction scheme (RC3). 4.2.2.1 Atomization Approach (RC0) In this method, the ∆ ° ,$¤¥ ′ were computed as follows: 94 ∆ ,1 298i = ! ˜∆ ^FG“° +· 1 0i − ² ! ˜ ,/ 298i − − ² ! ˜· ^FG“° / 1 ^FG“° 0i ¸ 298i − / / − 1 − ³´µ ¶ (4.1) 0i ¸¶ The enthalpy of formation of the molecule of interest (M) (∆ ,1 ) at 298 K is calculated using Eq. 4.1. The first term in Eq. 4.1 is the experimental enthalpy of formation of its constituent atoms (∆ ,/ ) at 0K multiplied by n, the number of each type of atom. The recommended experimental ∆ ° ,Q of the selenium atom, 57.9 kcal mol-1,66,67 was used in this study. Carbon and hydrogen ∆ ° ,Q ′ are the same as those used in Jorgensen’s study,40 which are 170.11 kcal mol-1 and 51.63 kcal mol-1,68,69 respectively. The second term is the atomization energy of the molecule of interest (¹Q = ∑^FG“° ˜ / − 1 − ³´µ ), where EA, EM, and EZPE are the total energies of the constituent atoms, the total energy of the molecule of interest, and the zero point energy of the molecule of interest, respectively. The energies are calculated at the specified level of theory. The last two components are the thermal corrections to the enthalpies to account for a temperature of 298 K for the atoms (experimental) and for the molecule of interest, calculated at a specified level of theory. The thermal corrections to enthalpies for atoms are 1.32 kcal mol-1 for selenium,66 0.25 kcal mol-1 for carbon,68 and 1.01 kcal mol-1 for hydrogen.69 4.2.2.2 Isodesmic (RC2) and Hypohomodesmotic (RC3) Reaction Schemes The ∆ ° ,$¤¥ ′ were computed using the RC2 and RC3 reaction schemes as follows: 95 ∆ In Eq. 4.2, the ∆ ,1 298i = − ! ∆ »\G¼ZDF ! \;^DF^