Mir: «'4‘: an.” aw. m... ‘ 1!- :. MFG}. ’ ‘60 “1' i'ysgiyirfi?sé’ifi.rn 3' WT." ' “if" u-d“ V ‘HESiS MICHIGANS III IIIIIII IIIIIIII III III III I‘IIIIIIII 01413 7768 ll This is to certify that the dissertation entitled The Study of Ion/Molecule Thermochemistry of Chiorotitanium Ions Utilizing a Hybrid Mass Spectrometer presented by Kurtis Richard Kneen has been accepted towards fulfillment of the requirements for Ph 4) . degree in Chem I. stny_ %/ i/e Major professor Date 1’4"“; 7‘ MSU i: an Affirmative Action/Eq ua! Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE IN RETURN BOX to remove thle checkout from your record. 70 AVOID F INES return on or bdore dete due. DATE DUE DATE DUE DATE DUE I I i i MSU Ie An Affirmative Action/Equal Opportunity lnetltulon Wanna-9.1 THE STUDY OF ION/MOLECULE THERMOCHEMISTRY OF CHLOROTITANIUM IONS UTILIZING A HYBRID MASS SPECTROMETER by Kurtis Richard Kneen A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1996 ABSTRACT THE STUDY OF ION/MOLECULE THERMOCHEMISTRY OF CHLOROTITANIUM IONS UTILIZING A HYBRID MASS SPECTROMETER By Kurtis R. Kneen A hybrid mass spectrometer has been modified to study ion/molecule thermochemistry. The principle components of the mass spectrometer are: electron-impact (EI) ionization source, magnetic sector for mass selecting the reagent ions, deceleration lens to reduce the reagent ion kinetic energy from approximately 1000 eV to 0-50 eV, an octopole to act as a beam guide to assure efficient collection of product ions, a collision cell in which the gas- phase chemistry occurs, a quadrupole for mass selecting both product and reagent ions exiting the octopole, and a daly-type detector with which standard pulse-counting techniques were employed. The principle of operation of each component is described here. Experimental procedures are presented as well as the method for data analysis. Experiments designed to evaluated both the accuracy of the reagent ion kinetic energy and for the collection efficiency of the beam guide are described. The gas-phase chemistry of chlorotitanium ions was studied using the mass spectrometer. Collision-induced dissociation (CID) experiments performed on TiClm+ (n = 1-4) ions generated by 70 eV E1 on TiC14 provided a measurement of the excess internal energy imparted to the ions during ionization. This information was then used to evaluate chemistry of the 'I‘iCln+ ions with oxygen-containing compounds and isobutene. This dissertation is dedicated to my grandfather, Bruce D. Caulkins. His curiosity of the natural world around him, along with his great love of teaching others, inspired me to fulfill my goal of attaining this advanced degree. My only regret is not to have graduated before he left us. iv ACKNOWLEDGMENTS Attaining my degree would not have been possible without the moral and technical support, guidance, and encouragement from a number of people. I would like to thank those who made the adjustment to graduate school during the first couple years a little more easier: Jeff and Kimberly Gilbert, Jon and Karen Wahl, Jason and Sue Rouse, Paul O'Connor, Gary Schultz, Dave Wagner, Mary Puzycki-Seeterlin, Dave Gale, Art Harms, Laura Pence and, Mike Waldo. Not only did they help me in my studies but also provided a release valve from the stress of school through activities such as attending football, hockey, and basketball games together, ski trips up north, Ceder Point each summer, parties at each other's apartment/house, and of course, "going out" on the town to experience some of the local taverns. I also would like to thank those who continued to show support throughout my career at MSU: Mark and Jan Sabo, Kris Kurtz and, Jim Ridge. These were the ones who helped me make it through the tough times by offering advice and moral support. Ed and Maria Townsend are special for a number of reasons. On top of the support at chemistry, I'll miss the early Tuesday and Thursday morning basketball games and runs across the MSU campus and going to MSU sporting events together. Ed and I seemed to always be the ones V planning and organizing and something tells me that it may continue if and when everyone from MSU gets together in the future. Maria has the distinction of introducing me to my wife. That is something for which I have no way of expressing my gratitude. Thank you Maria. Tracy and I met on a blind date while we both were attending MSU. She has provided me with the love and support only a best friend can offer. Her willingness to postpone her education and career while I attended MSU is the kind of support for which I know I'll never truly be able to return. I am grateful that my advisor, John Allison, was able to offer the kind of encouragement and guidance I needed to pursue my career at MSU. The weekly project meetings always seemed to offer a little humor which made the tougher times seem more bearable. JA, thank you for being a challenging and fair advisor. » TABLE OF CONTENTS LIST OF TABLES x LIST OF FIGURES xi CHAPTER 1 INTRODUCTION 1 1. Significance of Gas-Phase Organometallic Chemistry 1 2. Mass Spectrometry: A Discipline for Identification, Quantification, and Basic Research 3 3. Experimental Techniques for Studying Gas-Phase Ion/Molecule Chemistry 5 A. General Description 5 B. ICR, FTICR Mass Spectrometry 7 C. Metastable/Collision-Induced Decomposition Experiments 10 D. Ion-Beam Experiments 10 CHAPTER 2 INSTRUMENTATION 17 1. Overview 17 2. Ion Source 19 3. Magnetic and Electric Sectors 25 4. Deceleration Optics in 5. Collision Region 31 A. Beam Guide 31 B. Collision Cell 37 6. Ion Optics 43 7. Quadrupole 43 8. Detector System 45 9. Computer System 47 CHAPTER 3 EXPERIMENTAL SECTION 49 1. Initial Mass Spectrum 49 vii 2. Initial Identification of Products 50 3. Optimizing Conditions 51 4. Performing the Energy-Resolved Experiment 53 A. Stopping-Potential Scan 53 B. Energy-Resolved Scan 54 C. Background Scan 57 5. Data Analysis 57 A . Determining Ion Beam Energy Spread and Point of Zero Kinetic Energy in the Center-of-Mass Frame 57 B. Doppler Broadening 58 C. Incorporating Ion Beam Energy Spread 62 D. Experimental Cross Sections 63 E. Fitting the Data 64 CHAPTER 4 INSTRUMENT EVALUATION 67 1. Collision-Induced Dissociation of the Singly-Charged Manganese Dimer Ion 67 2. Reaction of Singly-Charged Argon with Molecular Deuterium 76 CHAPTER 5 CHEMISTRY OF CHLOROTITANIUM IONS 84 1. CID Experiments on 70 eV EI Chlorotitanium Ions 84 2. Chemistry Between TiClx+ and Oxygen-Containing Compounds 100 A. Results 101 B. Discussion 118 3. Chemistry Between TiClx+ ions and Isobutene 124 A. Thermochemistry 127 B. Mechanisms 135 1. Tile and isobutene chemistry 135 2. TiCl3+ and isobutene chemistry 138 3. Tile and isobutene chemistry 140 4. TiCl+ and isobutene chemistry 142 4. Conclusion 143 APPENDIX A ELECTRON ENERGY CALIBRATION 149 viii APPENDIX B DEFINITIONS 152 ix Table 4.1. Table 4.2. Table 5.1. Table 5.2. Table 5.3. Table 5.4. LIST OF TABLES Parameters of the Mn2+ CID Experiment. Parameters of the Ar+ + D2 Experiment. Heats of Formation for Ions Produced from TiCl4 [1]. Bond Dissociation Energies for Ionic and Neutral Chlorotitanium Molecules. List of products for the reaction TiClx+ (x: 1-4) + acetylaldehyde. Summary of Bond Dissociation Energies. 78 91 135 Figure 1. 1. Figure 1.2. Figure 1.3. Figure 1.4. Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. Figure 2.6. Figure 2.7. Figure 2.8. Figure 2.9. LIST OF FIGURES Block diagram of a conventional mass spectrometer 6 Schematic representation of a cubic ICR cell located in a magnetic field of strength B in the positive 2 direction. T and T' are the trapping, E and E' are the excitation, and D and D' are the analyzer plates. 8 Schematic diagram of a multisector mass spectrometer consisting of an ion source, three mass analyzers (which can be either magnetic or electrostatic), 2 collision cells, and a detector. 11 Block diagram of the MSU ion beam instrument. 14 Ion Beam Instrument Schematic Diagram. 18 Schematic and Voltage Diagram of Intensitron EI Source. 2) Illustration of a vertical transition. 21 SIMION model of the Intensitron EI Source. Equipotential lines are shown, separated by 1 V. 23 Dispersion of the ion beam by the magnetic sector. Fall is the focal point for ions with mass 3 and velocity 1. F82 is the focal point for ions with mass 3 and velocity 2. Fb1 and sz are the focal points for ions with mass b and velocities 1 and 2. % Schematic of Deceleration Lens System. I!) Comparison of restoring forces for a quadrupole (Quad), hexapole (Hexa), and an octopole (Oct). Calculated radial ion kinetic energy in rf multipole [3]. 35 Schematic diagram of the oct0pole and collision cell. 39 xi Figure 2.10. Figure 2.11. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 4.1. Figure 4.2. Figure 4.3. Figure 4.4. Figure 4.5. Figure 4.6. Figure 4.7. Figure 5.1. Figure 5.2. Ionization gauge calibration curve for argon. Schematic for transfer and extraction optics and the detector system. Potential energy diagram of the experiment. A stopping potential curve (TiCl3+) with the quadrupole operated in the rf-only mode. A stopping potential curve (TiCl4+) with the quadrupole operated in the rf-only mode. The first derivative curve was calculated from the stopping potential data plotted for TiCl4+ in fig. 3.3. The Gaussion curve was then fit to the first derivative. The FWHM is taken as the energy spread and the apex is taken as the point of zero kinetic energy. Comparison of three Mn2+ data curves. Eachcurve was normalized to its maximum value. Comparison of Mn2+ CID threshold regions. Fit of Run 1 of the Mn2+ experiment. ET, n, and m are defined in chapter 2. Fit of Run 2 of the an+ experiment. ET, n, and m are defined in chapter 2. Fit of Run 3 of the Mn2+ experiment. ET, n, and m are defined in chapter 2. Illustration of product collection efficiency of the MSU ion beam instrument. Energy dependence of the cross-section for the reaction: Ar+ 4» D2 --> ArD“ + D. Energy resolved CID of TiCl4+. Argon was used as the collision gas. Energy resolved CID of TiCl3+. Argon was used as the collision gas. xii it 8} 71 74 75 H) 81 87 88 Figure 5.3. Figure 5.4. Figure 5.5. Figure 5.6. Figure 5.7. Figure 5.8. Figure 5.9. Figure 5.10. Figure 5.11. Figure 5.12. Figure 5.13. Energy resolved CID of Tile. Argon was used as the collision gas. 89 Energy resolved CID of TiCl+. Argon was used as the collision gas. 90 Bond dissociation energies of chlorine atoms from TiClx (x = 1-4) where 1 represents the loss of the first chlorine (TiCl3--Cl), 4 represents dissociation of the last chlorine (Ti--C1), and 2 and 3 represent the intermediate cases. 92 Trends in successive bond dissociation energies (BDEs) [8]. The first bar in each series represents the energy required to remove the first atom, the second bar, the second atom, etc. 94 Comparison between proposed geometry after EI ionization and expected ground state geometry. % A) Product mass scan for TiC14+ + CH3CHO with the octopole DC offset set 4 Volts below Vacce]. B) Expanded view of the spectrum. 102 A) Product mass scan for TiCl4+ + CH3CHO with the octopole DC offset set 20 Volts below Vaccel- B) Expanded view of the spectrum. 103 A) Energy dependent reaction cross-sections for the reaction TiCl4+ + C2H4O (acetylaldehyde). B) Expanded view of plot (A). 104 A) Energy dependent reaction cross-sections for the reaction TiCl+ + CzH4O (acetylaldehyde). B) Expanded view of plot (A). 106 A) Energy dependent reaction cross-sections for the reaction TiC12+ + C2H4O (acetylaldehyde). B) Expanded view of plot (A). 107 A) Energy dependent reaction cross-sections for the reaction TiCl3+ + C2H4O (acetylaldehyde). B) Expanded view of plot (A). 107 xiii Figure 5.14. Figure 5.15. Figure 5.16. Figure 5.17. Figure 5.18. Figure 5.19. Figure 5.20. Figure 5.21. Figure 5.22. Figure 5.23. Figure 5.24. A) Energy dependent reaction cross-sections for the reaction TiCl+ + C3H60 (propanal). B) Expanded view of plot (A). 110 A) Energy dependent reaction cross-sections for the reaction TiC12+ + C3H60 (propanal). B) Expanded view of plot (A). 111 A) Energy dependent reaction cross-sections for the reaction TiCl3+ + C3H5O (propanal). B) Expanded view of plot (A). 112 A) Energy dependent reaction cross-sections for the reaction TiCl4+ + C3H60 (propanal). B) Expanded view of plot (A). 113 A) Energy dependent reaction cross-sections for the reaction TiCl“ + C3H60 (acetone). B) Expanded view of plot (A). 114 A) Energy dependent reaction cross-sections for the reaction TiC12+ + C3HGO (acetone). B) Expanded view of plot (A). 115 A) Energy dependent reaction cross-sections for the reaction TiCl3+ + C3H6O (acetone). B) Expanded view of plot (A). 116 A) Energy dependent reaction cross-sections for the reaction TiCl4+ + CgHGO (acetone). 117 A) Energy dependent reaction cross-sections for the reaction TiCl” + C2H6O (dimethyl ether). B) Expanded view of plot (A). 119 A) Energy dependent reaction cross-sections for the reaction T1012+ + CzHeO (dimethyl ether). B) Expanded view of plot (A). 120 A) Energy dependent reaction cross-sections for the reaction TiCl3+ + CgHeO (dimethyl ether). B) Expanded view of plot (A). 121 xiv Figure 5.25. Figure 5.26. Figure 5.27. Figure 5.28. Figure 5.29. Figure 5.30. Figure 5.31. Figure 5.32. Figure 5.33. Figure 5.34. A) Energy dependent reaction cross-sections for the reaction TiCl4+ + CZHGO (dimethyl ether). B) Expanded view of plot (A). 12 Reaction scheme developed by AR2 [41] for the chemistry between chlorotitanium ions and olefins. TiCl4+ was found to be unreactive with olefins. 125 A) Energy dependence of reaction cross-sections for the reaction TiCl4+ + C4H3 (isobutene). B) Expanded view of plot (A). 128 A) Energy dependence of reaction cross-sections for the reaction TiC13+ + C4H8 (isobutene). B) Expanded view of plot (A). 129 A) Energy dependence of reaction cross-sections for the reaction TiC12+ + C4H8 (isobutene). B) Expanded view of plot (A). 130 A) Energy dependence of reaction cross-sections for the reaction TiCl“ + C4H8 (isobutene). B) Expanded view of plot (A). 131 Proposed mechanism for the reaction between TiCl4+ and isobutene. 137 Proposed mechanism for the reaction between Tile and isobutene. 139 Proposed mechanism for the reaction between TiC12+ and isobutene. 141 Proposed mechanism for the reaction between TiCl+ and isobutene. 144 XV Chapter 1 Introduction *#*************************Il!#1.!3H"!*4!*III*4!I”!!!IIIIIIIMHO!************************* I . Significance of Gas-Phase Organometallic Chemistry Methods of synthetic chemistry have exploded both in number and complexity over the last 50 years. As a result, members of almost every class of organic compound can be considered starting material for the conversion to members of other classes or more complex materials such as polymers, biomolecules, or organometallic compounds. An exception to this generalization is the important class known as alkanes. Their lack of reactivity is often attributed to the high alkane C-H bond energies (~90-100 kcal/mole) and the low acidities and basicities of alkanes [1,2]. The stability of alkanes represents a fundamental challenge to chemists because the C-H and CC bonds represent the most prevalent forms of chemical linkage in nature. Therefore, resolving the requirements needed to break these bonds would greatly increase our understanding of chemical reactivity. Also, alkanes are relatively inexpensive due to their high concentrations in both natural gas and petroleum. They represent a significant source of carbon potentially available for chemical synthesis. 2 One area of chemistry that holds promise in achieving CH and C-C bond activation is that of organometallic chemistry. Understanding of the gas-phase chemistry of organometallic systems has undergone a rapid increase over the past decade. The motivation behind the increase in interest is the realization that knowledge about the intrinsic properties of the bare or ligated transition-metal ions can provide valuable insight into both the mechanisms active in the condensed phase and catalytic efficiencies in general. In fact, correlations have been made between results from gas-phase transition-metal ion studies and the reactivity of transition-metal-containing compounds in condensed phases [3]. Gas-phase studies are uniquely suited for the characterization of isolated molecules/ions as well as probing elementary reaction steps under well defined conditions. The studies are not complicated by solvent effects such as solvent-shell interactions, intermolecular processes that alter or destroy reactive intermediates, and associations by ion pairing [4]. The advantage of gas-phase studies stems from the spatial separation of the molecules/ions characteristic during high vacuum conditions. Many of the first studies on gas-phase organometallic chemistry were concerned with alkane fuctionalization to make use of the chemical potential of the natural gas and fuel feedstocks. Today, that interest continues as we look to uncover critical steps [5-8] and identify and characterize reaction intermediates [9-15] so that we may improve existing catalysts. For example, low temperature and pressure polymerization of 3 olefins is accomplished through the use of organo-aluminum/titanium compounds known as Ziegler-Natta catalysts [16]. Transition-metals have a number of low energy atomic orbitals that provide bonding opportunities for organic compounds. This ability to provide a favorable bonding environment may help to stabilize highly reactive intermediates. However, the details behind the process of bond cleavage of inactivated OH or C-C bonds of the metal-alkene derivatives are not yet fully understood. II. Mass Spectrometry: A Discipline for Identification, Quantification, and Basic Research Since its beginnings in 1913 as a method used for demonstrating the existence of isotopic forms of stable elements by J. J. Thomson [17], mass spectrometry has become one of the most widely applicable disciplines of scientific inquiry. Mass spectrometry is being used to gather qualitative and quantitative information in fields ranging from environmental trace analysis through the biological and medicinal sciences to studying the surface and atmospheric chemistry on other planets. Mass spectrometers are routinely used to analyze samples where the mass of the analytes range from one to near one hundred thousand Daltons, with an ever increasing upper limit. Extremely low detection limits allow for precise measurements when the analyte of interest may be present in sub-picomole levels. Mass spectrometers are also utilized in fields of basic research. They are used in nuclear and atomic physics as well as fundamental chemical 4 investigations. The low pressures under which the instruments are generally operated make them ideal for studying gas-phase chemistry. Mass spectrometers manipulate charged reactants and products through applied magnetic and electric fields. This provides additional advantages when studying ion/molecule or ion/ion chemistry because of the inherent electrical properties of the ions. When mass spectrometers are operated under proper conditions (e.g.. single-collision conditions), the experimental environments allow a significant amount of chemical information to be obtained. This dissertation describes the characteristics and performance of a hybrid mass spectrometer (ion beam instrument) designed to study gas- phase ion/molecule chemistry. The MSU ion beam instrument enables the study of ion/neutral chemistry under carefully controlled conditions. Cross-sections for ion/molecule reactions and collision-induced dissociation (CID) may be evaluated as a function of ion translational and internal energy. Information obtained in this manner may then be related to bond energies and heats of formation. Another technique being developed in this laboratory allows for the structural investigation of isolated ions [18]. This effort involves the deposition of mass-selected ions into a low temperature noble-gas matrix, whereafter Fourier transform infrared spectroscopy (FTIR) is applied. It is hoped that this technique will be developed to provide direct spectroscopic analysis of transition-metal-containing ions. The MSU ion beam instrument will provide complementary information. 5 The remaining section of this chapter describes other major techniques used to study gas-phase ion chemistry with respect to this research project. It also provides a general description of the MSU beam machine. The succeeding chapters detail the design characteristics, instrument evaluations, and experimental results of gas-phase chlorotitanium ion chemistry. I II. Experimental Techniques for Studying Gas-Phase Ion/Molecule Chemistry A . General Description Before experimental mass spectrometric techniques can be discussed, a general description of a mass spectrometer must be included. There is no unique mass spectrometer design. The wide diversity of mass spectrometers, however, all consist of four basic components: sample inlet system, ion source, mass analyzer, and detector. The entire instrument is operated under a vacuum; with the mass analyzer and detector pressures maintained below 10'6 torr. The sample inlet system and ion source are generally operated at slightly higher pressures (2 10‘5 torr). A general block diagram of a conventional mass spectrometer is shown in Figure 1.1. The sample is injected into the instrument, ionized, and extracted from the ion source with electrostatic optics. The ions are focused into the mass analyzer, separated according to their mass-to-charge ratio, and then detected. There are a wide variety of methods to perform the function of each basic component. .8588?on $38 32835.58 a mo Eobwmmc #85 AA 9363 58¢qu Al 553mg mmog mouoom :3 acid 395% B. ICR, FTICR Mass Spectrometry Currently there are three dominant instrumental techniques in mass spectrometry employed to study gas-phase ion/molecule chemistry. The first technique to be discussed utilizes ion cyclotron resonance mass spectrometers (ICR) [19]. When a Fourier transform is performed, the technique is referred to as FTICR mass spectrometry or FTMS [20]. When this type of mass spectrometer is employed, ions can be stored in crossed electric and magnetic fields generated in cells similar to the cell shown in Figure 1.2. The magnetic field restricts the motion of the ions to a circular path in the xy-plane. Static DC potentials placed on the trapping plates T and T' create a potential well between the two plates, thereby restricting the motion of the ions in the z-direction. Ions of a given mass move in a circular motion (known as the cyclotron . motion) and there is a characteristic frequency associated with this motion. The frequency, f, is a function of the charge, q, and mass, m, of the ion and the magnetic field strength, B, as expressed in the equation below: f = qB/21tm (1) The charge of the ion, q, is equal to the fundamental charge, 1.602x10’19 C multiplied by the number of charges. The angular frequency, to = 21tf, is known as the cyclotron frequency of the ion. The mass/charge value of the ion is thus related to the frequency of the ion orbit. The ion is detected by A ‘L . / Figure 1.2. Schematic representation of a cubic ICR cell located in a magnetic field of strength B in the positive 2 direction. T and T' are the trapping, E and E' are the excitation, and D and D' are the analyzer plates. 9 applying a signal at the cyclotron frequency across or between the analyzer plates. Ions orbiting at the applied frequency absorb power and experience an increase in the radii of their cyclotron motion. The power absorption is measured by the circuitry of the instrument with the absorption intensity being related to the ion population within the cell. FTMS is performed by applying a brief signal, referred to as a "chirp," which consists of a complex waveform usually containing all of the possible ion cyclotron frequencies over a given mass range. Ions within the ICR cell absorb power and begin to move in coherent ion "packets,' with each ion packet corresponding to a particular m/z value. These packets, in turn, induce a signal on the analyzer plates. This signal is a mix of all of the frequencies corresponding to each ion m/z value present in the ICR cell. A Fourier transform is applied to this signal which results in a plot of signal amplitude vs. frequency, which is converted to a mass spectrum once the relationship between frequency and m/z is determined (calibration). There are two important advantages of using an ICR to study gas-phase ion/molecule chemistry. Because of the instrument's ability to trap ions, ions can be stored for relatively long periods of time (millisecond range). This allows for long periods of interaction between ions and neutral reagent species (also present in the cell) which result in product abundances that are higher than those found when utilizing other mass spectrometers. Double resonance techniques employed on ICR's allow the identification of precursor ions that give rise to individual reaction products [21,22]. Therefore, primary, secondary, tertiary, etc. reactions can be determined unambiguously. 10 C. Metastable/Collision-Induced Decomposition Experiments The main difference between metastable/CID experiments and the other two techniques described in this section is that organometallic complexes are studied by investigating their decomposition pathways. The first application of this technique was performed by Freas and Ridge in 1980 [23]. Multisector instruments are generally employed although, in principle, every kind of tandem mass spectrometer could be used [24]. A schematic representation is shown in Figure 1.3. Complexes are generated through gas-phase chemistry in the ion source, extracted, and mass selected. In the field-free region (collision cell) following the first mass analyzer, products are generated either through unimolecular decomposition or CID. After drifting through the field-free region, both products and reagent ions are mass analyzed and detected. To obtain structural information on products, additional collision cells and mass analyzers can be used. An advantage of this technique is that the transition-metal-containing complexes are formed, decomposed, and detected in separate sections of the instrument. This decreases possible interferences such as additional gas- phase chemistry of the complexes before they can be analyzed. D. Ion-Beam Experiments Ion-beam experiments are performed in guided beam instruments (mass spectrometers) designed to measure product abundances as a function of kinetic energy. Reactions studied are between mass selected gas-phase 11 €230 E ~3wa mmmz N :oO pom—086D :oEEoO EH Al mouthing A. $32 _ _ _. lllllllll I_ $9298 monsoon—Sm 8:33 .8.“ 3:330 .5835“. a new i=8 oommmzoo m .Aosfimouuoflo .8 38838 .856 on :8 533 mansions mmoE 8.23 .858 com no mo wcsmmmcoo 888950QO mama 583538 a mo 8.2.me osofiofim .mA whomrm A H . Bahamas mmmz 8.25m :oH 12 reagent ions and stationary target neutral gases at ambient temperatures. Because the intent of these experiments is to measure the amount of energy required to "turn on" endothermic processes, conditions have to be set to allow single collisions. If multiple collisions were to occur, uncertainty in the measurements would arise from the inability to accurately measure the energy of interaction during any subsequent collisions. Therefore, to ensure single-collision conditions in the MSU ion-beam instrument (given the collision-cell dimensions), pressures in the collision cell were maintained at levels below 10’3 torr. The neutral reagent/collision gas is characterized by an isotropic Maxwellian velocity distribution. Also, the kinetic energy of the ions within the ion beam can be described by a distribution. These two properties combine to create an overall distribution of ion/neutral interaction energies. Techniques used to compensate for this distribution are described in Chapter 3. Theoretically, it is possible to increase the resolution of the measured ion/neutral interaction energy by incorporating a supersonic expansion neutral beam source in place of the static neutral collision gas cell. However, because of the low number density within the neutral beam and the small interaction volume (determined by the ion and neutral beam physical dimensions) [25,26], no successful crossed beam-guided ion beam instrument has been reported. Early pioneers of this technique were Futrell and coworkers in 1965 [27]. Their instrument consisted of two double focusing (magnetic and electric) mass spectrometers. Beauchamp and Armentrout were the first to apply this technique to organometallic chemistry [28-31]. Their instrument used a 13 quadrupole instead of a magnet or double sector mass spectrometer as the product mass analyzer. Armentrout further improved this method by employing an octopole beam guide-collision cell to ensure efficient collection of products [26,32]. The MSU ion beam instrument is of the guided beam design, incorporating an octopole as an ion beam guide and a static gas ion/neutral interaction region. A block diagram of the MSU instrument in shown in Figure 1.4. Detailed descriptions of the instrumentation are provided in Chapter. 2, including theory and operation. Chapter 3 covers experimental procedures and methods of data analysis. Instrument performance for both exothermic and endothermic chemical processes is given in Chapter 4. Chapter 5 concludes with experimental results of the gas-phase chemistry of chlorotitanium ions. 14 583mm 33 €585.5me :83 :3 sz 23 mo Eamon. x85 .vA warm ~3th mmmE Eomawwsfiu 22830 35.: I sumo nosmpflooom uBoom 032%sz 8.26m :8 10. 11. 12. 13. 14. 15. LIST OF REFERENCES Chapter 1 BA. Arndtsen, R.G. Bergman, T.A. Mobley, T.H. Peterson, Acc. Chem. Res, 28, 154 (1995). "Activation and Functionalization of Alkanes," C.L. Hill, Ed., John Wiley and Sons, New York, (1989). ED. Radecki, J. Allison, J. Am. Chem. Soc., 106, 946 (1984). AD. Ryabov, Chem. Rev., 90, 403 (1990). R.G. Bergman, P.F. Seidler, T.T. Wenzel, J. Am. Chem. Soc., 107, 4358 (1985). M. Hackett, J.A. Ibers, G.M. Whitesides, J. Am. Chem. Soc., 110, 1436 (1988). . T. Gregory, P. Harper, R.S. Shinomoto, M.A. Deming, T.C. Flood, J. Am. Chem. Soc., 110, 7915 (1988). W.A. Kiel, R.G. Ball, W.A.G. Graham, J. Organomet. Chem., 383, 481 (1990). A.H. Janowicz, R.G. Bergman, J. Am. Chem. Soc., 105, 3929 (1983). W.D. Jones, F.J. Feher, J. Am. Chem. Soc., 108, 4814 (1986). M.V. Baker, L.D. Field, J. Am. Chem. Soc., 108, 7436 (1986). G.L. Gould, D.M. Heinekey, J. Am. Chem. Soc., 111, 5502 (1989). RM. Bullock, C.E.L. Headford, K.M. Hennessy, S.E. Kegley, J.R. Norton, J. Am. Chem. Soc., 111, 3897 (1989). A.A. Bengali, R.H. Schultz, C.B. Moore, BB. Bergman, J. Am. Chem. Soc., 116, 9585 (1994). R.H. Schultz, A.A. Bengali, M.J. Tauber, B.H. Weiller, E.P. Wasserman, K.R. Kyle, C.B. Moore, RE. Bergman, J. Am. Chem. Soc., 116, 7369 (1994). 16. 17. 18. 19. 20. 21. 26. 27. 28. 29. 30. 31. 32. 16 K. Ziegler, Adv. Organometal. Chem., 6, 1 (1968). J .J. Thomson, "Rays of Positive Electricity," , Longmans, Green and Co., London (1913). T.M. Halasinski, J.T. Godbout, J. Allison, G.E. Leroi, J. Phys. Chem., 98, 3930 (1994). K.P. Wanczek, Int. J. Mass Spectrom. Ion Processes, 95, 1 (1989). AG. Marshall, P.B. Grosshans, Anal. Chem., 63, 215A (1991). MB. Comisarow, V. Grassi, G. Parisod, Chem. Phys. Lett., 57, 413 (1978). RT. McIver, Jr., W.D. Bowers, "Tandem Mass Spectrometry," F.W. McLafferty, Ed., Wiley-Interscience, New York (1983). RB. Freas, D.P. Ridge, J. Am. Chem. Soc., 102, 7129 (1980). F.W. McLafferty, Science, 214, 280 (1981). S. Anderson, F. Houle, D. Gerlich, Y. Lee, J. Chem. Phys., 75, 2153 (1981). K.M. Ervin, P.B. Armentrout, J. Chem. Phys., 83, 166 (1985). J .H. Futrell, C.D. Miller, Rev. Sci. Instrum., 37, 1521 (1966). PB. Armentrout, R.V. Hodges, J .L. Beauchamp, J. Chem. Phys., 66, 4683 (1977). RB. Armentrout, R.V. Hodges, J.L. Beauchamp, J. Am. Chem. Soc., 99, 3162 (1977). R.V. Hodges, P.B. Armentrout, J.L. Beauchamp, Int. J. Mass Spectrom. Ion Phys., 29, 375 (1979). RB. Armentrout, J .L. Beauchamp, Chem. Phys., 48, 315 (1980). J .L. Elkind, P.B. Armentrout, J. Am. Chem. Soc., 108, 2765 (1986). Chapter 2 Instrumentation ************************************************************************ A schematic diagram of the MSU ion beam instrument is shown in Figure 2.1. The primary ion beam is generated by using a saddle-field electron-impact ionization (EI) source and a magnetic sector from a VARIAN MAT CH5-DF double-focusing mass spectrometer. Ions are generated within the ion volume, extracted from the source, accelerated, and focused into the magnetic sector to be filtered according to their momentum. The mass-selected ions are then injected, with a desired kinetic energy, into a radio-frequency (rf) only octopole. The ions drift into a collision cell, through which the octopole passes, and collide with neutral reagent gas molecules. The product ions and transmitted reactant ions are then focused into a quadrupole, mass filtered, and detected using a scintillation ion detector (Daly detector) employing ion counting techniques. 17 Samoa osofionom Hannah—mom Soom :3 Ad 93mg oouoom :3 rlu is s is uSoSmQ Shana. ”86:30 $2 b0 mafia /E_.IPI____III\/___ L__||=_.H./_H_ Eel... J.LE. Q: 0%me houooeom mango ban usage—coon l______\ 19 II. Ion Source A schematic of the source is shown in the top panel of Figure 2.2. Electrons are boiled off a tungsten filament (0.007 in. in diameter) and accelerated to a kinetic energy of 70 electron-volts (eV). The electrons enter the source region via a small slit, pass through the source, exit through a slit on the opposite side of the source, and impinge on a collector. To calibrate the electron energy, appearance potential experiments using nitrogen as the standard were performed. This is described in Appendix A. When a neutral molecule interacts with a high energy electron, absorbs energy, and is ionized, the process is believed to follow the Franck-Condon principle. Since the nuclei are so much more massive than electrons, electronic transitions occur on a much shorter time scale than the response of the nuclei. This means that the electrons are promoted to higher energy states while the nuclei maintain their position relative to one another. Consider the potential energy curves in Figure 2.3. Quantum mechanics allows us to calculate the energy level of each vibrational state. The form of the vibrational wavefunction tells us the most probable distance between the nuclei, Yprob» by the position of the maximum amplitude of the wavefunction. The quantum mechanical version of the Franck-Condon principle tells us that the most probable electronic transitions are those transitions between vibrational states represented by wavefunctions that have maximum amplitudes occurring at approximately the same Ypmb. This means that electronic transitions are most likely to occur when the nuclei are separated by Ypmb. This type of transition is represented by the electron collector pusher W ionization box / filament M / —cI—- - - -- Ie< gas inlet W ionization region I x\ N\\ extraction plates///-——/—— —-fi shield plates / —J |__\\ z 1-deflectiorw? H9 II II lens plates r-def'lection plates /— m z 2-def'lection plates source slit a) section in z-direction b) section in r-direction U) 3 .‘3 o. c: O s a: ‘3 J: b s 5‘, Ionization Region 9. 1kV 990 V 990 V Figure 2.2. Schematic and Voltage Diagram of Intensitron EI Source 21 electronic energy level 1 - / \ X ‘7 vibrational energy _p‘ J7 levels A ‘ I E electronic energy level I 1 1 I j 7 j\ [I . . t\ \\ I; 1‘ 11 inbrlational energy Figure 2.3. Illustration of a vertical transition. 22 vertical line in Figure 2.3 and illustrates the origin of the term vertical transition. Other transitions can and do occur, but to a lesser extent. The consequence of vertical transitions is that the new electronic population densities, as a result of the electron-neutral interaction, cannot be modeled by a Maxwell-Boltzman distribution because the new population densities are more of a function of the shape of the vibrational wavefunctions rather than a simple distribution based strictly on the energetic separation of the vibrational and electronic states. Without the ability to determine the population densities of the ion after ionization we cannot predict the amount of internal energy it possesses. The lower panel of Figure 2.2 shows a voltage diagram of how the source is operated. During an experiment, the acceleration voltage would be set to approximately 1000 V. This means that 1000 V is applied to the ionization box. An adjustable resistive divider network within the instrument electronics (located in the BFD module of the CH5 instrument control panel) would then apply voltages on the extraction and focusing lenses of the source. Each lens is wired to a resistive potentiometer and can be individually adjusted (within a specified range determined by the divider network) to allow for optimal conditions. The pusher voltage is adjustable to allow the operator to select any value within the range from approximately 20 V below the acceleration voltage up to the acceleration voltage. Setting the pusher voltage to any value below the acceleration voltage would create a potential surface within the source that resembles a saddle. This is illustrated in Figure 2.4 when the acceleration voltage is set to 1000 V, the pusher: 990 V, and the extraction plates: 990 V. .> H .3 wSouoqom .8523 one 35—13583“:an .ooBom Hm coufimooufi 2: mo ESE ZOHEw .vfi 9:63 24 By creating a saddle-field within the source, the operator decreases the kinetic energy spread of the ion beam. Consider the voltage diagram in Figure 2.2. The beam of electrons within the source has a finite width. Neutral molecules that drift into this beam have an increased probability of being ionized. If the voltages are set to create a field in the source such that the maximum potential occurs within the electron beam and decreases as the field approaches both the exit slit and the back of the source, then only half of the ions created would be extracted from the source. This decreases the signal but, according to the product literature, the source is designed to produce ions with an initial kinetic energy spread of only approximately 0.3 eV. This has been verified experimentally (experiments to be presented later), using retarding potential field measurements with the rf-only octopole. The control circuitry for the electron filament is housed in the BER module. The nominal electron voltage is 70 volts. The emission of electrons is monitored by the collector located on the opposite side of the ion source. Electron currents are maintained by feedback circuits and can be adjusted from < 1 uA up to 1 mA. The module also contains controls which allow the operator to set the source to the electron energy mode. This mode allows the electron energy to be adjusted while maintaining a constant electron current of 15 [1A. The current is set to prevent space-charge burnout of the filament at low electron energies. 25 III. Magnetic and Electric Sectors The original MAT CH5-DF double-focusing mass spectrometer deployed a magnetic and electric sector configured in a reverse geometry. Reverse geometry means that the electric sector follows the magnetic sector. In this arrangement, ions are accelerated out of the source and focused into the magnetic sector. The magnetic field, of strength B, acts on the ions with a force perpendicular to their trajectory, causing them to follow a circular path with a radius, r, given by r = mv/Bez where m is the mass of the ion, v its velocity, e is the fundamental charge, 1.60219 x 10'19 C, and z is the number of charges. As seen from the equation, the radius of the circular path is dependent on both the mass of the ion and its velocity. Because the product of mass and velocity is momentum, ions are said to be dispersed according to their momenta, not just their m/z value. To a first approximation, the ion beam, as it exits the magnetic sector, is spatially dispersed for any given velocity (due to different mass ions) and diverging for any given mass (due to slight variations in velocity). This is illustrated in Figure 2.5. The image curve shown in Figure 2.5 is a curve along which ions with the same mass and velocity are brought into focus. The electric sector consists of two parallel curved plates between which an electric field, E, is applied. The ions pass between these plates and follow a curved trajectory with radius r, given by Velocity Dispersion +| Mass Dispersion +| F F 1 szFbl 82 8 Image Curve Ion beam Figure 2.5. Dispersion of the ion beam by the magnetic sector. Fall is the focal point for ions with mass a and velocity 1. F32 is the focal point for ions with mass a and velocity 2. Fb1 and sz are the focal points for ions with mass b and velocities 1 and 2. 27 r = 2(Kinetic Energy)/E The electric sector performs two functions. It acts both as a focusing device for diverging ion beams and as a dispersing device according to the kinetic energy of the ions [1]. Ions coming from a point source with a small angle of divergence enter the electric sector and are brought into focus and follow the same approximate radius, r. The sector acts as an energy filter in that ions with higher and lower kinetic energies are brought into focus at points other than the exit slit, and thereby are not allowed to pass through the device. One of the goals of the experiments performed on this instrument is to control the kinetic energy of the primary ion beam as precisely as possible. A double focusing mass spectrometer is designed to focus isomass ions to one point (usually the detector). Beyond that point, the beam is diverging and can have a significant kinetic energy spread. The result is a non- collimated beam of reagent ions that is inhomogeneous in energy. By removing the electric sector, the kinetic energy spread of the beam is decreased, the beam is more collimated and, as an added bonus, the ion beam current is increased due to the removal of a device with a less than perfect transmission coefficient. Experimental measurements of the kinetic energy spread, to be explained in later sections, before and after the removal of the electric sector have shown that the energy spread of the ion beam has been decreased from 0.31 i 0.02 eV to 0.25 i 0.02 eV. Quantitative measurements of the ion beam current did not show any marked improvement. Differences in ion currents were attributed more to B generating different numbers of ions in the source rather than the transmission through the electric sector. Operation of the magnet is controlled by circuits housed in the BYA module. This module allows the operator to manually scan the magnetic field (required for optimizing ion current) or to hold the magnet to a specific value during ion/molecule experiments. The BYD module allows the operator to initiate an electronically controlled scan of the magnet, where the rate at which the magnet is scanned is determined by the operator. The magnetic field strength is measured using a Hall effect probe. Spectra are recorded by a strip-chart recorder linked to the Hall probe. IV. Deceleration Optics The MSU ion beam instrument has been designed to study gas-phase ion/molecule chemistry and collision induced dissociation cross-sections as a function of translational energy. Because the translational energy must be accurately determined, the ion deceleration optics comprise one of the most critical components 0f the instrument. In order to establish a well-defined interaction energy between the ion and the neutral reagent particle, several criteria for the deceleration system have to be satisfied. Most importantly, the deceleration system must decelerate the ions from approximately 1000 eV to <2-100 eV in the laboratory frame. Z) The system should provide good focusing (small ion beam diameter), convert the beam symmetry from a rectangular shape (after the magnetic sector) to a circular shape to match the axially symmetric octopole and quadrupole sections of the instrument, and produce a highly-collimated beam nearly parallel to the main axis of motion. Satisfaction of these criteria help to minimize the effect rf-only multipole devices have on the transverse energy of ions as the ions pass through the oscillating field. A highly collimated beam also helps to decrease ion losses during collision processes [2]. A small beam diameter ensures the highest transmission of the primary ion beam through the multipole device. When employing ion optics, aberrations in the focal properties of the lenses can be the result of small imperfections in the construction and alignment of the various components. It is, therefore, advantageous to limit the number of elements in the deceleration assembly. A schematic diagram of the ion deceleration optics is shown in Figure 2.6. Details concerning construction and characterization can be found elsewhere [3,4]. A brief summation of the conclusions from that work is given below. To convert the beam symmetry from a rectangular shape to a circular profile, circular apertures were employed. Electrical potentials placed on circular apertures generate electric fields that exhibit steep gradients perpendicular to the central axis. These gradients result in over-focusing (high angular divergence) of the ion beam and is more pronounced as the initial beam diameter approaches the electrode aperture diameter. Einzel Focusing Deceleration Alignment Stage Stage Stage r'I I-'—I I ' l - I ”L. Guard Guard Cyl/inder Cylinder I" I l |\ Injection Element Lens R L\ens Z > Ion trajectory Figure 2.6. Schematic of Deceleration Lens System 31 Therefore, the most significant criterion in determining the performance of a deceleration lens system is the ratio of the initial beam diameter to the electrode aperture diameter. In order to obtain the best performance, the beam must be constrained to the central axis of the lens assembly and employ lenses with large apertures relative to the ion beam cross-sectional diameter. Monitoring and control of the applied voltages of the deceleration lens system are implemented through an in-house built power supply mounted on the CH5 electronic module frame. Each lens is independently controlled to provide maximum optimization. A description of each lens element is provided in Appendix B. V . Collision Region A. Beam Guide As the reagent ions drift into the collision cell they undergo collisions with neutral gas-phase atoms/molecules, resulting in the scattering of reactant and product ions. For this reason, the beam guide must provide highly efficient trapping to collect and focus the ions into the second mass analyzer. Another criterion the beam guide must satisfy is that it must have no or minimal mass discrimination effects. When the ions and neutral atoms/molecules react, many different products may be generated, representing a wide range of masses. To make quantitative measurements, all of the resulting ionic species must be efficiently collected and detected. And finally, in order to control the energy of interaction with precision and 32 accuracy, the trapping process generated by the beam guide must not perturb the initial ion kinetic energy distribution. Szabo and Hagg published a theoretical perspective on oscillatory electric field multipole devices and their comparative study of quadrupole, hexapole, and octopole ion traps concluded that the higher order multipole devices were superior as rf-only beam guides [5-8]. Radio frequency potentials are applied to the rods of the multipole device such that the potentials of adjacent rods are 180° out of phase. These potentials generate a dynamic heterogeneous electric field that creates a potential well in the center of the volume encircled by the rods. Ions entering the device are then acted upon by a potential restoring force, Ueff, towards the center. The relationship between Ueff and the radial distance from the center, r, is shown below nzqzvz r “‘2 U = __o _ 9" [4mw2rfi ][ ro ] where 2n is the number of poles, q is the charge of the ion, In is the mass of the ion, and r0 is the inner radius of the device [9]. The rf potential applied to the rods is Vocosz(wt). From the above expression, the potential restoring force can be plotted against the radial distance. The plot is shown in Figure 2.7. It can be seen from the plot that the higher order multipole devices generate potential wells that are broader near the minimum and have steeper slopes near the extremes. This has an advantage in that as the ions enter the beam guide, those ions that enter off-center experience a smaller restoring potential _ a 0. 5 4. A22 x .3 _a 1 7m ESP— moans.— 1 h. 0.0 -0.20 0.00 0.20 0.40 0.60 Relative Radial Distance (r/ro) p.40 -O.6O Figure 2.7. Comparison of restoring forces for a quadrupole (Quad), hexapole (Hexa), and an octopole (Oct). 34 force. This in turn means there is, on average, less transverse kinetic energy imparted to the ions as they enter higher order multipole devices. Another advantage of the higher order multipole devices is that they impart less transverse energy to the ions as the ions drift through the beam guide. The longitudinal kinetic energy is not affected. However, the rf field does impose a time-dependent transverse energy component on the ions perpendicular to the ions' trajectories causing the ions to oscillate as they drift through the device. This additional energy must be taken into consideration when measuring the total kinetic energy of interaction between the ions and the neutral reagent gas. Various factors that influence the transverse kinetic energy for both quadrupole and octopole beam guides were studied by O'Connor [3]. Using SIMION, an ion trajectory modeling program, the relative effects of rf phase, initial position, and initial angle of entry were investigated. The assigned variables for the study were: 20 ms residence time, ion m/z value of 100, initial kinetic energy near thermal (0.02 eV), initial position of 0.2 r0, and both beams guides operating at an rf amplitude of 300 volts with a frequency of 1 MHz. The transverse kinetic energy was plotted as a function of time shown in Figure 2.8. From the results, O'Connor concluded that quadrupole beam guides would introduce high levels of uncertainty in measured ion/neutral interaction energies. Also, O'Connor found that even though the initial rf phase relative to the point of ion injection was of little importance for the octopole O 00‘0000‘... .0... . ..... ..o- — A V v _ A --_..' A “‘... A- : ' fl - v emu-XI: _ A - ' .‘ H -‘C A “v L -- --_ 3 i Wan-z : . - --_ _ ::-===--:-c.=:”..;: { O 0.0... eoeeeth- - '- - ::."_ __v “30‘ 2 “$1.112- 3 A - A ‘ ii 14 m‘$-.'.::: ;' A A - - i“: - - - 'A ~ A ‘ iI> (I. A““‘4 It A A— A "‘1 ......oo.oeooo .. 12 Time in Multtpoic (ms) «“9" .00....OO‘.... o .l 10 A -A 5 )Il it Ib ‘9 “0:” : “Que 4 8 Octopole KE. .......-- Quadrupole ICE. A-‘ w Wo...oo.oooooeoo---- 0......0.0.0....o.-—=::=____ l, l I -A.A mllz:: 'A _ ______ - A v- v — v ' coo-3:3.....‘ @ -00.“..’ go M n H H 1 lb jij'ijvj'VTTV‘1‘Vf'W"V‘VVTV'YV1'1'VT'17“ a a v N O. “2 ‘9. ‘1 N. O. ' ' .4 -i .- O _O O O 0 (Ac) 16.13113 onaum asxaasuen, ‘02::00000 _ --:_--- A--‘.;e “1332........__‘ ' A':‘-"-::A VVVV v ‘ --_ ----_- _:...O.ooo....o.oo..-.o..oo.ooo .‘z. ”rag-:5 — :3 - ----- _ "WA :A A A i .e::.....:.. :3:-. -_-- —:-- v- 1 A _ A ._.‘. ‘ ...... . .e 1 N .....¢oo-O." ‘Coe-e..-_: - ‘ ‘A ‘ C Figure 2.8. Calculated radial ion kinetic energy in if multipole [3]. 36 device, the rf phase had an extreme effect on the transverse kinetic energy in the quadrupole beam guides. The octopole beam guide is operated by using a WAVETEK HF Sweep Generator (model # 144) to generate the primary rf signal. The resonant frequency of the octopole system is approximately 2.5 MHz. The rf signal is fed into an ENI Model A-150 power amplifier, amplified, and then fed into a modified Extrel Model E High-Q head. The High-Q head acts as a transformer supplying the rods of the beam guide with the amplified rf signal. The rod assembly of the octopole and the secondary coil of the High—Q head act as an inductor (L, the coil) coupled in series to a capacitor (C, the rod assembly). This arrangement is referred to as an LC tank circuit. If the circuit is operated under the correct resonance frequency conditions, power is alternately stored in the capacitor and the magnetic field of the inductor. Due to practical finite resistance of the circuitry, resistive losses occur which decrease the capacitor's ability to charge and discharge. This loss in power is compensated for by applying additional power to the system. Because an LC tank circuit operates within a specific frequency range, its ability to absorb power is frequency dependent. If the rf power being supplied to the circuit is not absorbed, the signal is reflected back to the signal source. To maximize the forward power and minimize the reflected power, a radio antenna tuner has been inserted in the line between the power amplifier and the primary coil of the High—Q head. The tuner (Heathkit Model 37 HFT-9A) consists of a variable inductor and two variable capacitors. By adjusting each component, a balance between the amplifier and High-Q head can be achieved. The relationship between the forward and reflected power is quantitated as the standing wave ratio (SWR). A perfect balance is reflected by a value of unity. To monitor the forward power and the SWR, a wattmeter (Heathkit Model HM-9 QRP) has been placed in the line between the power amplifier and the antenna tuner. The DC bias of the octopole (which determines the nominal interaction energy in the laboratory frame) is applied through a center tap on the secondary coil of the High-Q head. A floating power supply (referenced to the acceleration voltage) is used to generate the offset potential. The range of the power supply is O to -500 volts and it is programmed using a 0-5 volt driver which acts to insulate the floating power supply from the operator's computer. The potential of the driver is generated using a 12 bit digital-to- analog converter (DAC). The 12 bit DAC allows for 4096 different voltage settings. This corresponds to an increment of 0.1221 volts in the nominal ion/molecule interaction energy. The uncertainty in the interaction energy is therefore approximately i0.061 eV in the laboratory frame. B. Collision Cell To ensure efficient collection of products and transmitted reagent ions, the collision cell was constructed from a 7.62 cm long, 5.1 cm ID stainless steel cylinder placed over the central region of the beam guide. To limit the 38 conductance between the collision cell interior and the main chamber, an extension tube with an ID of 2.22 cm and length of 6.35 cm placed at each end of the collision cell. The ID of the extension tubes is slightly larger than the OD of the octopole beam guide. A schematic representation of the collision cell and beam guide is shown in Figure 2.9. The most significant contributions to the uncertainty of measured absolute reaction cross sections are the determination of the density of the neutral reagent gas within the collision cell and of the effective collision cell length [10]. Pressure within the collision cell was initially monitored by an MKS Baratron 310 capacitance manometer. The capacitance manometer has several advantages. First, a capacitance manometer is a direct-reading gauge. It measures pressure by calculating the force exerted on a diaphragm by the incident gas. This is a physical measurement in that the gauge monitors the degree of deflection of the diaphragm by observing the change in capacitance between the diaphragm and a fixed counter electrode. The advantage is that the pressure measurement does not rely on any individual property of the atoms/molecules (e.g.. ionization energy) that comprise the gas. Therefore, response is independent of the chemical nature of the gas. Another advantage of this manometer is that it does not require any net throughput of gas particles. The reference side of the gauge is sealed, thereby terminating the tube leading from the collision cell to the manometer. Also, the gauge does not consume any of the gas during the pressure measurement, which would require a net flux of gas particles. This means doc 5658 98 2338 23 .3 89%va oumfionum .md 98mg 925 mam manna. commcmaxm mvom £3390 _’ _ \ ‘7 acom :oU :omeoO $552 3535 (10: bass 40 the system reaches a steady-state whereby the pressures in the collision cell, the tube, and the manometer are equal. The flow of gas through a tubular-shaped object is a function of the pressure dr0p, AP, between the ends of the object as well as the item's geometry. Division of the throughput, Q, by the pressure drop across the object held at constant temperature yields a property known as the intrinsic conductance, C, of the object (C = Q/AP) [11]. Rearrangement of the equation gives the following relationship: Q/C=AP Because Q is zero and capacitance is zero only when the tube length approaches infinity, the difference in pressure between the two volumes must be zero. This means that the pressure measured at the gauge is equal to the pressure in the collision cell. The major disadvantage in using a manometer, however, is its lower pressure limit of 0.1 mtorr. Furthermore, the uncertainty in the pressure measurement near 0.1 mtorr may be greater than i25%, even with the best capacitance manometer currently available. In order to obtain single collision conditions during the experiments, the pressure within the collision cell must be no greater than 0.1 mtorr. In order to obtain more precise and accurate pressure measurements, a hot cathode ionization gauge was used. The operation of the ion gauge is based on the ionization of gas particles by electron impact and the 41 subsequent collection of the ions by an ion collector. The response of this type of pressure gauge is dependent on the species of gas present because of the ionization process. Pressures due to gas particles with higher ionization energies would result in lower pressure readings, because less of the gas population would be ionized. To improve the accuracy of the ionization gauge, the capacitance manometer was used to calibrate the ion gauge for each neutral reagent gas. An example of a calibration curve for argon is shown in Figure 2.10. Working curves, ionization gauge pressure readings verses capacitance manometer pressure readings, were generated under molecular flow conditions. Only during molecular flow conditions will the pressure drop across the extension tubes be linear. Knudsen's number, Kn, is used to determine the nature of a gas; it is a dimensionless ratio between the mean free path of a particle and a characteristic dimension of the instrumentation [11]. In this case, the characteristic dimension is the diameter of the extension tubes of the collision cell. Gas flow is termed molecular when Kn > 1.0. In order to maintain molecular flow conditions in the experimental apparatus for argon, the pressure in the collision cell must not exceed 2.36 mtorr. With working curves like Figure 2.10, the pressure reading from the ionization gauge can be used to measure the pressure within the collision cell. Using the ionization gauge with the corresponding working curve allows pressures at or below 0.1 mtorr to be monitored with greater precision and accuracy than with the capacitance manometer. downs “8 023 855:8 own—Wm 5:23:0— .o~ .N 8amE E3 we: 888252 mogmoaaao odm owfi 0.2 0.3 0.2 o.o_ o.» o6 o.v ON 06 _J W _ _ _ a n a _ d 0.9 x 0N O. V E II 8.5 x I Q 0 I o. no r o.- l O. v '— segmmma + x Eggs u a :3 S (.110; [’01) 33mg) uoynzguol l o. \O o— 1 o.w~ r odN 43 VI. Ion Optics Ions are extracted from the exit aperture of the octopole beam guide and focused into an Extrel quadrupole mass filter for subsequent mass analysis. The extraction and focusing of the ions are performed by transfer optics consisting of a stack of 5 independently controlled electrodes. Voltages applied to the optics are generated by a Spellman high voltage DC supply (model # UHR10P100) and a resistive divider network. Figure 2.11 shows the geometry of the lens system. SIMION modeling has shown that this lens configuration can provide excellent ion transmission over a broad energy range [3]. Extraction optics have been developed to extract ions from the quadrupole and direct them to either of the 2 detectors. Voltages applied to this set of optics are generated from the same voltage supply and divider network employed for the transfer optics. These optics are a modified einzel lens design and are also shown in Figure 2.11. VII. Quadrupole The quadrupole mass filter is a high transmission Extrel (model #4-324-9) combined with a Model 15 High-Q head. The device has an upper m/z limit of approximately 200 and the High-Q head has been modified to provide electric isolation at DC-bias potentials up to 5 kV. Principles of quadrupole mass filters are well documented in the literature [12]. The operating manual provides an excellent resource for theory, electronics, and 839$ 38989 on» can mow—no ~833be can Emma—duh. mo oflmfionom Afim 93mg omega 568280 m / \ a. a ‘ m 5.222.530 _ I: ..I I l I l Sam—550m '— — £3300 .. _ L 359 k.\ m ,5 anemomaG no a _uo mam." swam 38m \ FA 3 o a 9 Ba. 8:3 cozfixdfioaoam 2°53be 4.5 operation of the mass filter. The quadrupole mass filter is an ideal secondary mass analyzer for ion beam experiments because it performs mass discrimination independent of the ion energy. Although the DC offset voltage of the quadrupole is manually set, the rf amplitude and DC potentials associated with mass filtering are controlled through the quadrupole electronics. These control electronics are programmable and a second 12 bit DAC is used to drive the quadrupole. The setup allows the m/z value to be set to approximately :01 units. VIII. DetectorSystem Ions may be detected by either a continuous dynode electron multiplier or a Daly detector. Both detectors are shown in Figure 2.11. The continuous dynode detector consists of a Galileo 4800 series Channeltron detector with the input horn placed coaxial with the ion optical axis. The detector is normally operated in an analog current mode for measurement of relatively high ion currents and is used primarily for instrument diagnostics. When used with a picoameter, this detector can quantitate ion currents from 10'17 to 10'10 amperes. This range corresponds to count rates of approximately 100 ions/second to 109 ions/second. To quantitate lower ions currents (those typically seen during threshold measurements), ion counting techniques must be employed. The Channeltron is not an ideal detector for these measurements because it suffers from poor initial conversion efficiencies and mass discrimination. To perform ion counting techniques, a high efficiency Daly detector has 46 been developed. Concepts, operation, and characterization of this detector have been described elsewhere [13]. This design incorporates a conversion dynode constructed from a 1.25" diameter stainless steel disk with a thickness of 0.25". The surface of the dynode has been polished optically flat by using a 1 pm diamond polishing compound on a polishing wheel. Voltages applied to the dynode are approximately -27000 V. Ions exiting the quadrupole are accelerated by the electric field generated by the negative potential, impact the conversion dynode, and sputter secondary electrons off the surface. The electrons are accelerated through the same field to a plastic scintillator disk that has been coated by a 3 A thick aluminum film. The scintillator is optically coupled to a Hamamatsu R-425 photomultiplier tube (PMT). These high energy electrons stimulate photo emission in the scintillator which is detected by the PMT. In response to a single ion event, the PMT generates a negative pulse (<50 ns in duration). The signal from the PMT is amplified by an MIT F-lOOT amplifier with an integral discriminator circuit. Because of the high degree of polishing of the conversion dynode, the we were able to set the discriminator to its lowest setting. The amplifier generates a TTL pulse when the amplitude of the input signal exceeds the level set by the discriminator. These TTL pulses are then detected by a six digit ORTEC 775 pulse counter. The counter is digitally interfaced to a computer which controls the start and stops count states. Dark counts of this detection system are dominated by the field emission of electrons from the dynode and are typically <20 counts/sec when standard potentials of -27000 V to the dynode and 1000 V to the PMT are applied. 47 IX. Computer System The ion/molecule interaction energy, quadrupole mass filter setting, and the detector signal are all controlled/monitored using a computer system consisting of 10 MHz 80286 PC running MS DOS. Two IBM DACA interface boards (DACs) are controlled through a custom program [14] written in the ASYST 2.0 computer language. 3“ S°9°>39°Sn 11. 12. 13. 14. LIST OF REFERENCES Chapter 2 R.G. Cooks, J .H. Beynon, R.M. Caprioli, and GR. Lester, "Metastable Ions." Elsevier, Amsterdam (1973). JB. Hasted In Atomic and Molecular Processes; Vol. 13, DR. Bates, Ed.; Academic Press: New York, p. 696-720 (1962). P.J. O'Connor, Ph.D. Dissertation, Michigan State University (1991). P.J. O'Connor, GE. Leroi, J. Allison, J. Am. Soc. Mass Spectrom., 2, 322 (1991). I. Szabo, Int. J. Mass Spectrom. and Ion Proc., 73, 197 ( 1986). C. Hagg, I. Szabo, Int. J. Mass Spectrom. and Ion Proc., 73, 237 (1986). C. Hagg, I. Szabo, Int. J. Mass Spectrom. and Ian Proc., 73, 277 (1986). C. Hagg, I. Szabo, Int. J. Mass Spectrom. and Ian Proc., 73, 295 (1986). L. Landau, E. Lifschitz, "Mechanics." Permagon Press, Oxford (1960). K.M. Ervin, P.B. Armentrout, J. Chem. Phys., 83, 166 (1985). O'Hanlon, J .F., "A User's Guide to Vacuum Technology", 2nd ed., John Wiley & Sons, New York (1989). P. Dawson, Mass Spec. Rev., 5, 1 (1986). LA. Romero; M.S. Thesis, Michigan State Unversity (1989). The majority of the ASYST program was written by Scott Kuiphoff, an MSU senior majoring in computer science. Chapter 3 Experimental Section *****************************************5|!****************************** I . Initial Mass Spectrum As mentioned earlier, the field generated in the magnetic sector, of strength B, acts on an ion (of mass m and velocity v) with a force perpendicular to the ion's trajectory, forcing the ion to follow a circular path with a radius, r, given by: r = mv/Bez If we also consider the kinetic energy of the ion as the result of the acceleration voltage, V, the standard expression for the separation of ions by a magnetic sector is given as: m/ze = r2B2/2V From this expression it can be seen that a linear scan of B is proportional to ml/z. Therefore, as the magnet is scanned, peaks representing adjacent masses appear closer together and begin to overlap at higher masses. This means that it is important to optimize the source and magnetic sector over 49 50 the desired mass range prior to performing energy-resolved experiments. When the chemistry of an ion with a mass near the upper limit of the quadrupole is studied, the resolution of the magnetic sector must be increased to completely separate ions of adjacent masses. This decreases the ambiguity of the source of a product ion mass which could be due to isotopes of the reactant ion or to the presence of a hydrogen atom. When studying the chemistry of an ion with a lower mass, the resolution can be decreased to help increase the ion current while still completely separating ions of consecutive m/z values. I I . Initial Identification of Pmducts After a chemical system has been chosen for study, a series of experiments are performed to identify all products. When the beam instrument is used to study gas-phase reactions, three types of products can be identified: metastable, exothermic, and endothermic products. It is important to identify the nature of each product in order to determine the optimum lens [potentials required to monitor them. The experiments to identify the products begin by setting the octopole 5 volts below the acceleration voltage and scanning the quadrupole over its entire mass range using 1 m/z steps. Because the products are not known and therefore the instrumentation is not optimized for each product, the resolution of the quadrupole is decreased in order to increase the product ion current. After the spectrum is taken, the octopole is dropped another 5 volts and the quadrupole is again scanned. This is repeated until the octopole voltage has been dropped a total of 50 volts. 51 The spectra are then studied and possible product peaks are identified. To determine whether more experiments should be performed with the octopole set to lower voltages, the most intense peak at the lowest voltage is identified. If this is a collision-induced dissociation (CID) product generated from the reactant ion, then this indicates the energetic limit of the chemistry being studied. This will be explained in a later section. For each individual product, the octopole potential is then set to the voltage that generated the most intense peak for that product. The instrumentation is optimized and the resolution of the quadrupole is set to its maximum (approximately unit resolution). The quadrupole is then scanned over a range of 10 mass units, centered over the m/z value where the peak first appeared, using 0.1 m/z steps (the smallest possible step). Each quadrupole scan is done in triplicate. After all spectra are collected, each product is assigned an m/z value. It is these m/z values that are monitored during the energy-resolved experiments. III. Optimizing Conditions To begin the experiment, the instrumentation is set to monitor either the exothermic or endothermic products. Because optimization is very different for the two types of products, reliable energy dependence curves can only be generated one type at a time, not simultaneously. Optimization is different because the products have maximum ion currents at very different kinetic energies. 52 The appearance of the exothermic products coincides with the appearance of the reactant ion without any offset as the voltage on the octopole is stepped to lower voltages. The exothermic product signals then decrease rapidly as the energy of the system is increased. This means that the largest product ion currents occur at very low kinetic energies. Optimization must therefore be performed very near zero kinetic energy in the center-of-mass frame. Endothermic products begin to appear later during the experiment and therefore optimization for these products must be performed at voltages reached at the end of the experiment. This is in agreement with observations from other groups performing similar experiments [1]. There are three sets of lenses to be optimized: the lenses of the ionization source, the deceleration lenses between the magnet and octopole, and the transfer optics between the octopole and quadrupole. Voltages on the source and deceleration lenses are used to optimize the ion current of the reactant ion and the settings are independent of the kinetic energy range over which the experiment is conducted. The deceleration lenses have been designed to focus an impinging ion beam over a wide energy range, 0 - 100 Volts [2]. This is sufficient for all experiments performed on the instrument. The transfer optics between the octopole and quadrupole are used to optimize the ion currents of the products ions. Because these lenses are not computer controlled, the voltages on the lenses may be optimized simultaneously for product ions with masses within 20% of each other. 53 Product ions with a greater difference in mass must be monitored over two separate experiments. IV. Performing the Energy-Resolved Experiment For each experiment, a set of three types of scans must be made: the energy-dependent chemistry experiment and two supporting scans. The first‘scan performed is a stopping-potential scan. This scan is used to determine the point of zero kinetic energy and the kinetic energy spread of the reactant ion beam. The second supporting scan is the background scan. This scan is used to compensate for collisions that occur outside the collision cell during the main experiment. Each will be explained in the following section. A. Stopping-Potential Scan As mentioned earlier, a stopping-potential curve is generated to determine the point of zero kinetic energy and the kinetic energy spread of the reactant ion beam. Because the overall purpose of the experiments is to measure the energy of interaction between the reactant ion and the neutral particle, it is important to know the zero point from which all other kinetic energy values are determined. The experiment is performed by generating the ions of interest and injecting them into the beam guide. Initially, the octopole DC-bias voltage is set equal to the acceleration voltage. This prevents any ions from drifting through the beam guide. As the experiment is conducted, the bias voltage 54 placed on the octopole is decreased. This is schematically shown in Figure 3.1. The signal (counts/second) due to the ion beam first appears, increases as the oct0pole voltage continues to decrease, and finally levels off to a constant value. The ion intensity is then plotted as a function of the difference in electric potential between the ion source and the octopole DC- bias (Vsource-Voctopole). No gas is introduced into the collision cell and the quadrupole is operated in the rf-only mode. By using the octopole as a high- efliciency retarding field analyzer, the interaction region and the retarding region are physically the same. This has two advantages in that there are no uncertainties introduced into the kinetic energy measurements due to contact potentials or focusing aberrations. An example of a stopping curve is shown in Figure 3.2. B. Energy-Resolved Scan After it has been determined which m/z values must be monitored, the main experiment is performed. This experiment is equivalent to the stopping-potential scan except now a gas is introduced into the collision cell and the quadrupole is operated in the mass filter mode. The gas pressure within the collision cell is maintained at pressures below 0.1 mtorr to ensure single collision conditions. As the octopole DC-bias is stepped down, the ion currents of the different ions are individually monitored. This allows the onsets and reaction cross-sections for the different ionic species to be determined. .aqoatoaxo 23 mo 88»va 3.85 3593.5 Ad 3de LBoSoQ ran Lea—E 334 £32,320 8:90 .30 8.2.95. commmzoo :8 82:8 mam— ..mnbaé mam—z Camacho—moon 0955 Hill 8.532 8:80.:5 augm— mcoq mammaoch um comaoeaxm oocsom / l>xo INC: $318113 [etquaqod o6 .288 3:3: 2: E @8826 2883:: 05 53, C32,: 3.50 3688 9:3on < .~.m EsmE 32 .3 3.8m 2.25. ow oé o.m o.~ o; od “H + w iiiliiT od 1 md I o. F! I “n. F-I T o. N I V? N I C. M 1 m. m I oé (,01 x) oas/smnoo 57 C. Background Scan This scan is performed under the identical conditions under which the main experiment was conducted except for the collision gas. For this scan, the collision gas is diverted to an inlet which feeds into the chamber containing the octopole and collision cell. The gas is no longer being leaked directly to the collision cell. The pressure within the chamber is set equal to the pressure of the chamber (as monitored by an ionization gauge) during which the main experiment was performed. This scan is performed to compensate for any collisions that may occur outside the collision cell during the main experiment as well as any metastable products that may appear. V. Data Analysis A. Determining Ion Beam Energy Spread and Point of Zero Kinetic Energy in the Center-of-Mass Frame The kinetic energy scale used to analyze the data is in the center-of-mass (CM) frame. The data, however, are collected in the laboratory (lab) frame. In order to relate the two frames of reference, the energy distribution of the ion beam and the point of zero kinetic energy have to be determined. From the stopping potential data, a first derivative curve with respect to (Vsomce-Voctopole) is calculated. The resulting curve is Gaussian in shape and, through a least-squares method, a Gaussian curve is fitted to it. An example of a stopping potential curve and the Gaussian fit is shown in 58 Figures 3.3 and 3.4. The position of the apex of the Gaussian curve is then taken as the point of zero kinetic energy. The full width at half-maximum (fwhm) of the curve is taken as the kinetic energy spread of the ion beam. Absolute uncertainties in the energy scale due to the source, deceleration lens, and beam guide have been determined to be less than $0.1 eV lab [2]. B. Doppler Broadening l A larger effect than the absolute uncertainties described above, on the energy scale arises from the thermal motion of the neutral reagent gas (Doppler broadening). To address this problem, Chantry developed a three dimensional treatment of the interaction energy [3]. He began by assuming a monoenergetic ion beam interacting with neutral molecules having an isotropic Maxwellian velocity distribution. The total energy, Etot, available in the center-of-mass frame is defined as: Etot = (1/2)[m / (M + m)]MV2 (1) where V is the relative velocity of the incident ion and neutral reagent with masses M and m, respectively. If the assumption is made that the neutral molecule is a stationary target, the center-of-mass energy becomes: ECM = [m / (M + m)]E]ab (2) Chantry then presented an exact solution to the thermal distribution with the form: .ovoE bcof 2: 5 H3830 28233 05 53, +19 H .8 956 3:88; 9.3an .m.m SawE Gd 1r— - ‘ 32 $3 3:5 2.25— o6 H A o.m o.~ 0.“ od 0.0 md 04 n4 o.~ Wm o.m Wm o6 (901 X) pnoaasmunoo .335 unocc— 83 mo Eon 05 an :83 a 5% 2: can 38% 3.55 05 mm :85 fl 2:3”— BE. dzfizcov “a: 05 9 E 55 33 255 53330 2:. .m.m .wE 5 +10; he case—q 8% 3888 wcaqoa 95 Eat new—=28 33 9:3 393:3. SE 2F in 8:9”— 32 53 Beam 3.25. ll Noel o6 oé o.m o.~ o; TclolclfilIlhio o .p . L . iii . N O 0 “All 0 \ o O o .- No . 35cm £852. m. 2 o :5 A on a . m N. 1. v o m . _W “E 3.50 5:330 LT ..V A >0 R n 212E m 0333qu E cam—330 o l c o W w #1 wd m - . e o; 61 £12.80): (marl/21 exP[ - (elf2 - £0U2)2 1 - exp[ - (2V2 + 801/2)2 1 } de (3) where 'y = M/(m + M), e = Etot/(kaTgas) and, so = ECM/(ykagas). In most situations, when (IQtotEmvfll/2 > (1.15 ykagas). the contribution of the second exponential is less than 0.01% that of the first and therefore can be neglected. Thus, provided ECM is greater than kagas (4.1x10’21 J at 298 K), the solution, to a very good approximation, becomes: ) as, so) =(4neo)-1/2 exp[ - (.el/2 - 801/2)2 1 de (4) This distribution peaks at e = so and has a full width at half-maximum (FWHM) of: FWHM = (11.1 ykagasECMfl/Z (5) In general, if the product ion current is small compared to the incident ion current, the ratio of secondary to primary ion currents will be proportional to the effective cross section, Qem given by the following integral [4]: _, 1/2 8 an = J(—) a(£)f(e,eo)de (6) where the factor (9180)“ 2 is proportional to the effective path length, 0(a) is the cross section of interest, and f(€, £0)de is the distribution function defined by equation (4). 62 C. Incorporating Ion Beam Energy Spread To account for the energy spread of the primary ion beam, Lifshitz et al. expanded on Chantry's [3] work and developed an exact treatment for the distribution [5]. They assumed the initial ion energy spread could be represented by a Gaussian function. Given the nominal beam energy, Elab. the probability of an ion having energy E" is given by: P(E") = algexpl:-(§'&S£) ] (7) where S is related to the energy resolution of the instrument and is determined by the requirement that the full width at half-maximum be 0.3 eV (S = 0.2 eV). Inserting equation (7) into (6) gives the following double integral for the effective cross section: « ‘:;n.exp[-Ew(.- m —S— 0 l2 0 (exp{ _a[En1/2 _E11/2J2}_exp{-a[En1/2 +E'U2 2})£%£dE"dEd/2 (8) where E' = [(m + M)/m]Etot and a = m/(Mkagas). By neglecting the second term [exp(--a(E"1/2 + E'1/2)2], when E131, is large, and expanding E"1/2 of the first term in a Taylor series about E'Uz, equation (8) reduces to the following equation: _ 1 .. M W ‘(Eo'E')2 . . Qm—nm {(82+4E'la) exp SZ+4E'/a 0(E)dE (9) D. Experimental Cross Sections —ln[fI—] _. __°_ (10) 0’ 111 The total reaction cross section, 6, is calculated from the data using equation (10), where n is the number density of the neutral gas, 1 is the effective collision cell length, I is the measured intensity of the transmitted ion beam, and 10 is the sum of the intensities of all the ions I + ZIj where the subscript i refers to the a specific product signal. a, = o :I—‘I— (11) Cross sections for individual product ions, 6:, are calculated using equation (11). Accuracy of the determined cross sections are limited primarily by our ability to measure the neutral reagent gas density and to estimate the effective collision cell length. Uncertainties are increased in the determined cross section values near the threshold of an endothermic process due to the relatively small ion currents. This is primarily attributed to random counting noise (typically 5-10 counts/sec). Other sources of the background signal are reactions occurring outside the collision cell. These contributions are determined by subtracting the background scan from the corresponding main experiment. The result of the subtraction yields the signal of products generated within the collision cell. Random scatter in the 64 data results in some negative values for cross sections near the threshold. These are truncated to zero. As stated before, the diffusion of the collision gas cloud into the vacuum chamber outside of the literal 7.62 cm by 5.1 cm ID stainless steel collision cell results in uncertainties in the length of the effective collision cell. Althovgh the background subtraction is used to help compensate for reactions that occur outside of the effective collision cell, the exact dimensions of the effective collision cell are still undetermined. Therefore, relative cross sections (G/Gmax) were calculated as opposed to absolute cross sections. E. Fitting the Data After the experimental curve (relative cross section vs. kinetic energy (CM)) has been determined (using equations (10) and (11)), data analysis continues by solving equation (9) and comparing the theoretical curve to the experimental curve. Adjustments to the parameters that define C(E') of equation (9) are made through iterations involving least squares fitting which causes the theoretical curve to converge to the experimental curve. The best 6(E') function obtained using equation (9) is then used as the starting function for equation (8), which is then further optimized. The functional form used for 0(E') is: (12) 65 where ET is the desired threshold energy, 11 and m are adjustable parameters and A is a scaling factor. All four parameters, ET, n, m, and A, are adjusted to obtain the best fit. For the systems studied during our experiments, the best fits were obtained when the variable m was held equal to 1 and n allowed to vary between 1 and 2. All fits where carried out through the use of a software program "Crunch" given to our lab by Professor Peter Armentrout and his group [6]. 9‘99”.” LIST OF REFERENCES Chapter 3 LS. Sunderlin, personal communication. P.J. O'Connor, Ph.D. Dissertation, Michigan State University, ( 1991). P.J. Chantry, J. Chem. Phys., 55, 2746 (1971). H. Pauly, J .P. Toennies, Methods Exptl. Phys., 7A, 283 (1968). C. Lifshitz, R.L.C. Wu, T.O. Tiernan, D.T. Terwilliger, J. Chem. Phys., 68, 247 (1978). The "Crunch" program was written and continues to be modified by the Armentrout group of the University of Utah. Chapter 4 Instrument Evaluation ************************1"!III*****JIM!********III***************************** Two experiments have been performed to evaluate the MSU ion beam instrument. The first experiment involves the determination of the bond dissociation energy of Mn2+. This study illustrates the energy accuracy and precision of the instrument. The second experiment is a study of the energy dependence of the gas-phase reaction between Ar“ and deuterium. This investigation illustrates the product collection efficiency of the instrument. This section discusses each experiment in further detail. I . Collision-Induced Dissociation of the Singly-Charged Manganese Dimer Ion To demonstrate that the MSU ion beam instrument can be used to obtain reproducible, accurate relative reaction cross-section measurements for endothermic processes, the ion beam instrument was used to perform collision-induced dissociation (CID) experiments to determine the metal- metal bond energy of the singly-charged manganese dimer ion. These bond energies have been previously reported and will be used to evaluate the performance of the MSU ion beam instrument [1]. 67 68 Mn2++ Ar --> Mn" + Mn + Ar AH=O.85iO.2eV (1) The present CID experiment was conducted using argon as the collision gas with the chemistry occurring as shown in reaction (1). A complete list of the experimental parameters is documented in Table 4.1. The primary reactant ions, Mn2+, were generated by electron impact ionization of Mn2(CO)1o. An insertion probe was used to introduce the solid Mn2(CO)1o directly into the ionization chamber. Because of the low vapor pressure of the metal carbonyl at room temperature, the source diffusion pump had to be throttled to allow the pressure within the source to be maintained at approximately 8 x 10'6 torr. This generated a primary ion beam with a current corresponding 1.0 x 106 counts/second. To limit the amount of excess internal energy of the primary ion, 18.5 eV electronswere used to generate the primary ion, rather than the normal 70 eV. As mentioned in an earlier section, calibration of the electron energy is explained in Appendix A. With no gas introduced into the collision cell, the base pressure of the instrument is approximately 2.1 x 10'7 torr. The pressure is monitored by the ionization gauge mounted to the chamber containing the octopole and collision cell. The gauge was calibrated for nitrogen. The stopping potential experiment was performed by sweeping the octopole DC bias through the acceleration potential while the quadrupole was operated in the rf-only mode. The amplitude of the quadrupole rf signal was fixed to a setting that would normally be used to pass an m/z value of 20. The value of 20 is chosen because it corresponds to an m/z on the low end of the quadrupole range. If .m £289? 5 3255 98 £888qu 2: mo 822.62“ 82 a :3 33 88s 3 $3 on so 2 a 2 0? 258. 9v 8 28 2: E3 22 .=o> 283. :3 so to 8 spews: no 8-2: 538.2. 5 S m S 85cm :9 m-» a £8 828: m < cos; w: E a .NN 83% 23 E as on .36 as H .55; 8 E m 85am am dc a sums: a H 55:. «.2. E m .23 85cm 9720 3 w £85 3. 8.5.0.5; > .t as E n .23 850m 956 E s 598:. and £2 5882:. cm do a ._N 8:3 8888s 3 o 3955 NE 9v 6%; _85 one 5 A ._N 88m .9 94m sacs... 955 8.35m .88 m3 3 22.5 ma 3 4 .585 m2 8 a _88 82 E 5885 93 S m 923. 3.2 E S _85 82 3 Essex N8 5 N Ems; 858% 9V 5 _85 so 2: .35 7.1: 80 E _ sag; S: in :3 0.0 w @289 .EoEtoaxm DU +82 2: mo £20883 ._.v 033. 70 the amplitude value set too high, the quadrupole would perform low mass discrimination which may result in not detecting low mass products. The primary ion beam kinetic energy spread was calculated to be 0.27 :t .02 eV, as determined from three separate stopping potential curves. For the CID experiments, high purity Ar (99.999% as stated by AGA Gas Inc.) was used as the collision gas. Pressure within the collision cell was maintained at 0.09 mtorr as measured by the ionization gauge calibrated for argon (corresponds to a mean free path of 2.1 m/atom). Resolution of the quadrupole was decreased such that transmission of the mass selected ions was nearly 100%. The decrease in resolution was possible because of the large difference in mass between the primary ion, Mn2+, and the product ion, Mn+. The background pressure of the chamber was measured to be 6.4 x 10'6 torr. The reactions were monitored from one to 21 eV in the laboratory frame. After the zero kinetic energy point is determined and the energy scale is converted to the center-of-mass frame, the effective measurement is zero to 4.5 eV in the center-of—mass frame. A background scan was taken for each CID experiment, with each pair of scans performed in triplicate. Each background scan was subtracted from its corresponding CID experiment. Because of this background subtraction, some of the relative cross-section values are less than zero below the CID threshold. These negative values have been set equal to zero. The results of the experiments are shown in Figures 4.1-4.5. 71 .02? 53:88:. m: 8 33385: 33 023 scam .823 8% +82 09:: .«o near—3:50 A .v 2sz A20 53 mayo—é 9:25— mé oé m.m o.m Wm 9N m._ o._ m.o a a A a a a _ “m? 0.0 ‘ w _d 2;. h N Sam I 4| 0.4 fl 53 o freq 4 no: ‘ b“. 1 0 Q" 4 i... «no I 4 I 4 a“ on... I Nd I md r v.9 I md ‘ od .. 5.0 ; wd I ad I o; nouns-SSOJQ ”pupa Relative C ross-Sections Cross Section (10'1,6 cm2) 0.8 ~ MSU Ion Beam Instmment . R“ 0.7 a I54 A. F 0c 0.6 — IF! fin I z 0.5 - -’~ I o 0.4 ~ - . 'I' ‘M I lst 0.3 " . . A 0.2 ‘ Pie 1,. 0 2nd . “ A 3rd 0.1 r I F" O [#LLL%L1LLLIII‘1111+J I 1 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Kinetic Energy (eV, CM) Energy (eV, Lab) °~° 5- 0 mo . D Y I I ' ‘ * f T * ' Armentrout et. al. results [1] . . (. 15b Mn2++AraMn++Mn+Ar . 1.0 AIAAPLILJJ 0.5 frj'l'UVVTV'V—Y" l 1 A l A l l L 1 J l 1 j l l A I. l l J A A A 1 0.0 1.0 2.0 3.0 Energy (eV, CM) Figure 4.2. Comparison of an+ CID threshold regions. .N .5320 E wanton as E Ea .: .5 325393 +82 2: .8 fl and he an .mé 8:3”— 220 .>ov 38mm 0325. m.~ o.~ 2 o._ md od x . . _ + . . . . n 0.90.09.01.29 , fio E £95 I. 3.50 Sam +32 0 I «d 83 n a ,2 who.“ u a 86 u h.m. 11o 0 go lioness-$5013 M92193 74 .m .8320 5 33% 2a 8 can .: En.— .EoEtuaxo +32 05 mo N :3. mo HE .vé 2:3"— E 555 In 330 Sam +82 0 96 53 98m 085 o.~ —4._ coo; H 8 3m; H a 36 N um I md fl Yo 1 nd I ed l l‘. o Houses-8801;) analog .m 8320 E cacao—u 03 E can .: .hm .EoEtuaxo +82 05 mo m =3. .8 gm .m.w oSmE 95 .3 3.8m 2.85M Wm o.N _ m4 0; 5 £50 I 03:0 «:5 +32 0 75 83 n 8 mg; n a 85 n J - fio I 6! o M :5 [109308-88ng ”99193 1 Yo 76 Precision of the instrument is illustrated when the results of the three experiments are plotted together, as shown in Figure 4.1. An expanded plot of the threshold region is shown in the top panel of Figure 4.2. The energy accuracy of the instrument is illustrated by comparing the top panel of Figure 4.2 with the lower panel of Figure 4.2. The bottom panel of Figure 4.2 shows the results of a previous study on Mn2+ by Armentrout and coworkers [1]. Their results, plotted as absolute cross-section verses kinetic energy (CM), were calculated from data generated from an+ ions formed by ~18 eV electron ionization. Through curve-fitting and deconvolution, Ervin et al. determined the an+ bond dissociation energy, D°(Mn2+), to be 0.85 :t 0.2 eV [1]. The fitting of our data is shown in Figures 4.3-4.5. From these results D°(Mn2+) is determined to be 0.91 i 0.1 eV. The Mng+ CID experiments demonstrate that the MSU ion beam instrument can be used to produce reproducible, kinetic-energy-accurate relative cross-section measurements for endothermic processes involving ions generated through electron-impact ionization. I I. Reaction of Singiy Charged Argon with Molecular Deuterium The hydrogen transfer reaction of singly charged argon with molecular hydrogen and its isotopic counterpart, deuterium, has been extensively studied to the point that this system "represents one of the most thoroughly investigated systems in the history of ion-molecule chemistry" [2]. The reaction, shown below, has been found to be exothermic by 1.50 eV for ground state reactants and products. 77 Ar+ + D2 --> ArD+ + D AH=-1.50i0.03eV (2) The energy dependent reaction cross-sections of exothermic reactions are at their maximum at or near the lowest translational energies. Also, due to the release of the excess energy, products may possess high translational energies. These two factors contribute to a high degree of scattering in the collision cell of both the reactant and product ions. Because of this scattering, there is a great deal of difficulty in performing these types of experiments in a longitudinal ion beam instrument. Therefore, this system represents an ideal candidate to aid in the critical evaluation of the performance of the MSU ion beam instrument. The complete list of the experimental parameters is shown in Table 4.2. Argon ions were generated by 70 eV electron ionization at a source pressure of approximately 1 x 10'5 torr. Primary ion currents of 2.5 x 106 ions/second were easily obtainable. With the collision cell maintained at its base pressure of 2.5 x 10'7 torr, as measured by the ionization gauge calibrated for nitrogen, stopping potential analyses were performed. The quadrupole was operated in the rf-only mode with the amplitude corresponding to m/z 20. The primary ion beam kinetic energy spread was calculated to be 0.29 :t .02 eV as determined from three separate stopping potential curves. Technical grade deuterium was used for the energy resolved reaction studies. The pressure within the collision cell was maintained at 0.09 mtorr as measured by the ionization gauge calibrated for deuterium. The background pressure of the chamber was measured to be 6.2 x 10‘6 torr. The 78 .m 525%? E 3255 08 fiflofinan 05 mo macaw—con— 82 cc 5: :5 8:2. 3 as on so- 2 .8 3 up 258. cc 8 38 :2 23m. 22 .=o> ._§< :2 20 3 a 2222: no 235 .28: .t. a E m .3 82cm :2 2... a sou 228: m < .32; 3 cc A .NN 85% 23 E as on .230 3 H :25: S E m 8.8m 5 E a .283. a a .23 ”2“ E x .23 028m 9220 co m .2222 E 322:5 >: as... S 2 .23 825m 2220 S s 3:25 SN 52 5:83.: :- 2: 9v a 2m 858 28.28% E o 52:22 E E :83 _85 82 E A 2N 85% 2.: S m .2822 955 cc _ogm _85 :3 8 222m 5 9v 2. .222; 8. _- cc 2 s25 82 3 5822m— Eo E m 5.25:2 2 .2 E «N _85 82 $12.22.: as co N .222... 282.8% E _N .85 2o 2: .35 3: as E _ .2822 :2 __sm :2 m-» _ A332: .2585qu NO + +54. 2: «0 280823 NV 035. 79 reactions were monitored from 1 to 9.5 eV in the laboratory frame which resulted in a range from 0 to 0.63 eV in the center-of-mass frame. To achieve the highest ion beam current, the quadrupole was operated with the lowest resolution that completely separated the primary reactant ions from the product ions. Results of the experiments are shown in Figures 4.6 and 4.7. Stopping potential experiments were performed by monitoring the primary argon ion current as a function of kinetic energy with no gas introduced into the collision cell. Deuterium was then introduced into the collision cell. The reactant ion, Ar+, and product ion, ArD+, were then monitored as a function of kinetic energy. A background scan was then taken, as described in an earlier section. Each scan was performed in triplicate. As shown in Figure 4.6, the sum of the transmitted ion beam current and the product ion current is almost equal to the primary ion intensity when no gas was present in the collision cell. These results indicate that the dynamic trapping process of the octopole is effective at collecting highly scattered reactant and product ions even at low interaction energies. The data were collected and converted to the center-of-mass frame as described in an earlier section. The energy-dependent reaction cross- sections are shown in Figure 4.7. The curve is an average of three separate experiments. As expected, the cross-section is at a maximum at the lowest interaction energy and decreases exponentially as the energy is increased. This is consistent with Langevin—Gioumousis-Stevenson rate theories based on ion/neutral captures driven by ion/dipole and ion-induced/dipole attractive forces [3]. 2:2:ng Enos :2 DmE 05 mo .8565: 5:00:00 8385 mo 2.2852: 6.... BamE 32 .>2 2.522 2322 - We :8 222:8 2 6 .EE 2. +2 2 saw I... :8 222:8 2 Na .+:< + :8 222:8 2 d was. IT - 3 :8 82:8 2 5 oz .1... IT 1 m2 - o.~ 2 3 (,oI) pauses/saunas 81 .D + b? A: NO + .2. ”22828 2: 8.2 2.288-280 2: mo 85988: $225 .5“ 2:me Nmo 2o 0o 2o m.o No 2.0 0.0 _ O 0% w L. + m + o oooooooooooooooooooooooooo o .00 09909990009090... 9999909000 11 O. 00 0. r o.o o.N oé 0.0 o.w 0.3 92 0 fl' ‘I-t 0.3 (can "_01) uoyoas-ssoJQ 82 The monitoring of the reaction between singly-charged argon ions and molecular deuterium has demonstrated the high trapping and transmission efficiencies for exothermic ion-molecule reactions attainable with the MSU ion beam instrument. LIST OF REFERENCES Chapter 4 KM. Ervin, S.K. Loh, N. Aristov, P.B. Armentrout, J. Phys. Chem., 87, 3593 (1983). K.M. Ervin, P.B. Armentrout, J. Chem. Phys., 83, 166 (1985). G. Gioumousis, D. Stevenson, J. Chem. Phys., 29, 294 (1958). Chapter 5 Chemistry of Chlorotitanium Ions ***************************ItIIIIII*5!It*********1!*t************************** I I . CID Experiments on 70 eV EI Chlorotitanium Ions In 1968, Kiser, Dillard, and Dugger (KDD) performed ionization and appearance potential experiments to determine the heats of formation for all of the positive chlorotitanium ions generated from titanium tetrachloride [1]. Data obtained were plotted as ionization efficiency curves by the semi-logarithmic plot technique described by Lossing, Tickner, and Bryce [2]. Potentials were then extrapolated from the data using a procedure outlined by Warren [3]. A brief description of the procedure is given below. The instrument on which the experiments were performed was a time-of- flight mass spectrometer employing electron-impact ionization. Two gases were leaked into the instrument simultaneously; one was the compound under investigation, while the other, a standard (nitrogen), for which the ionization energy is already accurately known. Measurements of the ion current were made as a function of electron energy for both the unknown and standard. 85 Partial pressures of the gases were adjusted to give signals of approximately the same intensity for each gas when the electron energy is set to 50 eV. The peak heights were then plotted on a logarithmic scale as a percentage of the peak height at 50 eV as a function of electron energy. Because the curves were parallel within experimental error at the one percent level of the 50 eV peak or below, voltage differences were measured at the one percent level. The difference between the voltages required to generate a peak one percent of the height of the 50 eV peak for each compound was taken as the difference between the ionization and appearance energies of the two compounds. Results of the experiments are shown in Table 5.1. Table 5.1. Heats of Formation for Ions Produced from TiCl4 [1]. Ion AHf (ion) BDEl (kcal/mole) (kcal/mole) TiCl3*--C1 87 38 TiClz+--Cl 96 78 Tier--01 145 90 Tit-Cl 206 102 Ti“ 279 1Calculated from information provided in the paper. KDD calculated the AHf for each ion using the general reaction below: TiCl4 —. T101,+ + (4-x)Clo (1) 86 By including -181.6 kcal/mole for the AHf of TiCl4, 29.08 kcal/mole for the heat of formation of Cl- , and measuring the appearance energy of each 'I‘iCl,‘+ ion (which is equal to the Aern) the authors were able to generate values for the heats of formation for the four chlorotitanium ions. Bond dissociation energies (BDEs) were calculated using the general equation below for x = 0-3: AHf (1301;) + AHf(C1-) - AHf(TiCl(x+1)+) = BDE(TiClx*--Cl) (2) We have performed collision-induced dissociation experiments on chlorotitanium ions generated by 70 eV EI. Experimental procedures and data analysis were described in Chapter 3. Plots of our experimental results and curves of best fit are shown in Figures 5.1-5.4. Table 5.2 lists the results of our experiments, the BDEs calculated from the heats of formation determined by KDD, and the BDEs of the neutral chlorotitanium molecules calculated from the heats of formation reported in reference [4]. .me 222:8 05 mm :8: $3 coma 30:. mo 90 3282 Exam Ad 8:me :5 $2 22.2 2.25. E S 2 23 we to No 0 A We coco; H E 35m; n 2 >036 u 2m I °°. o uopoas-ssmg) M93193 7P .mmw 222:8 2: 3 cow: was» coma 30E. mo 90 3282 .35an .~.m onE 340 $3 .333.— 9:25— m.m m we v Wm m Wm N m._ 2 We ._ —s 521m mo 5. ‘0. 'n. “I "l O O O O O comes-smug axpnpu °°. 9 ad .mmw 222:8 2: mm :82 83 28.3. SUP mo 90 8388 Exam .m.m 22mm”: :5 $2 22.2 2.222 8.2 8.2 cos 8.: com cod cod cod co: — : a _ _ : LDC. o o 00 o COCOA H E .. . 22: u a oo o. oo >uwh.~flhm o o. o l 5.0 r ad uoyaas-ssmg analog .mmm 222:8 2: m2 .88 22>» 282 «5:. mo 90 8282 $82M in 22me :20 $2 .882: 2.25. 0.2 8 8 S. 8 8 . . . . _ . 3...; 8 . S O O O O T I No l m: I 2.: l m: l 0.0 coco; H E . om_w.— H 2 b o . . >o 3N u 2m r w.o l I O O O Q.o names-smug GAQBPH 91 Table 5.2. Bond Dissociation Energies for Ionic and Neutral Chlorotitanium Molecules. Ions Neutral BDEl 13m:2 BDE3 (kcal/mole) (kcal/mole) (kcal/mole) TiC13--Cl 8.3 38 82.5 , Ti012--Cl 64.7 78 101.2 TiCl--C1 41.0 90 122.6 Ti--Cl 63.8 102 105.1 1Results from our experiments. 2BDEs calculated from heats of formations determined by KDD. 3BDEs calculated from heats of formation reported in reference [4]. A graphical comparison of the bond dissociation energies is shown in Figure 5.5. The techniques used by the two studies to generate the BDEs for the ions are very different. The approach used during our study involved generating chlorotitanium ions and injecting a single TiClx+ species into the collision cell. This enabled us to measure the amount of energy required to remove one chlorine atom from the ion. This approach is limited by the unknown amount of internal energy imparted to the ion during the ionization process. The procedure employed by KDD generated ions through the successive loss of atoms following ionization. All the chlorotitanium ions were believed to be derived directly from TiCl4. The probability is low that each chlorine atom will depart with zero momentum and leave all the energy available for further dissociation of the TiClx.1+ fragment. Energy carried away from the system by the exiting atom must be compensated for and requires higher electron energies. The approach .880 o§3§§£ 05 882%: m can m 98 .cuib 8:25 .8. 05 .6 cocaoommE aaomoae v .A_U--M_U_.C onto—go 82c 05 .8 $2 2: $5338 _ 823 clubs +19 ._. Ea clans JOE. Eat mesa 2:820 .8 flamenco 223885 25m .n.n 959; 553982.. 88a 05..ch E 85.5% v m N 2 o _ a _ _ o I on m #53 £5 P T ow . / . ./. ,./ .I 8 m. 8:8 COM M -e m , m / - 82 w ./ m 8:8 .830: m 1 ofi ( [ OE 93 followed by KDD is also limited by the ability to accurately determine the electron energy and the efficiency of energy transfer from the electron to the molecule during the ionization process. A more complete explanation is provided in Appendix A. Therefore, only upper limit values can be derived from appearance potential experiments. The consecutive BDEs determined by our study follow an alternating "weaker-stronger" pattern. This differs from the pattern of increased BDEs determined by KDD. Below is a plot of trends in the central ion-atom BDEs (Figure 5.6) which shows there is only one known series (MoF3+) where the successive removal of each halogen atom requires a greater amount of energy. Titanium tetrachloride is known to be tetrahedral [5-7] which would suggest an sp3 hybridization. Ionization of TiCl4 most likely involves removal of an electron from one of the unshared pairs of electrons on one of the chlorine atoms, because these electrons are the least tightly bound to the molecule. Other valence electrons participate in the bonds between the metal center and the chlorine atoms. The ion is believed to retain the tetrahedral shape of the neutral molecule. This is shown in Figure 5.7. The ionization energy of a chlorine atom (12.967 eV [8]) is greater than that of titanium (6.82 eV [8]) . This could result in the transfer of an electron from the titanium atom to the chlorine atom with the partial charge now residing on the metal center. The loss of the electron decreases the electron density about the metal center which increases the repulsion between the titanium nucleus and the chlorine nuclei, thereby decreasing the bond strength. The average bond order would decrease to 7/8 (0.875). This may 1606 1401 120 A (atom/mu) 308 the energy series represents m each (BDE's) [8]. The first bar ton energies t ' bond dissoci 1n SUCCESSIVC Figure 5.6. Trends required to remove the first atom, the second bar, the second atom, etc. Configuration After EI Ionization \ TiCI4 /\ . l'e' \ .o“ TiCl4+ A l a) tetrahedral \/+\ TiC13+ i c) trigonal pyramid '+ lex Ti012+ C1 C1 l e) bent 'I‘i+ \ T'Cl+ C1 1 g) Proposed Ground State Configuration \ \ “go A b) tetrahedral )a d) planar C1 — Ti+— Cl f) linear Ti+- Cl h) Figure 5.7. Comparison between proposed geometry after EI ionization and expected ground state geometry. 96 explain why the BDE for the removal of the first chlorine atom determined by KDD and our experiments is lower than that of the neutral species. Our determined BDE is 29.7 kcal/mole lower in energy than that determined by KDD for the dissociation of the first chlorine atom. The nominal electron energy used during our experiments was 70 eV as compared to approximately 13 eV electrons used by KDD. This means that the energy of interaction between the electron and TiCl4 is much higher during our experiments and because TiCl4 has many degrees of freedom which can store energy, the resulting ion may be excited. An excited ion would have a lower apparent BDE. The BDE determined by KDD may also be slightly higher than the "true" BDE as a result of having to compensate for the departing chlorine atom's momentum. Titanium trichloride (TiCl3) is known to be planar [7] which implies an sz hybridization. The third p-orbital, which does not participate in the hybridization, would serve as an empty orbital into which the chlorine atoms could donate electrons through ligand 1: donation. TiCl4 has no opportunity for back bonding, which could explain the increase in BDE for dissociation of the second chlorine atom from the neutral molecule. Influence of the unpaired electron in one of the d-orbitals would not be significant because influence of d electrons decrease as the both the ligand atom size and electronegativity increase [6]. The titanium atom in TiC13+ is in the +4 oxidation state. Performing CID on TiCl3+ results in a change in the formal oxidation state of the metal from IV to III. This is similar to the neutral case when a chlorine atom is 97 removed from TiCl4 and, as seen from Figure 5.5, the BDEs of the Tile are approximately equal to the BDE of the neutral TiCl4 case. Both studies of the Tile ion, ours and that of KDD, found an increase in BDE when the second chlorine was dissociated from the metal center. The discrepancy between the BDE's, 13.3 kcal/mole, may again be attributed to the difference in interaction energy between the ionizing electron and TiCl4. KDD attempted to measure the onset at which TiCl3+ appeared. The result of their experiments may have been a TiCl3+ ion with a planar geometry as shown in Figure 5.7d. The experiments performed here at Michigan State may have resulted in an ion with a trigonal pyramid geometry, Figure 5.7c. This geometry would be similar to the geometry of a vibrationally excited TiCl3+ ion. However, TiCl3+ has fewer degrees of freedom with which to store energy than TiCl4+. This may contribute to the decrease in the difference between the two studies (13.3 verses 29.7 kcal/mole). The oxidation state of the titanium in Ti012 is II and the bonds may be essentially ionic [7]. Two anions ionically bound to a (+2) cation would explain the higher BDE of the third chlorine (third Cl atom removed from TiCl4) from the metal center. The titanium atom in TiClz+ has an oxidation state of III. Removing a chlorine atom from this ion is similar to cleaving the titanium-chlorine bond in TiC13 in that the metal center is reduced from an oxidation state of III to II. The difference in the determined BDEs for the ion, TiClz”, between our results and those of KDD may stem from the difference in the experiments. frm sel 98 KDD generated the ions directly from TiCl4, immediately extracted them from the source, and used a time-of-f'light mass spectrometer to mass select and detect the ions. Their experiments were essentially performed in one step. During our experiments, we first generated the precursor ion and then performed CID experiments on the ion. When TiC12+, a doublet, is generated from TiCl4, the resulting ion may initially resemble a tetrahedron without two ligands, i.e., would be bent, Figure 5.7e. If the ionization and dissociation processes occur before the remaining atoms can adjust, the resulting ion may initially have a highly strained geometry as shown. This geometry would be equivalent to a vibrationally excited linear ion. Performing CID experiments on this excited ion would result in low BDE's. 'I‘iCl has a lower BDE than TiClg because the ionic bond for TiCl is between two singly charged ions. As stated above, the bonds in TiC12 are between a (+2) ion and two (-1) ions. The higher charged state would produce stronger bonds. For the ion, TiCl”, breaking the bond would be similar to the neutral case when a chlorine atom is dissociated from TiClg in that the titanium atom would be reduced from an oxidation state of II to an oxidation state of I. However, in the case of the ion, KDD would be breaking covalent bonds of TiCl4, not ionic bonds and therefore would require less energy. During our experiments, TiCl+ is first generated and then CID experiments are performed on the ion. After the ion is generated, the geometry may resemble a vibrationally excited state. But, because TiCl+ has fewer 99 degrees of freedom, less energy is internally stored. Having less internal energy would require higher CID energies to break the bond. Also, the increase in BDE (from 'I‘iC12+ to TiCl‘”) may be explained by examining the nature of the bond in TiCl“. The energy required to generate Ti+2 from Ti+ is 313.9 kcal/mole [29]. The energy released when a gas-phase chlorine accepts an electron is 83 keel/mole [calculated as the AH for the reaction Cl(g) --> Cl'(g)] [29]. The difference between the ionization energy of Ti+ and the electron affinity of Cl is 231 kcal/mole. When we use Coulomb's Law to evaluate the potential energy between the two ions (Ti‘”2 and Cl'), we find a distance of <2.9 A between the two ions is required to generate a potential energy equal to the 231 kcal/mole energy barrier. This would lead us to believe that the bond is an ionic bond between Ti”2 and C1" instead of a covalent bond between 'I‘i+ and C1, the energy required to separate two oppositely ions is greater that the energy required to separate a neutral atom from an ion. As mentioned above, the ions studied during our experiments were generated by performing E1 on TiCl4. An attempt to mathematically correct for the BDEs determined during the present study would not be possible since the ionization process is believed to be dominated by vertical processes that do not allow the state populations to be characterized by a Maxwell-Boltzman distribution. It is also noted that the curves in Figures 5.2 and 5.4 are not well fit in the threshold region. The shape of the experimental plots resembles plots generated by Ervin, Loh, Aristov, and Armentrout [9] during their study on the effect electron energy has on determined thresholds. They concluded that higher electron energies 100 decrease the apparent onset. Reasons why data in Figures 5.1 and 5.3 do not possess the same characteristic are unknown. Therefore, considering that our determined BDEs may be those characteristic of excited ions and that the determined heat of formation for Ti+ by KDD agrees very well with others [9, 10], we conclude that the difference between our determined BDEs and those determined by KDD represent an estimate of the minimum amounts of internal energy the ions possess. Our results are consistent in that the experimentally determined BDEs for the excited chlorotitanium ions never exceed the BDE values determined by KDD. With these approximations, we are then able to gain more insight into the chemistry of chlorotitanium ions generated by 70 eV EI. I I . Chemistry Between 'l‘iClx“ and Oxygen-Containing Commands Titanium compounds containing oxygen have dominated the organometallic chemistry of this metal in the past [10,11]. The first organic derivative of titanium was an alkoxytitanium compound prepared in 1875 [12]. Oxygen-containing titanium compounds have been used in paints, in the treatment of textiles and paper, and as polymerization catalysts [10,11]. Titanium tetrachloride (TiCl4) has been used as a Lewis acid to catalyze condensation reactions involving enol acetates, acetals, carbonyl compounds, and other oxygen-containing organic molecules [13-16]. Studies of the gas-phase chemistry between chlorotitanium ions and 101 oxygen-containing compounds at room temperature have also been conducted [17]. More recently, the organometallic chemistry of titanium metal has focused on the chemistry of the bare-metal ion in the gas phase [18-40]. Here we present an attempt to investigate the endothermic and exothermic, gas-phase chemistry of chlorotitanium ions with oxygen- containing compounds. These studies were carried out in the hope that any discovered processes may help us to more fully understand reactions characteristic of condensed phase titanium compounds. A. Results Exothermic products are more abundant at lower kinetic energies while products generated through endothermic processes are more abundant at higher kinetic energies. For this reason, in order to identify all products for a given system, several product ion spectra are collected over a range of kinetic energies. Figures 5.8 and 5.9 show the product spectrum of TiCl4+ + CH3CHO (acetaldehyde) at two different kinetic energies. These two spectra are used to determine the mass values monitored during the energy-resolved experiment. The results of the energy-resolved experiment are shown in Figure 5.10. As seen from Figures 5.8-5.10, there are two exothermic products, TiCl302H4O+ (III/Z 197) and TiC12C2H3O+ (m/z 161). Only the product appearing at m/z 197 was previously reported by Allison and Ridge (ARl) [17]. The endothermic product is TiCl3+ (m/z 153). The appearance of 5.E+6 — 4.E+6 . L» m 3‘ A Counts/sec N m 3; 41 1.E+6 « 102 (A) all.“ h l l O.E+0 I 7.E+4 - 6.E+4 4 3'" m E 4.E+4 ~ Counts/sec ii; i .N m : LEM - 20 40 6O O.E+O ‘ 20 40 6O 80 100 120 140 160 180 200 m/z (B) 1 80 100 120 140 160 180 200 m/z Figure 5.8. A) Product mass scan for Tile + CH3CHO with the octopole DC offset set 4 Volts below V“. B) Expanded view of the spectrum. 4.E+6 ~ 3.E+6 4 Counts/sec 'm 3 1.E+6 - (A) 0.E+0 4.E+4 '1 3.E+4 4 Counts/sec .N m K 1.E+4 - 100 120 140 160 180 200 (B) 0.E+0 - Lt 100 120 140 160 180 200 Figure 5.9. A) Product mass scan for riot.“ + cnscno with the octopole DC offset set 20 Volts below v.00... B) Expanded view of the spectrum. 104 1.2E+04 -, (A) 1.01904 "s U 3 8.0E+03 S ‘5' 601,303 1 A m/z 197,TiC1_,C2H.,o+ f3 , x m/z 153, TiCl3+ I § . ‘ U ' 0.0E+00 ; ~ -------- 0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 Kinetic Energy (eV, CM) 4.0E+03 1.- 5 (B) 3.6E+03 ] 5 3.2E+03 2.8E+03 2.4E+03 2.0E+03 1.6E+03 l .2E+03 8.0E+02 Cross-Section (10“ cm’) 4.0E+02 0.0E+00 0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 Kinetic Energy (eV, CM) Figure 5.10. A) Energy dependent reaction cross—sections for the reaction TiCi,+ + c.1440 (acetylaldehyde). B) Expanded view of plot (A). 105 the collision-induced dissociation (CID) product, TiCl3+, signals the limit of the amount of energy that can be injected into the system through kinetic energy. As the kinetic energy is increased, the CID product abundance increases until its signal reaches a maximum and then decreases to zero. At higher kinetic energies, there are no product signals. Figures 5.11-5.13 show the results of the energy-resolved experiments for the reactions 'I‘iClx” (x = 1-3) + acetaldehyde. In each case, two exothermic products were observed with a CID product appearing at higher kinetic energies. For each experiment, the exothermic product with the higher reaction cross-section is the product observed by ARl. Allison and Ridge did not report the observation of the second exothermic product.- The experiments performed by ARI employed an ion cyclotron resonance (ICR) mass spectrometer which did not allow them to precisely inject additional energy beyond that of the ionization process. Motion within the mass spectrometer is governed by thermal energy and the magnetic fields generated by the magnet. The products they observed would have to be the result of either exothermic or endothermic processes with barriers small enough to be easily overcome by the energy imparted to the system during ionization. The exponentially decaying curves generated from our experimental data suggest these products to be the result of exothermic processes. We conclude there are no endothermic processes within this system available to the injection of kinetic energy. The next system we studied involved a slightly larger aldehyde: propanal. Results of the energy-resolved experiments from the TiClx+ (x = 1-4) + 106 1.6E+05 (A) 1.4E+05 F a 1.2E+05 U 2 b 1.0E+05 _ v 8 8.0E+04 0? r1.) 6.0E+04 . + , a o A m/Z 48, Ti 0 . . 5 4-0E+04 x m/z 64, “no+ 2.0E+04 1 O m/z 99. TiClO+ 0.0E+00 0.00 2.00 4.00 6.00 8.00 10.00 Kinetic Energy (eV, CM) 1.6E+03 1 (B) X A 1.4E+03 ~ 4“ X A x ‘i? "E 1.2E+03 ~ x. mu 0 2 2" b 1.0E+03 . ", —1 x5: 5* v X l“ 8 8.0E+02 - “2 x i i»: ' ‘2. . 5‘“ v.2 6.0E+02 ~ xii: “ n “will“ a 8 40E+02 . ‘W . “ A U °. 5.: * xw 2.013+02 ~ ‘9 558'; , £ . . .9. 9 o t’sit 5“ o.oo~.. °~’° 0:“ of. 0.0E+00 0.00 2.00 4.00 6.00 3.00 10.00 Kinetic Energy (eV, CM) Figure 5.11. A) Energy dependent reaction cross-sections for the reaction TiCl+ + C2H40 (acetylaldehyde). B) Expanded view of plot (A). 107 3.5E+05 (A) 3.0E+05 2.5E+05 2.013405 "5905 A m/z 126, Tic1c2H3o+ 1.0E+05 X m/Z 99. TiClOJr . m/z 83, TiCl+ Cross-Section (10.15 cmz) 5 .0E+04 0.0E+00 _—1 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 Kinetic Energy (eV, CM) 3.5E+03 - (B) 3.0E+03— ,, 2.5E+03 - e 2.0E+03 a 1.5E+03 1.0E+03 Cross-Section (10“ cm’) 5 .0E+02 0.0E+00 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 8.00 Kinetic Energy (eV, CM) Figure 5.12. A) Energy dependent reaction cross-sections for the reaction TiCiz+ + c.1140 (acetylaldehyde). B) Expanded view of plot (A). 108 2.5E+05 '1 (A) i) A 2.0E+05 ~ E U 3; 0 3 1.5305] 2: .2 fig . a.) 1.0E+05 i 5 J A m/z 118, Ticlz+ r . + U 5.03.04 . x m/z 126, T1C1C2H3O+ ' O m/z 161, T1C12C2H30 O 0 0.0E+00 w 0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00 Kinetic Energy (eV, CM) 5.0E+03 1 0 4.5E+03 1 (B) 4.0E+03‘1 . 3.5E+03 3.0E+03 2.5E+03 2.0E+03 1.5E+03 1.0E+03 Cross-Section (10“ cm’) 5.0B+02 0.0E+00 0.00 0.50 1.00 1.50 2.00 2.50 3.00 3.50 4.00 Kinetic Energy (eV, CM) Figure 5.13. A) Energy dependent reaction cross-sections for the reaction Tie],+ + C2H40 (acetylaldehyde). B) Expanded view of plot (A). 109 propanal reactions are shown in Figures 5.14-5.17. As with the acetaldehyde system, we observed two exothermic products as the result of the reactions involving TiClx+ (x: 1-3). And, as before, the products with the larger reaction cross-sections were the same products reported by ARI. A11 curves generated from the data of the propanal system exhibit an exponentially decaying signal which identifies them as being exothermic products. Again, we conclude there are no endothermic processes with the propanal system available to the injection of kinetic energy. The reaction TiCl4+ + propanal was carried out with only a single exothermic product being observed at m/z 175. ARI performed the reaction and reported the displacement of a Cl atom from the chlorotitanium ion resulting in the product TiCl3C3H60+ which has an m/z value of 211. This m/z value is beyond the range of our quadrupole. The next compound to be studied was acetone. This step in the neutral reagent series maintained a carbon-oxygen double bond, but placed the oxygen in the middle of the carbon backbone rather than at the end of the chain. Figures 5.18-5.21 show the experimental results for the reaction TiClx” (x = 1-4) + acetone. The TiCl4+ + acetone reaction generated results similar to those of the propanal system. As with the aldehydes, the other three chlorotitanium ions generated two exothermic products. Again, only the product with the higher reaction cross-section for each reaction was reported by ARI. In our attempt to uncover endothermic processes other than CID, we shifted our search from molecules containing a carbon-oxygen double bond 110 25E+5 l (A) A 2.01951 E U 2 a 1.5E+5 : e '5 8 10E+5 a.) . g ’. A m/z 48, “n" h . . + U 5.0E+4 , x m/z 64, T10 .0 . mlz 99, T1C10+ 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) 2.5E+3 n A "a 2.0E+3 " U .‘S 3 1.5E+3 « v ’5 3 1.0E+3 « ‘9 In (I: 5 5.01=.+2 ~ 00E+o 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) Figure 5.14. A) Energy dependent reaction cross-sections for the reaction m“ + C3H50 (propanal). B) Expanded view of plot (A). 111 5.0E+5 4.SE+5 1 (A) 4.0E+5 ' 3.5E+5 3.0E+5 2.5E+5 2.0E+5 A m/z 140, T1C1C3H50+ x m/z 99,TiC10+ LOB” o m/z 83, TiCi+ 5.0E+4 1.5E+5 Cross-Section (10'“5 cm’) 0.0E+0 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) (3) Cross-Section 00'" cm’) Kinetic Energy (eV, CM) Figure 5.15. A) Energy dependent reaction cross-sections for the reaction TiCiz+ + c.1160 (propanal). B) Expanded view of plot (A). 112 LOE+5 (A) 8.0E+4 6.0E+4 4.0E+4 A m/z 118, TiClz+ x m/z 140, TiClC3H50+ O m/z 175, TiC12C3H50+ Cross-Section (10“ cm’) 2.0E+4 0.0E+0 0.0 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) 2.0E+3 '1 1.8E+3 i (B) L6E+31 1.4E+3 ~ 1.2E+3 — :, LOB+33K 8.0E+2 1. ° 60E+2 4.0E+2 Cross-Section (10“ cm’) s 2.0E+2 0.0E+0 0.0 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) Figure 5.16. A) Energy dependent reaction cross-sections for the reaction TiCl3+ + c.1160 (propanal). B) Expanded view of plot (A). 113 3.0E+4 1 (A) 2.5E+4 . "s 3 2.0E+4 n 'c 3 .5 1.5E+4 « 8 "3 y, g "OE” A m/z118,'I'iC12C3H50+ U x m/z 140, TiCl3+ 5.0E+3 1+ 0...... M 0.0 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) 5.0E+3 - (B) h 4“ 4.0E 3 . E + 3 ”x: :51”; x: a 3.0E+3 «A xii-xix" "8 5"“ ’51?" = 8 X X x 3 A gnaw 8 2.0E+3 . A "a f 3 X41. 2 31‘ [A U 1.0E+3 32‘: A? 0.0 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) Figure 5.17. A) Energy dependent reaction cross—sections for the reaction 'ric1,+ + €311.30 (propanal). B) Expanded view of plot (A). 114 4.0E+5 } (A) 3.5E+5 4 3.0E+S . 2.5E+5 ~ 2.0E+5 . v 1.5E+5 4 Cross-Section (10“ cm’) A m/z 48, Ti+ 1.0E+5 4 . + t x m/z 64, T10 D 5.0E+4 o m/z 99, TiClO+ 0.0E+0 _ 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) 5.0E+3 ‘1 4.5E+3 « (B) 4.0E+3 ~ 3.5E+3 ~ )1 3.0E+3 d 2.5E+3 - 2.0E+3 .11 X 1.5F.+3 «1;, 1.0E+3 i ”1“,...” - M 5.0E+2 . Cross-Section (10“ cm’) 0.0E+O 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) Figure 5.18. A) Energy dependent reaction cross-sections for the reaction TiCl+ + €311.50 (acetone). B) Expanded view of plot (A). 115 1.5E+4 n 1.4E+4 a, (A) 1.2E-t-4 1 1.1E+4 3, 9.0E+3 . 7.SE+3 . 6-0E+3 I A m/z 83, TiCl+ 4,513+3 a x m/z 99, TiClO+ O m/z 140, TiClC3H50+ .evvv Cross-Section (10" cm’) 3.0E+3 d s M 1.5E+3 1 .~ x 0.0E+0 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) 0“. 2.0E+3 1 ‘ 1.8E+3 4 A“ 1.6F.+3 . ° 5413414 A“ A ‘ ‘ 1.4E+3 . A} 1.2E+3 . ° N 1.0E+3 '1 o 8.0E+2 . H Cross-Section (10“ cm’) 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) Figure 5.19. A) Energy dependent reaction cross-sections for the reaction Tim,“ + c.1160 (acetone). B) Expanded view of plot (A). 116 3.01~:+5 -. (A) 2.5E+5 ] E a 2.0E+5 * 'c fl V g 1.5E+5 « H 8 0') 10E 5 1 an . + a» . 8 g A m/z 118, TiClz+ i- U . X m/z 140, TIC]C3I‘I{,()+ .0 . 5 E” , o m/z 175, T1C12C3H50+ O O 0.0E+0 0.0 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) 3.013+3 . . (B) A 2.5E+3 « N E a 3; 2.0E+3 ~ , S V o S 1.5E+3 ~" . ’5 . 0 X o ‘2 1.0E+3 - 2.. g x “'0 U 50E+2 x O" a o . ”at“? .‘. fiM ODE-+0 W Kinetic Energy (eV, CM) Figure 5.20. A) Energy dependent reaction cross-sections for the reaction TiCl3+ + c.1160 (acetone). B) Expanded view of plot (A). 117 6:988 056 + +302. 5:88 05 no.2 8288-820 .6288 Eugene—u 328m .86 uSmE A20 $3 3.5:”..— 9385— me o.v mgm o.m Wm od m._ o._ md o.o 4 . Jami?! « D $8.2. m {83.3%fo m w . w 20:822.: x a $85 m b.8388: as < m... - $8.” a... mz a. $8.2 Alma.— 118 to one having two carbon-oxygen single bonds: dimethyl ether. Figures 5.22-5.25 show the experimental results for the reaction 'I‘iClx+ (x = I-4) + dimethyl ether. As seen from the plots, the results are similar to those of the ketones and aldehydes. Only exothermic products where observed. When the results of the dimethyl ether chemistry are compared to those of ARI, there appear to be two differences. Previously, when our results for the chlorotitanium ion chemistry are compared to those of ARI, we found that we observed one additional exothermic product for each reaction. For the reaction involving dimethyl ether, this is also true for the Ti012+ and TiCl4+ reactions. The reactions involving TiCl” and 'I‘iC13+ however, yield the same number of products in both studies. We did not observe any additional products from those reported by ARI. B. Discussion The two studies are in general agreement. For most of the individual reactions, we have observed one additional product to those reported by ARI. This may appear to be additional chemistry, previously not observed. However, when the results of the two studies are compared as series of chlorotitanium ions reacting with each compound, no new chemical species are discovered. In every case, the product with the larger reaction cross-section is the same product observed by ARI. The product with the smaller cross-section is the same product observed by ARI during the next reaction in the series. This is shown in Table 5.2 where the products of the chlorotitanium L2E+5 1.0E+5 8.0E+4 -I‘cm2) 6.0E+4 Ion (10 4.0E+4 Cross-Sect 2.0E+4 " 0.0E+0 0.0 2.0E+03 ~ 1.8E+03 * '5‘ 1.6E+03 « X m 3 w 1 1.2E+03 A 1.0E+O3 4 8.0E+02 - 6.0E+02 * ross-Section (10“‘cm C .e E: s .0. 2.0E+02 1 0.0E+00 % 119 (A) X m/z 114, TiClOCH3+ O m/z 48, Ti+ 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) 6.0 7.0 x O. X .00 x 5% up” ‘0 0.... «W it 3'} I“ ’0‘ x ”.3... J 1 i! l x X&VM‘M¥¢£¢&*¥ 1 0.00 1.00 2.00 3.00 4.00 5.00 Kinetic Energy (eV, CM) 6.00 7.00 Figure 5.22. A) Energy dependent reaction cross-sections for the reaction Tic:+ + CZHGO (dimethyl ether). B) Expanded view of plot (A). 2.0E+5 — 1.8E+5 1’ (A) «renews 1 U 2 1.4E+5j 31.2E+5 - 81.01435 4: fl 38.0E+4 - (I) $601544 d A m/Z 83, T1Cl+ 54 0E+4 3x x m/z 149,TiC120CH3+ 20W 9 m/z 114, TiCIOCH; w 0.0E+0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 Kinetic Energy (eV, CM) 2.0E+3 - 1.8E+3 e (B) .1" E 1.6E+3 u 1.4E+3 e 2 31.2193 — " V E x l.OE+3 fl 8 x m“ 88.01342 - V.) ‘ ‘5“ 36.0192 ~ “at O R x A 54.0E+2 U x “no“. 2.0E+2 « °°.. Kinetic Energy (eV, CM) Figure 5.23. A) Energy dependent reaction cross-sections for the reaction Tie]; + C2H60 (dimethyl ether). B) Expanded view of plot (A). 121 4.0E+4 (A) 15E+4 10E+4 2.5E+4 2.0E+4 L5E+4 A m/z 163,TiC12CH20CH3+ x m/z 149, TiC120CH3” o m/2118,TiC12+ LOE+4 Cross-Section (10“ cm’) 50E+3 QOE+0 00 05 ID 15 20 25 3D 35 Kinetic Energy (eV, CM) 2.0E+3 1| 1.8E+3 ,, (B) 1.6E+3 1.4E+3 " 1.2E+3 " LOE+3 . 8.0E+2 ,. 6.0E+2 ”k 4.0E+2 ‘, ”a x . . MW“: X 2.0E+2 “ . ...°o. xx °’ 0.0E+0 O.” l‘h’e‘efifiéw 00 05 ID 15 20 25 30 35 Kinetic Energy (eV, CM) Cross-Section (104‘ cm’) Figure 5.24. A) Energy dependent reaction cross-sections for the reaction TiCl3* + C2H60 (dimethyl ether). B) Expanded view of plot (A). 2.5E+4 - 4) (A) p 2.0E+4 ~ E U :2. , 2 1.5E+4 a = O '5 .J (I)? 1.0E+4 g x m/z 153, TiCl3+ 65 5,013,, _ e m/z 149,TiC120CH3+ W 0.0E+0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Kinetic Energy (eV, CM) 5.0E+3 — (B) .1" 4.0E+3 J E U 2 3 3.0E+3 - “XFXMW v g sag-9w .3 l) x W 8 2.0E+3 . ’9'”? m ’5! " a'a at,“ E . x U 1.0E+3 a "‘ :x D 0.0E+0 & - 7w" '. W fr“? T2“ -, w 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 Kinetic Energy (eV, CM) Figure 5.25. A) Energy dependent reaction cross-sections for the reaction Tia,” + (221160 (dimethyl ether). B) Expanded view of plot (A). 123 reaction series with acetylaldehyde are listed. Branching ratios for the reactions were calculated using the following formula: ratio value A = (peak intensityA/Z product peak intensities) x 100% Table 5.2. List of products for the reaction 'I‘iClx+ (x = 1-4) + acetylaldehyde. Ion ARI Results Our Results Branching (m/z) (m/z) ratios (%) TiCl+ TiClO” TiClO“ 63.6 ("V2 83) Ti0+ 36.4 Ti012+ TiCngH3O+ 'I‘iClC2H3O+ 99.4 (m/z 118) TiClO+ 0.6 Tile TiClgC2H3O+ TiClngH3O+ 98.3 (m/z 153) 'I‘iClC2H3O+ 1.7 Tile 'I‘iCl3CZH4O+ TiC13C2H4O+ 80.6 (m/z 188) 'I‘iC1202H3O+ 19.4 The difference between the two studies may be explained by our use of a more sensitive instrument. If the instrumentation used by ARI were more sensitive, they may have also observed a second reaction product. Our results indicate that there does not appear to be any endothermic processes available to increased reactant ion kinetic energy for the systems studied. Because the bond between titanium and oxygen is so strong, the interaction between the metal and oxygen releases a large amount of energy. This excess energy is then transferred throughout the newly formed ion. With relatively small compounds, the number of modes into 124 which this excess energy can be stored is limited. This results in the breaking of weaker bonds, to form both neutral and ionic fragments with a wide range of kinetic energies. It is of interest to note that the onset associated with the chlorotitanium CID product in every system studied occurred within the range 0.5-1.0 eV below the onset observed during the CID experiments with argon as the target gas. If we had been able to work with larger compounds, we may have been able to observe endothermic processes other than the CID of the reactant ion. I I I . Chemistry between 'I‘iClx+ ions and isobutene In the previous section, CID experiments performed on chlorotitanium ions generated by 70 eV EI were described. BDE's generated from the results of those experiments differed from BDE values previously determined Kiser, Dillard, and Duggar (KDD) [I]. The differences allow us to estimate the amount of internal energy present in the ions. In this context, we investigated the chemistry between TiCln+ ions and isobutene. This work refines previous conclusions and provides an important context for understanding the influence of internal energy on ion/molecule reactions of metal-containing ions generated by 70 eV EI. Gas-phase chemistry between positive chlorotitanium ions and olefins has previously been investigated by Allison and Ridge (AR2) [41]. The general reaction scheme developed from the results is shown in Figure 5.26. deco? 5m? 2,3283: 3 3 case 33 “Hop 85.33 28 meow 828359820 5253 humane”? 2.3 .8.“ SE mg .3 39233 08658 :oBowom find 3ng ammao + anaemmadoioa. AIIJ 5m + mammaoaeaa. I ammao + mam. mm + Jamaammaojofit All. 126 Chlorotitanium ions were found to react with small olefms through the elimination of molecular hydrogen and hydrogen chloride. Larger alkenes also reacted through the elimination of smaller olefins. TiCl4+ was found to be unreactive. Reactions (1-12) illustrate the observed chemistry for experiments performed at Michigan State. For comparison, previously observed chemistry is indicated by an asterisk (*). Also included with the reactions are the branching ratios. Reactions (6-8) contain two sets of branching ratios. The first set of ratios corresponds to ratios occurring at K.E.CM = 0. The second set corresponds to ion kinetic energy values where the endothermic product has a maximum intensity. Curves generated from our energy-dependence experiments are shown in Figures 5.27-5.30. Experimental procedures and data analysis are described in Chapter 3. Branching ratios TiClzC4H7+ + Cl- + HCl 92.1 (1) TiCl4++ 04H8 { TiClZ+ C4H7+ + C1- + HCl 7.9 (2) Ti012C4H7+ + H0] 94.0 * (3) TiCl3++ C4H8 TiClC4H6+ + 2HCl 3.1 (4) TiCl2+ C4H7+ + HCl 2.9 (5) TiClZC 4H8+ 20.3 8.0 (6) TiC12++ C4H8 TiClC4H7+ + H0] 79.6 60.2 * (7) TiCl + 04H?+ + HCl 0.1 31.8 (8) TiClC4H6+ + H2 78.8 * (9) LTiClC3H4“ + CH4 18.7 (10) TiCl" + 04H8 - '> TiC4H5+ + HCl + H2 0.9 (12) A. Thermochemistry Considering the results of the previous study performed by AR2, several conclusions could be drawn about the thermochemical information. 60000 a (A) 50000 .1 t.‘ 5 40000 — '5 I! § 30000 a 282 a a 20000 . -°-m/z 55 U + m/z 153 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 Kinetic Energy (eV, CM) 60000 . (B) 50000 d 40000 - 30000 a 20000 ~ Cross-section (10“ cm‘) 1 0000 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 Kinetic Energy (eV, CM) Figure 5.27 . A) Energy dependence of reaction cross-sections for the reaction Tich,+ + c411,. (isobutene). B) Expanded view of plot (A). 25000 20000 g—+%~Lu_aun-—J 15000 1‘ 10000 ~ Cross-section (10" cm’) 5000 ~ 2000 1 1600 - ‘5 U ’5, 1200 A E’ .2 ’ ¥ '5 E o 800 1, (A) 0 m/255 . m/z118 -+-m/z 137 x m/z 173 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) * (B) x o: O .0 o 00. . x9... .0. 0.. 0‘ . o . o. 00. O : O .000 . O O .9 in ° ° 0 1.0 2.0 3.0 4.0 5.0 Kinetic Energy (eV, CM) Figure 5.28. A) Energy dependence of reaction cross-sections for the reaction Tic13+ + C4118 (isobutene). B) Expanded view of plot (A). 130 9000 l (A) (It 8000 3, X 7000 - 6‘ X .5, 6000 . =5 3: c. 5000 - g x 3: m/z 138 '3 4000 ~ +an 174 g, x —Cnmch Fit (83) § 3000 . x -—Crunch Fit (55) U x o m/z 55 2°00 ' "x + m/z 83 X 1000 ( 0 w ~ I ~~W'a~¢;.‘a..ux....- .A . 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) 1000 ~ 2 (B) 900 — x I! X 300 ‘ x 12.05) = 1.73 eV ET(33) =2.98ev It _ = 5; 700 4 n - 1.0371 R 1.3841 9 x m = 1.0000 m = 1.0000 3 600 J "‘2, G x g 500 4 ,, ’9 3 400 - E'us.‘ firing ++ a +4. + ++ ++++ é’ - ++ Q 300 r I 200 - , , 100 - * 0 A...“ .. M 0.0 2.0 4.0 6.0 8.0 10.0 Kinetic Energy (eV, CM) Figure 5.29. A) Energy dependence of reaction cross-sections for the reaction Tile + C4113 (isobutene). B) Expanded view of plot (A). 131 S i Cross-section (10" cm’) 0) (A) +m/z 101 A m/z 110 2000 - +m/z 123 +m/z 137 x m/z 48 1°00 --Crunch En 0 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 Kinetic Energy (eV,CM) 120 - ‘ ‘ . (B) h 11 100 J 2 . ‘ ET = 1.97 eV 80 g ‘ n = 1.7326 + A m - 1.0000 1 r Cross-section (10“ cm’) 8 0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0 Kinetic Energy (eV, CM) Figure 5.30. A) Energy dependence of reaction cross-sections for the reaction TiCl+ + Cal-13 (isobutene). B) Expanded view of plot (A). 132 Reactions (1315) show three ways to generate TiCIZC4H7+ by reacting the various chlorotitanium ions with isobutene. Shown with each reaction equation is the limit of the BDE(TiClz+--C4H7) required for the chemistry to occur as described by the equation. BDE(TiClz+--C4H7) (kcal/mole) TiCl4+ + C4H8 --> TiClzC4H-7+ 4» HO] + Cl' <97.8 (13) Tile 4» C4H3 --> TiClzC4H7+ + HCl >59.9 (14) 'file + C4H8 --> TiClgC4H7+ + H' <85.0 (15) However, reaction (14) is the only chemistry observed by AR2 that generated the ion. Based on available thermochemical information and the fact that the reaction must be at least thermoneutral, a lower limit of 59.9 kcal/mole can be placed on the BDE; Reactions (13) and (15) were not observed, possibly due to an energy barrier to those reaction pathways. This allows an upper limit of 85.0 to be placed on the BDE. Therefore, the BDE can be bracketed: 59.9 kcal/mole < BDE(TiClz+--C4H7) < 85.0 kcal/mole In a similar manner, TiC1C4H6+ could be generated a number of ways which are shown by reaction equations (16-19). BDE(TiCl+--C4H6) (kcal/mole) TiCl4+ + C4H8 --> TiClC4H6+ + 2HC1 + Cl' <146.6 (16) 'I‘iCl3+ + C4H3 --> 'I‘iClC4H6+ + 2HCl <108.6 (17) 'I‘i012+ + C4H3 --> 'I‘iClC4H6+ + HCl + H' TlClC4H6+ + H2 >428 (19) 133 Reaction (19) was the only chemistry observed that generated the ion. Again, because this reaction is required to be thermoneutral/exothermic, a lower limit of 42.8 kcal/mole can be placed on the BDE(TiCl+--C4H6). Chemistry described by reactions (16-18) was not observed which places an upper limit on the BDE of 108.6 kcal/mole. Therefore: 42.8 kcal/mole < BDE(TiCl+--C4H6) < 108.6 kcal/mole Proposing different neutral products for reaction (18), such as H2 and Cl°, would raise the upper limit of the BDE to 134.7 kcal/mole for the chemistry described by that equation. We have found the Tile ion is reactive with isobutene as shown by reaction (1). This does not agree with the results generated by AR2, who found TiC14+ to unreactive with olefins. Our data indicate the product ion curve has the general form consistent with exothermic reactions, in that the curve peaks very close to K.E.CM = 0 and undergoes an exponential decrease. The results are shown in Figure 5.27. The peak of the curve, however, does not occur at K.E.CM = 0 which would indicate that a small energy barrier (possibly due to steric hindrances) does exist to the chemistry. The reaction pathway may be more accessible to increased K.E., rather than internal energy. This would explain why the product was not observed by AR2, because they had no means by which to inject additional energy into the system beyond that available during the ionization process. 134 Therefore, if reaction (1) is exothermic, the thermochemistry tells us that 97.8 kcal/mole are required for the reaction to occur. By approximating the internal energy of the Tile ion (shown in the previous section) to be 30 kcal/mole, the energy requirement is reduced to 67.8 kcal/mole. This means that the value for the BDE(TiC12+--C4H7) would have a lower limit of 67.8 kcal/mole. Therefore, bracketing of the BDE is adjusted to: 67.8 kcaJ/mole < BDE(TiC12+--C4H7)< 85.0 kcal/mole In a similar manner, our experimental results tell us that the TiCl3+ ion reacts with isobutene to eliminate I HCl (reaction (14)) and 2 HCls (reaction (17)). AR2 indicated only the elimination of a single HCl. The product ion curve generated from our data indicates an exothermic process. However, because the product ion current was low, the instrumentation employed by AR2 may not have been sensitive enough to detect the 'I‘iClC4H6+ ion. ,In order for the exothermic chemistry to occur as described by reaction (17), 108.6 kcal/mole of energy are required. By including the 13 kcal/mole excess internal energy estimated from the CID experiments, the lower limit of the BDE(TiCl+--C4H6) would be 95.6 kcal/mole. Reaction (18) would then determine the upper limit. Therefore, bracketing of the BDE(TiCl+-- C4H6) would be adjusted to: 95.6 kcal/mole < BDE('I‘iCl+--C4H6) < 133.7 kcal/mole 135 Table 5.3 summarizes the thermochemical information derived from the AR2 results and how our results make adjustments to those conclusions. Table 5.3. Summary of Bond Dissociation Energies. Ion Complex Lower BDE Limit Adjusted Lower BDE (kcal/mole)a Limit (kcal/mole)b 'I‘iClz+--C4H3 2 O 2 O 'I‘iClz+--C4H7 59.9 67.91t TiCl+--C4H8 72.0 23.0 TiCl+--C4H6 42.8 95.61 'I‘iCl+--C3H4 31.8 2 0 TiCl+--C2H3 95.4 56.4 'I‘i+--C4H5 24 2 0 a) determined by assuming chemistry to be exothermic b) calculated by subtracting the approximate internal energy due to the ionization process. i) discussed in text B. Mechanisms 1. Tile and isobutene chemistry As stated before, AR2 found TiCl4+ to be unreactive with unsaturated hydrocarbons, whereas the present study found isobutene to react through an apparent exothermic pathway resulting in two product ions (the initial complex, TiCl4C4H3+, may not have been observed due to the mass 136 limitations of the quadrupole). A proposed mechanism to help explain the observed chemistry is shown in Figure 5.31. The appearance of the products can be explained by an initial ligand exchange reaction which would suggest the TiCl3+--C1 bond to be a weak bond. Ligand exchange was observed during both our study and the study performed by AR2 of the chemistry between oxygen-containing compounds and Tile-. However, when other chlorotitanium ions (TiCl3,2,1+) were used as reactants, ligand exchange was not observed. This chemistry supports the findings of the KDD study, which determined the TiCl3+--Cl bond to be the weakest chlorotitanium bond. Following the ligand exchange, the resulting complex may be in an excited state. Excess internal energy (as a result of the ionization process and/or interaction between the titanium atom and the double bond of the olefin) may then result in electron rearrangement. As a consequence of this rearrangement, HCl is eliminated, generating a complex (represented by the peak at m/z 173) consisting of a dichlorotitanium ion bonded to a hydrocarbon through a four electron, three center bond. If this final complex still contains excess energy, it may further decompose. The charge of the resulting products is governed by the ionization energies (I.E.) of the neutrals. The fragment with the lower ionization energy is the fragment most likely to be charged. As shown in Figure 5.31, I.E.(TiC12) = 9.3 eV and I.E.(C4H7) = 7.9 eV which explains why C4H7+ (m/z 55) and not 'I‘iC12+ (m/z 118) is detected. 137 C1\+ /Ti +<4:) >—C3H + HCl M I..E(C4H7)=7.9eV I..E(T1012 )=9.3eV TiCl2 12:2._&7 Figure 5.31. Proposed mechanism for the reaction between TiCl: and isobutene. 138 2. TiCl3+ and isobutene chemistry AR2 found TiCl3+ to react with isobutene through the elimination of HCl. During our investigation we found evidence of a loss of a second HCl as well. A proposed mechanism is shown in Figure 5.32. As with the chemistry of Tile, the initial complex, TiCl3C4H8+, was not observed, possibly due to the mass limitations of the quadrupole. Without the steric hindrance of a fourth chlorine, the titanium is more exposed and therefore able to react more readily. This is evidenced in the fact that AR2 did observe chemistry and reported an ion at m/z 173. Although they were able to detect an ion at m/z 173, it is of interest that they did not report an ion at m/z 209 representing TiCl3C4H8+, the assumed precursor. One explanation is that the initial complex is unstable and decomposes very quickly. The lifetime of an ion must exceed the period of one cycle to be detected in an ICR instrument. Because ICRs generally operate in the 100- 200 kilohertz range, the lifetime of an ion must exceed 5-10 microseconds. During our investigation, we also observed an ion at m/z 137. This can be explained by the elimination of HCl from the precursor represented by the peak at m/z 173. A possible mechanism is illustrated in Figure 5.32. One driving force behind this further decomposition may be the stability of the product ion. Finseth, Sourisseau, and Miller (FSM) [42] obtained complete vibrational spectra for tricarbonyl(trimethylenemethane)iron. They found that the terminal carbons of the ligand should be described as having appreciable sp2 character and that trimethylenemethane behaves as a 139 01 + '4. ..... T1013 + —> Cl—Ti'j c':1\> H H01 H01 C A 1>Ti+(\4’>—CH3 H2 Cl v m/z 173* 04H; + T1012 + HCl Figure 5.32. Proposed mechanism for the reaction between 1301; and isobutene. 140 strong ligand by making substantial bonding contributions through 1: donations. According to the authors, the bonding description of the iron complex in terms of a single bond between the metal and the central carbon of the ligand is a misleading oversimplification. It must also be pointed out that FSM found that less than 0.3% of the molecules were in the ground state at 300 K. The majority of our experiments were carried out at temperatures ranging from 298-300 K. Further decomposition was not observed possibly because of the already low ion current of the chlorotitanium-trimethylenemethane complex. The ion appearing at m/z 55 (C4H7+) is thought to be the result of the decompostion of the ion at m/z 173. This follows the same reasoning provided in the previous section concerning the chemistry between Tile and isobutene. 3. TiClz” and isobutene chemistry AR2 found TiClz+ to react with isobutene through the elimination of HCl. During our investigation, we also observed the HCl elimination. And, because the mass of the initial complex, TiClzC4H8+, is within the mass range of our quadrupole, we were able to observe a peak at m/z 174. A proposed mechanism is shown in Figure 5.33. As before, the product represented by the peak at m/z 138 can be explained by the formation of a four center, four electron bond resulting in the elimination of HCl. Because of the relatively high product ion current, further decomposition was observed. As shown in Figure 5.33, the CID i 01 —Ti+<’4‘>—0H3 + H01 \vl m/z 138* I.E.(TiCl) = 7.0 eV I.E.(C4H7) = 7.9 eV 1301+ m/z 33 + C4H7 m/z 55 Figure 5.33. Proposed mechanism for the reaction between TiCl; and isobutene. 142 ionization energies favored the formation of the TiCl” ion. Observation of this ion was complicated by the fact that 'I‘iCl” is also the CID product of the reactant ion TiC12+. We were, however, able to generate our first measurable onset (other than CID) by observing the formation of C4H7+ from the reactants Tile and C4H8. This is shown in Figure 5.29. The proposed mechanism for this reaction predicts C4H7+ is the result of the dissociation of the ion at m/z 138. Ionization energies of TiCl and C4H7 favor the formation of TiCl+ and suggest that this reaction pathway is not exothermic. Therefore, the assumed reaction is: '1‘1012+ + C4H8 —' H01 + TiCl + 04H7+ (20) Using available thermochemical information, reaction (20) is endothermic by 2.93 eV. Our experimental measurement generated a value of 1.73 eV. The low experimental value is consistent with performing experiments on "hot" reactant ions. 4. TiCl+ and isobutene chemistry AR2 found TiCl” to react with isobutene through the elimination of H2. During our investigation, we also observed a more rich chemistry through the elimination of H2, CH4, C2H5, and HCl. It is noted that even though the mass of the initial complex, TiClC4,H3+ at m/z 139, falls within the mass range of the quadruple, the ion was not observed. Repeated attempts to increase the resolution of the quadruple to separate the possible peak from 143 that appearing at m/z 137 were unsuccessful. An added complication may be that the peak is very small due to a short lifetime of the ion. A proposed mechanism is shown in Figure 5.34. Because of the relatively high product ion currents of the ions represented by the peaks at m/z 137 and 123, it is believed that there are two competing reaction pathways rather than a sequential mechanism. This competition may contribute to the small signal of the initial complex. Although there were no measurable onsets (other than by CID) it is believed that the chemistry represented by the mechanism that generates the ion at m/z 123 has some small energy barrier. This is supported by the fact that AR2 did not report the observation of this ion, even though our experiments suggest a relatively high reaction cross-section. The reaction between TiCl+ and isobutene generated a higher ion current of the proposed chlorotitanium-trimethylenemethane complex than any other reaction between a chlorotitanium ion and isobutene. Because of this higher current it is believed that observation of the decomposition of this complex is now possible. Product ions believed to be the result of this decomposition are represented by the peaks at m/z 110 and 101, however, a reasonable mechanism has yet to be proposed. IV. Conclusion It has been demonstrated that the MSU ion beam instrument can be used to generate thermochemical information from gas-phase chemistry studies. The utility of the instrument would be increased greatly by the 144 +,* C1—Ti(\4> + CH4 \ > I V m/z 123 Figure 5.34. Proposed mechanism for the reaction between TiCl+ and isobutene. 145 replacement of the quadrupole with a higher performance mass analyzer. Also, the installment of a drift cell between the source and the magnet to facilitate collisional relaxation of excited states of the reactant ion, would enable the investigation of ground state chemistry. To increase the transmission efficiency of the intrument for product ions which differ greatly in mass, programable power sources could be employed to apply potentials to the ion optics. Coupling these power sources to the quadrupole would allow for the monitoring of each ion under ideal transmission conditions. 10. 11. 12. 13. 146 LIST OF REFERENCES Chapter 5 R.W. Kiser, J.G. Dillard, D.R. Dugger, "Mass Spectrometry in Inorganic Chemistry”, Adv. Chem. Ser., No. 72, 153 (1968). RF. Lossing, A.W. Tickner, W.A. Bryce, J. Chem. Phys., 19, 1254 (1951). J .W. Warren, Nature, 165, 810 (1950). O. Kubaschewski, O. Kubaschewski-von Goldbeck, P. Rogl, H.F. Franzen, "Atomic Energy Review-Special Issue No. 9; Titanium: Physico-Chemical Properties of its Compounds and Alloys," ed. K.L. Kamarek, International Atomic Energy Agency, Vienna, Austria (1983). G. Skinner, H.L. Johnston, C. Beckett, "Titanium and Its Compounds," ed. H.L. Johnston, Herrick L. Johnston Enterprises, Columbus, OH, p. 43 (1954). U. Muller, "Inorganic Structural Chemistry," 2nd ed., John Wiley & Sons, New York, NY, p. 70 (1990). FA. Cotton, G. Wilkinson, "Advanced Inorganic Chemistry," 5th ed., John Wiley & Sons, New York, NY, p. 653 (1988). SO. Lias, J .E. Bartmess, J .F. Liebman, J .L. Holmes, R.D. Levin, and WC. Mallard, J. Phys. Chem. Refi Data, 17, Suppl. 1 (1988). K. Ervin, S.K. Loh, N. Aristov, P.B. Armentrout, J. Chem. Phys., 87 , 3593 (1983). R.S.P. Coutts, P.C. Wailes, Adv. Organomet. Chem., 9, 136 (1970). R. Feld, "The Organic Chemistry of Titanium", Butterworths, Washington, DC, (1965). C.R. Demarcay, C. R. Acad. Sci., 80, 51 (1875). K. Saigo, M. Osaki, T. Mukaiyama, Chem. Lett., 989 (1975). 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 25. 26. 27. 30. 31. 32. 147 T. Mukaiyama, Chem. Lett., 1189 (1974). T. Mukaiyama, T. Izawa, K. Saigo, Chem. Lett., 323 ( 1974). T. Mukaiyama, M. Hayashi, Chem. Lett., 15 ( 1974). J. Allison, D.P. Ridge, J. Am. Chem. Soc., 78, 163 (1977). R. Kinser, J. Allison, T.C. Dietz, M. deAngelis, D.P. Ridge, J. Am. Chem. Soc., 100, 2706 (1978). J. Allison, R.B. Freus, D.P. Ridge, J. Am. Chem. Soc., 101, 1332 (1979). J.S. Uppal, R.H. Staley, J. Am. Chem. Soc., 102, 4144 (1980). GD. Byrd, R.C. Burnier, B.S. Freiser, J. Am. Chem. Soc., 104, 3565 (1982). RC. Burnier, G.D. Byrd, T.J. Carlin, M.B. Wiese, R.B. Cody, B.S. Freiser In Ion Cyclotron Resonance Spectrometry II; Lecture Notes in Chemistry Vol. 31; H. Hartmann, K.-P. Wanczek, Eds.,Springer- Verlag, Berlin, p. 98 (1982). M.A. Tolbert, J .L. Beauchamp, J. Am. Chem. Soc., 108, 7509 (1986). T.C. Jackson, T.J. Carlin, B.S. Freiser, J. Am. Chem. Soc., 108, 1120 (1986). H. Kang, D.B. Jacobson, S.K. Shin, J.L. Beauchamp, M.T. Bowers, J. Am. Chem. Soc., 108, 5668 (1986). M.A. Tolbert, J .L. Beauchamp, J. Phys. Chem., 90, 5015, (1986). J .L. Elkind, P.B. Armentrout, Int. J. Mass Spect. Ion Processes, 83, 259 (1988). R. Tonkyn, M. Ronan, J .C. Weisshaar, J. Phys. Chem., 92, 92 (1988). LS. Sunderlin, P.B. Armentrout, Int. J. Mass Spectrom. Ion Processes, 94, 149 (1989). K.K. Irikua, J .L. Beauchamp, J. Am. Chem. Soc., 111,75 (1989). H. Higashide, T. Oka, K. Kasatani, H. Shinohara, H. Sabo, Chem. Phys. Lett., 163, 485 (1989). T.J. Carlin, M.B. Wise, B.S. Freiser, Inorg. Chem., 20, 2743 (1989). 33. 36. 37. 39. 40. 41. 42. 148 P. Braunstein, D. Nobel, Chem. Rev., 89, 1927 (1989). K. Eller, W. Zummack, H. Schwarz, J. Am. Chem. Soc., 112, 621 (1990). RB. Armentrout, Ann. Rev. Phys. Chem., 41, 313 (1990). DE. Clemmer, L.S. Sunderlin, P.B. Armentrout, J. Phys. Chem., 94, 3008(1990). P.B. Armentrout, Int. Rev. Phys. Chem., 9, 115 ( 1990). K. Eller, H. Schwarz, Chem. Rev., 91, 1121 (1991). BC. Guo, A.W. Castleman, Jr., J. Am. Chem. Soc., 114, 6152 (1992). G.M. Daly, M.S. El-Shall, J. Phys. Chem., 98, 696 ( 1994). J. Allison, D.P. Ridge, J. Am. Chem. Soc., 99, 35 (1977). D.H. Finseth, C. Sourisseau, F.A. Miller, J. Phys. Chem., 80, 1248 (1976). APPENDICES Appendix A Electron Energy Calibration The electron energy calibration was performed under identical source conditions to those under which the Mng+ CID experiments were performed. Nitrogen was used as the calibration compound. Ion currents (counts/sec) of N2+ (m/z 28) were recorded for as a function of nominal electron energy. The plot of the experimental results is shown in Figure A.1. The extrapolated onset for ion production is approximately 16.3 eV. The accepted literature value for the ionization energy of nitrogen is 15.57 eV. Therefore, a correction of -0.7 eV must be applied to the measured nominal electron energy. The nominal electron energy is determined by the potential difference between the filament and the ionization box. As the electrons boil off the filament wire, they experience an electric field that accelerates them toward the ionization box. As the electrons enter the ionization region they experience an electric field generated by the potentials applied to the pusher, ionization box, and the extraction lens. The design of the source is such that these applied potentials generate a potential energy surface resembling a saddle (discussed in chapter 2). The potential energy surface (following the trajectory of an electron from the filament to the electron collector) has a minimum value near the center of the source with the field reaching a maximum at the sides of the 149 Counts/sec 1.6E+6 a 1.4E+6 ~ 1.2E+6 4 1.0E+6 J 8.0E+5 — 6.0E+5 * 4.0E+5 J , 2.0E+5 4 . O 0.0E+0 11.1.1.1.1.Wl+4144 0 2 4 6 810121416182022 -1 Nominal Electron Energy (eV) Figure A1 Electron energy calibration performed by using nitrogen as the calibration compound. Mass monitored was that associated with m/z 28. 151 ionization box. The purpose of this field is to force positively charged particles toward the center of the source, where they are more efficiently extracted. The field experienced by the electrons, however, has an opposite effect due to the negative charge of the electron. Instead of the potential surface having a minimum at the center of the source (along the electron trajectory), it would have a maximum. Stopping potential experiments have shown that positively charged ions have been generated at potentials as much as 5 volts below the acceleration potential (the voltage applied to the ionization box). Electrons experiencing the same electric field would have to overcome a 5 volt barrier. As the electrons approach the potential energy surface apex, their kinetic energy decreases. Electrons interacting with neutral molecules near this point would be at some energy less than the potential difference between the filament wire and the ionization box. When appearance potential experiments are performed, the measured electron energy thus would be greater than the "true" electron energy, creating false high reading. Adding to this high reading would be the energy transfer efficiency between the electron and the molecule/atom being ionized. A 13 eV electron most likely would not transfer 13 eV, but some amount less than 13 eV. This would also result in a measured electron energy higher than the "true" electron energy. AppendixB Definitions P(source) Gas pressure within the source housing as measured by a Penning gauge. I(e-) Amps. Measured electron current (Amperes) at the electron collector. Used for regulating the filament electron emission. See Figure 2.2. Repeller(V) Voltage applied to the repeller. The repeller is also referred to as the pusher. Used to create a saddle-field-shaped potential energy surface within the source. See Figure 2.2. Extractor (V) Voltage applied to the extraction plate. Used to create a field that extracts ions out of the source. See Figure 2.2. Shield (V) Voltage applied to the shield plate. Used to focus the ion beam into the magnetic sector. See Figure 2.2. Source Zl,L(V) Voltage applied to one of the two 2 l-deflection plates. Used to focus the ion beam into the magnetic sector. See Figure 2.2. Source Z], R (V) Voltage applied to the second of two 2 l-deflection plates. Used to focus the ion beam into the magnetic sector. See Figure 2.2. Source Lens, L (V) Voltage applied to one of two lens plates. Used to focus the ion beam into the magnetic sector. See Figure 2.2. Source Lens, R (V) Voltage applied to the second of two lens plates. Used to focus the ion beam into the magnetic sector. See Figure 2.2. 152 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 153 SourceR(V) Voltage applied to the r-deflection plate. Used to focus the ion beam into the magnetic sector. See Figure 2.2. SourceZZ,L(V) Voltage applied to one of the two 2 2-deflection plates. Used to focus the ion beam into the magnetic sector. See Figure 2.2. Source 22,11 (V) Voltage applied to the second of two 2 2-deflection plates. Used to focus the ion beam into the magnetic sector. See Figure 2.2. Accel. Volt. Voltage applied to the ionization box. This defines the nominal potential energy of the ions. See Figure 2.2. TIC Total ion current as measured at the exit slit of the source. The current is given as the meter reading with the corresponding meter scale. The meter is on the BDG module of the CH5. Hall Hall probe value of the magnetic sector. This provides a measure of the magnetic field strength. DecelZl (V) Voltage applied to one of two z-deflection plates of the deceleration lens system. Used to focus the ion beam into the deceleration stage. See Figure 2.6. DecelZZ (V) Voltage applied to the second of two z-deflection plates of the deceleration lens system. Used to focus the ion beam into the deceleration stage. See Figure 2.6. DecelR(V) Voltage applied to the r-deflection plate of the deceleration lens system. Used to focus the ion beam into the deceleration stage. See Figure 2.6. Decel Einzel(V) Voltage applied to the middle plate of the Einzel focusing stage. Used to focus the ion beam into the deceleration stage. See Figure 2.6. 20. 21. 22. 26. 27. 29. 154 Decel Inject.(V) Voltage applied to the injection element of the deceleration stage of the deceleration lens system. See Figure 2.6. ranequencyMHz Frequency of the rf signal applied to the octopole. Reported in megahertz. rf V (ms-amp) Value of the readout on the rf power amplifier. This is the root-mean—square amplitude of the rf signal applied to the octopole. Tuner,T Setting for the first capacitor of the radio antenna tuner used to create a balance between the amplifier and High-Q head of the octopole. See chapter 2. Tuner, I Setting for the inductor of the radio antenna tuner used to create a balance between the amplifier and High-Q head of the octopole. See chapter 2. Tuner,A Setting for the second capacitor of the radio antenna tuner used to create a balance between the amplifier and High-Q head of the octopole. See chapter 2. rf Power HM-9 A measure of the forward power from the amplifier to the octopole High-Q head. Readout provided by the Heathkit Model HM-9 QRP Wattmeter. See chapter 2. SWR Standing wave ratio providing a measure of the relationship between the forward and reflected power between the amplifier and the octopole High-Q head. See chapter 2. Oct. DC bias (V) DC bias applied to the octopole. Computer controlled. See chapter 2. Transfer] (V) Voltage applied to the transfer lens 1. This lens helps to extract ions from the beam guide and focus them into the quadrupole. See Figure 2.11. 30. 31. 32. 33. 35. 36. 37. 38. 39. 40. 155 Transfer2(V) Voltage applied to the transfer lens 2. This lens helps to focus the ion beam into the quadrupole. See Figure 2.11. ’I‘ransfer3(V) Voltage applied to the transfer lens 3. This lens helps to focus the ion beam into the quadrupole. See Figure 2.11. ’I‘ransfer4(V) Voltage applied to the transfer lens 4. This lens helps to focus the ion beam into the quadrupole. See Figure 2.11. Transfer5(V) Voltage applied to the transfer lens 5. This lens helps to focus the ion beam into the quadrupole. See Figure 2.11. 'l‘ransfer6 (V) This lens was found to be unnecessary and was removed. Transfer lenses 7-9 where never renamed. Transfer7 (V) Voltage applied to the transfer lens 7. This lens is used to help focus the ion beam into the detector system. See Figure 2.11. 'l‘ransferS (V) Voltage applied to the transfer plate 8. This plate is used to help focus the ion beam into the detector system. See Figure 2.11. Transfer9 (V) Voltage applied to the transfer plate 9. This plate is used to help focus the ion beam into the detector system. See Figure 2.11. Quad. DCbias (V) DC bias applied to the quadrupole. This is manually set using a Hewlett-Packard DC Power Supply model 6110A. P(Collision Cell) Gas pressure within the collision cell as measured by a calibrated ion gauge. See chapter 2. P(chamber B) Gas pressure of the instrument chamber housing the beam guide and collision cell. The pressure is monitored by the same ion gauge used to monitor P(Collision Cell). 41. 42. 43. 44. 45. 156 Daly CD(V) Voltage applied to the conversion dynode of the Daly detector. See chapter 2. Daly PMT (V) Voltage applied to the photomultiplier tube of the Daly detector. See chapter 2. Er Energy of onset. This value is generated during the curve fitting routine. See chapter 3. n Curve fitting parameter. See chapter 3. m Curve fitting parameter. See chapter 3.