PLACE IN RETURN Box to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJClRC/DateDuep65-p. 15 EXPERIMENTAL AND THEORETICAL STUDIES OF THE ADSORPTION AND PHOTOCHEMISTRY OF DIBROMODIFLUOROMETHANE (HALON-1202) ON A MODEL CARBONACEOUS AEROSOL SURFACE By Michael John Dorko A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2003 ABSTRACT EXPERIMENTAL AND THEORETICAL STUDIES OF THE ADSORPTION AND PHOTOCHEMISTRY OF DIBROMODIFLUOROMETHANE (HALON-1202) ON A MODEL CARBONACEOUS AEROSOL SURFACE By Michael John Dorko The adsorption and photochemistry of dibromodifluoromethane, CFzBrz, on highly ordered pyrolytic graphite (HOPG) was studied by theoretical and experimental methods as a model for heterogeneous photochemical reactions occurring on carbonaceous aerosols in the upper tIOposphere and lower stratosphere. The monolayer- covered HOPG surface was characterized using temperature-programmed desorption (TPD), X-ray photoelectron spectroscopy (XPS), and electron energy loss spectroscopy (EELS) both before and after broadband ultraviolet (UV) irradiation from ~ 225-350 nm. In the absence of UV radiation, molecular adsorption and desorption of CFzBrz was observed (~ 140 K) while irradiation of the monolayer resulted in photodissociation of the adsorbate to yield CFzBr and Br atoms. The most prevalent reaction observed in these experiments was the recombination of a large fraction of the photolyzed molecules (~70%) to reform CFzBrz during TPD measurements. Minor channels revealing the formation of Brz (~6%) and C2F4Br2 (~l.4%) were present as well. Comparison of the estimated integrated photodissociation cross section (1.9 x 10'19 cm2) to the cross section calculated from the TPD data (4.5 x 10'19 cm2) suggested that photodissociation of CF2BI'2 occurred by direct photoabsorption by the adsorbate and not by dissociative electron attachment mediated by the surface. Differences in the distributions of Br; and C2F4Br2 when compared to results of gas phase and matrix isolation experiments were attributed to possible structural constraints or reduced surface diffusion for one or more of the adsorbates due to the high density of photogenerated species trapped in the surface layer. Theoretical studies of the preferred adsorption sites and orientations of CFzBrz and CFzBr on the HOPG surface were examined using a cluster model approach to gain additional insight into the experimental data. Second-order Moller-Plesset perturbation theory (MP2) calculations were performed using 6-31G* and 6-31G(2d) basis sets and a single layer slab approximation consisting of either a four (pyrene) or seven (coronene) ring polycyclic aromatic hydrocarbon to model the HOPG surface. The preferred adsorption site for CFzBrz was found to be located at the hollow site of a six-membered ring for each combination of surfaces and basis sets examined. The interaction between the molecule and the surface was dipolar in nature with a small amount of charge transfer occurring from the surface to the adsorbate as determined from a Mulliken population analysis. For the CFzBr radical, different preferred adsorption sites were found for the four and seven ring substrates independent of the basis set used. The atop site was the preferred location on the four ring system while the bridge site was preferred on the seven ring system. The radical favored these sites because it contains a singly occupied molecular orbital to which electron density can be added stabilizing it on the surface. Rotation of both species on the seven ring surface revealed binding energies similar to those Obtained for the starting geometries suggesting that free rotation occurs at each adsorption site. These results were used to rationalize the product distributions Obtained from the TPD experiments and the decrease in background intensity of the EELS data after UV irradiation. To the one who brought me into the world and the one who takes care of me now. ACKNOWLEDGMENTS As just about everyone knows the completion of a thesis involves more than just the person whose name appears under the title. It is for these people that I wish to give thanks. First and foremost, I have to thank God. There were numerous times along this journey when I thought I would not make it. Through His strength, guidance, and love, I found a way to persevere and see this process to completion. I would like to thank my adviser, Dr. Simon J. Garrett, for allowing me the opportunity to join the group and giving me the freedom to delve into the world of theory. His door was always open for helpful discussions, advice, and chats about things other than research. I can now proudly say that I use “fiddly bits” and “bloody” as a regular part of my vocabulary. The members of my committee also deserve a round of thanks as well. Drs. Marcos Dantus, Lynmarie Posey, and Piotr Piecuch often provided helpful suggestions in all areas of my graduate career. I am grateful for the time and, at times, overwhelming amounts of knowledge and advice that they freely gave to me. I would also like to acknowledge the members of the Garrett group during my tenure (Todd, Lili, Heather, and Jason) for the numerous conversations that we had about work, politics, and anything else that was of current interest. Their support and friendship were valuable during both difficult as well as good times. All the members of both the machine (Sam, Glenn, and Tom) and electronics (Dave, Scott, and Ron) shops need a lot of thanking. From laser emergencies to meltdowns of the liquid nitrogen reservoir, they handled all that I could throw at them and more in the timeliest of fashions. Along the way, you meet many people but there are only a few that you can call truly good friends. I would be extremely remiss if I didn’t thank John and Mike for their friendship for the last six years. From our exploits in various fantasy leagues, MSU hockey games with the crew, and discussions during lunches at the International Center, I feel like I have gained a couple of brothers who I know I can always count on. GO GREEN! GO WHITE! My family (both biological and extended) has had an indirect role in the completion of this thesis as well. Even though they usually had no clue as to what I was doing and think that graduate students work forty hours a week, they would always express much needed love and support whenever I would talk to them. I greatly appreciate all that they have done for me. Last, but most certainly not least, I must thank my wife Holly for every single thing she has ever done for me. I know that I’ll never be able to truly express my thanks enough but I’m happy that she gave me the privilege of having a lifetime with her to try. With her love and support, I know that anything is possible. vi TABLE OF CONTENTS LIST OF TABLES ............................................................................................................. ix LIST OF FIGURES ............................................................................................................. x KEY TO SYMBOLS AND ABBREVIATIONS ............................................................. xiv CHAPTER 1 INTRODUCTION ............................................................................................................... l 1.1 Heterogeneous Atmospheric Chemistry ............................................................ 1 1.2 Aerosol Properties ............................................................................................. 1 1.3 Role of Carbonaceous Aerosols in Heterogeneous Atmospheric Chernistry.... 5 1.4 Halons ............................................................................................................... 9 1.5 Proposed System of Study .............................................................................. 11 1.6 Heterogeneous Reactions of Br-Containing Molecules .................................. 12 1.7 Theoretical Studies of Heterogeneous Atmospheric Chemistry ..................... 13 References ............................................................................................................. 16 CHAPTER 2 EXPERIMENTAL ............................................................................................................ 22 2.1 Ultrahigh Vacuum (UHV) Apparatus ............................................................. 22 2.2 Sample Preparation and Dosing ...................................................................... 23 2.3 X-ray Photoelectron Spectroscopy (XPS) ....................................................... 23 2.4 Temperature-Programmed Desorption (TPD) ................................................ 26 2.5 Electron Energy Loss Spectroscopy (EELS) .................................................. 30 2.6 Photochemistry ................................................................................................ 31 References ............................................................................................................. 32 CHAPTER 3 THEORETICAL METHODS ........................................................................................... 33 3.1 Introduction ..................................................................................................... 33 3.2 Perturbation Theory ........................................................................................ 33 References ............................................................................................................. 37 CHAPTER 4 ADSORPTION AND PHOTOCHEMISTRY OF CFzBrz (HALON 1202) ON HIGHLY- ORDERED PYROLYTIC GRAPHITE (HOPG) ............................................................. 38 Abstract ................................................................................................................. 38 4.1 Introduction ..................................................................................................... 39 4.2 Results and Discussion .................................................................................... 40 4.2.1 Temperature-Programmed Desorption (TPD) ................................. 40 4.2.2 X-ray Photoelectron Spectroscopy (XPS) ........................................ 47 4.2.3 Electron Energy Loss Spectroscopy (EELS) ................................... 50 4.2.4 Adsorption and Desorption of CFzBr2(ad)/HOPG ........................... 53 vii 4.2.5 Photolysis of CFzBrz on HOPG ....................................................... 55 4.2.6 Surface Recombination Reactions ................................................... 57 4.2.7 Loss of Fluorine Species .................................................................. 60 4.2.8 Quantification of Surface Reactions ................................................ 62 4.3 Conclusions ..................................................................................................... 65 References ............................................................................................................. 67 CHAPTER 5 THEORETICAL INVESTIGATION OF THE ADSORPTION SITES AND ORIENTATION OF CFzBrz AND CFzBr ON MODEL GRAPHITE SURFACES ........ 71 Abstract ................................................................................................................. 71 5.1 Introduction ..................................................................................................... 72 5.2 Computational Details ..................................................................................... 74 5.3 Results and Discussion .................................................................................... 80 5.3.1 Adsorption Sites and Geometries of CFzBrz .................................... 82 5.3.2 Adsorption Sites and Geometries of CFzBI' ..................................... 96 5.4 Conclusions ................................................................................................... 110 References ........................................................................................................... 1 11 CHAPTER 6 CONCLUSIONS AND FUTURE WORK ..................................................................... 116 6.1 Experimental ................................................................................................. 116 6.2 Theoretical ..................................................................................................... 123 References ........................................................................................................... 127 APPENDIX ..................................................................................................................... 130 APPENDIX A ................................................................................................................. 131 viii LIST OF TABLES Table 5.1 Comparison of experimental and calculated frequencies with various basis sets (absolute and relative differences between calculation and experimental data shown in brackets) ............................................................................................................................ 81 Table 5.2 Atomic charges, dipole moments and charge differences for CFzBrz on each surface ............................................................................................................................... 93 Table 5.3 CFzBr atomic charges, dipole moments, and total charges on each surface. 106 ix LIST OF FIGURES Images in this dissertation are presented in color. Figure 1.1 Types and sources of some common atmospheric aerosols .............................. 2 Figure 1.2 Schematic showing the growth modes, sources and size distribution of atmospheric aerosols ........................................................................................................... 3 Figure 1.3 Formation and composition of carbonaceous aerosols ..................................... 7 Figure 1.4 Schematic of gas-phase and heterogeneous reactions of brominated compounds in the atmosphere. Thicker lines indicate heterogeneous processes ............. 10 Figure 2.1 Surface sensitivity enhancement by variation of the electron take-off angle. 25 Figure 2.2 Surface coverage, desorption rate, and temperature dependence on time during TPD ........................................................................................................................ 28 Figure 2.3 TPD simulations of zero (a), first (b), and second (c) order desorption processes ............................................................................................................................ 29 Figure 4.1 Temperature-programmed desorption data (m/z = 129, CFzBr+) for increasing exposures of CFzBrz on HOPG. Exposures were (a) 0.4 langmuir, (b) 1.5 langmuir, (c) 2.5 langmuir, (d) 3.6 langmuir, (e) 4.5 langmuir, (f) 5.1 langmuir, (g) 6.1 langmuir, (h) 8.1 langmuir, and (i) 10.1 langmuir. Inset shows integrated area Of the feature at ~ 140 K (monolayer) as a function of exposure .............................................................................. 41 Figure 4.2 Temperature-programmed desorption data for 4.8 langmuir CFzBrz on HOPG. Monitored masses were m/z = 31 (CF), 50 (CFf), 79 (Br+), 100 (C2F4+), 129 (CFzBr+), 160 (Brf) and 179 (C2F4Br+) ............................................................................................ 43 Figure 4.3 Temperature-programmed desorption data for 4.8 langmuir CFzBrz on HOPG after 5 minutes of UV irradiation (filtered Hg-arc, 225-350 nm, incident power 8.6 mW/cmz). Monitored masses same as in Figure 4.2 ........................................................ 45 Figure 4.4 Temperature-programmed desorption data for 4.8 langmuir CFzBrz on HOPG after 15 minutes of UV irradiation. Conditions same as in Figure 4.3 ............................ 46 Figure 4.5 X-ray photoelectron spectra of F ls region (a) and Br 3d region (b) of CFzBrz monolayer as a function of UV irradiation. Vertical lines are drawn at 688.4 eV and 71.6 eV BE ................................................................................................................................ 48 Figure 4.6 (a) X-ray photoelectron spectra of the C Is region of bare HOPG (solid line), 4.8 langmuir CFzBrz-covered HOPG (solid circles) and the normalized difference spectrum (crosses). (b) X-ray photoelectron spectra of the C ls region of 4.8 langmuir CFzBrz-covered HOPG (solid line), 4.8 langmuir CFzBrz-covered HOPG after 30 min UV irradiation (solid circles) and the normalized difference spectrum (crosses) ............. 49 Figure 4.7 Integrated area ratios (F ls/C ls and Br 3d/C ls) from X-ray photoelectron spectroscopy of 4.8 langmuir CFzBrz adsorbed on HOPG, as a function of UV irradiation time (conditions as in Figure 4.3). The solid curve in the F ls/C ls ratio is a single- exponential fit to the data. The solid curve in the Br 3d/C ls ratio is a linear fit to the data .................................................................................................................................... 51 Figure 4.8 High-resolution electron energy loss spectra of (a) clean HOPG, (b) 4.8 langmuir CFzBrz-covered HOPG and (c) 4.8 langmuir CFzBrz-covered HOPG after 30 minute UV irradiation ....................................................................................................... 52 Figure 4.9 Relative percentages of the different reaction channels present on the surface as a function of UV irradiation. The lines through the data are guides for the eye ......... 63 xi Figure 5.1 Four (a) and seven ring (b) model graphite surfaces with adsorption positions indicated. The atop site occurs over a substrate atom, the bridge site is between to substrate atoms and the hollow site is at the center of a six-membered substrate ring ..... 75 Figure 5.2 CFzBrz on the four ring model surface at the (a) atop, (b) hollow, and (0) bridge sites. Carbon atoms are green, hydrogen atoms are white, bromine atoms are purple, and fluorine atoms are gray in all figures ............................................................. 78 Figure 5.3 CFzBr on the four ring model surface at the (a) atop, (b) hollow, and (0) bridge sites ......................................................................................................................... 79 Figure 5.4 Potential energy curves for CFzBrz approaching the four ring substrate at the MP2/6-31G* level of theory ............................................................................................. 83 Figure 5.5 Potential energy curves for CFzBrz approaching the four ring substrate at the MP2/6-31G(2d) level of theory ......................................................................................... 85 Figure 5.6 CFzBrz adsorbed at the (a) atop, (b) hollow, and (c) bridge positions on the model seven ring surface ................................................................................................... 87 Figure 5.7 Rotated CFzBl‘z adsorbed at the (a) atop, (b) hollow, and (c) bridge positions on the model seven ring surface ........................................................................................ 88 Figure 5.8 Potential energy curves for CFzBrz approaching the seven ring substrate at the MP2/6-31G* level of theory ............................................................................................. 89 Figure 5.9 Comparison of the effects of basis sets on the binding energies and adsorption sites for CFzBrz on the model seven ring surface. The squares denote the atop site, circles denote the hollow site, and the triangles denote the bridge site ........................................ 91 Figure 5.10 Potential energy curves for CF2Br approaching the four ring substrate at the MP2/6-31G* level of theory ............................................................................................. 97 xii Figure 5.11 Potential energy curves for CF2Br approaching the four ring substrate at the MP2/6-3 1G(2d) level of theory ......................................................................................... 99 Figure 5.12 CFzBr adsorbed at the (a) atop, (b) hollow, and (c) bridge positions on the model seven ring surface ................................................................................................. 100 Figure 5.13 Rotated CF2Br adsorbed at the (a) atop, (b) hollow, and (c) bridge positions on the model seven ring surface ...................................................................................... 101 Figure 5.14 Potential energy curves for CFzBr approaching the seven ring substrate at the MP2/6-31G* level of theory ........................................................................................... 103 Figure 5.15 Comparison of the effects of basis sets on the binding energies and adsorption sites for CFzBr on the model seven ring surface. The squares denote the atop site, circles denote the hollow site, and triangles denote the bridge site ......................... 104 Figure 6.1 Band structures of an insulator, metal, semimetal, and semiconductor. Shaded areas represent filled bands ............................................................................................. 117 xiii KEY TO SYMBOLS OR ABBREVIATIONS EC = Elemental Carbon OC = Organic Carbon CFCs = Chlorofluorocarbons HOPG = Highly Ordered Pyrolytic Graphite UV = Ultraviolet PSC = Polar Stratospheric Cloud PAH = Polycyclic Aromatic Hydrocarbon UHV = Ultrahigh Vacuum QMS = Quadrupole Mass Spectrometer LN; = Liquid Nitrogen XPS = X-ray Photoelectron Spectroscopy TPD = Temperature-Programmed Desorption EELS = Electron Energy Loss Spectroscopy FWHM = Full Width at Half Maximum CSF = Configuration State Function HF = Hartree Fock SCF = Self-Consistent Field LCAO = Linear Combination of Atomic Orbitals MO = Molecular Orbital STO = Slater-Type Orbital GTO = Gaussian-Type Orbital xiv CGTO = Contracted Gaussain-Type Orbital CI = Configuration Interaction CISD = Configuration Interaction Singles and Doubles MPn = M ller-Plesset Perturbation Theory of Order n CC = Couple Cluster DFT = Density Functional Theory LDA = Local Density Approximation LSDA = Local Spin Density Approximation GGA = Generalized Gradient Approximation VDW = Van Der Waals ODP = Ozone Depletion Potential BE = Binding Energy MD = Molecular Dynamics IRAS = Infrared Absorption Spectroscopy DEA = Dissociative Electron Attachment UPS = Ultraviolet Photoelectron Spectroscopy LDOS = Local Density of States CASPT2 = Complete Active Space Second-order Perturbation Theory MRCISD = Multi-Reference Configuration Interaction Singles and Doubles XV Chapter 1 Introduction 1.1 Heterogeneous Atmospheric Chemistry. The Earth’s atmosphere is a complex mixture containing all three phases of matter. While the atmosphere is largely gaseous in nature, a small portion contains liquid and solid particulate material. These particulates can form stable suspensions in the atmosphere and are known as aerosols.l It has been determined that reactions occurring in or on aerosols are responsible for such processes as acid rain2 and the destruction of stratospheric ozone above Antarctica.3 (see Figure 1.1)4 Also, these particulates have been shown to contribute to such processes as the scattering or absorption of solar radiation, the formation of cloud condensation nuclei and can serve as possible sites for heterogeneous chemical reactions."4 A majority of the reactions that occur in the Earth’s atmosphere are between gas- phase molecules. However, gas-phase molecules can adsorb on the surfaces of liquid droplets or solid particles where they can undergo reactions with each other or with the substrate. Reactions of this type are termed heterogeneous because they are limited to surface reactions as opposed to those that take place in the bulk.5 The mechanisms by which heterogeneous chemical reactions take place are not as well understood as those in the gas phase, and a need exists to be able to quantify their role in such processes as the photochemical production of air pollution, climate changes and catalytic ozone destruction.l 1.2 Aerosol Properties. Aerosols encompass a wide variety of sizes and chemical compositions depending upon their source and history in the atmosphere (see Figures 1.1 and 1.2).1'4 Typical aerosol diameters range in size from ~ 0.002 to ~ 100 um Combustion Process Emissions Primary OC-EC so2 Emissions HO 2 Gas-Phase Photochemistry Primary HCl 3 (IV) H280. H280, Emissions \ \ HCl 1180;, sto, - H’ so '2 Cl+ Primary OC—EC ' 3 NH ‘ NH: Na NH‘K, OH' 3 Emissions Sea-Salt Emission Dust,Fly Ash Condensible Secondary Metals . oc Organics 1 NO{,H* T Ca+2’ Mg+2 Gas-Phase FefiletC- HNOL‘LQas—Phase Photochemistry Photochemistry “\\‘Gaseous Or anics // . . g Dust,F1y Ash NO? . Em1881ons Emissions EmlSSlonS o , 4 Figurel.l Types and sources of some common atmospheric aerosols. Chemical conversion of gases to low WW We Condeiisation [Primary vparticles] Coagulation Condensation growth [Chain aggregates duet + Chemical conversion Emissions of gases to low volatility vapors Sea Spray _ . Coagulation Volcanoa I Low volatility I Coagulation Plant p+artlcles Homogeneous nucleation /\ I I / Sedimentation // L L 0.001 0.01 0.1 1 .2 10 100 Particle diameter (um) Ultra fine 3%.? w: Accumulation Mechanicaly generated particles | range ' range t aerosol range * 4 Fine particles t 1 Coarse particles —" Figure 1.2 Schematic showing the growth modes, sources and size distribution of atmospheric 1 aerosols. in the troposphere (10-15 km altitude) and stratosphere (~ 20-50 km altitude) with the largest fraction occurring between 0.002 and 10 urn.l The lower bound of this range is approximate due to the lack of a set definition that distinguishes between a cluster of molecules and an aerosol6 while the upper bound corresponds to aerosols whose size is comparable to fine sand or pollen.l Classification of aerosols according to their size distributions can reveal some insight into the process by which they were formed and their chemical composition. A typical classification scheme based on mass or volume divides particles into three different ranges: nuclei (d < 0.1 um), accumulation (0.1 < d < 1 pm) and coarse (d > 1 um).7 Aerosols that are classified as belonging to the nuclei mode are usually formed by condensation of hot vapors in high temperature combustion processes and are emitted directly into the atmosphere (primary aerosol) or are formed from gas-to-particle conversion (secondary aerosol) depending on atmospheric conditions. They are typically present in large concentrations (~10-1000 cm’3)4 but account for only a small amount of the total aerosol mass. These aerosols do not have long atmospheric lifetimes because they are often involved in processes that form larger aerosols. Hence, their role in heterogeneous atmospheric processes is likely to be restricted, at best. Accumulation range aerosols are formed during combustion processes by the condensation of hot vapors onto nuclei mode aerosols and by coagulation of small nuclei mode aerosols with themselves or with larger particles."‘7 While high temperature combustion processes can form both nuclei and accumulation aerosols, the predominant type that is formed depends upon the type of fuel, combustion method (spark, explosion, etc.), and the amount of atmospheric mixing the particles undergo after they are formed.1 Aerosols in this size range have been found to contain a significant fraction of organic matter as well as some soluble inorganics such as ammonium, nitrate, and sulfate ions."4 Accumulation mode aerosols are usually present in smaller concentrations compared to nuclei mode aerosols; however, accumulation mode aerosols account for a larger fraction of the total aerosol mass (~50%).l These aerosols tend to have the longest atmospheric lifetimes because of inefficient methods of particle removal,7 such as washout during precipitation events or by dry deposition due to eddy diffusion and advection."4 Due to their long atmospheric lifetimes, these aerosols provide the greatest contribution to heterogeneous atmospheric chemistry, cloud formation, air pollution, etc. Aerosols in the coarse particle range are produced by a variety of mechanical processes such as grinding, erosion, or wind. These aerosols usually contain inorganic material (metal oxides, heavy metals, sea salt, etc.) due to their method of production. Coarse aerosols do not have long atmospheric lifetimes because sedimentation and washout by precipitation occurs rapidly for aerosols of this size. As a result, their atmospheric concentration is on the order of ~ 1 cm'3 or less and their role in atmospheric heterogeneous reactions is limited.4 1.3 Role of Carbonaceous Aerosols in Heterogeneous Atmospheric Chemistry. While a number of chemically diverse aerosols are present in the atmosphere, the most abundant particulate type over continental landmasses is carbonaceous aerosolf”9 Two different types of carbonaceous aerosols are present in the atmosphere: primary (or elemental) and secondary. Primary carbonaceous aerosols are mainly produced during incomplete combustion of biomass and fossil fuels. Other sources of primary carbonaceous aerosols come from chemical, geological and biogenic sources. They are composed of tiny spherules of impure graphite that are emitted directly into the atmosphere.7 These spherules eventually combine to form the visible clusters termed soot and smoke (See Figure 1.3).1 Secondary carbonaceous aerosols are formed by the coagulation of the photo-oxidation products of low vapor pressure hydrocarbons or from emissions by vegetative matter. As a newly formed primary carbonaceous aerosol ages in the atmosphere, it will become coated with secondary carbon increasing the complexity of the surface. The resultant aerosol can then be viewed as consisting of a small core of elemental carbon covered by a liquid-like layer of secondary carbon.lo A variety of different functional groups (quinones, carboxylic acids, ethers, etc.)”'12 will be present on the surface due to both the oxidizing nature of the atmosphere and as a result of the process by which it was formed.9 The role that carbonaceous aerosols play in heterogeneous atmospheric reactions ”'33 Because has been the subject of numerous studies over the past two decades. carbonaceous aerosols are often quite diverse in the atmosphere, graphite and amorphous carbon have been used as model surfaces in addition to natural aerosol samples. A large portion of the work done in this area has involved characterizing heterogeneous reactions of atmospherically important gases (802, 03, HNO3, N02, and H2804) and assessing what role the surface played in the reaction. For example, studies of the interaction between 802 and various carbon surfaces revealed that little or no reaction took place implying that carbonaceous aerosols are probably not a major sink for 802.”29 The addition of water to this system revealed no change in SO; oxidation but an increase in the uptake coefficient for H20 was observed. ”'30 Studies of the reaction of ozone with Solid elemental carbon spheres ': a .:. .0} Gas phase 4,513" organics ' Surface-absorbed / adsorbed organics . . . . 1 Figure 1.3 Formation and composrtron of carbonaceous aerosols. carbonaceous surfaces have shown that rapid uptake of O3 occurred and that reaction with the surface resulted in 02, CO, and C02 being produced as seen in equations 1.1— 1.3.31'33 O3+Csoo.—>02+co(g) (1.1) 203 -i- C500; —) 302 + C500! (1.2) Csoot oxidized + 03 —> Csoot reduced + 2 C02 (13) These investigations showed that increasing ozone concentrations poison the surface to further reaction and a number of surface bound carbonyl-containing species were produced. Reactions of N02 with various graphite surfaces have shown that uptake of N02 occurred rapidly with several surface species (C-NOz, C-ONO, and C-N-NOz) being formed.15'“5'21 Reduction of N02 to HONO has been found to occur on most graphite surfaces in the presence of water (see equations 1.4—1.5). NO; + CSOOJHZOWS) —> HONO + CSoot oxidized (1.4) 2N02 + HzOmS) ——> HONO + HNO3 (1.5) The amount of HONO that was produced varied but was not enough to account for the high levels of HONO observed in urban air parcelsm’lg‘z"26 Comparison between the various studies conducted to date is difficult due to the wide variety of model or collected surfaces that have been used. While most studies involving heterogeneous reactions on carbonaceous aerosols have focused on using common atmospheric pollutants, fully halogenated molecules have been studied less intensively. Chlorofluorocarbons (CFCs) are species that contain only carbon, chlorine and fluorine and are non-toxic and chemically inert. These species were used for such applications as propellants, refrigerants, and blowing and cleaning agents before being restricted by the Montreal Protocol in 1987.34 Chlorofluorocarbons have been shown to be involved in heterogeneous atmospheric reactions that result in stratospheric ozone depletion as well as being greenhouse gases with long tropospheric lifetimes."2 Knorr and co-workers have studied Chlorofluorocarbons (mainly CFzClz and its analogs) adsorbed on graphite surfaces utilizing X-ray diffraction and infrared absorption spectroscopy techniques.”39 Their studies were focused mainly on phase transitions and structural features of CFC monolayers as model two-dimensional systems. The relevant literature from this group is discussed below. 1.4 Halons. Halons are an additional series Of halogenated molecules with properties similar to CFCs. These molecules also contain carbon, chlorine, and fluorine as well as one or more bromine atoms. Halons have been used for fire and explosion suppression, in gas and oil production, and in military aircraft applications and play a similar role in the stratosphere as CFCs in terms of catalytic ozone destruction.l Their removal from the atmosphere occurs by short wavelength photolysis in the upper troposphere and lower stratosphere. Halons have higher ozone depletion potentials than CFCs because Br atoms do not react as rapidly with organic molecules as Cl atoms (these reactions provide for temporary removal of halogen atoms from the ozone destruction cycle) and have larger photodissociation cross sections in the visible region of the solar spectrum. A diagram of both the gas-phase and heterogeneous reactions of brominated compounds in the atmosphere is shown in Figure 1.4.1 Studies Of the interaction of Halon molecules with surfaces have not received as much attention as those of CFCs. A small number of investigations have focused on the adsorbate structure and phase transitions Of CF3BI‘.35’40 Knorr and co-workers have ?? hv OPP) HOCI (het) CIN02 (i161) * BION02 H20 (het) HCI (het) HCI (bet) BrCI Figure 1.4 Schematic of gas-phase and heterogeneous reactions of brominated compounds in the atmosphere. Thicker lines indicate heterogeneous processes. 10 hi dd 02.; C0” performed experiments similar to those for CFCs adsorbed on graphite. It was determined that CF3Br probably resides with the F3 tripod in contact with the surface and the monolayer formed a structure that was commensurate with the graphite lattice.35’40 Robinson, et. al. have examined the decomposition of CFzBrz (and CFzClz) on alumina surfaces at stratospheric temperatures.“ The alumina surfaces were used as a model for solid-propellant rocket motor exhaust particles. They found that dissociative chemisorption of both species occurred resulting in a surface-bound halomethyl fragment and ejection of halogen-containing species during decomposition. Pre-dosing the alumina surface with water was found to deactivate the surface only moderately with respect to halogen uptake. 1.5 Proposed System of Study. The system presented in this thesis is the adsorption and photochemistry of dibromodifluoromethane, CFzBrz, on the surface of highly ordered pyrolytic graphite (HOPG). The HOPG surface will serve as a model for primary carbonaceous aerosol as well as a source of sub-vacuum electrons or photoelectrons upon exposure to ultraviolet (UV) radiation. It is well known that UV irradiation of metal and semiconductor surfaces produces sub-vacuum or photoelectrons that may initiate chemical reactions through dissociative electron attachment.42 Conversely, direct absorption of UV radiation by the adsorbate can occur and the resulting photochemistry will be the similar to that observed in the gas phase. A determination of which mechanism is operative will help elucidate the possible role of carbonaceous aerosols in heterogeneous atmospheric chemistry of bromine-containing compounds. Also, an examination of the nascent photoproducts will reveal if the HOPG ll surface traps or releases Br atoms and whether additional reaction products are generated on the surface. Dibromodifluoromethane was chosen as the adsorbate in this study for a variety of reasons. This molecule currently is not regulated by the Montreal Protocol of 1987,34 and its concentration in the atmosphere has been increasing at a rate of 17% a year.43 Additionally, it has a short tropospheric lifetime of ~ 1-3 years, which allows for the possibility of significant amounts of Br atoms to be introduced into the atmosphere quickly. The initial step in the photodissociation of this molecule between 200-300 nm is the cleavage of a C-Br bond by a single photon."""59 CF2Br2 + hv —> CFzBr + Br (1.6) Additional photodissociation pathways are available that result in the second C-Br bond breaking and the production of C2F4Br2 and Br2.4648' 50‘55’59 Each of these products can be involved in the destruction of stratospheric ozone as well. 1.6 Heterogeneous Reactions of Br-Containing Molecules. Numerous studies have been performed examining the interaction between Br-containing molecules and such atmospherically important surfaces as polar stratospheric clouds (PSCs) and sulfuric acid-water surfaces.“67 The main effect Of heterogeneous reactions involving bromine on these surfaces is to increase the conversion of Cl reservoir species to produce free C1 atoms that participate in catalytic ozone destruction. The Br atoms facilitate this by forming products that are easily photolyzed by visible solar radiation after reaction occurs on a surface that contains a chlorine reservoir species or by generating OH radicals that can react with HCl to produce free Cl atoms. 12 The most common brominated reservoir species in the upper troposphere and lower stratosphere are HBr, BrO, HOBr, and BrONOz.60 Of these species, HOBr and BrONOz are the ones most responsible for catalytic destruction of stratospheric ozone. The formation of BrONOz occurs in the gas phase from the reaction of BrO and N02. When it becomes adsorbed on a PSC or sulfuric acid-water surface, destruction occurs in the following manner: BrONOz + H20 —9 HOBr + HN03 (1.7) This reaction produces HOBr, which is easily photodissociated in the visible region of the solar spectrum, providing a source Of OH radicals and Br atoms. The HOBr that is formed can be heterogeneously converted into BrCl if HCl is present on the PSC. 1.7 Theoretical Studies of Heterogeneous Atmospheric Chemistry. A majority of theoretical calculations investigating heterogeneous atmospheric reactions have studied interactions between chlorine—containing species and water clusters as model PSC surfaces to obtain mechanistic information about processes that are involved in stratospheric ozone destruction.68’74 McNamara and Hillier examined the hydrolysis reactions of XON02 (X=Cl, Br) and N205 on bare water clusters and ClONOz and N205 with HCl-containing water clusters as shown in reactions 1.8-1.11.“72 XON02 + H20 —> HOX + HNO3 (1.8) N205 + H20 -—> 2 HNO; (1.9) ClONOz + HCl ——> C12 + HN03 (1.10) N205 + HCl -—> ClNOz + HNO3 (1.11) They observed that the mechanism for each reaction occurred by coupled proton transfer/8N2 nucleophilic attack on ClONOz or N205 by a water molecule contained 13 within the cluster. As the size of the water cluster increased, most reactions were observed to change from a mechanism where molecular products were formed to an ionic mechanism where solvation of ionic products occurred. It also was observed that the barrier heights for the reactions decreased with an increasing number of waters in the cluster and most reactions proceeded spontaneously at stratospherically relevant temperatures. Solvation of the ions that are formed in the larger water clusters appeared to facilitate the reactions. Bianco and Hynes studied reactions 1.7 and 1.8 and arrived at similar conclusions.73 Studies of other chlorine-containing molecules (HOCl, HCl, and C12) interacting with model PSC surfaces have been performed as well.74 These studies have shown that HOCl and HCl interact more strongly with the water surface than C12 due to the dipole-dipole interaction present between these species. Most theoretical studies on graphite surfaces have been limited to determining binding energies and preferred adsorption sites for atoms and diatomic molecules.”88 Cluster models have been used for the majority of these studies. Graphite surfaces are typically approximated by the use of a single graphene layer or by a polycyclic aromatic hydrocarbon (PAH) due to the weak interaction between graphite sheets.75"76‘73'79'8”33'85 Properties calculated using these cluster models often agreed well with experimental values. To date, there have been no theoretical studies investigating interactions between brominated molecules and atmospherically significant surfaces. A variety of studies have been performed using molecular simulations to model the adsorption of numerous small molecules on graphite surfaces.”101 These adsorption studies have used extended graphite surfaces, graphite slit pores as well as carbon nanotubes to address such phenomena as structural and phase transitions of mono- and 14 multi-layers and diffusion on the surface. Pair-wise potentials, usually of the Lennard- J ones form, were used to describe the interaction between the molecule and the surface as well as between the carbon atoms within the surface. Results of these simulations often provided a clearer physical picture of the processes (growth mode, orientational ordering, etc.) occurring on the surface and compared well with experimental data. Thus far, no simulations have been carried out on the structure, growth mode, or reactions of Halons on graphite surfaces. 15 8. 9. References . Finlayson-Pitts, B. J.; Pitts, J. N., Jr. Chemistry of the Upper and Lower Atmosphere; Acdemic Press: San Diego, 2000. Finlayson-Pitts, B. J.; Pitts, J. N., Jr. Atmospheric Chemistry: Fundamental and Experimental Techniques; John Wiley and Sons: New York, 1986. Solomon, S. Nature 1990, 347, 347. Pandis, S. N.; Wexler, A. S.; Seinfeld, J. H. J. Phys. Chem. 1995, 99, 9646. . Ravishankara, A. R.; Longfellow, C. A. Phys. Chem. Chem. Phys. 1999, I, 5433. Wayne, R. P. Chemistry of Atmospheres; Oxford University Press: Oxford, 1991. Seinfeld, J. H.; Pandis, S. N. Atmospheric Chemistry and Physics; John Wiley and Sons: New York, 1998. Andreae, M. O.; Crutzen, P. J. Science 1997, 276, 1052. Lary, D. J .; Shallcross, D. E.; Toumi, R. J. Geophys. Res. 1999, 104, 15929. 10. McDow, S. R.; Jang, M.; Hong, Y.; Kamens J. Geophys. Res. 1996, 101, 19593. 11. Ahkter, M. S.; Chughtai, A. R.; Smith. D.M. Applied Spec. 1985, 39, 143. 12. Ahkter, M. s.; Chughtai, A. R.; Smith. D.M. Applied Spec. 1985,39, 154. 13. Rogaski, C. A.; Golden, D. M.; Williams, L. R. Geophys. Res. Lett. 1997, 24, 381. 14. Lary, D. J.; Lee, A. M.; Toumi, R.; Newchurch, M. J.; Pirre, M.; Renard, J. B. J. Geophys. Res. 1997, 102, 3671. 15. Smith, D. M.; Akhter, M. S.; Jassin, J. A.; Sergides, C. A.; Welch, W. F.; Chughtai, A. R. Aerosol Sci. Tech. 1989, 10, 311. 16. Kirchner, U.; Scheer, V.; Vogt, R. J. Phys. Chem. A 2000, 104, 8908. 17. Saathoff, H.; Naumann, K.-H.; Riemer, N .; Kamm, S.; MOhler, O.; Schurath, U.; Vogel, H.; Vogel, B. Geophys. Res. Lett. 2001, 28, 1957. 16 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. Chughtai, A. R.; Jassim, J. A.; Peterson, J. H.; Stedman, D. H.; Smith, D. M. Aerosol Sci. Tech. 1991, 15, 112. Schurath, U.; Naumann, K.-H. Pure and Appl. Chem. 1998, 70, 1353. Al-Abadleh, H. A.; Grassian, V. H. J. Phys. Chem. A 2000, 104, 11926. Akhter, M. S.; Chughtai, A. R.; Smith, D. M. J. Phys. Chem. 1984, 88, 5334. Gerecke, A.; Thielmann, A.; Gutzwiller, L.; Rossi, M. J. Geophys. Res. Lett. 1998, 25, 2453. Kalberer, M.; Tabor, K.; Ammann, M.; Parrat, Y.; Weingartner, E.; Piguet, D.; ROssler, E.; Jost, D. T.; Tiirler, A.; Gaggeler, H. W.; Baltensperger, U. J. Phys. Chem. 1996, 100, 15487. Kleffmann, J .; Becker, K. H.; Lackhoff, M.; Wiesen, P. Phys. Chem. Chem. Phys. 1999, 1, 5443. Longfellow, C. A.; Ravishankara, A. R.; Hanson, D. R. J. Geophys. Res. 1999, 104, 13833. Stadler, D.; Rossi, M. J. Phys. Chem. Chem. Phys. 2000, 2, 5420. Grassian, V. H. J. Phys. Chem. A 2002, 106, 860. Choi, W.; Leu, M.-T. J. Phys. Chem. A 1998, 102, 7618. Baldwin, A. C. Int. J. Chem. Kinet. 1982, I4, 269. Britton, L. G.; Clarke, A. G. Arm. Environ. 1980, 14, 829. Sergides, C. A.; Jassim, J. A.; Chughtai, A. R.; Smith, D. M. Applied Spec. 1987, 41, 482. Disselkamp, R. S.; Carpenter, M. A.; Cowin, J. P.; Berkowitz, C. M.; Chapman, E. G.; Zaveri, R. A.; Laulainen, N. S. J. Geophys. Res. 2000, 105, 9767. POschl, U.; Letzel, T.; Schauer, C.; Niessner, R. J. Phys. Chem. A 2001, 105, 4029. United Nations Environmental Program. Montreal Protocol on Substances That Deplete the Ozone Layer; Montreal, 1987. Knorr, K. Phys. Rep. 1992, 214, 113. 17 36. 37 38 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53 54. Knorr, K.; Civera-Garcia, E. Surf Sci. 1990, 232, 203. Nalezinski, R.; Bradshaw, A. M.; Knorr, K. Surf. Sci. 1995, 333, 255. Volkmann, U. G.; Knorr, K. Phys. Rev. B 1993, 47, 4011. Warken, A.; Enderle, M.; Knorr, K. Phys. Rev. B 2000, 61, 3028. FaBbender, S.; Enderle, M.; Knorr, K.; Noh, J. D.; Rieger, H. Phys. Rev. B 2002, 65, 165411. Robinson, G. N .; Freedman, A.; Kolb, C. E.; Worsnop, D. R.; Geophys. Res. Lett. 1994, 21, 377. Zhou, X.-L.; Zhu, X.-Y.; White, J. M. Surf. Sci. Rep. 1991, I3, 73. Fraser, P.J.; Oram, D. E.; Reeves, C. E.; Penkett, S. A.; McCulloch, A. M. J. Geophys. Res. 1999, 104, 15985. Jacox, M. E. Chem. Phys. Lett. 1977, 53, 192. Johnson, C. A. F.; Ross, H. J . J. Chem. Soc., Faraday Trans. 1 1978, 74, 2930. Sam, C. L.; Yardley, J. T. Chem. Phys. Lett. 1978, 61 , 509. Wampler, F. B.; Tiee, J. J.; Rice, W. W.; Oldenborg, R. C. J. Chem. Phys. 1979, 71, 3926. Molina, L. T.; Molina, M. J. J. Phys. Cem. 1983, 87, 1306. Krajnovich, D.; Zhang, Z.; Butler, L.; Lee, Y. T. J. Phys. Chem. 1984, 88, 4561. Gosnell, T. R.; Taylor, A .J.; Lyman, J. L. Chem. Phys. 1991, 94, 5949. Talukdar, R. K.; Vaghijani, G. L.; Ravishankara, A. R. J. Chem. Phys. 1992, 96, 8194. Vatsa, R. K.; Kumar, A.; Naik, P.D.; Rama Rao, K. V. S.; Mittal, J. P Chem. Phys. Lett. 1993, 207, 75. Van Hoeymissen, J .; Uten, W.; Peeters, J. Chem. Phys. Lett. 1994, 226, 159. Vatsa, R. K.; Kumar, A.; Naik, P.D.; Rama Rao, K. V. S.; Mittal, J. P. Bull. Chem. Soc. Jpn. 1995, 68, 2817. 18 55.Ta1ukdar, R. K.; Hunter, M.; Warren, R. F.; Burkholder, J. B.; Ravishankara, A. R. Chem. Phys. Lett. 1996, 262, 669. 56. Dhanya, S.; Saini, R. D.; Naik, P.D.; Vatsa, R. K. Chem. Phys. Lett. 2000, 318, 125. 57. Park, M. S.; Kim, T. K.; Lee, S.-H.; Jung, K.-H.; Volpp, H.-R.; Wolfrum, J. J. Phys. Chem. A 2001, 105, 5606. 58. Felder, P.; Yang, X.; Baum, G.; Huber, J. R. Israel J. Chem. 1993, 34, 33. 59. Cameron, M. R.; Johns, S. A.; Metha, G. F.; Kable, S. H. Phys. Chem. Chem. Phys. 2000, 2, 2539. 60. Chaix, L.; Allanic, A.; Rossi, M. J. J. Phys. Chem. A 2000, 104, 7268. 61. Hudson, P.K.; Foster, K. L.; Tolbert, M. A.; George, S. M.; Carlo, S. R.; Grassian, V. H. J. Phys. Chem. A 2001, 105, 694. 62. Barone, S. B.; Zondlo, M. A.; Tolbert, M. A. J. Phys. Chem. A 1999, 103, 9717. 63. Chu, L.; Chu, L. T. J. Phys. Chem. A 1999, 103, 8640. 64. Chu, L. T.; Chu, L. J. Phys. Chem. A 1999, 103, 384. 65. Lary, D. J.; Chipperfield, M. P.; Toumi, R.; Lenton, T. J. Geophys. Res. 1996, 101, 1489. 66. Gane, M. P.; Williams, N. A.; Sodeau, J. R. J. Phys. Chem. A 2001, 105, 4002. 67. Fan, S.-M; Jacob, D. 1. Nature 1992, 359, 522. 68. McNamara, J. P.; Hillier, I. H. J. Phys. Chem. A 1999, 103, 7310. 69. McNamara, J. P.; Hillier, I. H. J. Phys. Chem. A 2001, 105, 7011. 70. McNamara, J. P.; Hillier, I. H. Phys. Chem. Chem. Phys. 2000, 2, 2503. 71. McNamara, J. P.; Hillier, I. H. J. Phys. Chem. A 2000, 104, 5307. 72. McNamara, J .P.; Tresadem, G.; Hillier, I. H. J. Phys. Chem. A 2000, 104, 4030. 73. Bianco, R.; Hynes, J. T. Int. J. Quant. Chem. 1999, 75, 683. 74. Geiger, F. M.; Hicks, J. M.; de Dios, A. C. J. Phys. Chem. A 1998, 102, 1514. 19 75 76. 77 78 79. Ishida, H.; Palmer, R. E. Phys. Rev. B 1992, 46, 15484. Ancilotto, F.; Toigo, F. Phys. Rev. B 1993, 4 7, 13713. Lamoen, D.; Persson, B. N. J. J. Chem. Phys. 1998, 108, 3332. Lou, L.; Osterlund, L.; Hellsing, B. J. Chem. Phys. 2000, 112, 4788. Jeloaica, L.; Sidis, V. Chem. Phys. Lett. 1999, 300, 157. 80. Arellano, J. S.; Molina, L. M.; Rubio, A.; Alonso, J. A. J. Chem. Phys. 2000, 112, 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91 92 93 94 95 96. 8114. Feller, D.; Jordan, K. D. J. Phys. Chem. A 2000, 104, 9971. Ishikawa, S.; Madjarova, G.; Yamabe, T. J. Phys. Chem. B 2001, 105, 11986. Sorescu, D. C.; Jordan, K. D.; Avouris, P. J. Phys. Chem. B 2001, 105, 11227. Lushington, G. H.; Chabalowski, C. F. J. Molec. Struc. 2001, 544, 221. Tran, F.; Weber, J.; Wesolowski, T. A.; Cheikh, F.; Ellinger, Y.; Pauzat, F. J. Phys. Chem. B 2002, 106, 8689. Ferro, Y.; Marinelli, F.; Allouche, A. J. Chem. Phys. 2002, 116, 8124. Letardi, S.; Celino, M.; Cleri, F.; Rosato, V. Surf Sci. 2002, 496, 33. Tanaka, H.; El-Merraoui, M.; Steele, W. A.; Kaneko, K. Chem.Phys. Lett. 2002, 352, 334. Nyman, G.; Holmlid, L. J. Chem. Phys. 1986, 85, 6163. Hansen, F. Y.; T aub, H. J. Chem. Phys. 1987, 87, 3232. Hammond, W. R.; Mahanti, S. D. Surf. Sci. 1990, 234, 308. Hentschke, R.; Schilrmann, B. L.; Rabe, J. P. J. Chem. Phys. 1992, 96, 6213. Hentschke, R.; Winkler, R. .G. J. Chem.Phys. 1993, 99, 5528. Pinches, M. R. S.; Tildesley, D. J. Surf Sci. 1996, 367, 177. Kim, Y. H.; Rec, 1.; Shin, H. K. Chem. Phys. Lett. 1999, 314, 1. Jackson, B.; Lemoine, D. J. Chem. Phys. 2001, 114, 474. 20 97. Meijer, A. J. H. M.; Farebrother, A. J.; Clary, D. C.; Fisher, A. J. J. Phys. Chem. A 2001, 105, 2173. 98. Sha, X. Jackson, B; Lemoine, D. J. Chem. Phys. 2002, 116, 7158. 99. Sha, X.; Jackson, B. Surf. Sci. 2002, 496, 318. 100. Kwon, 8.; Russell, J.; Zhao, X.; Vidic, R. D.; Johnson, J. K.; Borguet, E. Langmuir 2002, I8, 2595. 101. Zhao, X.; Kwon, S.; Vidic, R. D.; Borguet, E.; Johnson, J. K. J. Chem. Phys. 2002, 117, 7719. 21 Chapter 2 Experimental 2.1 Ultrahigh Vacuum (UHV) Apparatus. All experiments described here were carried out in two connected stainless steel ultrahigh vacuum chambers. Each chamber was separately pumped by a 270 US titanium ion sublimation pump (Perkin Elmer Ultek D — 1). The first of the two chambers contained a molecular leak valve (Varian 9515106), a 1 — 200 amu quadrupole mass spectrometer (QMS) (VG Masstorr 200 DX) a dual anode (Mg/Al K013) X—ray source (VG 3EXR2) and a hemispherical electron energy analyzer (VG CLAM2). A nude ion gauge was used to monitor the chamber pressure. The second chamber, which was double u-metal shielded to decrease stray magnetic fields inside the chamber, contained a high-resolution electron energy loss Spectrometer (EELS) (L K Technologies LK3000). Baking the entire apparatus at 100 °C for ~ 48 hours allowed for a base pressure Of ~ 2 x 10’10 Torr to be obtained. A long-travel manipulator (500 mm) capable of x, y, z translation and 0 rotation (Thermionics 910438NW) was mated to the first chamber and allowed for mounting of the sample. The sample mount consisted of a molybdenum block that was secured to a liquid nitrogen (LNz) cooled copper reservoir by a threaded molybdenum rod and nut. The copper reservoir was connected to a Macor spacer that was used to provide for both thermal and electrical isolation of the sample. The sample was secured to the molybdenum mount by two tungsten clips, under one of which was placed an E-type thermocouple directly in contact with the sample face. This arrangement insured good thermal contact between the mount and the sample. The sample could be indirectly heated to > 800 K by a tungsten filament (Alfa Aesar, 0.25 mm diameter) embedded in 22 the molybdenum mount. Cooling the sample to < 85 K was accomplished by drawing liquid N2 through the copper reservoir with a diaphragm pump.l 2.2 Sample Preparation and Dosing. A 10 x 10 x 1 mm highly ordered pyrolytic graphite (HOPG) sample (Grade SPI-2, SPI Supplies) was repeatedly cleaved using adhesive tape to expose a fresh, visibly flat C(OOOI) surface and then outgassed at 720 K in UHV for 5—6 hours. Such a procedure is known to produce clean, ordered C(OOOI) surfaces.2 Prior to each day’s experiments, the sample was flashed to ~ 800 K for 2—3 minutes to remove any accumulated adsorbates. X-ray photoelectron spectroscopy measurements showed up to 1% oxygen-containing species were present on the surface after flashing. These species could influence the adsorption and monolayer formation of CF2Br2 on the HOPG surface. However, reproducible results were obtained between experiments on subsequent days. The clean HOPG surface was cooled to 85 i l K and exposed to CF2Br2 (~ 98%, PCR Research Chemicals, Inc.) by backfilling the UHV apparatus for a predetermined time using the molecular leak valve. The CF2Br2 was used without further purification. All CF2Br2 exposures are given in langmuirs (1 langmuir = 1L = 10'6 Torr - s) and are uncorrected for ion gauge gas sensitivity. 2.3 X-ray Photoelectron spectroscopy (XPS). X-ray photoelectron spectroscopy was used to determine the chemical composition of the clean and adsorbate- covered surface. This technique works by irradiating the surface with X-rays of sufficient energy such that photoelectrons are produced according to equation 2.13 EK=hV—EB-¢ (21) 23 where Ex is the electron kinetic energy, hv is the energy of the X-ray photon, E3 is the binding energy of the electron in the sample and (l) is the sample work function. Collection and energy analysis of the photoelectrons reveals the binding energies of the core levels of the atoms contained within the sample. Because each binding energy can be uniquely associated with a core level in an atom, XPS can be used to determine both the number and type of atoms in the sample. Changes in the chemical environment of an element will produce a concomitant Shift in the core level binding energy relative to that of a pure sample. This effect is commonly called the core-level shift. Information about oxidation states and changes in bonding can be obtained by observing these shifts. X-ray photoelectron spectroscopy is a surface sensitive technique because most of the photoelectrons are limited to escape from sample depths of a few tens of A due to inelastic scattering processes within the sample.4 The surface sensitivity can be increased by changing the surface/sample geometry. This can be appreciated by looking at Figure 2.1.3 The vertical sample depth d, is described by equation 2.2 d = 3 A sinor (2.2) where A is the electron attenuation length and or is the angle of electron ejection with respect to the surface plane (take—off angle). The maximum vertical sampling depth will be 3}. when or = 90°. This configuration yields 95% of the signal intensity. If, however, the take-off angle is decreased, the vertical sampling depth becomes smaller and only those atoms contained within this depth will be able to emit photoelectrons. The reduction in sampling depth provides for the further increase in surface sensitivity. This angular dependence on the XPS signal is useful for structural determination as well as “depth profiling” in a sample. 24 +0. 01. 3.1 A hv Figure 2.1 Surface sensitivity enhancement by variation Of the electron take-off angle.3 25 All XP spectra were collected using the Mg Kong X-ray line (hv = 1253.6 eV) Operated at 300 W (15kV, 20 mA) and an analyzer pass energy of 100 eV. To maximize surface sensitivity, photoelectrons were collected at a take-off angle of 35° from the surface plane. Spectra were referenced to the C Is peak from HOPG at 284.7 eV binding energy.5 No changes in the photoemission data (C IS, F Is, and Br 3d) were observed during extended periods (up to 1 hour) of X-ray irradiation and data typically were acquired in less than 20 minutes. After each spectral acquisition, the surface was flashed to 450 K and a fresh layer Of CF2Br2 was dosed. A temperature of 450 K was found to be sufficient for removal of all F and Br from the surface after adsorption and UV exposure experiments as evidenced from XPS measurements. 2.4 Temperature-Programmed Desorption (TPD). Characterization of the adsorbate-covered surface was undertaken by TPD. In these experiments, the HOPG surface was cooled to 85 i l K and the UHV apparatus was backfilled with CF2Br2 to a 4.8 L exposure (~ 1 monolayer). The surface was subsequently heated with a linear heating rate of 10 K/s over a temperature range of 85 — 450 K. The desorbing molecules were transmitted to the QMS through a 2 mm diameter aperature in a stainless steel shroud enclosing the QMS, ionized by impact with 70 eV electrons, and separated according to their mass—to-charge ratio. The QMS line-of-site was coincident with the surface normal to insure that the largest number of desorbing molecules would be collected. The data was then plotted as QMS intensity vs. temperature to determine the desorption profile and kinetics. Analysis of TPD data is often done in terms of the Polanyi-Wigner equation6 dN rate = — dta = va NJ“ exp(—Ea /RT) (2.3) 26 where X, is the reaction order from the state a, v, is a frequency factor, Ea is the desorption energy, N8 is the coverage of molecules in state a, R is the gas constant, T is the temperature and t is time. A plot of the heating rate, surface coverage and the desorption rate are shown in Figure 2.27 which simulates a typical TPD experiment. A majority of TPD analyses focus on Obtaining the reaction order, xa and the desorption energy, E... A variety of methods have been employed to obtain these parameters.8 The desorption energy indicates the strength of the interaction between an adsorbate and the surface. An adsorbate may be physisorbed on the surface, implying a weak adsorbate- substrate interaction with a low desorption temperature. A chemisorbed species forms a chemical bond to the surface and desorbs at a higher temperature compared to a physisorbed molecule. The difference between the two species is not clearly defined but a value of 40 kJ/mol is Often used. The reaction order is normally an integer such as 0,1, or 2 but fractional values are possible as well.9 Plots showing typical zero, first, and second order processes are shown in Figure 2.3 with an E3 of 40 kJ/mol, v0 Of 1 x 1028, v; of 1 x 10”, v; of 1 x 10", a heating rate of 1.5 K/s, and an N2, of 1 — 6 x 1013 molecules/cmz. Each curve in Figure 2.3 shows an increase in Na starting from 1 x 1013 molecules/cm2 to 6 x 1013 molecules/cm2 in increments of l x 1013 molecules/cmz. The zero order process Shows a common low-temperature leading edge with increasing exposure and can be indicative of bulk evaporation of the adsorbate from the surface. In this case, the rate is independent of coverage.9 Zero-order desorption kinetics also have been observed for the desorption of adsorbates from 2-D islands modified by attractive 10-12 interadsorbate interactions and for desorption of multilayers.l3 A first order process can be distinguished by a constant maximum desorption temperature with increasing 27 Rate Coverage Temperature Time Figure 2.2 Surface coverage, desorption rate, and temperature dependence on time during TPD.7 28 a) Intensity I 100 I 120 I I I 140 160 180 Temperature (K) I 200 220 Figure 2.3 TPD simulations of zero (a), first (b), and second (c) order desorption processes. 29 coverage, while a second order process has a common high temperature trailing edge with additional exposure. Bond formation between two adsorbates prior to desorption from the surface accounts for this high temperature trailing edge feature. Two of the more common methods used for obtaining X, and E3 from TPD data are Arrhenius plots and Redhead’s peak maximum method.14 In an Arrhenius plot, a graph of ln(rate) vs. l/T for x = 0, 1, 2 is made and the order is determined by the value that yields a linear plot with Ea being obtained from the slope of the line. The Redhead method is usually only applicable for first-order desorption kinetics and has the following form E = RTm[ln(\/I‘m lli) — 3.46] (2.4) where Tm is the temperature of the peak maximum, B is the heating rate, v is the pre- exponential factor, and R is the gas constant. General values for v ranging between 108 and 1013 s'1 are chosen and errors of less than 2% are obtained with this method. Many other methods for analysis of TPD data are available but will not be discussed here. 2.5 Electron Energy Loss Spectroscopy (EELS). Further characterization of both the clean and adsorbate-covered surface was performed by electron energy loss (spectroscopy. In EELS, a monoenergetic beam of electrons scatters from a surface and the scattered electrons are energy analyzed. Most electrons are elastically scattered in the specular direction (9m = 00."). However, some electrons are inelastically scattered because they excite vibrational modes in the surface or an adsorbate. These vibrational modes can only be excited if a component of the dynamic dipole moment lies along the surface normal. If the dipole lies parallel to the surface plane then it will be canceled by an equal but opposite image dipole in the surface (surface selection rule).15 The 30 IE dc‘ SC. fu} U\ 1311 Nif Hm Win ape mm the: inelastically scattered electrons will be collected close to the specular direction due to the small change in their momentum. In this mechanism, the electron is scattered from long- range dipole fields that are caused by the vibrations of adsorbates or surface atoms. While other scattering mechanisms may occur, the long-range dipole scattering mechanism is assumed to be the dominant scattering mechanism for the experiments described here. All EELS experiments were performed in the specular scattering geometry (em = 00... = 55°) with a primary beam energy of 6.09 eV. Typical resolution of the elastically scattered beam from the clean and adsorbate-covered surface was 51-54 cm'1 (6-7 meV) full width at half maximum (FWHM). Count rates from the clean HOPG surface were >106 Hz while those from the adsorbate-covered surface were >105 Hz. 2.6 Photochemistry. The adsorbate-covered surface was exposed to unpolarized UV radiation from a medium-pressure Hg-arc lamp (Oriel 6286) Operated at 350 W. The lamp was equipped with a condenser lens and a visible/infrared liquid filter (~ 1 M NiSO4 solution) that primarily transmitted wavelengths in the range between 225 and 350 nm. The UV radiation was introduced into the UHV apparatus through a fused quartz window such that its angle of incidence was 45° with respect to the surface normal. An aperature (12.7 mm diameter) affixed to the window minimized irradiation of the sample mount. A power of 8.6 mW/cm2 was measured at the sample-lamp distance by a therrnopile detector. All UV irradiation of the adsorbate-covered surface was performed at a surface temperature of S 85 K. Upon exposure to UV irradiation, the surface temperature increased by ~ 6 K. No evidence was Observed for thermally driven reactions under these modest temperature increases. 31 10. 11. 12. 13. 14. 15. References . Bryden, T. R. Ph.D. Thesis, Michigan State University, 2001. Musket, R. G.; McLean, W.; Colmenares, C. A.; Makowiecki, D. M.; Siekhaus, W. J. Appl. Surf Sci. 1982, 10, 143. Briggs, D. in Practical Surface Analysis; Briggs, D.; Seah, M., Eds.; Wiley and Sons: New York, 1990; Vol. 1. Woodruff, D. P.; Delchar, T. A. Modern Techniques of Surface Science; Cambridge University Press: Cambridge, 1994. Barr, T. L. Modern ES CA: The Principles and Practice of X -ray Photoelectron Spectroscopy; CRC Press: Boca Raton, FL, 1994. King, D. A. Surf Sci. 1975, 47, 384. Falconer, J. L.; Schwarz, J. A. Catal. Rev. Sci. Eng. 1983, 25, 141. de long, A. M.; Niemantsverdriet, J. W. Surf Sci. 1990, 233, 355. Vollmer, M.; Trager, F. Surf Sci. 1987, 187,445. Golze, M.; Grunze, M.; Hirschwald, W. Vacuum 1981, 31, 697. Niemantsverdriet, J. W.; Markert, K.; Wandelt, K. Appl. Surf Sci. 1988, 31, 211. Zhou, X.-L.; White, J. M.; Koel, B. E. Surf Sci. 1989, 218, 201. Yates, J. T., Jr. Methods of Experimental Physics; Academic Press: San Diego, 1985; Vol. 22. Redhead, P. A. Vacuum 1962, 12, 203. Pemble, M. in Surface Analysis-The Principle Techniques; Vickerrnan, J. C., Ed.; Wiley and Sons: West Sussex, 1997. 32 II) Sr Chapter 3 Theoretical Methods 3.1 Introduction. Ab initio electronic structure methods were used to Obtain information about the localized interactions of CF2Br2 and its possible photofragments with the HOPG surface. Calculations were performed examining the preferred adsorption sites and geometries of the adsorbed molecules to aid in further explaining the experimental data. To accomplish this, the electronic SchrOdinger equation must be solved for the adsorbate/substrate system. The main interaction between the molecules and the graphite surface occur through dispersion forces. Because these forces are weak in nature compared to chemical bonds, ab initio methods that describe electron correlation must be used. The simplest method that accounts for electron correlation is second order Mallet-Plesset (MP2) perturbation theory.1 A description of the general method of perturbation theory follows in the next section along with a more detailed discussion of MP methods. 3.2 Perturbation Theory. In perturbation theory, the system to be studied is assumed to be slightly perturbed in some manner compared to a known unperturbed reference system.2 This assumption allows for the solutions of the unperturbed system, typically from a Hartree-Fock calculation, to be used as a starting point to obtain solutions for the perturbed system. If this is the case, the Hamiltonian of the perturbed system can be written as f1 = [10 +119, (3.1) where [I 0 is the Hamiltonian Of the unperturbed problem, [-1, is the perturbed Hamiltonian, and A is a parameter that determines the strength of the perturbation.2 The solutions to the unperturbed Hamiltonian can be chosen such that they form a complete 33 orthonormal set. For the discussion here, the perturbed Hamiltonian is assumed to be time-independent and non-degenerate. Values for A. will range between zero and one. Since A. can vary continuously within this range, the energy and wavefunction for the system will change in a similar fashion and can be written as Taylor series expansions in powers of A E = AOEO +AE, + 22E, + 23E, + ‘1’ = 29% + 2‘11, + 22‘1’2 + 23‘113 + (3.2) where E0 and ‘P0 are the unperturbed energy and wavefunction when 2t=0 and all other terms are corrections to the energy and wavefunction. Putting the Taylor series expansions for the energy and wavefunction into the Schrtidinger equation and collecting like powers of A (through second order) yields to: now, = EO‘I’O A‘: Help, + 119?, = EO‘I’, + 5311,, A2: now, + Apr, = EO‘I’2 + 13911, + 15311,, (3.3) The term in A0 is the solution of the unperturbed problem. The equation for the first order corrections contains two unknowns (E1 and W1) that can be found if the first order wavefunction is expanded in the complete set of functions produced by the unperturbed wavefunction (‘1’, = Zed), ). The first and second order energy corrections can be found by multiplying by (Do on the left of the equations for the Al and 7&2 terms in equation 3.3, respectively, and integrating. The corrections to the wavefunction will be found by multiplying by <1),- on the left of the equations for the Al and 2.2 terms in equation 3.3, 34 respectively. If the unperturbed Hamiltonian is chosen as a sum of one-electron Fock operators, the method is known as Moller-Plesset (MP) perturbation theory.1 By making this choice for the unperturbed Hamiltonian, the sum of the zerO-order energy and first order correction to the energy will be the Hartree-Fock energy. The higher energy corrections correspond to wavefunction corrections that contain singly, doubly, etc. excited Slater determinants relative to the Hartree-Fock reference configuration. These excited determinants describe the many-electron correlation effects. The correlation energy is not Obtained until the second order corrections (the lowest order terms due to doubly excited Slater determinants) are added.2 The second order energy correction in this case becomes we <O|I§r,|cr>;b)(c;b £322 ab i—<¢i¢il¢b¢a> ] i 4.8 L lead to the appearance of a lower temperature desorption peak centered at 109 K. This peak did not saturate with increasing exposure and also exhibited a common leading edge. Such zero-order desorption kinetics are characteristic of 18 A leading edge analysis of this peak, gave a desorption energy multilayer desorption. of 22.4 i 1.9 kJ/mol. Figure 4.2 shows TPD spectra for 4.8 L CF2Br2 on HOPG measured at m/z = 31 (CF), 50 (CF2+), 79 (Br+), 100 (C2F4+), 129 (CF2Br+), 160 (Br2+), and 179 (C2F4Br+). These ions were chosen because they are representative of the fragmentation pattern of gas phase CF2Br2 and possible photolysis products. Each spectrum was obtained from a separate CF2Br2 exposure. Except for the m/z = 100 and 179 data, each ion had a similar 42 4.8 L CFzBrleOPG Unirradiated "it/L *5 m/z = 50 x5 2‘ .5 _ ac) m/z - 79 X5 E c m/z = 100 x200 2 m/z = 129 / = 1 m z 60 x50 m/z = 179 x200 L1llllllllJlLJJllllllllllllllllllllllllllllllllll'IJlllllJl 120 130 140 150 160 170 180 Temperature (K) Figure 4.2 Temperature-programmed desorption data for 4.8 langmuir CF2Br2 on HOPG. Monitored masses were m/z = 31 (CF), 50 (CF2I), 79 (Br‘), 100 (C2F4*), 129 (CF2BrI), 160 (Br2+) and 179 (C2F4Br1). desorption profile and peak temperature (138 K) suggesting a common origin. We concluded that molecular desorption occurred by comparison of the relative ion abundances for the TPD data presented here at m/z = 31, 50, 79, 129, and 160 (21:24:17:100:1.5) with the mass spectrum of gas phase CF2Br2 (18:21:172100z2) as measured by residual gas analysis. A small amount of m/z = 100 (< 1% of the area of m/z = 129) was observed both in TPD measurements and residual gas analysis of the CF2Br2 admitted to the chamber indicating the presence of an unidentified contaminant. No signal was observed at m/z = 179 in residual gas analysis of CF2Br2. A significant change in the TPD spectra was seen after UV irradiation of the adsorbate. Figure 4.3 shows a similar set of TPD spectra to those in Figure 4.2 after 5 minutes of UV irradiation. The monolayer peak originally present at 138 K broadened and shifted to 141 K and a second desorption peak centered at 155 K appeared for m/z = 31, 50, 79, 129, and 160. Relative ion abundances Of 23:27:20:100:l.4 and 22:27:22:100:1.3 (m/z = 31, 50, 79, 129, and 160) for the 141 K and 155 K peaks, respectively, indicated that each peak was associated with molecular CF2Br2. The m/z = 100 and 179 TPD spectra showed a broad desorption peak at ~165 K. These features are associated with the formation of a photolysis product, C2F4Br2, which will be discussed later. Figure 4.4 shows TPD spectra of the previously monitored ions after 15 minutes of UV irradiation. The peak originally present at 141 K for m/z = 31, 50, 79, and 129 in Figure 4.3 shifted to a higher desorption temperature (~ 146 K) and had started to merge with the leading edge of the peak at 155 K. The peak at 155 K was present for all UV 44 4.8 L CFzBrZIHOPG Irradiation Time = 5 minutes m/z = 31 x5 m/z = 50 x5 2: FEB m/z = 79 3 x5 E c m/z = 100 . ' x200 _q - _ m/z = 129 m/z = 160 x50 m/z = 179 ' I , X200 120 130 140 150 160 170 180 Temperature (K) Figure 4.3 Temperature-programmed desorption data for 4.8 langmuir CF2Br2 on HOPG after 5 minutes of UV irradiation (filtered Hg-arc, 225-350 nm, incident power 8.6 mW/cmz). Monitored masses same as in Figure 4.2. 45 4.8 L CFzBr2/HOPG Irradiation Time = 15 minutes M X5 3‘ 2 / - 79 9 m - x§_ E a g m/z 100 , , - - _ x200 m/z = 129 m/z = 160 "50 lz =179 : I x 00 JIIllllllIlllllllliJlllIllIllllljlllllljlllllllllllllllllll 120 130 140 150 160 170 180 Temperature (K) Figure 4.4 Temperature—programmed desorption data for 4.8 langmuir CF2Br2 on HOPG, after 15 minutes of UV irradiation. Conditions same as in Figure 4.2. 46 exposures up to 60 minutes irradiation time (data not shown). Furthermore, a comparison of the m/z = 160 TPD spectra in Figures 4.3 and 4.4 revealed the growth of a new peak at 165 K with continued UV exposure. This new peak in the m/z = 160 TPD spectrum was observed only for UV irradiation times greater than 5 minutes. 4.2.2 X-ray Photoelectron Spectroscopy (XPS). Figure 4.5 shows the F ls and Br 3d regions for increasing UV exposure. For the unirradiated sample, the F ls binding energy was observed at 688.4 eV similar to that of CF2Br2 adsorbed on sapphire.19 As the UV irradiation time increased, the F ls peak decreased in intensity and shifted to lower binding energy (687.5 eV after 60 minutes UV exposure) but the peak width (FWHM) remained constant at 3.0 eV. The Br 3d peak maximum (unresolved spin-orbit split doublet) was observed at 71.6 eV for the unirradiated surface. This value is relatively high compared to simple metal bromide solids (68.3-69.3 eV)20 and about 1 eV higher than CH3Br(ad) on Ag(111).21 The Br 3d peak shifted to 70.4 eV and the peak broadened from 3.9 eV to 4.8 eV (FWHM) after 60 minutes of irradiation. The intensity of the Br 3d peak remained nearly constant. Figure 4.6(a) shows the C 1s region of the clean HOPG surface and that of a 4.8 L CF2Br2-covered HOPG surface. The main photoemission peak in both cases is located at 284.7 eV. Furthermore, the clean graphite surface also shows the well-known shake-up feature (Ti—Mt“) centered at about 292 eV. The addition of the CF2Br2 monolayer resulted in the appearance of a feature at 292.8 eV. This feature was most clearly Observed in the normalized difference spectrum (bare HOPG minus 4.8 L CF2Br2(ad)/HOPG), and was shifted by +8.1 eV BE relative to the main photoemission peak. We assign this feature to 47 678 680 682 684 686 688 690 692 694 696 698 700 702 ' . I ' I . I ' I ' I ' I . a) _ d - fl - = 1000 cps s 60 min 30 min 20 min 10 min IT‘I‘U l T 0 min fr! Br 3d b) 60 min ['11 (II C O O K '0 (D E 75E g- yd” 3: 20min to” CL '93- kan E? E 0min Fi.r.I.I.I.I.I.I.I.I.I.I 58 60 62 64 66 68 7O 72 74 76 78 80 82 Binding Energy (eV) Figure 4.5 X-ray photoelectron spectra of F ls region (a) and Br 3d region (b) of CF2Br2 monolayer as a function of UV irradiation. Vertical lines are drawn at 688.4 eV and 71.6 eV BE. 48 I T r I l I I fi I I I I I I l I I I 1 I — Bare HOPG """"" 4.8 L CFzBrleOPG a) . + Normalrzed Difference Tr? 5 8 .4341}- -+ I" W.-------E v : 1 “we > 'I l l l I I l l l l I l j 1 j J l l l I t : 4 a L or B /HOPG a) L . r c C b) 2 O2 . q, C —— min UV E E- -------- 30 min UV — ~_- + Normalized E‘ Difference Fae. El: I l l J I l l l l I L l l I I l l I I I 280 285 290 295 300 Binding Energy (eV) Figure 4.6 (a) X-ray photoelectron spectra of the C Is region of bare HOPG (solid line), 4.8 langmuir CF2Br2-covered HOPG (solid circles) and the normalized differmce spectrum (crosses). (b) X-ray photoelectron spectra of the C ls region of 4.8 langmuir CFzBry-covered HOPG (solid line), 4.8 langnuir CF2Br2-covered HOPG after 30 min UV irradiation (solid circles) and the normalized diffa'ence spectrum (crosses). 49 photoemission from the adsorbate carbon atom bonded to four electronegative atoms (2 F atoms and 2 Br atoms), with the majority of the observed chemical shift expected to originate from the F atoms (according to literature, approximately 2.9 eV per F atom, 1.0 eV per Br atom).22 Figure 4.6(b) shows XPS data of the C Is region Of a 4.8 L CF2Br2 exposed graphite surface and an identical surface after 30 minutes of UV irradiation. Each spectrum represents a separate dose. After irradiation, the 292.8 eV feature was observed to decrease in intensity and a new feature appeared at approximately 290.0 eV. This is illustrated in the normalized difference spectrum shown in Figure 4.6(b). Figure 4.7 shows the integrated raw area ratios F ls / C 1s and Br 3d / C Is as a function Of UV exposure. Each data point corresponds to a separate 4.8 L CF2Br2 dose (saturated first layer). The F ls/C ls ratio decreased by 25% between 0 and 60 minutes irradiation time, suggesting a net loss of F atoms from the adlayer. In contrast, the Br 3d/C ls ratio remained unchanged indicating that no Br atoms were expelled from the surface during UV exposure. Error bars represent 1 o for two separate experiments. 4.2.3 Electron Energy Loss Spectroscopy (EELS). Figure 4.8 shows electron energy loss spectra of 4.8 L of CF2Br2 on the HOPG surface with and without exposure to UV radiation. A spectrum of the bare HOPG surface is also shown for comparison. The spectrum for bare graphite (a) exhibited the characteristic background of a semimetal caused by excitation of low-energy electron-hole pair transitions across the Fermi level.23 NO Other loss features were present in this spectrum. Spectrum (b) shows the surface exposed to 4.8 L of CF2Br2 without UV irradiation. A weak loss was observed at 1157 cm". We assign this feature to the CF2 asymmetric stretch in the adsorbed molecule 50 - 4.8 L CFzBrleOPG 1.1- . I F1s/C1s e Br3d/C1s 1.0- 0.9- 0.8- 0.7- lntegrated Raw Area Ratio 0 (II 1 l '—.—.l H___1 Irradiation Time (min) Figure 4.7 Integrated area ratios (F ls/C Is and Br 3d/C ls) from X-ray photoelectron spectroscopy of 4.8 langmuir CF2Br2 adsorbed on HOPG, as a function of UV irradiation time (conditions as in Figure 4.3). The solid curve in the F lle ls ratio is a single-exponential fit to the data. The solid curve in the Br 3d/C ls ratio is a linear fit to the data. 51 4o. l x20 4.8 L CFZBrleOPG 30 in as b) 5,? ml 0- t 8 E: '6 c .03. 5 x20 10 -» C) a) 01.1. I'.-._:,=_I.i 0 500 1000 1500 2000 2500 3000 3500 Energy Loss (cm’1) Figure 4.8 High-resolution electron energy loss spectra of (a) clean HOPG, (b) 4.8 langmuir CF2Br2-covered HOPG and (c) 4.8 langmuir CF2Br2-covered HOPG after 30 minute UV irradiation. 52 since it is close to the reported value for CF2Br2(g/s) of 1153 cm'l.24'25 Spectrum (0) shows the 4.8 L CF2Br2/HOPG surface after 30 minutes of UV irradiation. Again, the only loss peak present was a C-F-like stretch that had red-shifted to 1129 cm'1 and gained intensity. Furthermore, a noticeable decrease in the overall background intensity was Observed. 4.2.4 Adsorption and Desorption of CF2Br2(ad)/HOPG. The TPD data in Figures 4.1 and 4.2 show that CF2Br2 undergoes molecular adsorption and desorption from a clean HOPG surface in the absence of UV irradiation. Measured ion abundances for m/z = 31, 50, 79 and 129 in our desorption data closely match those of CF2Br2(g). For the adsorbed layer, apparent zero-order desorption kinetics were observed. Such behavior can arise either through true zero-order desorption kinetics, first-order desorption kinetics modified by attractive interadsorbate interactions, or half-order desorption kinetics from 2-dimensional islands modified by attractive interadsorbate interactions.”28 Knorr and coworkers”33 have proposed monolayer structural models based on X- ray diffraction data for many halocarbons (but not CF2Br2) on graphite. Their data for CF2C12 indicated coexistence Of a two-dimensional island (8 phase) and lattice gas phase (a disordered sub-monolayer adsorbate coverage) at the temperatures and coverages appropriate to our measurements. It was suggested that the presence of the permanent dipole moment in CF2C12 favored either an antiparallel or zig-zag arrangement of dipoles within the islands. We believe that the zero-order desorption kinetics observed in our TPD experiments for the first layer of CF2Br2 are also consistent with the formation of 2- D islands modified by attractive interadsorbate interactions.”28 Similar TPD behavior 53 has also been observed for H2O adsorption on HOPG surfaces, and attributed to the creation of 2-D islands nucleated at defect sites.34 We speculate that the lower temperature desorption feature (~110 K), characteristic of multilayer desorption, appears after the 2-D islands in the monolayer coalesce. To our knowledge, the structure of the CF2Br2 monolayer on graphite has not been determined. However, some information on a possible structure can be obtained by comparing the fractional coverage, <1>CF2Br2, (defined as the number of adsorbate molecules per surface C atom) with the fractional coverage of CF2Cl2 on graphite. In our case, the fractional coverage was calculated for CF2Br2 on HOPG using the XPS integrated raw area ratios shown in Figure 4.7 and expressions developed in the literature.22 Using sensitivity factors appropriate for our electron analyzer (C Is = 0.205, F IS = 1.00 and Br 3d = 0.5920), a mean value of (DCF2Br2 = 0.14 i 0.01 (4.6 x 1014 molecules/cmz) was obtained (see Appendix A). The calculated fractional coverage compares favorably to the maximum fractional coverage of the [3 phase of CF2Cl2 on graphite (<1>CF2C12=0.137).30'33 Similar surface structures for monolayers of CF2Cl2 and CF2Br2 may be anticipated based on the similarity of their dipole moments (CF2C12 = 0.51 D, CF2Br2 = 0.66 D)35 and projected surface areas (CF2Cl2 = 27.9 A2, CF2Br2 = 31.7 A2). Areas were estimated from van der Waals’ radii assuming both molecules have one F and two X atoms (X = Cl, Br) in contact with the surface (CFX2 “tripod”) as suggested by the theoretical studies in Chapter 5.29'33 The EELS data shown in Figure 4.8(b) provides additional evidence for this adsorption geometry. The appearance Of a single vibrational mode at 1157 cm'l, corresponding to the CF2 asymmetric stretch (1153 cm", b1 symmetry in CF2Br2(g)), is 54 consistent with a CF2 plane that is not parallel to the HOPG surface (surface selection rule). The companion vs(CF2) mode Of CF2Br2 expected at ~1090 cm'1 was not observed, presumably due to a small component of the dynamic dipole parallel to the surface normal. No loss features attributable to CBr2 vibrations were observed. 4.2.5 Photolysis of CF2Br2 on HOPG. Numerous studies have shown the initial photodissociation step operative at 200—300 nm in CF2Br2(g), is C-Br bond cleavage in a 36-47 single photon process. CF2Br2 + hv —+ CF2Br + Br (4.1) Similarly, upon exposure to UV radiation, photolysis of the CF2Br2 adsorbed on graphite was observed. The m/z = 129 TPD data of Figures 4.3 and 4.4 showed a decreased intensity of the 138-141 K peak associated with molecularly adsorbed CF2Br2 for increasing UV irradiation times. The dissociation cross-section for CF2Br2(ad)/graphite was estimated by measuring this loss in intensity as a function of UV exposure. For the range of wavelengths generated by our filtered arc lamp, the apparent total cross-section (the integrated cross-section for 225-350 nm) was 1.9 x 10'l9 cm2 (using a mean wavelength of 287.5 nm in the calculation). Similarly, we estimated the integrated cross- section for CF2Br2(g) over the same range of wavelengths by multiplying the scaled irradiance curve for our lamp,48 the NiSO4 filter transmission spectrum,49 and the gas phase cross-section data.50'51 The integrated cross—section (225-350 nm) for gas-phase CF2Br2 was 4.5 x 10'19 cm2, very similar to our CF2Br2(ad)/HOPG value. In fact, greater than 95 % of the contribution to the cross-section occurs in the 240-250 nm region (where the Hg lamp irradiance is high and the cross-section for CF2Br2(g) is large). Absorption cross-sections for wavelengths greater than 250 nm decrease rapidly 55 (~ 10'19 cm2 at 250 nm to ~ 10’23 cm2 at 320 nm).50'51 Essentially, no contribution to the photolysis of CF2Br2 is expected for wavelengths greater than 320 nm. There was no evidence for photochemical modification of CF2Br2 monolayers by ambient (visible) light entering the UHV chamber. The EELS data shown in Figure 4.8(c) indicated that after UV irradiation the loss originally present at 1157 cm", due to CF2 vas in CF2Br2(ad), became more intense and red-shifted to 1129 cm“. This value is very close to that observed in a study Of the 254 nm photolysis of CF2Br2 in an AI matrix (1138 cm”).36 In that study, a definitive symmetry could not be assigned to this mode but it was assigned to one of the fundamental C—F stretches of CF2Br. The similarity of the integrated cross-sectional data for CF2Br2(ad)/HOPG and CF2Br2(g), and the observation of a vibrational mode of CF2Br(ad), implies that UV irradiation of dibromodifluoromethane produces surface- bound CF2Br in a similar fashion to equation 4.1. Further evidence for photochemistry of the adsorbed CF2Br2 is observed by a shift in the C 1s photoemission peak (Figure 4.6), from 292.8 to 290.0 eV BE following UV exposure. The direction of the shift suggests a net reduction in the number of electronegative substituent atoms attached to the C atom of CF2Br2(ad). The magnitude of the shift (-2.8 eV) is consistent with loss of an F atom to produce CFBr2(ad), but we believe this to be misleading. The shift must correspond to loss of a single Br atom to form CF2Br(ad), as clearly observed in EELS and TPD measurements. The large C ls BE shift observed following photolysis is likely to be complicated by changes in the intensity of the C Is shake-up feature at ~292 eV. This feature is sensitive to the electronic structure of the surface. Indeed, based on analogy with bromine-intercalated 56 graphite compounds, the presence of Br on the graphite surface is expected to cause some charge transfer from the graphite to the Br atoms.52 This would cause a decrease in the population of the filled graphite Ir-band and an increase in electronic conductivity.53 The decrease in the electron-hole pair background in our EELS data (Figure 4.8) also supports the idea of a more metallic surface for UV-irradiated CF2Br2 on graphite. It is somewhat surprising that electron-induced chemistry due to photon absorption by graphite does not appear to significantly contribute to the observed photochemistry of CF2Br2(ad). Graphite strongly absorbs in the UV generating free photoelectrons with kinetic energies from 0 to about 0.7 eV (hv250 um = 5.0 eV, graphite work-function, (I) = 4.35 eV54). These do not appear to cause dissociative electron attachment-type reactions for the adsorbed molecule despite the fact that for gas phase 2 is observed CF2Br2, a large dissociative electron attachment cross-section of >10’15 cm for ~ 0 eV electrons.55 There is substantial evidence in the literature that, in general, low energy electrons generated by photon irradiation of metallic and semiconductor surfaces can cause dissociation of an adsorbed molecule.16 The apparent lack of electron-induced chemistry for adsorbed CF2Br2 is likely due to either an increased work-function for graphite upon adsorption (> 0.7 eV), poor spatial/energetic overlap between the graphite and adsorbate orbitals or competitive quenching by the surface. The theoretical studies presented in Chapter 5 provide an indication that a change in the work function of graphite could occur as evidenced by transfer of charge from the surface to adsorbed CF2Br2. 4.2.6 Surface Recombination Reactions. In static cell studies of the photolysis Of CF2Br2(g) a variety of radicals and stable molecules are generated, including CF2Br, 57 CF2, Br and C2F4Br2 (formed by biradical reaction of CF2Br).3840’42'MJ'7 Reactions generating Br2 and C2F4 do not appear to be major product channels at wavelengths greater than about 248 nm and at low photon fluences. The high temperature (155 K) TPD feature observed in our data is likely due to recombination of photogenerated CF2Br(ad) and Br(ad) atoms during the TPD experiment. The photodissociation event imparts kinetic energy to the nascent photofragments (total up to about 2.0 eV at our shortest wavelength), separating them and preventing immediate recombination. Prompt recombination would generate CF2Br2 that would be indistinguishable in TPD measurements from the unphotolyzed molecules. We assume CF2Br and Br are weakly chemisorbed and not significantly mobile, since they are stable on the graphite surface up to ~155 K. Diffusion of one or both of these species apparently occurs only at the elevated surface temperatures experienced during TPD. Another recombination product observed on the HOPG surface was Br2(ad) as indicated by the appearance of a desorption peak centered at 165 K for m/z = 160 (Figure 4.4). This feature is observed only after extended UV irradiation periods (210 minutes). The expected m/z = 79 ion (Br+) at 165 K from fragmentation of Br2 in the ionizer has a small abundance, preventing detection in our data (see Figure 4.4). It should be noted that the appearance of coincident features in the m/z = 100 and 179 are not related to the production of Br2(ad) but are associated with another photolysis product, C2F4Br2 (see below). Molecular bromine can result from concerted or sequential elimination from a single CF2Br2 molecule CF2Br2 + hv —> CF2 + Br2 (4.2) 58 OI' CF2Br2 + hv —-> CF2Br + Br CF2Br + hv —> CF2 + Br (4.3) Br + Br —> Br2 or from the recombination of Br atoms produced from photolysis of two CF2Br2 molecules 2CFzBr2 + 2hv —> 2CF2Br + 2Br (4.4) Br + Br —> Br2 Even at the shortest wavelengths employed in this study, there is insufficient energy in a single photon (hv250 m = 5 .0 eV) to simultaneously cleave two C-Br bonds (2 x Do (C-Br) E 6.0 eV).35 Therefore, we discount reaction 4.2 as the source of Br2. The photodissociation cross-section of CF2Br(g) is (0.5—4.4) x 10‘18 cm2 at 248 nm,“'44416 about an order of magnitude greater than the value for CF2Br2(g) at the same wavelength. As such, we anticipate significant photodissociation of the CF2Br photoproduct to occur. We found no evidence for the formation of CF2(ad); however, the formation of CF2(g) which is expelled from the adlayer following photolysis Of CF2Br(ad), does not preclude the formation of adsorbed Br. We are, therefore, unable to comment on the relative contributions of reactions 4.3 and 4.4 to the production of Br atoms. The XPS data indicated Br-containing species formed during UV irradiation Of CF2Br2 remained adsorbed on the surface, as shown in Figure 4.7. Consistent with the production of several species (CF2Br, Br, Br2, C2F4Br2) during photolysis, the Br 3d peak broadened from 3.9 to 4.8 eV FWHM between 0 and 60 minutes irradiation, respectively. The peak center also shifted from 71.6 eV to 70.4 eV BE. The shift implies a net reduction in the number of Br atoms in an electronegative environment as expected for Br 59 bi pi. fr and Br2 compared with CF2Br2 (we assume the binding energies for Br in CF2Br2 and C2F4Br2 are similar). A second photoproduct formed was C2F4Br2 (Halon 2402) by dimerization of two photogenerated CF2Br radicals. The dimer is indicated by the appearance of desorption peaks at ~165 K for m/z = 100 (C2F4I) and 179 (C2F4Br+), as shown in Figures 4.3 and 4.4. The measured TPD intensity ratio at these two masses (47:100 for m/z = 100 and 179, respectively) agrees with our measured QMS ion abundances for C2F4Br2(g) (43: 100 for m/z = 100 and 179, respectively), confirming the presence of C2F4Br2(ad). It should be noted that the unirradiated CF2Br2 monolayer showed a small TPD peak at ~ 141 K for m/z = 100. This is believed to be an impurity; there is no peak at this m/z in the mass spectrum of CF2Br2(g).56 It is unlikely that any Br (or Br2) is contributed by the direct photolysis of C2F4Br2. Ultraviolet photodissociation cross sections for gas- phase C2F4Br2 at the wavelengths used in our experiments (225-350 nm) are quite small (~ 10.20 _ 10.21 cmz).50'51 4.2.7 Loss of Fluorine Species. The total loss of parent CF2Br2 as determined from TPD measurements (~ 140 K peak), was approximately 14 i 3 % after 15 minutes irradiation. This value is almost identical to the decrease in the F ls/C ls XPS intensity ratio after the same UV irradiation period (13 i 5 %). However, it will be recalled that there was essentially no change in the Br 3d/C ls XPS intensity ratio over these (or extended) UV exposures. This immediately implies that the Observed loss of fluorine- containing species cannot be due to photodesorption of CF2Br2 or CF2Br, both of which would reduce the total Br concentration on the surface. 60 In addition to CF2Br, both CF2 and C2F4 products have been Observed in matrix isolation experiments of CF2Br2 photochemistry.36 Although we see no evidence for adsorbed CF2 in our EELS data, it is possible that the CF2 formed during photolysis of CF2Br is expelled from the surface as CF2(g). The formation of CF2(ad) would be signaled by the appearance of vibrational modes at ~1225 cm"(vs — a1 symmetry) and ~1106 cm'l (vas — b1 symmetry).56 Alternatively, CF2 photofragments may combine to produce C2F4(ad) or C2F4(g). We can discount the formation of appreciable quantities of C2F4(ad) since at no time did we observe losses in the EELS data due to C2F4(ad) (~ 1180 and 1330 cm")56 or in TPD data attributable to the cracking of QR; (notably at m/z = 31 and 100)?6 In principle, it should be possible to determine the overall stoichiometry of the departing fluorine-containing species by examining the reduction in the ~293 eV feature in the C Is XP spectrum (Figure 4.6) with UV exposure. However, the difficulties associated with background subtraction and the relative weakness of the feature (the number of C atoms possibly leaving the surface as CF2(g) is very small compared to the number of C atoms sampled in the HOPG) made the results unreliable. A shift in binding energy (-0.9 eV after 60 minutes UV irradiation) was observed for the F ls peak, presumably due to net loss of the Br atom. In contrast to the Br 3d peak, the F ls peak exhibited no discemable broadening, implying that only one type Of F chemical environment was present on the surface after extended photolysis. The remaining species, CF2Br and C2F4Br2, both contain similar bonding arrangements for F so different binding energies for these adsorbates are not expected. 61 Several authors have studied the photodissociation of CF2Br2(g) and determined that approximately 20-30% of the CF2Br formed following the initial CF2Br2 photolysis undergoes a second C-Br scission via a vibrationally excited CF2Br* radical.42‘43’45 This process generates a second Br atom and a CF2 radical. Such a scenario is also consistent with our Observations for loss of fluorine from CF2Br2(ad)/graphite (see Figure 4.7) if the vibrationally excited CF2Br* subsequently fragments to produce CF2(g) and Br(ad). Indeed, the rate of loss of F-containing species is, on average, 25% the rate Of dissociation of CF2Br2(ad)/graphite (Figure 4.9), which is in good agreement with the value of Gosnell, et. al.42 Experimental geometry prevented us from directly monitoring species desorbing during irradiation, and we cannot independently confirm desorption of CF2(g) from the adlayer in our case. 4.2.8 Quantification of Surface Reactions. Calculations were performed based on our TPD data in order to quantify the various products on the HOPG surface after UV irradiation. Electron impact ionization cross sections were calculated for CF2Br2, Br2, and C2F4Br2 using Deutsch-Mark (D-M) theory57 in order to correct for ionization efficiency of the products at the electron energy of our QMS (70 eV). Deutsch-Mark theory is a semi-classical approach based on quantum-mechanically optimized molecular geometry, population weighting factors, and summed contributions arising from electron ejection by each occupied molecular orbital. Of the various calculation schemes used to determine absolute electron impact ionization cross-sections, Deutsch-M'ark theory appears to give the most reliable results for molecules containing heavy atoms.58 Calculated electron impact ionization cross-sections for CF2Br2, Br2, and C2F4Br2 were 62 100 - 80 _ A A Total CF2Br2 ‘ I 60 - CFZBr E . 8 . a) I 0- e ‘ 4o 4 . ‘ CFzBr+Br . . 0 20 - - Br2 0 _ ‘__ ___ 4 CZF48r2 Time (minutes) Figure 4.9 Relative percentages of the different reaction channels present on the surface as a function of UV irradiation. The lines through the data are guides for the eye. 63 11.25 A2, 7.78 A2, and 14.48 A2, respectively. 59 Corrections were also made for ion fragmentation using published and measured mass spectra for CF2Br2(g), Br2(g) and C2F4Br2(g) by multiplying each cross-section by the sum of the ion intensities of each fragment divided by the intensity of the fragment of interest .56 Figure 4.9 summarizes the relative rates of the various processes observed during UV irradiation of 4.8 L CF2Br2(ad)/HOPG, after correction for electron impact ionization cross-sections and fragmentation. The most prominent reaction channel for all UV exposures studied was the C-Br photodissociation event as shown in equation 4.1. For example, after 15 minutes of UV irradiation, the monolayer is composed of approximately 45 i 7 % unphotolyzed and 55 :l: 7 % photolyzed CF2Br2 molecules. During temperature programmed desorption experiments, most of the photolyzed molecules (70 i 10 % of the photolyzed fraction or 38 i 6 % of the total CF2Br2 molecules) recombine to regenerate parent molecules (155 K peak in TPD data). The remainder of the photolyzed molecules (30 i 5% of the photolyzed fraction or 17 j: 3% of the total CF2Br2 molecules) formed other reaction products. The fluorine-containing component of this remaining 17 2t 3% appears to be almost completely lost from the adlayer (as indicated by both F ls/C ls XPS and m/z = 129 TPD data). The formation of C2F4Br2 and Br2 increased to 1.4 :l: 0.6% and 6 i 2% of the total number of molecules, respectively, after 15 minutes of irradiation, indicating that recombination of CF2Br radicals and Br atoms are relatively minor channels. We observe various recombination reactions during TPD measurements due to the formation of an adlayer with a high density of photogenerated radicals. However, we observe reaction probabilities that are significantly different to those observed for the photolysis of CF2Br2(g). The most prevalent reaction observed in our experiments was the recombination of a large fraction of the photolyzed molecules to reform the parent molecule during heating in TPD measurements, a channel not detected in static-cell gas 3340:4144“ or (low temperature) matrix isolation36 studies. Vatsa et. al.46 have phase suggested that the majority (90%) of the CF2Br radicals produced by 248 nm photodissociation of CF2Br2(g) in a cell formed C2F4Br2 through dimerization reactions. We Observed that the formation of C2F4Br2 was a minor channel on the surface reaching a maximum after 10 minutes of photolysis. The proportions of the various products are not simply statistical if only CF2Br and Br are produced. In this case, we would expect the formation of equimolar amounts of Br2 and C2F4Br2. The formation of a higher amount of Br2 than expected (by about a factor of four compared to C2F4Br2) is likely due to efficient photolysis of CF2Br(ad), spontaneous C-Br scission in CF2Br* or other controlling factors operative on the surface. Surface processes may include reduced reaction probability for a particular recombination, perhaps due to structural (orientational) constraints, or reduced surface diffusion for one of the species. 4.3 Conclusions. Dibromodifluoromethane (CF2Br2) adsorbed molecularly on an HOPG surface at 85 K, with monolayer saturation corresponding to a fractional coverage of 0.14 :l: 0.01 CF2Br2 per surface C atom. Molecular desorption occurred from the monolayer at approximately 138-141 K. Photolysis of CF2Br2 was Observed upon exposure to 225-350 nm Hg arc lamp irradiation. The major products formed during photolysis were CF2Br and Br atoms. The estimated integral cross-section (225-350 nm) 2 for this process was ~ 1.9 x 10'19 cm , similar to the integrated UV photolysis cross- section for CF2Br2(g) at these wavelengths. This implies that dissociative electron 65 attachment (DEA) does not significantly contribute to the adsorbate photochemistry. A large fraction Of the CF2Br and Br recombine to produce CF2Br2 during TPD experiments. Additionally, minor channels for Br2 and C2F4Br2 (Halon 2402) formation were Observed. It should be noted that the adsorption/desorption temperatures Observed for all species in our work are 30-50 K lower than the minimum temperatures of the upper troposphere or lower stratosphere. Within the validity of our HOPG surface as an accurate model for primary carbon aerosols, the observed reactions will be negligible in the atmosphere. 66 References Quinlan, M. A.; Reihs, C. M.;. Golden, D.M; Tolbert, M.A. J. Phys. Chem. 1990, 94, 3255. Tolbert, M. A.; Rossi, M. J .; Malhotra, R.; Golden, D. M. Science 1987, 238, 1258. Donsig; H. A.; Herridge, D.; Vickerrnan, J. C. J. Phys. Chem. A 1999, 103, 9211. Berland, B. S.; Tolbert, M. A.; George, S. M. J. Phys. Chem. A 1997, 101, 9954. Finlayson-Pitts, B. J .; Pitts, Jr., J. N. Chemistry of the Upper and Lower Atmosphere; . Academic Press: San Diego, 2000. 10. ll. 12. 13. 14. 15. 16. 17. Schauffler, S. M.; Atlas, E. L.; Flocke, F.; Lueb, R. A.; Stroud, V.; Travnicek, W. Geophys. Res. Lett. 1998, 25, 317. Molina, M. J.; Molina, L. T.; Kolb, C. E. Annu. Rev. Phys. Chem. 1996, 47, 327. United Nations Environmental Program, Montreal Protocol on Substances That Deplete the Ozone Layer; Montreal, 1987. Fraser, P. J.; Oram, D. E.; Reeves, C. E.; Penkett, S. A.; McCulloch, A. M. J. Geophys. Res. 1999, 104, 15985. Zhang, R.; Jayne, J. T.; Molina, M. J. J. Phys. Chem. 1994, 98, 867. Roberts, J. T. Acc. Chem. Res. 1998, 31 , 415. Oum, K. W.; Lakin, M. J.; DeHaan, D. O.; Brauers, T.; Finlayson-Pitts, B. J. Science 1998, 279, 74. Hemminger, J. C. Int. Rev, Phys. Chem. 1999, 18, 387. DeHaan, D. O.; Brauers, T.; Oum, K.; Stutz, J.; Nordmeyer, T.; Finlayson-Pitts, B. J. Int. Rev. Phys. Chem. 1999, I8, 343. Andreae, M. O.; Crutzen, P. J. Science 1997, 276, 1052. Zhou, X.-L.; Zhu, X.—Y.; White, J. M. Surf Sci. Rep. 1991, 13, 73. de Jong, A. M.; Niemantsverdriet, J. W. Surf Sci. 1990, 233, 355. 67 18. Yates, Jr., J. T. Methods of Experimental Physics, VOl.22.; Academic Press: San Diego, 1985; p. 425. 19. Robinson, G. N.; Freedman, A.; Kolb, C. E.; Worsnop, D. R. Geophys. Res. Lett. 1994. 21 , 377. 20. Wagner, C. D.; Riggs, W. M.; Davis, L. E.; Moulder, J. F.; Mullenburg, G. E. (Ed.) Handbook of X-ray photoelectron Spectroscopy; Perkin-Elmer Corporation: Eden Prairie, 1979; p 94. 21. Zhou, X. —L.; White, J. M. Surf Sci. 1991, 241, 259. 22. Briggs, D. In Practical Surface Analysis, Briggs, D., Seah, M., Eds.; Wiley and Sons: New York, 1990; Vol. 1, p 444. 23. Palmer, R. E.; Annett, J. F.; Willis, R. F. Phys. Rev. Lett. 1987, 58, 2490. 24. Eix, S. L.; Schlueter, S. A.; Anderson, A. J. Raman Spectrosc. 1992, 23, 495. 25. Baldacci, A.; Gambi, A;. Giorgianni, S.; Visinoni, R. and Ghersetti, S. Spectrochim. Acta 1987, 43A, 455. 26. Golze, M; Grunze, M.; Hirschwald, W Vacuum 1981, 31 , 697. 27. Niemantsverdriet, J. W.; Markert, K.; Wandelt, K. Appl. Surf Sci. 1988, 31 , 211. 28. Zhou, X. —L.; White, J. M.; Koel, B. E. Surf Sci. 1989, 218, 201. 29. Knorr, K. Phys. Rep. 1992, 214, 113. 30. Knorr, K.; Civera-Garcia, E. Surf Sci. 1990, 232, 203. 31. Nalezinski, R;. Bradshaw, A.M.; Knorr, K. Surf Sci. 1995, 333, 255. 32. Volkmann, U. G.; Knorr, K. Phys. Rev. B 1993, 47, 4011. 33. Warken, A.; Enderle, M.; Knorr, K. Phys. Rev. B 2000, 61, 3028. 34. Chakarov, D. V.; Osterlund, L.; Kasemo, B. Vacuum 1995, 46, 1109. 35. Lide, D. R. Ed.; CRC Handbook of Chemistry and Physics, 75'” ed.; CRC Press: Boca Raton, 1995. 36. Jacox, M. E. Chem. Phys. Lett. 1977, 53, 192. 37. Johnson, C. A. F.; Ross, H. J. J. Chem. Soc. Farad. Trans. 1 1978, 74, 2930. 68 (II (_II 38. 39. 40. 41. 42. 43. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. Sam, C. L.; Yardley, J. T. Chem. Phys. Lett. 1978, 61 , 509. Wampler, F. B.; Tiee, J. J .; Rice, W. W.; Oldenborg, R. C. J. Chem. Phys. 1979, 71 , 3926. Molina, L. T.; Molina, M. J. J. Phys. Chem. 1983, 87, 1306. Krajnovich, D.; Zhang, 2.; Butler, L.; Lee, Y. T. J. Phys. Chem. 1984, 88, 4561. Gosnell, T. R;. Taylor A. J .; Lyman, J. L. J. Chem. Phys. 1991, 94, 5949. Talukdar, R. K.; Vaghijani, G. L.; Ravishankara, A. R. J. Chem. Phys. 1992, 96, 8194. . Vatsa, R. K.; Kumar, A.; Naik, P. D.; Rama Rao, K. V. S.; Mittal, J. P Chem. Phys. Lett. 1993, 207, 75. Van Hoeymissen, J .; Uten, W.; Peeters, J. Chem. Phys. Lett. 1994, 226, 159. Vatsa, R. K.; Kumar, A.; Naik, P. D.; Rama Rao, K. V. S.; Mittal, J. P. Bull. Chem. Soc. Japan 1995, 68, 2817. Talukdar, R. K.; Hunter, M.; Warren, R. F.; Burkholder, J. B.; Ravishankara, A. R. Chem. Phys. Lett. 1996, 262, 669. Esposito, E.; Femandes, N.; FitzSimons, C.; Marino III, E. Eds.; Light Sources, Monochromators and Spectrographs, Detectors and Detection Systems, and Fiber Optics, Vol. 2; Oriel Corporation: Stratford, 1994; p. 1-42. Calvert, J. G.; Pitts, Jr., J. N. Photochemistry; Wiley and Sons: New York, 1966; p. 729. Orkin, V. L.; Kasimovskaya, E. E. J. Atmos. Chem. 1995, 21, 1. Burkholder, J. B.; Wilson, R. R.; Gierczak, T.; Talukdar, R.; McKeen, S. A.; Orlando, J. J.; Vaghjiani, G. L.; Ravishankara, A. R. J. Geophys. Res. 1991, 96, 5025. Ellis, A. B.; Geselbracht, M. J.; Johnson, B. J.; Lisensky, G. C.; Robinson, W. R. Teaching General Chemistry: A Materials Science Companion; American Chemical Society: Washington D. C., 1993; p. 193. Burdett, J. K. Chemical Bonding in Solids; Oxford: New York, 1995; p. 71. Somorjai, G. A. Introduction to Sudace Chemistry and Catalysis; Wiley and Sons: New York, 1994; p. 381. 69 55. Underwood-Lemons, T.; Gergel, T. J .; Moore, J. H. J. Chem. Phys. 1995, 102, 119. 56. N. M. S. D. Center; S. Stein, In NIST Chemistry WebBook, NIST Standard Reference Database Number 69; W. Mallard, P. Linstrom, Eds.; National Institutes of Standards and Technology: Gaithersburg, MD, 1998 (http://webbook.nist.gov). 57. Deutsch, H.; Becker, K.; Matt, S.; Mark, T. D. Int. J. Mass Spec. 2000, 197, 37. 58. Vallance, C.; Harris, S. A.; Hudson, J. E.; Harland, P. W. J. Phys. B 1997, 30, 2465. 59. P. W. Harland (personal communication). 70 Chapter 5 Theoretical Investigation of the Adsorption Sites and Orientation of CF2Br2 and CF2Br on Model Graphite Surfaces Abstract Adsorption sites and orientations of CF2Br2 and CF2Br on model graphite surfaces were determined using a cluster model approach. Second-order Mdller- Plesset perturbation theory (MP2) calculations were performed using 6-3lG* and 6-3lG(2d) basis sets with four ring (pyrene) and seven ring (coronene) systems as models for the graphite surface. The preferred adsorption site for CF2Br2 was found to be located at the hollow site of a six-membered ring for each combination of surfaces and basis sets examined. For the CF2Br radical, different preferred adsorption sites were found for the four and seven ring substrates. The atop site was the preferred location on the four ring system while the bridge site was preferred on the seven ring system. Rotation of the molecules on the seven ring surface revealed similar binding energies compared to those Obtained for the initial geometries suggesting that free rotation may occur at each adsorption site. 71 5.1 Introduction. Adsorption and reaction Of small molecules on graphite and various carbonaceous surfaces has attracted a great deal of attention due to their role in such fields as astrophysics,” air and water purification,5 and atmospheric chemistry.(*26 The ubiquity of carbonaceous surfaces in such diverse environments make them a likely surface upon which reactions can take place. For example, the formation of molecular hydrogen in the interstellar medium is thought to occur by recombinative desorption of two H atoms on particles coated with a graphite-like layer."4 Heterogeneous reactions occurring on carbonaceous aerosols in the troposphere have been implicated in the reduction of such adsorbed species as HNO3, NO2, and O; as well.‘5'14’19’24’26 However, due to the diverse nature of carbonaceous surfaces (porosity, surface functional groups, etc.), the mechanisms by which these reactions take place are not well understood. In an effort to determine the role of the surface in such processes, a number of ab initio and molecular dynamics calculations have been performed examining the binding energies, sites, orientations, and mono- and multi-layer structures of various 2742 The dominant interaction between most molecules and adsorbates on graphite. graphite occurs through dispersion forces which result in the molecules being physisorbed on the surface. Numerous studies have been performed on the adsorption - - - 27.29-3239 properties of some homonuclear dratonucs (H2, N2, and 02) as well as larger adsorbates such halogenated methanes, ethane, and acetone.3‘3‘38'4°’42 For calculations involving the homonuclear diatomics, it was observed that adsorption energies of less than 10 kJ/mol were obtained in each case with a majority finding the preferred 27.29—32.39 adsorption site to occur over the center of a six-membered ring on the surface. The preferred adsorption orientation for most of these species was observed to occur with 72 the axis of the molecule parallel to the surface due to steric and electrostatic effects. Molecular dynamics (MD) calculations on a series of dipolar halogenated methanes revealed that the global energy minimum occurred for a position slightly removed from the atop site with the molecules in a “tipped” orientation (the molecular axis forms an angle with the surface normal).36 Simulations of the formation of mono- and multi-layers Of acetone41 revealed that islanding of the adsorbate occurred (Volmer-Weber growth mode) in contrast to the layer-by-layer growth for non-polar species.37'38‘42 The particular molecules of interest in this study are CF2Br2 and CF2Br and their interaction with a graphite surface. Dibromodifluoromethane belongs to a class of molecules known as Halons (CF2Br2 is termed Halon 1202) that are commonly used in fire suppression and gas and oil production.43 To date, only one experimental study has been performed on the adsorption (and photochemistry) of CF2Br2 on a HOPG surface.44 It was determined that molecular adsorption occurred on the surface with a structure similar to that of CF2Cl2 adsorbed on HOPG and that the initial photochemistry of CF2Br2(ad/HOPG closely resembled that of gas phase CF2Br2.45’57 CF2Br2 + hv —> CF2Br + Br (5.1) The present study aims to address the preferred adsorption sites and orientations for both pre- and post-irradiated CF2Br2 adsorbed on HOPG. Cluster model calculations have been performed examining the interaction between CF2Br2 and CF2Br with model graphite surfaces as a way of obtaining information on the local environment of each adsorbate. Comparison of the results presented here with those obtained from the experimental study will allow for a more complete description of this system. 73 5.2 Computational Details. Calculations of the preferred adsorption sites and orientations of CF2Br2 and CF2Br on model graphite surfaces were carried out using the cluster model approach.58 A cluster model was adopted to obtain information on the localized interaction between an isolated adsorbate and the substrate. A single adsorbate interacting with the substrate corresponds to a zero-coverage limit for each species. The graphite substrate was modeled using a single layer slab approximation consisting of either a four-ring (pyrene) or seven-ring (coronene) polycyclic aromatic hydrocarbon (PAH), as shown in Figure 5.1. Previous studies have shown that the interaction between the first and second layers in graphite is weak compared to the in-plane bonding between carbon atoms and that a single layer is often sufficient to provide a reasonable representation of the electronic properties of the three-dimensional system.28‘3°""9'63 Furthermore, the use of small PAHs allows for reduced computational effort due to a smaller number of basis functions involved for each calculation. Both ring systems were stabilized by terminating the edge carbon atoms with hydrogen atoms eliminating the presence of all carbon dangling bonds. The adsorption sites used for calculating all of the potential energy curves are shown in Figure 5.1. The indicated positions will be referred to as the atop, hollow, and bridge sites. Two basis sets, 6-31G* and 6-31G(2d), were used to calculate the potential energy curves between each adsorbate and substrate. These medium-sized basis sets were chosen because they contain polarization functions that are essential in describing weakly bound systems.“65 The 6-31G(2d) contains a second d function on all heavy atoms that allows for added flexibility in the valence region. An attempt was made at 74 b) Figure 5. 1 Four (a) and seven ring (b) model graphite surfaces with adsorption positions indicated. The atop site occurs over a substrate atom, the bridge site is between to substrate atoms and the hollow site is at the center of a six-membered substrate ring. 75 using the 6-31+G* basis set in these calculations because it has been suggested that diffuse functions are important in obtaining significant portions of the interaction energy.3°'64’65 However, convergence of the SCF equations could not be achieved with the addition of diffuse functions. All calculations were performed with the core electrons (C Is, F ls, Br Is, 25, 2p, 3s, 3p, and 3d) held frozen.66 Calculations of the optimized molecular geometries and potential energy curves were performed using second-order Moller-Plesset perturbation theory (MP2). MP2 yields qualitatively similar interaction energies for weakly bound systems when 30 This theory fundamentally includes compared to higher level correlated techniques. contributions to the interaction energy that are due to electrostatic, induction, dispersion and exchange effects present in weakly bound systems64‘65 and has the benefit of being size-consistent. Also, due to the size of the systems being studied here, MP2 provides for dynamic electron correlation at a reduced computational expense compared to multi- reference methods. Geometry optimizations for each molecule were performed at the MP2 level of theory using the 6-3lG* basis set. All calculations of the preferred adsorption sites and geometries for the adsorbates on both the four and seven ring substrates (as determined from the largest negative value for the interaction energy) were carried out using the previously mentioned basis sets with the optimized 6-31G* geometries. The four and seven ring systems were chosen to investigate changes in the binding energies, adsorption sites and orientations as the size of the substrate was increased. The minimum height of the molecules above the surfaces was determined by the distance between the C atom of the molecule and the plane containing the surface. 76 SL CZ DE (SD; .333 The Optimized geometries were kept rigid during the calculation of the interaction energies and were assumed to remain unchanged during cluster formation. The geometries of CF2Br2 and CF2Br on the four ring substrate were arranged so that the maximum molecular surface area/electron density of the adsorbate was interacting with the substrate as seen in Figures 5.2 and 5.3, respectively. Rotation of the adsorbed molecules on this surface resulted in one or more atoms not being in contact with the substrate. The use of the seven ring system allowed for rotation of the adsorbates such that all atoms remained on the substrate and new orientations could be investigated for _ calculation of the interaction energy. New molecular orientations were obtained with the F and Br atoms of the adsorbates placed either over the center of a ring in the substrate or nearly on top of an atom of the substrate. The orientation of CF2Br2 interacting with the model graphite surfaces (Br2F “tripod” approaching the surface) was determined by a previous experimental study of the adsorption of this molecule on a highly-ordered pyrolytic graphite surface (HOPG)44 and by comparison with the adsorption of CF2Cl2 on HOPG.“71 Partial geometry optimization calculations confirmed this geometry. The geometry of CF2Br with respect to the substrate was determined by partial geometry optimization calculations as well. Calculations of the interaction energies were performed for both basis sets by subtracting the energy ' of each isolated fragment from the energy of the cluster (superrnolecular approach) at every separation distance with basis set superposition error (BSSE) being accounted for by the counterpoise method of Boys and Bemardi72 as implemented in Gaussian 98.73 77 8a macaw Emcee»: .803 2a mfiofi 8930 3 .moaawm as a bum one macaw onion—w Ba .2959 2a .6.an 05.55 .223 some emote A8 v8 .Bcsoa 3V doom 3 05 a 08,35 338 met .58 05 no ~5me «d 9:5...— § 3 78 use swede so as .332 so see 3 as a 8&5 .868 we: as ea :6 Hmemo on as»... 3 so 79 5.3 Results and Discussion. An initial series of calculations were undertaken to determine an appropriate basis set for the study of these systems. Geometry optimization and frequency calculations were performed on CF2Br2 at the MP2 level of theory with a variety of basis sets. Comparison of bond lengths and angles between the calculated and experimental values could not be made because the crystal structure of CF2Br2 has not been determined to date. However, vibrational frequencies of this molecule have been experimentally determined74 and can be compared to the calculated values. The results are summarized in Table 5.1. The computed frequencies obtained from calculations using each basis set were comparable to the experimental values. The 6-31G* basis set produced frequencies that matched well with lower frequency stretches Of CF2Br2 but yielded larger differences for the higher frequency Stretching modes. The larger basis sets (cc-pVDZ and 6-311G*) gave improved values for the higher frequency stretches, but comparison with the lower frequency stretches showed increasing deviations. From these data, it was concluded that the 6-31G* basis set would be used as the starting basis set for all calculations: this basis set produced the smallest average error in the calculated and experimental vibrational frequencies. Further evidence to support the use of this basis set comes from the geometry optimization of the 4 ring substrate (pyrene). The 6-31G* basis set yielded values for C-C bond lengths that were on average only 0.007 A larger than values obtained from a neutron diffraction study.75 A larger basis set (6-31G(2d)) was used in later calculations to determine the effects that additional basis functions in the valence region would have on adsorption sites and geometries. 80 3.8a 3.3.. 3. m: _eoem omega 3.2+ it me: 3.3+ .NE 82 3.9? .83 22 mm: seas 6.56855 .5 3.2+ .emt 2 : 3.2+ em; 2 : 3a.? it a: 88 Ease erases do 32+ Es wee 3.3+ at new 3.3+ .3 ems E as... outside emo 3; .N+ it em... 3.0.: at E. 3.3- .3 2e 8e .88... mo 3.4.? .3 E 3.3+ at E E %m as .62 m6 3; .N+ .3 as 3.2+ .3 9a 3.2 .E E as ease assess emu 33+ it ”2 3a.? at one 3.3+ at E am we... Nmo 3.3+ 63 am 3.3+ .3 as 3%.? .3 as am as: «mo 33+ .3 E 3.»: .3 E 3.2+ .3 E we seem emu p.63 .9 _ SE: 9.83 33-6282 pee 5:382 pee 38586 8:85 5953 80:th 3.5—2 23 03°33 88. names Botg 5.5» momoeuaaoa 88328 new 3585950 mo ecmEEoO fin 93:. .9385 a gone 83 agatonxo 93 83328 81 5.3.1 Adsorption Sites and Geometries of CF2Br2. The results of calculations of the interaction energies between CF2Br2 and each of the model graphite surfaces are shown in Figures 5.4-5.9. An examination of Figure 5.4 shows broad, shallow potential energy curves were obtained for each of the adsorption sites. For the 6-31G* basis set, it was observed that the minimum height of CF2Br2 above the 4 ring surface for the atop and bridge sites occurs near 4.25 A while that for the hollow site is approximately 4.1 A. These values are larger than those obtained from molecular dynamics simulations of a number of Chlorofluorocarbons interacting with graphite.”39 Hammond and Mahanti36 found that vertical adsorption heights ranged between 3.23 A (for CH3C1) and 3.61 A (for CF3Cl) for dipolar molecules while Pinches and Tildesley39 observed a distance of 3.55 A for CF4. The vertical adsorption heights calculated for CFzBrz above the graphite surface are expected to be larger due to the increase size of the van der Waals radius of a Br atom compared to a Cl atom and the restriction that the adsorbate and surface structures were frozen in the calculations. Figure 5.4 shows all three adsorption sites have comparable binding energies on the four ring substrate. The preferred adsorption site occurs over the hollow of a six- membered ring with a binding energy of at least -1 1.79 kJ/mol. In comparison, the atop (-10.48 kJ/mol) and bridge (-9.93 kJ/mol) sites were found to have smaller binding energies. Errors of 10% or less are typical for conformational energies using MP2 and double zeta basis sets so the actual preferred adsorption site cannot be determined accurately from these calculations.76 The molecular dynamics studies of Hammond and Mahanti (using Lennard-Jones potentials) found that dipolar molecules preferred an adsorption site slightly removed from the atop position with the molecules in a “tipped” 82 N 01 : A 205 j I CFZBr2 + 4 Ring 1 MP2/6-31G* 15~ I ' I Atop A 10: o Hollow 73 ‘ A Brid e E : g 2 s- >. Z 9 . 0:) J “J 0.. A A 2 I I -5; A 3 o I j A -10-1 I f a 2 o o '15 I I I I I I I I I I I 4 5 6 C - 4 Ring Distance (Angstroms) Figure 5.4 Potential energy curves for CF2Br2 approaching the four ring substrate at the MP2/6- 316* level of theory. 83 geometry (molecular dipole not aligned with surface normal).36 These molecules tended to shift slightly from the atop site towards a bridge site, i.e. the carbon atom of the adsorbate moved to a position over a CC substrate bond which was the global minimum on the potential energy surface. As the adsorbate changed position, shifts in the binding energy and vertical distance above the surface occurred. For example, the largest shift in energy (0.11 kJ/mol) and distance (0.29 A) occurred for CH3C1. The energy differences between the hollow and bridge sites calculated here are larger than those of the MD simulations of Hammond and Mahanti.36 However, typical diffusion barriers of adsorbates on metal and semiconductor surfaces range between 10-25% of the binding energy77 so diffusion between all of these positions may readily occur. The results of Figure 5.5 revealed a similar trend in the preferred adsorption site and binding energies as the size of the basis set was increased for the system. The potential energy curves calculated with the 6-310(2d) basis set are not as broad as those calculated with the 6310* basis set and the minimum distance above the surface for each site has shifted to 4.0 A compared to values of 4.1 A (hollow) or 4.25 A (atop and bridge) for the 6-31G* basis set. Also, an increase in the binding energies as well as the energy differences between the hollow and atop and the hollow and bridge sites occurred. Diffusion of the adsorbate on the surface will probably occur even though the differences between the binding energies at each site have increased. It should be noted that the goal of this study is not to obtain the most accurate binding energy for the adsorbate to the surface but to gain information about preferred adsorption sites. The larger basis set provides a better description of the electronic properties of both the adsorbate and the 84 _s U" 3 CFZBr2 + 4 Ring 10 l A MP2/6-31G(2d) 5 j - Atop . ' o Hollow 3 A Bridge 0 _ 9 ' A O .- g 1 ° - .2. -5- >. 1 9 - A 1% -1o-: ' 4 A : I f -15 - o I A i . . I . -20 1 D '25 a ' I I I ' I I I I . I 4 5 6 C - 4 Ring Distance (Angstroms) Figure 5.5 Potential energy curves for CF2Br2 approaching the four ring substrate at the MP2/6- 31G(2d) level of theory. 85 substrate so the interaction energy will increase and the distance above the substrate will decrease. Alternatively, the size of the substrate was increased while fixing the basis set (6- 316*) to examine variations in the binding energy or sites that occurred and also to observe changes due to rotation of the molecule on the surface (Figures 5.6-5.9). The molecule was rotated about an axis perpendicular to the surface that contained the C atom of the adsorbate and a point directly below the molecule on the surface (see Figures 5.6 and 5.7). Comparison of the 7-ring and 4-ring surfaces utilizing the same basis set revealed that the potential energy curves were not as broad and the binding energies at all sites increased in the former case (Figure 5.8). The values obtained for the binding energies on the 7-ring system using the 6-31G* basis set fall between those of the 4 ring system using the 6-BIG* and 6-31G(2d) basis sets and follow the same trend in adsorption site energies (hollow, atop, bridge) as seen previously. An increase in the binding energies is expected due to the larger van der Waals attraction between the surface and the molecule. As a first approximation, these values demonstrate that changing the size of the substrate had less of an effect than changing the size of the basis set. On the 7 ring system, the minimum vertical distance of CF2Br2 above the surface occurs at 4.0 A which is the same for the pyrene/6-31G(2d) system but slightly shorter than the value for the pyrene/6-310* system. The energy difference between the hollow and atop sites for the coronene/6-3lG* system becomes noticeably smaller (0.58 kJ/mol) . compared to the other systems discussed previously (1.31 kJ/mol for 4 ring/6-31G* and 86 .83.. we: 82... _oeoa as 8 Season .33 3 e8 52.2. 3. do... E 2e 2. Been? 3.5 3. 23$ 8V 3V 3 87 data 9E 8% 32: ea 8 2528a $63 9 2a .252 Q .33 3 2a a Bag £50 833 E 2:53 88 Energy (kJ/mol) N O .. 10- . ‘ I a + . 0 . CF28r2 7 Rlng g 5. Rotated 15 _ MP2/6-31G* g 0; j I Atop “2:34;: a C t A 10 _ A C HC.)||OW “1.10. ; ‘ - A Bridge . . . ‘ . ~15. . a . I 3.'5 * 4.‘o ' 45 5 - C - 7 Ring Distance (Angstroms) : O o .: ‘ A ‘ U -5 _ f A .J ‘ . -10 _' ‘ ‘ I : o ‘ ‘ I -15 5 I - I 1 ‘20 I I I I I I I I I I I I I 3 4 5 6 C - 7 Ring Distance (Angstroms) Figure 5.8 Potential energy curves for CFzBrz approaching the seven ring substrate at the MP2/6816* level of theory. 89 2.55 kJ/mol for 6-3lG(2d)). From this comparison, it may be concluded that the actual adsorption site may lie between these sites. Rotation of the molecule on the surface revealed similar magnitudes of binding energies compared to non-rotated systems. It can be seen that the hollow and atop sites are now equal in energy with the bridge site still having the smallest binding energy. From these values, it appears that the molecule can freely rotate at each adsorption site because there is almost no energy barrier to rotation. The molecular orientation does not seem to make much difference for the binding energy to the substrate for the two orientations tested here. Additional calculations were performed on the 7-ring system using the 6-3lG(2d) basis set with the molecule in a “non-rotated” orientation. A smaller number of calculations were done on this system around the minimum of the potential energy surface due to the large number of basis functions involved. Figure 5.9 shows a comparison between the system using the 6-3lG* and the 6-31G(2d) basis sets. Similar results were obtained for the energy ordering of the adsorption sites (hollow site preferred) and the minimum vertical distance above the surface (4 A). Using a comparatively larger double zeta basis set for the calculation yielded ~ 58% of the binding energy as determined from temperature programmed desorption data.44 Use of larger basis sets than those here would obtain a greater binding energy but it would become computationally prohibitive to do so. From the data presented here, it appears that the most probable adsorption site for CF2Br2 on graphite occurs at the hollow site with free rotation of the molecule. 90 _ej onBr2 + 7 Ring -16; CFzBrz + 7 Ring . 6-31G* . 6-31G(2d) -7- -17- . A i -8- -18— . ‘ ‘ A -9- -19- ’3‘ -10- 320- g - 2 4 fi '7 5, -11 d - é—Z‘l - A > , >s ‘ I 9 -12- 9122‘ o a) . A ‘ 8 . ‘fi -13- ° ”-234 I -14- -24- . . . -15. ’ -25- . , I I g -16.. o -26- -17 . . . -27 . . . 3.5 4.0 4.5 3.5 4.0 4.5 C - 7 Ring Distance (Angstroms) C - 7 Ring Distance (Angstroms) Figure 5.9 Comparison of the effects of basis sets on the binding energies and adsorption sites for CF2Br2 on the model seven ring surface. The squares denote the atop site, circles denote the hollow site, and the triangles denote the bridge site. 91 An analysis of the Mulliken atomic charges and dipole moments of the complex revealed the preference for the hollow site. Table 5.2 shows atomic charge and dipole moment values for gas-phase CFzBrz and the 4-ring and 7-ring complexes for each basis set. Calculated dipole moments of 0.7051 D (6-3lG*) and 0.6004 D (6-31G(2d)) were obtained for gas phase CFzBrz. These values compare well to the experimental value of 0.66 D.78 When the molecule is adsorbed over the hollow site in each of the systems studied, an increase in dipole moment occurred compared to the gas-phase value, while the dipole moments for the atop and bridge sites decreased. This trend was not observed for rotation of the molecule on the 7-ring substrate. Comparison of the dipole moments for the atop and hollow sites for the rotated molecule on the 7-ring complex revealed similar values and nearly identical binding energies. Upon adsorption of the molecule on the substrate, charge is transferred from the surface to the adsorbate. The amount of charge transferred is dependent upon the adsorption site of the molecule. In the structures of Figure 5.6 for example, it was observed that the largest amount of charge transfer occurred at the bridge site. At this location, three atoms of the adsorbate (C and both Br atoms) are over sites of increased electron density. An increase in charge on the “up” F atom (F atom farthest from the surface) and both Br atoms compared to the gas-phase molecule can be observed in Table 5.2. The increase in charge on these atoms leads to a subsequent lowering of the dipole moment because the charge separation within the molecule is reduced. A comparatively small amount of charge transfer occurred at the hollow site. Here, most of the adsorbate atoms are over sites of lower electron density except for the “down” F atom (F atom 92 35895 05 no mom—Eu o5 .«o 83 05 Ba macros—m 05 :o woman—0 05 mo 88 08 5933 “Canasta u a «88.—0&5 «$8.? £33. 333. 383. $83- $.83. ~53- 883. 0956 «23 83° 33 E3 53 33 383 33 é a $2.3. $33. 833. 8:3. ”53- 233. 5 S3. $2.3. .._ ago? 8283 2883 2383 $83. $33. 383. 333 :33. am 233. 88m? 8333. Read. 3E3. SE3. 2: s3. ”33. e :9... 8283 8383 2383 :53. 833. 833. 82.83 :33. am 8233 3.33 ”833 «$83 $33 $283 353 3:83 0 seem 323m 82 3pm 326: 82 939 me .9 3 32 £32 3 was. 2.23 can; so use so 6an» :03 no «ammo .8.“ Bony—obfi guano 93 85:58 38% .3930 0:52 dd 02am. 93 008.5me 823- 383. 833- $33. 383. $83- 283- $83- :83- ”wane :33 ~93 323 323 833 33 «$3 383 «33 a: 8E3. 3:33. 283. 3883- 283- “53. $2.3- 383. 833. m .38.. $383 :53 :883 $83- 2 2.3- $33- a 53- S 2.3. 383- cm 8333. 2&3- $33- 2&3- 2 $3. $23. 3333. 233. 3&3. m :9: $283 $383 3233 $33. 2:33. ”$83. 3283. :53. 333- am 383 8333 $833 2333 383 $883 2283 «883 3283 o owvtm 323: QS< owvtm Bo=om n03. 03:5 30:03 n02 may. 2959 3 38335 $02-3 use 2.93 $.83 .3 2.5. 94 closest to the surface). Most of the transferred charge is localized on this F atom and the C atom of the adsorbate and augments the already existing dipole moment. The atop site showed the smallest amount of charge transfer between the surface and the molecule even though the molecular orientation is similar to the bridge site. The two Br atoms of CF2Br2 are over C atoms in the substrate, but now the C atom of the molecule is over a surface C atom as well. In this configuration, the charge on the Br atoms and the “up” F atom were observed to increase (become more negative) in most cases analogous to the bridge site. A smaller amount of charge transfer occurred because the C atom in the adsorbate is now over a region of lower electron density compared to the bridge site. Therefore, the dipole moment decreases slightly due to a small change in the charge separation in the molecule. This shift of charge within the molecule can best be seen by observing the difference between the sum of the charges on the fluorines and the sum of the charges on the bromines as seen in Table 5.2. This difference in charges shows that the atop and bridge sites have comparable values while the hollow site has a larger difference. However, the differences are reversed for the atop and hollow sites when the molecule is present on the 7 ring substrate even though the dipole moment is slightly larger at the hollow site. A variety of studies exist that have examined the structure of polar molecules 28'36'4L67'7‘ Knorr and co-workers have investigated the adsorbed on graphite surfaces. structure of CFzClz on HOPG at a variety of coverages and temperatures.”71 They determined from x-ray diffraction and IRAS data that 6 out of 9 CFzClz molecules in the unit cell sit on the PC]; tripod at an atop site for the commensurate on phase. The other three molecules are determined to be positioned above hollow sites and are 95 orientationally disordered with respect to the other molecules in the unit cell. The 0L phase represents a low coverage-low temperature phase that should be comparable to the structures calculated here. The results obtained from these calculations show that at very low coverages the CF2Br2 molecule will sit over a hollow site which is slightly preferred over the atop site and compares well to the structure obtained by Knorr.”71 Caution must be exercised when comparing the preferred adsorption site from these calculations to higher adsorbate coverages. It is believed that these dipolar molecules do not wet the graphite surface but instead form 2-dimensional islands because of stronger inter- molecular interactions.“ 5.3.2 Adsorption Sites and Geometries of CF2Br. The results of the interaction between CF28r and the model graphite surfaces are shown in Figures 5.10-5.15. Comparison of the potential energy curves between CF2Br and CF2Br2 adsorbed on the model graphite surfaces showed similar characteristics (broad shallow minima and similar vertical adsorption heights). For example, Figure 5.10 shows the interaction between CFzBr and the 4-ring substrate using the 6-31G* basis set. The minimum vertical adsorption height of the radical occurred at 4.0 A, which is comparable to the value obtained for CFzBrz interacting with the 4 ring surface (~ 4.1 A) using the 6-31G* basis set. Also, it was observed that the binding energies of the radical to the model surfaces were smaller than those calculated for CFzBrz on the model surfaces. The decrease in binding energies can be explained by noting that the dominant interaction between each species and the surface occurs through dispersion forces. The radical “borrows” electron density from the surface to increase its stability (as discussed below), but no covalent bond is formed due to the strong C-C bonding within the substrate. 96 1O Energy (kJ/mol) -10.. CF28r + 4 Ring MP2/6-31G* I Atop O Hollow A Bridge Figure 5.10 Potential energy curves for CFzBr approaching the four ring substrate at the MP2/6- 310* level of theory. 1 l 5 C - 4 Ring Distance (Angstroms) 97 A change in the preferred adsorption site was observed to occur between CFzBr and CF2Br2. The radical favors either the bridge or atop site (depending on the size of the substrate) while the molecule preferred the hollow site. The radical contains a singly occupied molecular orbital that is directed perpendicular to the plane containing the fluorine and bromine atoms of the molecule. Addition of electron density to this orbital stabilizes the interaction between the radical and the surface. Therefore, positions that contain higher amounts of electron density, such as the atop and bridge sites, are favored by the radical. Density functional theory calculations of the bare seven ring surface by Arellano, et. a1. show that the atop and bridge sites contain higher amounts of electron density compared to the hollow site.27 At present, no experimental evidence or theoretical calculations have been performed on this or any similar system so the true geometry and orientation of the molecule with respect to the surface are unknown. Expanding the size of the basis set used for CF2Br adsorbed on the 4-ring system to the 6-31G(2d) basis set increased the binding energy at each adsorption site. A more accurate description of the system is obtained with this basis set when compared to the calculations using the 6-31G* basis set. It can be seen from Figure 5.11 that a similar trend in the preferred adsorption site occurred when compared to the results obtained with the 6-31G* basis set. Now, the atop and bridge sites are separated by only 0.29 kJ/mol with the atop site having a slightly greater binding energy. A similar energy difference was observed for the 6-31G* basis set (0.33 kJ/mol). Also, the minimum vertical adsorption height above the surface for each site has shifted to a smaller value of 3.75 A and an increased curvature of the potential energy curve near the equilibrium bond distance occurred when compared to the 6-31G* calculation. 98 CFZBr+4Ring MP2/6-31G(2d) I Atop 0' o Hollow A Bridge I ‘5‘ E -5- i 3 V O >~. 9’ (D O C A UJ I -10. A o O I I A A I ~15- ‘ 3' "21' ' 'é' ' '63 C - 4 Ring Distance (Angstroms) Figure 5.11 Potential energy curves for CF2Br approaching the four ring substrate at the MP2/6- 31G(2d) level of theory. 99 .m 2:5 Hammo «— 3&8? . E a. a 3 2a .323 Q 8a a 350; 0&3.— 2: no me . . >8 $608 6an wat do 3 3 3 data we. :28 38a 2: 8 Sousa owes 3 Ea 98:2 5 den 3 2c a 89.83 ammo 850m 2.... an»... A3 3 3 101 Additional calculations were performed using the 6-31G* basis set to investigate changes in the binding energies and adsorption sites as the size of the substrate was increased. A further series of calculations were carried out on this system with rotation of the radical on the surface analogous to the CF2Br2/7 ring system (see Figures 5.12 and 5.13). It can be seen from Figure 5.14 that the values for the binding energies of the radical fell between those of the 6-31G* and 6-31G(2d) basis sets on the 4 ring surface. Similar behavior was observed in the molecular system as explained previously. It is interesting to note that the preferred adsorption site switched from the atop to the bridge site and the minimum vertical adsorption decreased to ~ 3.75 A compared to 4.0 A. The other adsorption sites remain unchanged with minimum vertical distances above the surface of 4.0 A. A closer approach of the radical to the surface at a bridge site occurs because more electron density is present at this position and the atoms of the adsorbate are not directly over those of the substrate (reduced steric hindrance). Rotation of the radical on the surface showed slight changes in the adsorption energies. A similar trend was observed for the energy ordering of the adsorption sites compared to the non-rotated system, but now the minimum vertical distance for the bridge site returned to 4.0 A. This effect is probably due to the atoms in the radical taking positions over top of the atoms in the surface. It was observed that the binding energy at the atop site decreased indicating an unfavorable orientation of the radical on the surface possibly due to a smaller amount of electron density transferred at this site. Because of the small energy differences between the rotated and non-rotated orientations, the radical may rotate freely at the preferred adsorption site. 102 10 -6 o Rotated CFZBr + 7 Ring .5 -7* ' x MP2/6-31G* 5 -8‘ _ jig-9‘ O 0 . . A 54 I Atop ““10 J O Hollow 1“ ‘ ‘ ‘ Bridge '12 3:5 ' 420 ' 415 C - 7 Ring Distance (Angstroms) 3 0- E .3 i as I >. 0) L 8 LIJ _5_ - ‘ A ‘ A -10- ' A D 3 4 5 6 C - 7 Ring Distance (Angstroms) Figure 5.14 Potential energy curves for CFzBr approaching the seven ring substrate at the MP2/6616* level of theory. 103 4“ -3i CFzBr + 7 Ring ‘ crzar + 7 Ring , 6-31G* 42* 6-31G(2d) -4. ‘ A . -13- u -5- a . A 6‘ A44- 5 - _ '5 g . g I 2 -7. ‘ 3.15. {=3 ‘ § .05) '8‘ ljag-w- . i -9. . -17~ ‘ 401 : J l -11- ‘ . '18- ‘ ‘ 4 -12 I . r n ‘19 l ' l ' 3.5 4.0 4.5 3.5 4.0 4.5 C - 7 Ring Distance (Angstroms) C - 7 Ring Distance (Angstroms) Figure 5.15 Comparison of the effects of basis sets on the binding energies and adsorption sites for CF2Br on the model seven ring surface. The squares denote the atop site, circles denote the hollow site, and triangles denote the bridge site. 104 Figure 5.15 shows a comparison of the binding energies of the radical on the 7- ring surface using both the 6—3IG* and 6-3lG(2d) basis sets. It was observed that the binding energies obtained using the 6-31(3(2d) basis set were larger than those with the 6- 31G* basis set. Similar values were obtained for the binding energies (differing by a maximum of 0.55 kJ/mol) with the bridge site preferred over the hollow and atop sites. The minimum height above the surface remained at 3.75 A which is the same value for the bridge site when compared to the 6-3 16* basis set on 7-ring system. It appeared that the bridge site is the preferred adsorption position for the radical on HOPG because it is lowest energy site on 7-ring system using either basis set. Depending on the magnitude of the diffusion barriers between sites, the radical may be able to easily diffuse across the surface because of the similarity in binding energy of each site. It may be recalled that these calculations represent the zero-coverage limit and the preferred adsorption sites and orientation may change with increased coverage. In comparison to the adsorption of the molecule on the surface, the radical preferred adsorption sites where the dipole moment of the complex has its lowest value as seen in Table 5.3. This table shows the atomic and total charges and dipole moments of the radical in the gas phase as well as on the 4- and 7-ring substrates calculated using both basis sets. The interaction between surface and radical is not as dipolar in nature when compared to the molecule-surface interaction. It will be recalled that the radical favors the at0p and bridge sites because it has a singly occupied molecular orbital to which it can add this extra electron density. The decrease in dipole moment is the result of the radical obtaining charge from the surface and spreading it evenly over all the atoms in the radical. 105 £83. «:83. E33. 883. 383. 883.3. 3 o 3&5 Bob 33 «33 883 53 :33 £93 33 ~83 1 $83. 863. SS3. 233. 883. $33. $83. 333. m ~33. “$23. :83. :33. $83. $33. $83. 883. m 3:3 3:33 3.23 32.3. 32.3. 383. 38.3 853. am $33 833 E83 :33... 2833 83.33 8383 «333 o 085 30%: 82. 3.5 323m .5... £503 .9 3 8.2m 9.23 max 352 3 max 3.9 3 so no 68.35. :98 no Banana 38 v5 £8808 28.6 .8920 283“ umumo .n.m 033—. 106 owaao ~33. 283. $83. ~88? 383. :83- 9883 8883 :83. :89 E3 083 ~83 :33 ~33 33 323 333 333 1 $33- 883- :33- 333. 833. $33- 233. $33. 833. m 833. 583. :33- :33. 833- 833. 3383- 833. $33. .._ 83:3 853 S383 333- ~33- R83- 3823- :33- 353. .m 333 :33 333 8333 333 8833 £833 253 2:33 o 3cm 323m 82 3cm 328m 8% 08cm 3%: 82 max .2892 3 333 max 8.9 3 max 2.9 3 €383 .3 2.3. 107 Comparison of the gas phase atomic charges to those of the radical on both surfaces provided further insight into the interaction of the radical with the graphite surface. The C and Br atoms of the adsorbed radical (with the 6-31G* basis set) had slightly more negative values than in the gas phase while the F atoms became slightly more positive. As a result, the charge separation in the radical is reduced and more electron density goes into the unfilled orbital. Use of the larger basis set showed that the F atoms gained charge from the surface but the amount that they acquired was nearly offset by the amount of positive charge on the C and Br atoms. Therefore, the radical gained more total charge but the dipole moment was reduced to half of its gas-phase value. Adsorption of the radical on the 7-ring surface with the 6-31G* basis set showed that the preferred binding site was over the bridge position as seen in Figure 5.14. A decrease in the dipole moment was observed due to a slightly larger negative charge on the Br atom and a slight positive charge on the C atom (compared to the gas phase values) when adsorbed at this site. It appeared that the radical donated charge to the surface as evidenced by the decreased total charge on CF2Br. Rotation of the radical on the surface caused a change in the charges on the Br and C atoms to occur. The Br atom was slightly more positive in this orientation and the C atom, as well as one of the F atoms, became more negative compared to the non-rotated case. In this position, the F atom that is part of the OP bond that is over a C atom in the substrate is closer to a site of higher electron density which allowed both the F and C atoms to become more negative and charge shifted away from the Br atom. As a result, a return of charge transfer from the surface to the radical was observed. 108 Calculations of the radical on the 7-ring substrate with the 6-316(2d) basis set showed the same preferred binding site as the 6-31G* basis set and now the signs and magnitude of the atomic charges are comparable to those of the gas phase and adsorption on the 4-ring surface. The C atom became more positive, and the Br and F atoms were more negative than in the gas phase so charge became more evenly distributed. A return of charge from the surface to the radical was observed to occur helping to stabilize the interaction between the radical and the surface. The data presented here can be compared to that obtained from our study of the photochemistry of CF2Br2 on a highly ordered pyrolytic graphite (HOPG) surface.44 In this photodissociation study, a slight peak shift and change in intensity as well as a change in background in EELS data were observed after irradiation of a monolayer of the molecule. This was attributed to Br atoms being produced on the surface that were involved in charge transfer from the surface to the Br atoms. From the data presented here, it appears that the molecule and radical are involved in this charge transfer process as well. The combination of two open shell systems (CF2Br and Br) on the surface can account for the large change in EELS background that is characteristic of a metal surface. In addition, the increase in intensity and shift of the EELS peak also can be explained. For longer irradiation times, more radicals will be present on the surface. The dipole moment CF2Br is nearly aligned with the surface normal in the adsorption geometry of this study. The asymmetric stretch of the radical will have a larger component of its dipole moment perpendicular to the surface than the molecule resulting in an increase in the intensity of the peak. The red shift in the peak occurred because the radical is not as strongly bound to the surface compared to CF2Br2. A comparison of the results presented 109 here with the values obtained from the TPD data is limited. Approximately 58% of the experimental binding energy of CF2Br2 on HOPG was obtained from these calculations utilizing the largest surface and basis set. With the knowledge of the preferred adsorption sites for each species, future simulations of both the pre- and post-irradiated monolayer will allow comparison to the TPD data. 5.4 Conclusions. It was found that the preferred adsorption site for the CF2Br2 molecule occurred over the hollow position of a six-membered ring on the surface for each basis set and model graphite surface studied. The dominant interaction between the molecule and the surface was dipolar in nature with charge transfer occurring from the surface to the molecule. It was observed that the site with the largest dipole moment for the complex was the preferred one. Rotation of the molecule on the surface showed that the molecule probably undergoes free rotation on the surface. The preferred adsorption site for the CFzBr radical occurred at the atop site on the 4 ring substrate and the bridge site on the 7 ring substrate for each of the basis sets used. The interaction between the radical and the surface was not as dipolar in nature compared to CF2Br2 even though charge was transferred from the surface to the radical. Sites of higher electron density (bridge and atop) were found to be preferred due to the presence of a singly occupied orbital in the radical. Rotation of the radical on the surface showed no significant changes in binding energy at the preferred site so free rotation of the species probably OCCUI'S . llO 10. 11. 12. 13. 14. 15. 16. 17. References Gould, R. J .; Salpeter, E. E. Astrophys. J. 1963, I38, 393. Hollenbach, D.; Salpeter, E. E. Astrophys. J. 1971, I63, 155. van der Hulst, H. C. Rec. Astron. Obs. 1949, XI(II). Williams, D. A.; Herbst, E. Surf. Sci. 2002, 500, 823. Ruthven, D. M. Principles of Adsorption and Adsorption Processes; Wiley: New York, 1984. Rogaski, C. A.; Golden, D. M.; Williams, L. R. Geophys. Res. Lett. 1997, 24, 381. Lary, D. J .; Lee, A. M.; Toumi, R.; Newchurch, M. J .; Pirre, M.; Renard, J. B. J. Geophys. Res. 1997, 102, 3671. Smith, D. M.; Akhter, M. 8.; Jassin, J. A.; Sergides, C. A.; Welch, W. F.; Chughtai, A. R. Aerosol Sci. Tech. 1989, 10, 311. Kirchner, U.; Scheer, V.; Vogt, R. J. Phys. Chem. A 2000, 104, 8908. Saathoff, H.; Naumann, K.-H.; Riemer, N.; Kamm, S.; M6hler, 0.; Schurath, U.; Vogel, H.; Vogel, B. Geophys. Res. Lett. 2001, 28, 1957. Chughtai, A. R.; Jassim, J. A.;.Peterson, J. H.; Stedman, D. H.; Smith, D. M. Aerosol Sci. Tech. 1991, 15, 112. Schurath, U.; Naumann, K.-H. Pure and Appl. Chem. 1998, 70, 1353. Al-Abadleh, H. A.; Grassian, V. H. J. Phys. Chem. A 2000, 104, 11926. Akhter, M. S.; Chughtai, A. R.; Smith, D. M. J. Phys. Chem. 1984, 88, 5334. Gerecke, A.; Thielmann, A.; Gutzwiller, L.; Rossi, M. J. Geophys. Res. Lett. 1998, 25, 2453. Kalberer, M.; Tabor, K.; Ammann, M.; Parrat, Y.; Weingartner, E.; Piguet, D.; Rossler, E.; Jost, D. T.; Tfirler, A.; Gaggeler, H. W.; Baltensperger, U. J. Phys. Chem. 1996, 100, 15487. Kleffmann, J.; Becker, K. H.; Lackhoff, M.; Wiesen, P. Phys. Chem. Chem. Phys. 1999, I, 5443. 111 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Longfellow, C. A.; Ravishankara, A. R.; Hanson, D. R. J. Geophys. Res. 1999, 104, 13833. Stadler, D.; Rossi, M. J. Phys. Chem. Chem. Phys. 2000, 2, 5420. Grassian, V. H. J. Phys. Chem. A 2002, 106, 860. Choi, W.; Leu, M.-T. J. Phys. Chem. A 1998, 102, 7618. Baldwin, A. C. Int. J. Chem. Kinet. 1982, 14, 269. Britton, L. 6.; Clarke, A. G. Am. Environ. 1980, I4, 829. Sergides, C. A.; Jassim, J. A.; Chughtai, A. R.; Smith, D. M. Applied Spec. 1987, 41 , 482. Disselkamp, R. S.; Carpenter, M. A.; Cowin, J. F.; Berkowitz, C. M.; Chapman, E. G.; Zaveri, R. A.; Laulainen, N. S. J. Geophys. Res. 2000, 105, 9767. Poschl, U.; Letzel, T.; Schauer, C.; Niessner, R. J. Phys. Chem. A 2001, 105, 4029. Arellano, J. S.; Molina, L. M.; Rubio, A.; Alonso, J. A. J. Chem. Phys. 2000, 112, 8114. Feller, D.; Jordan, K. D. J. Phys. Chem. A 2000, 104, 9971. Sorescu, D. C.; Jordan, K. D.; Avouris, P. J. Phys. Chem. B 2001, 105, 11227. Lushington, G. H.; Chabalowski, C. F. J. Molec. Struc. 2001, 544, 221. Tran, F.; Weber, J .; Wesolowski, T. A.; Cheikh, F.; Ellinger, Y.; Pauzat, F. J. Phys. Chem. B 2002, 106, 8689. Ferro, Y.; Marinelli, F.; Allouche, A. J. Chem. Phys. 2002, 116, 8124. Tanaka, H.; El-Merraoui, M.; Steele, W. A.; Kaneko, K. Chem.Phys. Lett. 2002, 352, 334. Nyman, G.; Holmlid, L. J. Chem. Phys. 1986, 85, 6163. Hansen, F. Y.; Taub, H. J. Chem. Phys. 1987, 87, 3232. Hammond, W. R.; Mahanti, S. D. Surf. Sci. 1990, 234, 308. 112 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. Hentschke, R.; Schilrmann, B. L.; Rabe, J. P. J. Chem. Phys. 1992, 96, 6213. Hentschke, R.; Winkler, R. G. J. Chem.Phys. 1993, 99, 5528. Pinches, M. R. S.; Tildesley, D. J. Surf Sci. 1996, 367, 177- Kim, Y. H.; Ree, J.; Shin, H. K. Chem. Phys. Lett. 1999, 314, 1. Kwon, 8.; Russell, J.; Zhao, X.; Vidic, R. D.; Johnson, J. K.; Borguet, E. Langmuir 2002, 18, 2595. Zhao, X.; Kwon, S.; Vidic, R. D.; Borguet, E.; Johnson, J. K. J. Chem. Phys. 2002, 117, 7719. Finlayson-Pitts, B. J.; Pitts, Jr., J. N. Chemistry of the Upper and Lower Atmosphere; Academic Press: San Diego, 2000. Dorko, M. J .; Bryden, T. R.; Garrett, S. J. J. Phys. Chem. B 2000, 104, 11695. Jacox, M. E. Chem. Phys. Ben. 1977, 53, 192. Johnson, C. A. F.; Ross, H. J. J. Chem. Soc. Farad. Trans. 1 1978, 74, 2930. Sam, C. L.; Yardley, J. T. Chem. Phys. Lett. 1978, 61 , 509. Wampler, F. B.; Tiee, J. J.; Rice, W. W.; Oldenborg, R. C. J. Chem. Phys. 1979, 71, 3926. Molina, L. T.; Molina, M. J. J. Phys. Chem. 1983, 87, 1306. Krajnovich, D.; Zhang, 2.; Butler, L.; Lee, Y. T. J. Phys. Chem. 1984, 88, 4561. Gosnell, T. R;. Taylor A. J .; Lyman, J. L. J. Chem. Phys. 1991, 94, 5949. Talukdar, R. K.; Vaghijani, G. L.; Ravishankara, A. R. J. Chem. Phys. 1992, 96, 8194. Vatsa, R. K.; Kumar, A.; Naik, P. D.; Rama Rao, K. V. S.; Mittal, J. P Chem. Phys. Lett. 1993, 207, 75. Van Hoeymissen, J.; Uten, W.; Peeters, J. Chem. Phys. Lett. 1994, 226, 159. Vatsa, R. K.; Kumar, A.; Naik, P. D.; Rama Rao, K. V. S.; Mittal, J. P. Bull. Chem. Soc. Japan 1995, 68, 2817. 113 56. Talukdar, R. K.; Hunter, M.; Warren, R. F.; Burkholder, J. B.; Ravishankara, A. R. Chem. Phys. Lett. 1996, 262, 669. 57. Felder, P.; Yang, X.; Baum, G.; Huber, J. R. Isr. J. Chem. 1993, 34, 33. 58. Surjan, P. R.; Csaszar, P.; Poirier, R. A.; van Lenthe, J.-H. Acta Phys. Hungarica 1990, 67, 387. 59. Ishida, H.; Palmer, R. E. Phys. Rev. B 1992, 46, 15484. 60. Ancilotto, F.; Toigo, F. Phys. Rev. B 1993, 47, 13713. 61. Lou, L.; Osterlund, L.; Hellsing, B. J. Chem. Phys. 2000, 112, 4788. 62. Jeloaica, L.; Sidis, V. Chem. Phys. Lett. 1999, 300, 157. 63. Ishikawa, S.; Madjarova, G.; Yamabe, T. J. Phys. Chem. B 2001, 105, 11986. 64. Chalasinski, G.; Szczesniak, M. M. Chem Rev. 1994, 94, 1723. 65. Hobza, P.; Selzle, H. L.; Schlag, E. W. Chem Rev. 1994, 94, 1767. 66. Rassolov, V. A.; Pople, J. A.; Redfem, P. C.; Curtiss, L. A. Chem. Phys. Lett. 2001, 350, 573. 67. Knorr, K. Phys. Rep. 1992, 214, 113. 68. Knorr, K.; Civera-Garcia, E. Surf. Sci. 1990, 232, 203. 69. Nalezinski, R.; Bradshaw, A. M.; Knorr, K. Surf. Sci. 1995, 333, 255. 70. Volkmann, U. G.; Knorr, K. Phys. Rev. B 1993, 47, 4011. 71. Warken, A.; Enderle, M.; Knorr, K. Phys. Rev. B 2000, 61 , 3028. 72. Boys, S. F.; Bemardi, F. Mol. Phys. 1970, 19, 553. 73. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Zakrzewski, V. G.; Montgomery Jr., J. A.; Stratmann, R. E.; Burant, J. C.; Dapprich, S.; Millam, J. M.; Daniels, A. D.; Kudin, K. N.; Strain, M. C.; Farkas, 0.; Tomasi, J.; Barone, V.; Cossi, M.; Cammi, R.; Mennucci, B.; Pomelli, C.; Adamo, C.; Clifford, S.; Ochterski, J.; Petersson, G. A.; Ayala, P. Y.; Cui, Q.; Morokuma, K.; Salvador, P.; Dannenberg, J. J .; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Cioslowski, J.; Ortiz, J. V.; Baboul, A. G.; Stefanov, B. B.; Liu, 0.; Liashenko, A.; 114 74. 75. Piskorz, P.; Komaromi, I.; Gomperts, R.; Martin, R. L.; Fox, D. J .; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, 8.; Chen, W.; Wong, M. W.; Andres, J. L.; Gonzalez, C.; Head-Gordon, M.; Replogle, E. S.; Pople, J. A. Gaussian 98 (Revision A”); Gaussian Inc.: Pittsburgh, PA, 2001. N. M. S. D. Center; S. Stein, In NIST Chemistry WebBook, NIST Standard Reference Database Number 69; Mallard, W., Linstrom, P., Eds.; National Institutes of Standards and Technology: Gaithersburg, MD, 1998 (http://webbook.nist.gov). Hazel], A. G; Larsen, F. K.; Lehmann, M. S. Acta Cryst. 1972, 828, 2977. 76. Jensen, F. Introduction to Computational Chemistry; John Wiley and Sons: 77. 78. Chicester, 1999. Seebauer, E. G.; Allen, C. E. Prog. Surf. Sci. 1995, 49, 265. Lide, D. R., Ed.; CRC Handbook of Chemistry and Physics, 75th ed.; CRC Press: Boca Raton, FL, 1995. 115 Chapter 6 Conclusions and Future Work 6.1 Experimental. In the experimental investigation of the photochemistry of CF2Br2 adsorbed on HOPG it was revealed that direct excitation of the adsorbate occurred to produce photoproducts similar to those observed from the gas-phase molecule. It appeared that the HOPG surface was not directly involved in the photodissociation event however; the surface did provide a medium upon which other products (most notably C2F4Br2 and Br2) were formed. Comparison of the cross sections of CF2Br2(g) at the photolysis wavelengths used in this study to those calculated from TPD and other data showed that dissociation of the adsorbate by attachment of low- energy photoelectrons produced by ultraviolet irradiation of the surface did not occur to any measurable degree. This finding is significant because it is well known that metal surfaces produce bou nd sub-vacuum or free photoelectrons when irradiated with ultraviolet light.1 These electrons can be captured by the anti-bonding orbitals of an adsorbate, and dissociation of the molecule may occur. This process is known as dissociative electron attachment (DEA). It would be expected that a similar DEA-type process might occur for adsorbates on an HOPG surface because it has a band structure similar to that of a metal. Graphite is a Setnimetal and its band structure, as well as those of a typical metal, semiconductor, and insulator is illustrated in Figure 6.1.2 In a semimetal, the valence band overlaps slightly in energy with the conduction band to produce a small number of holes in the valence band and a corresponding number of electrons in the conduction band.2 Theoretically, 116 Energy Figure 6.1 Band structures of an insulator, metal, semimetal, and semiconductor.2 Shaded areas represent filled bands. H7 electrons present in the graphite conduction band, which terminates at the Fermi level, can be excited with ultraviolet radiation and captured by the anti-bonding orbitals of CF2Br2(ad) resulting in dissociation of the molecule. The apparent lack of a DEA-type process could be the result of a change in the surface properties or band structure of the graphite. Results obtained from ab initio calculations performed on this system suggest a small amount of charge transfer occurred from the surface to the molecule upon adsorption. This charge transfer process would create a graphite surface with a net positive charge that would appear as an increase in the surface work function. Experiments performed on bromine intercalated graphite compounds have shown that charge is transferred from graphite to the bromine with a concurrent change in the graphite work function between 0.5 and 1.5 eV.?"4 As a consequence, the number of photoelectrons with sufficient energy to cause dissociation in the range of wavelengths used here would decrease. Therefore, an accurate measurement of the work function of graphite as a function of CF2Br2 coverage should be determined. An increase in work function is evident as an increase in the separation of the adsorbate and substrate bands and so charge-transfer leading to work function changes also affect processes driven by sub-vacuum level electrons. Poor energetic or spatial overlap of the anti-bonding orbitals of CF2Br2 with the 1:. band of graphite could generate a low probability for excitation of the adsorbate by substrate electrons. Alternatively, favorable overlap may produce rapid quenching of any adsorbate excited state formed through intra-adsorbate excitation. The rate of quenching is dependent upon the lifetime of the hole generated by irradiation of the surface.5 If the 118 hole lifetime is long compared to bond breaking in the adsorbate, then dissociation of the molecule will occur. The hole lifetime is determined by the resonance between the adsorbate and substrate orbitals that are involved in the excitation process. Charge transfer can occur between energetically resonant orbitals more easily than between non- resonant orbitals. If a resonant charge transfer process cannot occur between the adsorbate and substrate, then the likelihood of dissociative electron attachment occurring will be drastically reduced. Dissociative electron attachment cross sections for Halons and CFCs are largest for electrons with near zero electron kinetic energy.6 Large numbers of these electrons should be available at all the wavelengths used in this experiment. The relative positions of the graphite valence bands with respect to the occupied orbitals of the adsorbate can be deduced from ultraviolet photoelectron spectroscopy (UPS). Further studies of other Halons, CFCs, and their possible replacements also should be attempted to compare possible photodissociation pathways for this family of molecules while adsorbed on HOPG. Dichlorodifluoromethane (CFC-11) has been extensively studied by Knorr and co-workers and, they have proposed a structure for a monolayer of CFzClz on graphite.7'll Ultraviolet irradiation of this monolayer may or may not cause a similar photodissociation event to occur when compared to that obtained for CF2Br2. Similar results would imply that direct excitation of CFzClz occurred, while the appearance of dissimilar photoproducts could indicate that a DEA-type process may have occurred. Photodissociation studies of other molecules in this family such as CF3Br (Halon 1301), CHzBrz and C2F4Br2 (Halon 2402) should be investigated to determine their photodissociation pathways as well. 119 Although crude measurements of the total photodissociation cross section for the adsorbed and gaseous molecule were found to be comparable, excitation within the substrate should not be dismissed. Additional experiments should be conducted to confirm that direct adsorbate excitation predominates in this system. If the direct excitation mechanism is the prevailing one, then the photodissociation cross section will depend on lp. - E|2 where E is the electric field vector and p. is the transition dipole moment.5 Experiments could be performed as a function of radiation intensity or polarization to measure if an increase in photodissociation occurs. Similar experiments have been used by Zhu et. al. to separate excitation in the substrate, where the E field intensity is determined by the Fresnel equations, from excitation in the adsorbate, exhibiting a simple c0326 dependence on polarization angle 9.12 An additional series of experiments could be performed where the excitation wavelength is varied and the degree of photodissociation or product yield probed. If direct intra-adsorbate excitation occurs, then the wavelength of maximum absorption for a physisorbed molecule should be comparable to that of the gas phase species. The wavelength of maximum absorption for CF2Br2(g) occurs at 226 nm.”"4 This wavelength was available from our excitation source but the lamp fluence in this wavelength region was very low, and no filters were available to isolate this wavelength. An ideal radiation source for these studies would be a continuously tunable source such as a Xe arc lamp and monochromator or a dye laser system. Although graphite has been used as a simple model for a carbon aerosol particle surface, the model can be improved in a number of ways. Hence, future experiments should be performed by utilizing modified graphite surfaces. While carbonaceous 120 IvVL'll'llr aerosols may have small domains that resemble HOPG, the majority of the surface contains a large number of defects.15 One way to introduce defects on the surface is by ion sputtering. Ion sputtering of the graphite surface will disrupt the pure sp2 conjugation present in the unsputtered surface and sites of increased electron density will be created. Previous studies of graphite surfaces sputtered with noble gas ions have shown that both vacancy (removal of surface C atoms) and interstitial defects (trapping of the noble gas within the graphite) were formed.”20 Molecules may preferentially adsorb at these sites, and changes in the photochemistry may occur as a result of the change in the local density of states (LDOS) of HOPG. Experiments where the HOPG surface is sputtered with a reactive gas such as molecular oxygen should be undertaken as well. Ion sputtering studies performed with molecular oxygen revealed defects similar to those obtained with the noble gases, and incorporation of oxygen at defects occurred to produce carbonyls and other oxygen- containing functional groups.”22 Further oxidation of these sputtered surfaces revealed new functional groups, such as quinones and ethers, were present which could alter both the adsorption properties as well as photochemistry occurring on the surface. Alternatively, etch pits of controlled sizes and depths also can be prepared on an HOPG surface by heating the sample in air at 650 °C.23'24 It is expected that aged carbonaceous aerosols will have oxidized defects that will influence any photochemistry that occurs on the surface. Carbonaceous particles become oxidized as they age in the atmosphere and heterogeneous reactions that once occurred on a newly formed particle may no longer take place. By introducing defects in each of these ways, the photochemistry can be 121 observed as a function of the number of defects and type of functionalities present on the surface. Modification of the HOPG surface also can be accomplished by the adsorption of straight chain alkanes and polycyclic aromatic hydrocarbons (PAHs). A large fraction of aged carbonaceous aerosols have been found to contain varying amounts of both polycyclic aromatic hydrocarbons and long-chain alkanes present on their surface.15 Addition of these adsorbates will alter the electronic properties of the surface. Possible photo-oxidation of the PAHs may lead to an enhanced photolysis rate for CF2Br2 or possible reaction between them. Bond cleavage in a PAH will lead to sites of higher electron density (“dangling bonds”) that can be easily donated to the adsorbate. To date, there have been no studies investigating changes in the adsorption or the photochemistry of adsorbates on graphite using hydrocarbon spacer layers. Additional studies using atmospherically important small molecule co-adsorbates on the photochemistry of CF2Br2 adsorbed on HOPG could be performed. One molecule that is present in large atmospheric quantities is water. Adsorption of water ice on HOPG has been studied previously,”27 but its influence on the photochemistry of co-adsorbates has not been examined. It is well known that polar molecules, such as water ice, can act as low energy electron traps.”31 Madey and co-workers reported a large increase in the amount of F" and Cl' ions produced when they electron irradiated water and ammonia ice surfaces covered with CF2C12.32'36 Secondary electrons produced from the irradiation of the surface were shown to increase dissociative electron attachment to CFzClz adsorbed on the condensed H20 and NH3 layers by transfer of the secondary electrons in precursor 7 States to the CF2C12 molecules.3 If sub-vacuum or photoelectrons are produced by 122 ultraviolet irradiation of the HOPG surface, then an increase in the photodissociation of CF2Br2 should be observed. The thickness of the water layer, as well as its crystal structure, also can be varied. Roberts and co-workers have shown different adsorption behaviors for polar H-bonding molecules on amorphous and crystalline ice surfaces.”41 It also has been observed that the degree to which DEA occurs in ice films varies with film thickness, crystallinity, and temperature.30 Lastly, soot samples obtained from a variety of natural sources could be used as the adsorption substrate. Carbonaceous aerosols that are produced from the combustion of different types of fossil fuels will have differing chemical compositions and particle size distributions depending on the combustion conditions.15 These factors may influence the number and type of adsorption sites on the surface and any possible photochemistry that may occur. Newly formed carbonaceous aerosols would be expected to have a larger number of adsorption sites compared to an aged aerosol. However, an aged aerosol may be coated with adsorbates that provide routes for alternative photochemical processes to occur. 6.2 Theoretical. It was shown that the preferred adsorption site for the CF2Br2 molecule occurred over the hollow site of a six-membered ring on each surface independent of the basis set used. The interaction between the molecule and the surface was observed to be dipolar in nature with the largest dipole occurring over the hollow site. Examination of the Mulliken atomic charges showed that (negative) charge was transferred from the surface to the molecule. The amount of charge transferred depended upon the position of the atoms in the adsorbate with respect to the atoms of the substrate. At the hollow site, a larger amount of charge was transferred to the F atom closest to the 123 surface than at the other positions due to its proximity to larger amounts of electron density. It was determined that the interaction between CF2Br and the model HOPG surfaces involved borrowing of electron density from the surface by the singly occupied orbital of the radical. For the four ring substrate, the preferred adsorption site on HOPG was located at the atop position. However, the preferred adsorption site shifted to the bridge position when CF2Br was adsorbed on the seven ring substrate. Both sites contain larger amounts of electron density than at the hollow site. To obtain the most accurate description of both the molecule and the radical on the HOPG surface, complete geometry optimization calculations should be done using larger basis sets and the seven ring substrate to represent the surface. Bigger PAHs could be used to model the graphite surface however, this would result in an increase in the number of basis functions used in the calculations and a large increase in the amount of computational resources as well. Larger basis sets allow for better correlation of the valence electrons in the system and would yield binding energies comparable to those obtained from TPD experiments. Vibrational frequencies obtained from the complete optimization calculations could be compared to both pre- and post-irradiated EELS data to assess the results of the calculations. To further investigate the role of charge transfer in this system, a series of calculations should be performed on the interaction of Br adsorbed on a model HOPG surface. Considerable amounts of charge transfer are expected for this system as evidenced by experimental studies of intercalation compounds involving bromine.3‘4 As a result, single-reference wave function methods are inappropriate for this system.42 124 Initial MP2 calculations performed on Br interacting with both the four and seven ring surfaces using the 6-3lG* and 6-31G(2d) basis sets revealed that the Br atom was not bound to the substrate. Use of multi-reference wave function methods (CASSCF, CASPT2, MRCISD) and larger basis sets will allow for a bound, charge-transfer interaction to occur between Br and the surface. By examining the preferred adsorption site for Br, additional information about the post-irradiated surface can be obtained. Placing two CF2Br2 molecules on a sufficiently large surface slab and performing geometry optimizations would allow for additional information about the structure of the monolayer to be obtained. Experimentally, the growth of the monolayer is thought to occur by two-dimensional islanding of CF2Br2 on the surface due to stronger intermolecular interactions compared to the adsorbate-substrate interaction.““15 The strength of each of these interactions as well as changes in preferred adsorption sites and geometries should be studied. Similar information could be obtained for the post- irradiated surface by performing calculations on CF2Br2 and CF2Br co-adsorbed on a model HOPG surface. Information concerning variations in the adsorption enthalpy with changes in the molecular orientation and coverage would be obtained as well. Molecular dynamics calculations may lend further insight into the structure of the monolayer both before and after irradiation, and would provide information on the processes leading to formation of specific product molecules. In order to study the structure of the monolayer, accurate intermolecular potential energy surfaces need to be developed for various orientations between species using the same methods as described previously. These potentials would be fit to a functional form that will be used in modeling both mono- and multilayer growth of CF2Br2 as well as the structure of the 125 monolayer after each period of irradiation with UV light. The structure of the un- irradiated monolayer obtained from these calculations can be compared to that proposed for CFzClz by Knorr and co-workers.7'll The results obtained from the post-irradiated surface can be used to rationalize the experimental TPD data. 126 10. ll. 12. 13. 14. 15. 16. References . Zhou, X.-L.; Zhu, X.-Y.; White, J. M. Surf Sci. Rep. 1991, 13, 73. Kittel, C. Introduction to Solid State Physics( 7th Edition); John Wiley and Sons, Inc.: New York, 1996. Cox, P. A. The Electronic Structure and Chemistry of Solids; Oxford University Press: Oxford, 1987. Ellis, A. B.; Geselbracht, M. J .; Johnson, B. J .; Lisensky, G. C.; Robinson, W. R. Teaching General Chemistry: A Materials Science Companion; American Chemical Society: Washington, DC, 1993. Zhou, X.-L.; White, J. M. In Laser Spectroscopy and Photochemistry on Metal Surfaces, Part II; Dai, H.-L.; Ho, W., Eds.; World Scientific: New Jersey, 1995; Advanced Series in Physical Chemistry, Vol. 5. Underwood-Lemons, T.; Gergel, T. J.; Moore, J. H. J. Chem. Phys. 1995, 102, 119. Knorr, K. Phys. Rep. 1992, 214, 113. Knorr, K.; Civera-Garcia, E. Surf Sci. 1990, 232, 203. Nalezinski, R;. Bradshaw, A.M.; Knorr, K. Surf Sci. 1995, 333, 255. Volkmann, U. G.; Knorr, K. Phys. Rev. B 1993, 47, 4011. Warken, A.; Enderle, M.; Knorr, K. Phys. Rev. B 2000, 61 , 3028. Zhu, X.-Y.; White, J. M.; Wolf, M.; Hasselbrink, E.; Ertl, G. Chem. Phys. Lett. 1991, 176, 459. Orkin, V. L.; Kasimovskaya, E. E. J. Atmos. Chem. 1995, 21 , l. Burkholder, J. B.; Wilson, R. R.; Gierczak, T.; Talukdar, R.; McKeen, S. A.; Orlando, J. J.; Vaghjiani, G. L.; Ravishankara, A. R. J. Geophys. Res. 1991, 96, 5025. Seinfeld, J. H.; Pandis, S. N. Atmospheric Chemistry and Physics; John Wiley and Sons: New York, 1998. Li, T.; King, B. V.; MacDonald, R. J .; Cotterill, G. F.; O’Conner, D. J .; Yang, Q. Surf Sci. 1994, 312, 399. 127 17. An, B.; Fukuyama, S.; Yokogawa, K.; Yoshimura, M. J. Vac. Sci. Technol. B 1999, 1 7, 2439. 18. Kappel, M.; Ktippers, J. Surf Sci. 1999, 440, 387. 19. Hahn, J. R.; Kang, H.; Lee, S. M.; Lee, Y. H. J. Phys. Chem. B 1999, 103, 9944. 20. Hahn, J. R.; Kang, H. Surf Sci. 2000, 446, L77. 21. Nowakowski, M. J.; Vohs, J. M.; Bonnell, D. A. Surf Sci. 1992, 271, L351. 22. Nowakowski, M. J .; Vohs, J. M.; Bonnell, D. A. J. Am. Ceram. Soc. 1993, 76, 279. 23. Chang, H.; Bard, A. J. J. Am. Chem. Soc. 1990, 112, 4598. 24. Stevens, F.; Kolodny, L. A.; Beebe, Jr., T. P. J. Phys. Chem. B 1998, 102, 10799. 25. Chakarov, D. V.; Osterlund, L.; Kasemo, B. Langmuir 1995, 11, 1201. 26. Chakarov, D. V.; Osterlund, L.; Kasemo, B. Vacuum 1995, 46, 1109. 27. Chakarov, D. V.; Kasemo, B. Phys. Rev. Lett. 1998, 81, 5181. 28. Gilton, T. L.; Dehnbostel, C. P.; Cowin, J. P. J. Chem. Phys. 1989, 91, 1937. 29. Bass, A. D.; Sanche, L. J. Chem. Phys. 1991, 95, 2910. 30. Simpson, W. C.; Orlando, T. M.; Parenteau, L.; Nagesha, K.; Sanche, L. J. Chem. Phys. 1998, 108, 5027. 31. Lu, Q.-B.; Sanche, L. J. Chem. Phys. 2001, 115, 5711. 32. Lu, Q.-B.; Madey, T. E. J. Chem. Phys. 1999, 111, 2861. 33. Lu, Q.-B.; Madey, T. E. Phys. Rev. Lett. 1999, 82, 4122. 34. Lu, Q.-B.; Madey, T. E. Surf Sci. 2000, 451, 238. 35. Lu, Q.-B.; Madey, T. E. J. Phys. Chem. B 2001, 105, 2779. 36. Lu, Q.-B.; Madey, T. E.; Parenteau, L.; Weik, F.; Sanche, L. Chem. Phys. Lett. 2001, 342, l. 37. Lu, Q.-B.; Sanche, L. Phys. Rev. B 2001, 63, 153403. 128 38. Schaff, J. E.; Roberts, J. T. J. Phys. Chem. 1994, 98, 6900. 39. Graham, J. D.; Roberts, J. T. Geophys. Res. Lett. 1995, 22, 251. 40. Schaff, J. E.; Roberts, J. T. Langmuir 1999, 15, 7232. 41. Schaff, J. E.; Roberts, J. T. Surf Sci. 1999, 426, 384. 42. Vila, F. D.; Jordan, K. D. J. Phys. Chem. A 2002, 106, 1391. 43. Golze, M; Grunze, M.; Hirschwald, W Vacuum 1981, 31 , 697. 44. Niemantsverdriet, J. W.; Markert, K.; Wandelt, K. Appl. Surf Sci. 1988, 31 , 211. 45. Zhou, X. —L.; White, J. M.; Koel, B. E. Surf. Sci. 1989, 218, 201. 129 Appendix 130 Appendix A Calculation of CF2Br2 Fractional Coverage on HOPG The fractional coverage of CF2Br2 was calculated as the number of adsorbate molecules per surface C atom. The fraction of surface-adsorbed CF2Br2 was calculated —a (I) 1_ex _i LAX§A= ABS[ {’1‘ C089]] 13 SA —a 1— +ex A where ass is the effective “size” or height of the adsorbate (CF2Br2) and is determined as follows: (A.1) by dividing the molecular mass of the adsOrbate by its density multiplied by Avagadro’s number and then taking the cube root, AA is the inelastic mean free path of photoelectrons ejected from atom A of the adsorbate, AB is the inelastic mean free path of photoelectrons ejected from atom B of the adsorbate, (DABS is the fractional coverage of the adsorbate, IA is the area as measured from XPS data of peak A, 13 is the area as measured from XPS data of peak B, SA and 8;; are atomic sensitivity factors for atoms A and B, and 9 is the angle between the surface normal and the ejected photoelectrons. The following values were used in calculating the fractional coverage given atom B is carbon and atom A is bromine: aABs = 5.33 A, ltc = 14 A, 21.3, = 15 A, 0 = 65°, 13, = 63044, [C = 109456, 83, = 0.83, and Sc = 0.25. A value of 0.24877 is obtained that must be divided by 2 (there are two bromine atoms in CF2Br2) to obtain a fractional coverage of 0.124. A similar calculation for the fractional coverage using the fluorine (atom A) and carbon atoms 131 (atom B) with the following values (aABs = 5.33 A, Ac 2 14 A, A}: = 11 A, 0 = 65°, Ic = 109456, 1;: = 116761, Sc = 0.25, 8;: = 1.00) yielded a value of 0.317. Dividing by 2 (two F atoms in CF2Br2) gives 0.159. An average was taken for the two values obtained from the calculations and gave a fractional coverage of 0.14. 132 IIIIIIIIIIIIIIIIIIIIIIIIIIIIII WillllWWW.l‘l‘llllillllllllll'l1| 3 1293 02493 8551 I